You are on page 1of 983

Seismic Performance of Slender Reinforced Concrete Structural Walls

Anna C Birely

A dissertation submitted in partial fulfillment of


the requirements for the degree of

Doctor of Philosophy

University of Washington

2012

Reading Committee:

Laura N. Lowes, Chair

Dawn E. Lehman, Chair

Marc O. Eberhard

Gregory R. Miller

Program Authorized to Offer Degree:


Department of Civil and Environmental Engineering
University of Washington

Abstract

Seismic Performance of Slender Reinforced Concrete Structural Walls

Anna C Birely

Co-Chairs of the Supervisory Committee:


Associate Professor Laura N. Lowes
Civil and Environmental Engineering
Associate Professor Dawn E. Lehman
Civil and Environmental Engineering

Reinforced concrete structural walls are one of the most common lateral-load resisting sys-
tems found in mid-rise buildings. They are stiff and strong, easily incorporated into ar-
chitectural layouts, and, when well designed and detailed, generally considered to perform
well under earthquake loading. However, damage to mid-rise walled buildings in the 2010
Chilean earthquake has reminded the engineering community that structural walls can sus-
tain serious damage and that consequently there is a need to improve understanding of
wall performance. The research presented here seeks to address this need through exper-
imental testing of slender planar walls and evaluation of performance-assessment tools for
performance-based earthquake engineering.
Despite the engineering community’s reliance on reinforced concrete structural walls,
relatively few experimental tests have been done to investigate the seismic performance of
modern, code-compliant walls. Those tests that have been conducted provide a limited
amount of data to support development of performance-based seismic design tools. To
address this lack of data, a large experimental test program was undertaken by researchers
at the Universities of Washington, Illinois, and California. As a portion of this program, four
large-scale planar (rectangular) walls representative of mid-rise West Coast construction
were tested and a large number of data were collected. Data analysis, presented here,
was done to provide improved understanding of earthquake response and performance of
rectangular walls and support the validation of numerical models.
Collected experimental data included detailed damage data. These data, along with
documented damage from previous experimental tests, were used to develop performance-
prediction tools. These tools, known as fragility functions, relate engineering demand pa-
rameters such as strain, rotation, or drift, to the likelihood of specific damage occurring.
Damage sustained by buildings during the 2010 Chile earthquake provided a unique
opportunity to evaluate performance-based design tools. Several mid-rise buildings that
sustained damage in the earthquake were studied, with a focus on evaluating the fragility
functions developed from experimental data and evaluating the ASCE/SEI 31/41 standards
for the seismic evaluation of existing structures. Evaluation of the ASCE standards involved
the use of both linear and nonlinear models. Results of the building evaluations indicate
aspects of the procedures that require improvements.
TABLE OF CONTENTS

Page

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xxxvii

Chapter 1: Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 The Earthquake Response and Design of Reinforced Concrete Walls . . . . . 3
1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
1.3 Outline of Document . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

Chapter 2: Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7


2.1 Code-Based Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.1.1 Design Loads and Strength Reduction Factors . . . . . . . . . . . . . . 8
2.1.2 Material properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.1.3 Shear Strength and Distributed Web Reinforcement . . . . . . . . . . 9
2.1.4 Boundary Elements . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.1.5 Splices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.1.6 Considerations for Bi-Directionally Loaded Walls . . . . . . . . . . . . 12
2.2 Performance-Based Seismic Design of Walls . . . . . . . . . . . . . . . . . . . 12
2.2.1 IBC Seismic Design Recommendations . . . . . . . . . . . . . . . . . . 13
2.2.2 Fragility Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.3 Building Inventory Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4 Wall Damage Observed Following Earthquakes . . . . . . . . . . . . . . . . . 19
2.4.1 Earthquakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.4.2 Types of Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.4.3 Building Damage Summary . . . . . . . . . . . . . . . . . . . . . . . . 61
2.5 Experimental Research on Performance of Walls . . . . . . . . . . . . . . . . 66
2.5.1 Portland Cement Association . . . . . . . . . . . . . . . . . . . . . . . 68
2.5.2 Shiu et al . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
2.5.3 Morgan et al . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77

i
2.5.4 Lefas, Kotsovos, and Ambraseys . . . . . . . . . . . . . . . . . . . . . 80
2.5.5 Lefas and Kotsovos . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 82
2.5.6 Ali and Wight . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
2.5.7 Pilakoutas and Elnashi . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
2.5.8 Thomsen and Wallace . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
2.5.9 Tupper . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
2.5.10 Tasnimi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
2.5.11 Mobeen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96
2.5.12 Riva, Meda, and Giuriani . . . . . . . . . . . . . . . . . . . . . . . . . 99
2.5.13 Liu . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
2.5.14 Khaliil and Ghobarah . . . . . . . . . . . . . . . . . . . . . . . . . . . 104
2.5.15 Ile and Reynouard . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
2.5.16 Beyer, Dazio and Priestley . . . . . . . . . . . . . . . . . . . . . . . . 107
2.5.17 Brueggen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110
2.5.18 Dazio et al . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
2.5.19 Other Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 116
2.5.20 Failure modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
2.6 Summary of Experimental Research . . . . . . . . . . . . . . . . . . . . . . . 124
2.6.1 Evaluation of Code Compliance . . . . . . . . . . . . . . . . . . . . . . 125
2.6.2 Comparison to Building Inventory . . . . . . . . . . . . . . . . . . . . 126
2.6.3 Evaluation of Failure Mechanism . . . . . . . . . . . . . . . . . . . . . 132
2.6.4 Comparison of Laboratory and Field Damage . . . . . . . . . . . . . . 136
2.6.5 Impact of Design Parameters . . . . . . . . . . . . . . . . . . . . . . . 137
2.7 Summary of Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . 158

Chapter 3: NEESR Complex Wall Project: Planar Wall Test Program . . . . . . . 160
3.1 Test Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 160
3.1.1 Comparison of Design Parameters to Building Inventory and Other
Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
3.1.2 Dimensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 162
3.1.3 Lateral Load Distribution . . . . . . . . . . . . . . . . . . . . . . . . . 165
3.1.4 Reinforcement Design . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
3.1.5 Splice . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
3.2 Material Properties and Construction Details . . . . . . . . . . . . . . . . . . 171
3.2.1 Concrete Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

ii
3.2.2 Reinforcing Bar Properties . . . . . . . . . . . . . . . . . . . . . . . . 173
3.3 Expected Strengths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
3.3.1 Shear Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
3.3.2 Moment Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
3.4 Specimen Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 179
3.5 Experimental Test Setup . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 180
3.6 Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
3.6.1 Non-Contact Instrumentation . . . . . . . . . . . . . . . . . . . . . . . 184
3.6.2 External Instrumentation . . . . . . . . . . . . . . . . . . . . . . . . . 185
3.6.3 Steel Strain Gauges . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
3.7 Loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190
3.7.1 Displacement History . . . . . . . . . . . . . . . . . . . . . . . . . . . 192
3.7.2 Axial Load . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
3.7.3 Lateral and Overturning Loads . . . . . . . . . . . . . . . . . . . . . . 195
3.8 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 196

Chapter 4: Summary of Experimental Results . . . . . . . . . . . . . . . . . . . . 199


4.1 Specimen PW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
4.1.1 General Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 200
4.1.2 Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
4.2 Specimen PW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
4.2.1 General Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 210
4.2.2 Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 211
4.3 Specimen PW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
4.3.1 General Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 217
4.3.2 Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
4.4 Specimen PW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.4.1 General Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
4.4.2 Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 225
4.5 Comparison of Walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
4.5.1 General Response . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
4.5.2 Damage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
4.6 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247

iii
Chapter 5: Methods for Analysis of Experimental Data . . . . . . . . . . . . . . . 249
5.1 Displacement and Rotation Calculations . . . . . . . . . . . . . . . . . . . . . 249
5.1.1 Displacements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 250
5.1.2 Rotations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 252
5.2 Calculation of Effective Stiffness Values . . . . . . . . . . . . . . . . . . . . . 253
5.2.1 Axial Stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
5.2.2 Shear and Flexural Stiffness: Timoshenko Beam Theory . . . . . . . . 255
5.2.3 Flexural Stiffness: Euler-Bernoulli Beam Theory . . . . . . . . . . . . 258
5.3 Drift Estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
5.4 Yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
5.4.1 Stress Calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 260
5.4.2 Definitions of Yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 262
5.4.3 Yield Maps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 264
5.5 Strain Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
5.5.1 Calculation Using Finite Element Method . . . . . . . . . . . . . . . . 267
5.5.2 Calculation Using Geometry . . . . . . . . . . . . . . . . . . . . . . . 268
5.6 Deformation Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
5.6.1 Flexural Deformations . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
5.6.2 Shear Deformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 271
5.6.3 Fixed-end Deformations . . . . . . . . . . . . . . . . . . . . . . . . . . 273
5.6.4 Total Drift Contributions . . . . . . . . . . . . . . . . . . . . . . . . . 273
5.7 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 275

Chapter 6: Analysis of Experimental Data . . . . . . . . . . . . . . . . . . . . . . 276


6.1 Summary of Experimental Observations . . . . . . . . . . . . . . . . . . . . . 276
6.2 Displacement and Rotation Profiles . . . . . . . . . . . . . . . . . . . . . . . . 278
6.3 Stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
6.3.1 Axial Stiffness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 282
6.3.2 Flexural and Shear Stiffness . . . . . . . . . . . . . . . . . . . . . . . . 284
6.3.3 Secant Stiffness at Effective Height . . . . . . . . . . . . . . . . . . . . 287
6.4 Deformation Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 287
6.5 Yield . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 293
6.5.1 Yield of Longitudinal Reinforcement . . . . . . . . . . . . . . . . . . . 293
6.5.2 Yield of Horizontal Reinforcement . . . . . . . . . . . . . . . . . . . . 301
6.6 Strain Fields . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302

iv
6.6.1 Vertical Strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 304
6.6.2 Shear Strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 308
6.6.3 Second Principal Strain . . . . . . . . . . . . . . . . . . . . . . . . . . 313
6.7 Strain Demands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
6.7.1 Nominal Strains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
6.7.2 Failure Strains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
6.8 Stress Demands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
6.9 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322

Chapter 7: Fragility Functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 326


7.1 Experimental Data Used to Develop Fragilities . . . . . . . . . . . . . . . . . 326
7.1.1 Assembled Test Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
7.1.2 Design Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
7.2 Demand Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
7.2.1 Drift . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
7.2.2 Rotation Demand . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
7.2.3 Drift at Effective Height . . . . . . . . . . . . . . . . . . . . . . . . . . 339
7.3 Damage States and Methods of Repair . . . . . . . . . . . . . . . . . . . . . . 341
7.3.1 MOR 1: Cosmetic Repair . . . . . . . . . . . . . . . . . . . . . . . . . 342
7.3.2 MOR 2: Epoxy Injection and Patching . . . . . . . . . . . . . . . . . . 345
7.3.3 MOR 3: Replace Concrete . . . . . . . . . . . . . . . . . . . . . . . . . 347
7.3.4 MOR 4: Replace Steel and Concrete . . . . . . . . . . . . . . . . . . . 348
7.4 Development of Fragility Functions . . . . . . . . . . . . . . . . . . . . . . . . 351
7.4.1 Analysis of the Full Dataset . . . . . . . . . . . . . . . . . . . . . . . . 353
7.4.2 Removal of Outliers . . . . . . . . . . . . . . . . . . . . . . . . . . . . 353
7.4.3 Reduction of Dataset . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
7.4.4 Impact of Design Parameters on Damage Progression . . . . . . . . . 357
7.4.5 Fragility Function Theory . . . . . . . . . . . . . . . . . . . . . . . . . 362
7.4.6 Fragility Functions for Slender Walls . . . . . . . . . . . . . . . . . . . 363
7.5 Recommendations and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . 364

Chapter 8: Evaluation of Buildings Damaged in 2010 Maule Earthquake: Overview367


8.1 Overview of Standards Used . . . . . . . . . . . . . . . . . . . . . . . . . . . . 368
8.1.1 Evaluation Requirements . . . . . . . . . . . . . . . . . . . . . . . . . 370
8.1.2 Tier 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
8.1.3 Tier 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 373

v
8.1.4 Tier 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
8.2 Overview of Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 380
8.2.1 Tier 2 Evaluation Models . . . . . . . . . . . . . . . . . . . . . . . . . 380
8.2.2 Tier 3 Evaluation Models . . . . . . . . . . . . . . . . . . . . . . . . . 389
8.3 Overview of Buildings Studied . . . . . . . . . . . . . . . . . . . . . . . . . . 397
8.3.1 Plaza del Rio (Buildings A & B) . . . . . . . . . . . . . . . . . . . . . 399
8.3.2 Centro Mayor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
8.3.3 Concepto Urbano . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 408

Chapter 9: Evaluation of Building Damaged in 2010 Maule Earthquake: Results . 412


9.1 Summary of Tier 1 Evaluations . . . . . . . . . . . . . . . . . . . . . . . . . . 413
9.1.1 Additional Qualitative Analysis . . . . . . . . . . . . . . . . . . . . . . 414
9.2 Summary of Tier 2 Evaluations . . . . . . . . . . . . . . . . . . . . . . . . . . 415
9.2.1 Evaluation of Unfactored DCR Values . . . . . . . . . . . . . . . . . . 416
9.2.2 Evaluation of Factored DCR Values . . . . . . . . . . . . . . . . . . . 418
9.2.3 Impact of Building Period . . . . . . . . . . . . . . . . . . . . . . . . . 424
9.2.4 Impact of Coupling Effects . . . . . . . . . . . . . . . . . . . . . . . . 426
9.3 Summary of Tier 3 Evaluations . . . . . . . . . . . . . . . . . . . . . . . . . . 427
9.3.1 General Building Response . . . . . . . . . . . . . . . . . . . . . . . . 428
9.3.2 Evaluation of Acceptance Criteria . . . . . . . . . . . . . . . . . . . . 432
9.3.3 Damage Progression . . . . . . . . . . . . . . . . . . . . . . . . . . . . 455
9.3.4 Impact of Axial Load . . . . . . . . . . . . . . . . . . . . . . . . . . . 456
9.4 Recommendations for Changes to ASCE 31/41 Evaluations . . . . . . . . . . 457
9.4.1 Comments on Usability of ASCE 31/41 Provisions . . . . . . . . . . . 460
9.4.2 Local Vertical Discontinuities . . . . . . . . . . . . . . . . . . . . . . . 461
9.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 470

Chapter 10: Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473


10.1 Conclusions from Literature Review . . . . . . . . . . . . . . . . . . . . . . . 474
10.2 Conclusions for Experimental Test Program . . . . . . . . . . . . . . . . . . . 475
10.3 Conclusions from Development of Fragility Functions . . . . . . . . . . . . . . 476
10.4 Conclusions from Study of Damaged Chilean Buildings . . . . . . . . . . . . . 477

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 480

List of Symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 490

vi
Appendix A: Summary of Experimental Damage . . . . . . . . . . . . . . . . . . . . 493
A.1 PW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 493
A.2 PW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 509
A.3 PW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 527
A.4 PW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 545

Appendix B: Additional Data Analysis Results . . . . . . . . . . . . . . . . . . . . . 567

Appendix C: Experimental Data Supporting Fragility Function Development . . . . 594


C.1 Database Properties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 596
C.2 Damage States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 608
C.3 Damage Data for Individual Specimens . . . . . . . . . . . . . . . . . . . . . 609
C.3.1 PilaSW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 610
C.3.2 PilaSW5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 615
C.3.3 PilaSW6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 618
C.3.4 PilaSW7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 621
C.3.5 PilaSW8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
C.3.6 PilaSW9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 627
C.3.7 TasSHW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
C.3.8 TasSHW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 631
C.3.9 TasSHW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 632
C.3.10 TasSHW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
C.3.11 ThomRW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634
C.3.12 ThomRW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 636
C.3.13 ThomTW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
C.3.14 ThomTW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 640
C.3.15 DazioWSH1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 642
C.3.16 DazioWSH2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 644
C.3.17 DazioWSH3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 646
C.3.18 DazioWSH4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 648
C.3.19 DazioWSH5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 650
C.3.20 DazioWSH6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 652
C.3.21 LefasSW21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 654
C.3.22 LefasSW22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 656
C.3.23 LefasSW23 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 657
C.3.24 LefasSW24 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 658

vii
C.3.25 LefasSW25 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 660
C.3.26 LefasSW26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 661
C.3.27 LefasSW30 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 663
C.3.28 LefasSW31 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 665
C.3.29 LefasSW32 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 667
C.3.30 LefasSW33 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 669
C.3.31 OestR1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 671
C.3.32 OestR2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
C.3.33 OestB1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675
C.3.34 OestB2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 677
C.3.35 OestB3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 679
C.3.36 OestB4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 681
C.3.37 OestB5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683
C.3.38 OestB6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
C.3.39 OestB7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
C.3.40 OestB8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689
C.3.41 OestB9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
C.3.42 OestB10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 693
C.3.43 OestF1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695
C.3.44 OestF2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
C.3.45 Morgan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 699
C.3.46 LiuW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
C.3.47 LiuW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 708
C.3.48 AliW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
C.3.49 TupperW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 716
C.3.50 MobeenW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 719
C.3.51 RivaW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 721
C.3.52 ShiuC1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723
C.3.53 KhaC1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 726
C.3.54 ElnCW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 727
C.3.55 ElnCW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 728
C.3.56 IleX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 729
C.3.57 IleY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 730
C.3.58 IleXY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 731
C.3.59 TUA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733

viii
C.3.60 TUB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 734
C.3.61 NTW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 735
C.3.62 NTW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 738
C.3.63 PW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 741
C.3.64 PW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742
C.3.65 PW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 743
C.3.66 PW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 744

Appendix D: Building Summaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . 745


D.1 Plaza del Rio (Buildings A & B) . . . . . . . . . . . . . . . . . . . . . . . . . 746
D.2 Centro Mayor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 768
D.3 Concepto Urbano . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 774

Appendix E: Tier 1 Evaluation of Chilean Buildings . . . . . . . . . . . . . . . . . . 782


E.1 Comprehensive List of Deficiencies . . . . . . . . . . . . . . . . . . . . . . . . 782

Appendix F: Tier 2 Evaluation of Chilean Building . . . . . . . . . . . . . . . . . . 785


F.1 Summary of Adequate Walls . . . . . . . . . . . . . . . . . . . . . . . . . . . 785
F.1.1 Plaza del Rio Building A (PR-A) . . . . . . . . . . . . . . . . . . . . . 785
F.1.2 Plaza del Rio Building B (PR-B) . . . . . . . . . . . . . . . . . . . . . 791
F.1.3 Centro Mayor (CM) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 797
F.2 Demand Capacity Ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 803
F.3 Axial Demands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 814

Appendix G: Tier 3 Evaluation of Chilean Building . . . . . . . . . . . . . . . . . . 820


G.1 Concepcion (Unscaled) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 821
G.2 Concepcion (DBE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 828
G.3 Concepcion (MCE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 835
G.4 San Pedro de la Paz (Unscaled) . . . . . . . . . . . . . . . . . . . . . . . . . . 842
G.5 San Pedro de la Paz (DBE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 849
G.6 San Pedro de la Paz (MCE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 856
G.7 Llolleo (DBE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 863
G.8 Llolleo (MCE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 870
G.9 Matanzas (DBE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 877
G.10 Matanzas (MCE) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 884
G.11 Valparaiso Almendral (DBE) . . . . . . . . . . . . . . . . . . . . . . . . . . . 891

ix
G.12 Valparaiso Almendral (MCE) . . . . . . . . . . . . . . . . . . . . . . . . . . . 898
G.13 Vina del Mar - Centro (DBE) . . . . . . . . . . . . . . . . . . . . . . . . . . . 905
G.14 Vina del Mar - Centro (MCE) . . . . . . . . . . . . . . . . . . . . . . . . . . . 912
G.15 Vina del Mar - El Salto (DBE) . . . . . . . . . . . . . . . . . . . . . . . . . . 919
G.16 Vina del Mar - El Salto (MCE) . . . . . . . . . . . . . . . . . . . . . . . . . . 926

x
LIST OF FIGURES

Figure Number Page


2.1 Fragility functions developed by Gulec et al. [40, 41] for rectangular squat
walls. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
2.2 Fragility functions developed by Brown [12] for reinforced concrete walls. . . . 18
2.3 Indian Hills Medical Center plan. Building was damaged in both the 1971
San Fernando and 1994 Northridge earthquakes. . . . . . . . . . . . . . . . . 23
2.4 Typical detail of Chilean wall [91]. . . . . . . . . . . . . . . . . . . . . . . . . 24
2.5 Examples of diagonal cracks in structural walls. . . . . . . . . . . . . . . . . . 30
2.6 Examples of diagonal cracks in structural walls. . . . . . . . . . . . . . . . . . 32
2.7 Horizontal and diagonal cracks in wall of 13-story building in Concepcion
following the 2010 Chile earthquake [69] . . . . . . . . . . . . . . . . . . . . . 32
2.8 Crack map of Mt. McKinley core walls following the 1964 earthquake [74] . . 33
2.9 Indian Hills Medical Center following the 1971 San Fernando earthquake [57]. 33
2.10 Fractured longitudinal reinforcement in 16-story Indian Hills Medical Center
following the 1971 San Fernando earthquake [74]. . . . . . . . . . . . . . . . . 34
2.11 Fractured longitudinal reinforcement in 16-story Santiago building following
the 2010 Chile earthquake [69]. . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.12 Boundary region cover spalling and minor bar buckling . . . . . . . . . . . . . 36
2.13 Boundary region core crushing and bar buckling . . . . . . . . . . . . . . . . 37
2.14 Boundary region core crushing and bar buckling following the 2010 Chile
earthquake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.15 Shear failure of walls following the 1960 Chile earthquake. . . . . . . . . . . . 39
2.16 Shear failure of 3-story Paloma building in Sendai, Japan following the 1978
Japan earthquake. Structural walls were only located on one end of the
building [31]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
2.17 Shear failure of walls in Kobe, Japan, 1995. . . . . . . . . . . . . . . . . . . . 40
2.18 Shear failure of wall in an unknown building following the 1999 Turkey earth-
quake [74]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.19 Shear failure in a building under construction following the 1999 Chi-Chi
earthquake [60]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41
2.20 Shear failure of walls following the 2010 Chile earthquake [69]. . . . . . . . . 42

xi
2.21 Web crushing/uniform failure plane in walls following the 1960 Chile earth-
quake [74]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 42
2.22 Web crushing in structural walls . . . . . . . . . . . . . . . . . . . . . . . . . 43
2.23 South face of the 1200 L Street building following the 1964 Alaska earthquake
[74]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
2.24 Second story of West Anchorage High School following the 1964 Alaska earth-
quake. Note the lack of confining reinforcement in the boundary regions [10]. 45
2.25 15-story building in Oakland, CA following the 1989 Loma Prieta earthquake
[33]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.26 Horizontal failure plane in walls following 2010 Chile earthquake [69]. . . . . 47
2.27 Construction joints in the Mt. McKinley and 1200 L Street buildings (1964
Alaska). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
2.28 Damage at construction joints following the 1964 Alaska earthquake. . . . . . 50
2.29 Holy Cross Hospital construction joint damage following the 1971 San Fer-
nando earthquake . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
2.30 Damage to construction joints of exterior walls of the Indian Hills Medical
Center following the 1971 San Fernando earthquake [57] . . . . . . . . . . . . 53
2.31 Nethercutt construction joint damage following the 1971 San Fernando earth-
quake [57] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.32 Crushing and bar buckling at construction joint of core-wall in 16-story lift
slab building following the 1988 Armenia earthquake [118]. . . . . . . . . . . 54
2.33 Maryknoll retirement home in Mountain View, CA following the 1989 Loma
Prieta earthquake. Horizontal cracks at construction joints can be seen in
multiple locations [33]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
2.34 Damage to the Indian Hills Medical Center in the Northridge earthquake was
located at the construction joint of the elevator tower [74]. . . . . . . . . . . . 55
2.35 Construction joint damage following the 1995 Kobe earthquake [66]. . . . . . 55
2.36 Web crushing at construction joint following the 1999 Turkey earthquake [94]. 56
2.37 Damage at construction joint following the 2001 Nisqually earthquake [30]. . 56
2.38 Spalling exposing splice at 4th story level of the Indian Hills Medical Center
following the 1994 Northridge earthquake [74]. . . . . . . . . . . . . . . . . . 58
2.39 Spalling at 1st floor in 16-story Santiago building following the 2010 Chile
earthquake [69]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.40 Damaged core concrete at lap splice locations following the 2010 Chile earth-
quake [46]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59
2.41 Buildings with structural walls that collapsed. . . . . . . . . . . . . . . . . . . 60
2.42 Percent of buildings in which a particular type of damage was observed, by
region. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

xii
2.43 Primary damage type for buildings of known height (All regions) . . . . . . . 64
2.44 Primary damage type for buildings of known height (Asia) . . . . . . . . . . . 64
2.45 Primary damage type for buildings of known height (United States) . . . . . 65
2.46 Primary damage type for buildings of known height (Chile) . . . . . . . . . . 65
2.47 Cross-sections of Oesterle Phase I wall specimens [80]. . . . . . . . . . . . . . 69
2.48 Test set-up for walls tested by Oesterle et al. [80]. . . . . . . . . . . . . . . . 70
2.49 Cross-sections of Oesterle et al. [80] Phase II wall. . . . . . . . . . . . . . . . 72
2.50 Test setup axial load modification for Phase II tests by Oesterle et al. [78]. . 73
2.51 Typical crack patterns in Oesterle et al. walls [80]. . . . . . . . . . . . . . . . 74
2.52 Shiu [100] solid wall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
2.53 Cross-section of wall tested by Morgan et al. [72]. . . . . . . . . . . . . . . . . 78
2.54 Crack map after testing by Morgan et al. [72]. . . . . . . . . . . . . . . . . . 79
2.55 Monotonic wall tests by Lefas et al. [56]. . . . . . . . . . . . . . . . . . . . . . 80
2.56 Typical damage during monotonic wall tests by Lefas et al. [56]. . . . . . . . 82
2.57 Damage to cyclic wall tests by Lefas et al. [55]. . . . . . . . . . . . . . . . . . 83
2.58 Ali [3] specimen W1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
2.59 Reinforcement layouts of Pilakoutas and Elnashai [88] wall tests. . . . . . . . 86
2.60 Pilakoutas and Elnashai [88] test setup. . . . . . . . . . . . . . . . . . . . . . 87
2.61 Pilakoutas and Elnashai [88] wall test damage at failure. . . . . . . . . . . . . 88
2.62 Rectangular specimens tested by Thomsen and Wallace [104]. . . . . . . . . . 89
2.63 T-wall specimens tested by Thomsen and Wallace [104]. . . . . . . . . . . . . 90
2.64 Experimental test setup of Thomsen and Wallace walls [103]. . . . . . . . . . 92
2.65 Damage at failure of walls tested by Thomsen and Wallace [103]. . . . . . . . 92
2.66 Cross-section of conventionally reinforced wall tested by Tupper [106]. . . . . 93
2.67 Failure of Tupper conventionally reinforced wall [106]. . . . . . . . . . . . . . 94
2.68 Wall tests by Tasnimi [102]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
2.69 Typical damage at base of Tasnimi wall specimens [102]. . . . . . . . . . . . . 97
2.70 Conventional wall specimen tested by Mobeen [68]. . . . . . . . . . . . . . . . 97
2.71 Damage to wall tested by Mobeen [68]. . . . . . . . . . . . . . . . . . . . . . . 99
2.72 Wall tested by Riva et al. [92]: (a) wall supports, (b) ribs simulating ground
and basement floor diaphragms, (c) post tensioned strands and bars, (d) steel
frame to avoid lateral instability, and (e) additional frames for improving
safety in the test set-up. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
2.73 Load history for wall tested by Riva et al. [92] . . . . . . . . . . . . . . . . . 101
2.74 Damage to large scale wall tested by Riva et al. [92] . . . . . . . . . . . . . . 101

xiii
2.75 Cross-section of walls tested by Liu [59]. . . . . . . . . . . . . . . . . . . . . . 102
2.76 Damage to walls tested by Liu [59]. . . . . . . . . . . . . . . . . . . . . . . . . 103
2.77 Cross-section of wall tested by Khalil and Ghobarah[53]. . . . . . . . . . . . . 104
2.78 Walls tested by Khalil and Ghobarah [53]. . . . . . . . . . . . . . . . . . . . . 105
2.79 Cross section of walls tested by Ile and Reynouard [49]. . . . . . . . . . . . . 106
2.80 Flanged after bi-directional loading of U-shape wall tested by Ile and Rey-
nouard [49]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
2.81 Cross-section of walls tested by Beyer et al[11]. . . . . . . . . . . . . . . . . . 108
2.82 Beyer et al. wall tests [11]. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
2.83 Cross-sections of walls tested by Brueggen [15]. . . . . . . . . . . . . . . . . . 111
2.84 Displacement history of walls tested by Brueggen [15]. . . . . . . . . . . . . . 112
2.85 First floor of Brueggen wall tests following strong-axis bending failure [15]. . 113
2.86 Cross-sections of walls tested by Dazio et al. [25] . . . . . . . . . . . . . . . . 114
2.87 Typical stress-strain response of steel used in walls tested by Dazio et al. [25] 115
2.88 Final damage to walls tested by Dazio et al. [25]. . . . . . . . . . . . . . . . . 117
2.89 Cross-section of walls tested by Paterson and Mitchell [85]. . . . . . . . . . . 119
2.90 Final damage of walls tested by Paterson and Mitchell [84]. . . . . . . . . . . 121
2.91 Strain profiles at splice in walls tested by Paterson and Mitchell [84]. . . . . . 121
2.92 Failure mode of walls studied by Wood [115]. . . . . . . . . . . . . . . . . . . 123
2.93 Reinforcement ratios for building inventory and experimental test programs. . 128
2.94 Geometry and demand characteristics for building inventory and experimen-
tal test programs. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
2.95 Reinforcement ratios for building inventory and experimental test programs
(ACI 318-08 compliant walls only). . . . . . . . . . . . . . . . . . . . . . . . . 130
2.96 Geometry and demand characteristics for building inventory and experimen-
tal test programs (ACI 318-08 compliant walls only). . . . . . . . . . . . . . . 131
2.97 Distribution of failure modes for experimental test specimens. . . . . . . . . . 133
2.98 Failure mode of experimental test specimens as a function of shear span ratio,
M/(V `w ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
2.99 Drift capacity for rectangular walls as a function of material strengths. . . . . 140
2.100Drift capacity as a function of wall geometry. . . . . . . . . . . . . . . . . . . 141
2.101Drift capacity as a function of reinforcement ratios. . . . . . . . . . . . . . . . 143
2.102Drift capacity as a function of shear span. . . . . . . . . . . . . . . . . . . . . 145
2.103Drift capacity as a function of axial load. . . . . . . . . . . . . . . . . . . . . 146
2.104Drift capacity as a function of shear demand and capacity values. . . . . . . . 148

xiv
3.1 Reinforcement ratios for building inventory and experimental test programs
(ACI 318-08 compliant walls only). . . . . . . . . . . . . . . . . . . . . . . . . 163
3.2 Geometry and demand characteristics for building inventory and experimen-
tal test programs (ACI 318-08 compliant walls only). . . . . . . . . . . . . . . 164
3.3 Dimensions of 10-story prototype wall and 31 scaled wall. The tested speci-
mens were the lower 3 floors of the scaled 10-story wall. . . . . . . . . . . . . 165
3.4 ASCE 7 ELF lateral load distributions. . . . . . . . . . . . . . . . . . . . . . 166
3.5 Uniform latera load distributions. . . . . . . . . . . . . . . . . . . . . . . . . . 167
3.6 Cross-section of boundary element longitudinal steel distribution. Used for
Specimens PW1, PW2, and PW4. See Chapter 4 for cross-sections of in-
dividual walls as cover dimensions of the constructed specimens may vary
slightly. Dimensions shown are in inches [44]. . . . . . . . . . . . . . . . . . . 169
3.7 Cross-section of uniform longitudinal steel distribution. Used for specimen
PW3. Dimensions shown are in inches [44]. . . . . . . . . . . . . . . . . . . . 170
3.8 Typical detail of splice in boundary regions. Dimensions are in inches [44]. . . 172
3.9 Concrete cylinder test stress-strain response. . . . . . . . . . . . . . . . . . . 174
3.10 Reinforcement stress-strain response with critical values . . . . . . . . . . . . 176
3.11 Construction of foundation block and wall cap. . . . . . . . . . . . . . . . . . 180
3.12 Load and Boundary Condition Box (LBCB) at the UIUC MUST-SIM Facility
for control of load point in 6 degrees of freedom (http://nees.uiuc.edu/
facilities/LBCBs.html). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181
3.13 Plan view of experimental test set-up for Specimen PW1 [44]. Dimensions
are in inches. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182
3.14 Front view of experimental test set-up [44]. Dimensions are in inches. . . . . 183
3.15 Range available for Krypton non-contact measurement system [44]. . . . . . . 184
3.16 Krypton LED locations for specimen PW1 [44]. Units in inches. . . . . . . . . 185
3.17 Location of fixed cameras used to capture high-resolution images [44]. . . . . 186
3.18 Layout of instruments that measured relative displacements on the north
(back) face of the wall [44]. Units are in inches. . . . . . . . . . . . . . . . . . 187
3.19 Locations of in-plane displacement measurements [44]. Units are in inches. . . 187
3.20 Locations of absolute out-of-plane displacement measurements [44]. Units
are in inches. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188
3.21 Location of concrete surface gauges. Gauges are located on the front and
back faces of the wall [44]. Dimensions are in inches. . . . . . . . . . . . . . . 189
3.22 Locations of strain gauges on horizontal reinforcement for PW1 [44]. Units
are in inches. Horizontal strain gauge locations for specimens PW2, PW3,
and PW4 are similar. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 190

xv
3.23 Locations of strain gauges on vertical reinforcement [44]. Units are in inches. 191
3.24 Applied displacement histories . . . . . . . . . . . . . . . . . . . . . . . . . . 193
3.25 Notation for loads applied to specimens. . . . . . . . . . . . . . . . . . . . . . 195
4.1 Specimen PW1 cross-section. Drawing provided by Chris Hart [44]. . . . . . . 201
4.2 Specimen PW1 elevation. Drawing provided by Chris Hart [44]. . . . . . . . . 201
4.3 Notation for loads applied to specimens. . . . . . . . . . . . . . . . . . . . . . 202
4.4 Specimen PW1 load-drift response. . . . . . . . . . . . . . . . . . . . . . . . . 203
4.5 Specimen PW1 crack pattern. . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
4.6 Specimen PW1 damage at 1.0% drift. . . . . . . . . . . . . . . . . . . . . . . 206
4.7 Specimen PW1 damage along wall edges. . . . . . . . . . . . . . . . . . . . . 206
4.8 Specimen PW1 bar fracture. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207
4.9 Specimen PW1 final damage. . . . . . . . . . . . . . . . . . . . . . . . . . . . 208
4.10 Specimen PW2 cross-section. Drawing provided by Chris Hart [44]. . . . . . . 209
4.11 Specimen PW2 elevation. Drawing provided by Chris Hart [44]. . . . . . . . . 209
4.12 Specimen PW2 load-drift response. . . . . . . . . . . . . . . . . . . . . . . . . 211
4.13 Specimen PW2 crack pattern. . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
4.14 Specimen PW2: Comparison of damage above splice in west (left) and east
(right) boundary elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
4.15 Specimen PW2: Damage to in the first floor before and after loss of lateral
load carrying capacity. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
4.16 Specimen PW2 final damage. . . . . . . . . . . . . . . . . . . . . . . . . . . . 215
4.17 Specimen PW3 cross-section. Drawing provided by Chris Hart [44]. . . . . . . 216
4.18 Specimen PW3 elevation. Drawing provided by Chris Hart [44]. . . . . . . . . 216
4.19 Specimen PW3 load-drift response. . . . . . . . . . . . . . . . . . . . . . . . . 218
4.20 Specimen PW3 crack pattern. . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
4.21 Specimen PW3 damage at second WL- peak at 1.0% drift (Step 1230). . . . . 221
4.22 Specimen PW3 final damage. . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
4.23 Specimen PW4 cross-section. Drawing provided by Chris Hart [44]. . . . . . . 223
4.24 Specimen PW4 elevation. Drawing provided by Chris Hart [44]. . . . . . . . . 223
4.25 Specimen PW4 load-drift response. . . . . . . . . . . . . . . . . . . . . . . . . 225
4.26 Specimen PW4 crack pattern. . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
4.27 Specimen PW4: Damage to wall following three cycles to 0.75% drift (Step
1070). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
4.28 Specimem PW4 damage at base of wall at the second ER+ peak at 1.0%
drift (Step 1230). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 229

xvi
4.29 Specimen PW4 final damage. . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
4.30 Base moment (Mb ) envelopes normalized by nominal moment strength (Mn ). 232
4.31 Normalized moment envelopes at the top of the spliced region of the wall
specimens. Specimen PW4 did not have a splice; the envelopes at the same
height are shown as a reference. . . . . . . . . . . . . . . . . . . . . . . . . . . 233
4.32 Critical moment (Mcrit ) envelopes normalized by the computed moment
strength at the critical section (Mcompcrit ). . . . . . . . . . . . . . . . . . . . . 235
4.33 Base shear (Vb ) envelopes normalized by nominal shear strength (Vn ). . . . . 237
4.34 Crack patterns in lower two floors of the walls following three cycles to 0.35%
drift. Crack maps provided by Chris Hart [44]. . . . . . . . . . . . . . . . . . 244
4.35 Crack and spall patterns in lower two floors of the walls at end of test. Crack
maps provided by Chris Hart [44]. . . . . . . . . . . . . . . . . . . . . . . . . 245
5.1 Comparison of string pot and Krypton measured lateral displacements in
Specimen PW2 at the top of the first floor (ER+ peaks). . . . . . . . . . . . 251
5.2 Specimen PW2 comparision of 10th floor drifts calculated using combinations
of shear and flexure stiffness estimated from experimental data. . . . . . . . . 259
5.3 ER+ loading direction envelopes for the specimen, effective, and roof drifts. . 261
5.4 Hoehler-Stanton cyclic material model [45] . . . . . . . . . . . . . . . . . . . . 263
5.5 Example stress-strain response for yield definitions. . . . . . . . . . . . . . . . 264
5.6 Sample yields maps for longitudinal reinforcement. . . . . . . . . . . . . . . . 266
5.7 Specimen PW2 vertical strain fields at the first ER+ peak at 0.75% drift. . . 268
5.8 Notation used for Krypton elements. . . . . . . . . . . . . . . . . . . . . . . . 269
5.9 Notation used for calculating element shear strains. A shear strain value is
calculated for each corner of the element. . . . . . . . . . . . . . . . . . . . . 270
6.1 Profiles at first ER+ and WL- peaks at 0.50% drift. . . . . . . . . . . . . . . 279
6.2 Profiles at first ER+ and WL- peaks at 1.0%. . . . . . . . . . . . . . . . . . . 281
6.3 Profiles of residual displacements and rotations (approximately zero lateral
load) following the first ER+ peak at 1.0% drift. . . . . . . . . . . . . . . . . 282
6.4 Calculated axial stiffness at zero lateral force for cycles to 0.50% drift or less
(minimum stiffness for each displacement level is shown). . . . . . . . . . . . 283
6.5 Calculated effective stiffness values for the first floor assuming the base of
the wall is fixed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
6.6 Calculated effective stiffness values for the first floor. . . . . . . . . . . . . . . 286
6.7 Calculated effective stiffness values for the second floor. . . . . . . . . . . . . 288
6.8 Secant stiffness at the effective height of loading and roof. Stiffnesses were
calculated using estimated displacement values. . . . . . . . . . . . . . . . . . 289

xvii
6.9 Average deformation components for the full wall specimens. . . . . . . . . . 291
6.10 Average deformation components in the first floor. . . . . . . . . . . . . . . . 292
6.11 Tensile yield in vertical reinforcement during final four drift levels. . . . . . . 295
6.12 Compression yield in vertical reinforcement during final four drift levels. . . . 296
6.13 Gauges indicating tension yield and strain hardening in longitudinal bars at
end of test. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 297
6.14 Gauges indicating compression yield and strain hardening in longitudinal bars
at end of test. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 298
6.15 Final yield and strain hardening pattern in foundation bars for spliced walls. 299
6.16 Tensile yield in horizontal reinforcement during final four drift levels. . . . . . 303
6.17 Vertical strain fields at first ER+ peak at 0.75% drift. . . . . . . . . . . . . . 304
6.18 Vertical strain for Specimens PW1 and PW2 at last WL- peak at 1.0% drift. 305
6.19 Vertical strain for Specimen PW1 at first and last WL- peak at 1.5% drift. . 306
6.20 Vertical strain for Specimens PW2 and PW3 at the last two peaks prior to
failure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 307
6.21 Vertical strain for Specimens PW1 and PW2 at last WL- peak at 1.0% drift. 308
6.22 Shear strain fields at first ER+ peak at 0.75% drift. . . . . . . . . . . . . . . 309
6.23 Specimens PW1 and PW2 shear strain fields at last WL- peaks at 0.50% and
1.0% drift. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 311
6.24 Shear strain fields at first WL- peak at 0.75% drift. . . . . . . . . . . . . . . . 312
6.25 Specimen PW4 shear strain fields at last peaks at 0.50% drift. . . . . . . . . . 312
6.26 Shear strain fields at first WL- peak at 0.75% drift. . . . . . . . . . . . . . . . 313
6.27 2nd principal strain fields at first ER+ peak at 0.75% drift. . . . . . . . . . . 314
6.28 Specimen PW1 and PW2 2nd principal strain fields at first WL- peak at 1.0%
drift. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 315
6.29 Specimen PW2 and PW3 2nd principal strain fields at last ER+ peak at 1.0%
drift. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
6.30 Specimen PW2 and PW4 2nd principal strain fields at first WL- peak at 1.0%
drift. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316
7.1 Histograms of material strengths . . . . . . . . . . . . . . . . . . . . . . . . . 329
7.2 Reported crack widths . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 333
7.3 Histograms of reinforcement ratios . . . . . . . . . . . . . . . . . . . . . . . . 334
7.4 Histogram of wall scale . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
7.5 Histograms of load characteristics. . . . . . . . . . . . . . . . . . . . . . . . . 336
7.6 Histograms of shear demand and capacity . . . . . . . . . . . . . . . . . . . . 338

xviii
7.7 Equal base reactions for a real wall (white box) can be achieved for different
specimen (gray box) heights (hw ) by changing the configuration of the applied
loading. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
7.8 Model used to estimate the drift at the effective height of the wall. . . . . . . 341
7.9 Damage state (DS) versus engineering demand parameter (EDP). Outliers
are indicated by filled marker. . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
7.10 Examples of spalling damage states. . . . . . . . . . . . . . . . . . . . . . . . 348
7.11 Examples of damage states associated with MOR4. The damage states shown
are severe examples to provide clear illustration of the damage states. . . . . 350
7.12 Examples of shear failures associated with MOR4. . . . . . . . . . . . . . . . 352
7.13 Fragility functions for cyclically loaded walls . . . . . . . . . . . . . . . . . . . 365
8.1 Centro Mayor (CM) response spectra. ASCE and NCh. spectra are design
spectra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 383
8.2 Plaza del Rio Building A (PR-A) response spectra. ASCE and NCh. spectra
are design spectra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 384
8.3 Plaza del Rio Building B (PR-B) response spectra. ASCE and NCh. spectra
are design spectra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 385
8.4 Nonlinear shear backbone curve used in Tier 3 OpenSees model. . . . . . . . 391
8.5 Generic wall model results to demonstrate the variation in rotation demands
resulting from lumped and distributed plasticity models. . . . . . . . . . . . . 398
8.6 Photo of Plaza del Rio buildings. . . . . . . . . . . . . . . . . . . . . . . . . . 400
8.7 Floor plan of Plaza del Rio. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401
8.8 Plaza del Rio Building A: Layout of walls in typical floors. Color is used to
indicate the level of observed damage. . . . . . . . . . . . . . . . . . . . . . . 402
8.9 Damage at south end of Plaza del Rio Building A. . . . . . . . . . . . . . . . 403
8.10 Damage at center T-walls in Plaza del Rio Building A. . . . . . . . . . . . . . 403
8.11 Plaza del Rio Building B: Layout of walls in typical floors. . . . . . . . . . . . 404
8.12 Photo of Centro Mayor. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 405
8.13 Plans of typical floors of Centro Mayor. . . . . . . . . . . . . . . . . . . . . . 406
8.14 Concepto Mayor: Layout of walls in basement levels. Perimeter walls are not
shown. Color is used to indicate the level of observed damage. . . . . . . . . . 406
8.15 Concepto Mayor: Layout of walls in typical floors. Color is used to indicate
the level of observed damage. . . . . . . . . . . . . . . . . . . . . . . . . . . . 407
8.16 Centro Mayor (CM): Damage at intersection of coupled wall piers [27]. . . . . 408
8.17 Sketch of Concepto Urbano . . . . . . . . . . . . . . . . . . . . . . . . . . . . 409
8.18 Plan of Concepto Urbano. The various shades of indicate the plan at different
floors. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 410

xix
8.19 Concepto Urbano: Layout of walls in basement levels. Perimeter walls are
not shown. No damage was recorded. . . . . . . . . . . . . . . . . . . . . . . . 410
8.20 Concepeto Urbano: Layout of walls in typical floors in lower portion of build-
ing. No damage was recorded. . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
8.21 Concepto Urbano: Layout of walls in floors in the upper portion of the build-
ing. No damage was recorded. . . . . . . . . . . . . . . . . . . . . . . . . . . . 411
9.1 PR-A Floor 1: Unfactored demand capacity ratios (DCRu ). . . . . . . . . . . 417
9.2 PR-A Floor 1: Controlling action using unfactored demand capacity ratios
(DCRu ). Blue = Moment; Red = Shear; Black = Axial. . . . . . . . . . . . . 417
9.3 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under ASCE 7 response spectrum. . . . . . . . . . . . . . . . . . . . . . 421
9.4 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under Concepcion response spectrum. . . . . . . . . . . . . . . . . . . . 422
9.5 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under San Pedro response spectrum. . . . . . . . . . . . . . . . . . . . . 423
9.6 Plaza del Rio Building A Floor 1: # of damage state occurances in walls
achieving various performance objectives. Color indicates the observed damage.425
9.7 PR-A Floor 1: Ratio of maximum axial load due to earthquake loading to
axial load due to gravity loading from different response spectra. . . . . . . . 427
9.8 PR-A Floor 1: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial
capacity of the wall (Ag fc0 ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 428
9.9 Summary of walls exceeding acceptance criteria for unscaled ground motion
records. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 438
9.10 Floor 1 walls exceeding ASCE 41 rotation limits (Concepcion ground motion).439
9.11 Floor 2 walls exceeding ASCE 41 rotation limits (Concepcion ground motion).439
9.12 Floor 1 walls exceeding fragility function probabilities of 0.50 (Concepcion
ground motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
9.13 Floor 2 walls exceeding fragility function probabilities of 0.50 (Concepcion
ground motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 440
9.14 Floor 1 walls exceeding strain limits (Concepcion ground motion). . . . . . . 441
9.15 Floor 2 walls exceeding strain limits (Concepcion ground motion). . . . . . . 441
9.16 Floor 1 walls exceeding ASCE 41 rotation limits (San Pedro de la Paz ground
motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442
9.17 Floor 2 walls exceeding ASCE 41 rotation limits (San Pedro de la Paz ground
motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 442

xx
9.18 Floor 1 walls exceeding fragility function probabilities of 0.50 (San Pedro de
la Paz ground motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
9.19 Floor 2 walls exceeding fragility function probabilities of 0.50 (San Pedro de
la Paz ground motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 443
9.20 Floor 1 walls exceeding strain limits (San Pedro de la Paz ground motion). . 444
9.21 Floor 2 walls exceeding strain limits (San Pedro de la Paz ground motion). . 444
9.22 Summary of walls exceeding acceptance criteria with consideration for shear
strength limit for unscaled ground motion records. . . . . . . . . . . . . . . . 445
9.23 Floor 1 walls exceeding ASCE 41 rotation limits with consideration for shear
strength limit (Concepcion ground motion). . . . . . . . . . . . . . . . . . . . 446
9.24 Floor 2 walls exceeding ASCE 41 rotation limits with consideration for shear
strength limit (Concepcion ground motion). . . . . . . . . . . . . . . . . . . . 446
9.25 Floor 1 walls exceeding fragility function probabilities of 0.50 with consider-
ation for shear strength limit (Concepcion ground motion). . . . . . . . . . . 447
9.26 Floor 2 walls exceeding fragility function probabilities of 0.50 with consider-
ation for shear strength limit (Concepcion ground motion). . . . . . . . . . . 447
9.27 Floor 1 walls exceeding strain limits with consideration for shear strength
limit (Concepcion ground motion). . . . . . . . . . . . . . . . . . . . . . . . . 448
9.28 Floor 2 walls exceeding strain limits with consideration for shear strength
limit (Concepcion ground motion). . . . . . . . . . . . . . . . . . . . . . . . . 448
9.29 Floor 1 walls exceeding ASCE 41 rotation limits with consideration for shear
strength limit (San Pedro de la Paz ground motion). . . . . . . . . . . . . . . 449
9.30 Floor 2 walls exceeding ASCE 41 rotation limits with consideration for shear
strength limit (San Pedro de la Paz ground motion). . . . . . . . . . . . . . . 449
9.31 Floor 1 walls exceeding fragility function probabilities of 0.50 with consider-
ation for shear strength limit (San Pedro de la Paz ground motion). . . . . . 450
9.32 Floor 2 walls exceeding fragility function probabilities of 0.50 with consider-
ation for shear strength limit (San Pedro de la Paz ground motion). . . . . . 450
9.33 Floor 1 walls exceeding strain limits with consideration for shear strength
limit (San Pedro de la Paz ground motion). . . . . . . . . . . . . . . . . . . . 451
9.34 Floor 2 walls exceeding strain limits with consideration for shear strength
limit (San Pedro de la Paz ground motion). . . . . . . . . . . . . . . . . . . . 451
9.35 Ratio of maximum compressive axial demand to axial compressive capacity
(Ag fc0 ). Values shown are percents. Note: All walls are colored and apparent
differences in line thickness do not indicate different results. . . . . . . . . . . 458
9.36 Ratio of minimum (most tensile) axial demand to axial tensile capacity
(As fy ). Values shown are a percent. Note: All walls are colored and ap-
parent differences in line thickness do not indicate different results. . . . . . . 458

xxi
9.37 Examples of geometric discontinuities with damage level indicated. . . . . . . 462
9.38 Number of local geometric discontinuities, sorted by rank. A rank of 0 in-
dicates the highest quality data; a rank of 6 indicates the lowest quality of
data. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 464
9.39 Change in global parameters vs discontinuity ID. . . . . . . . . . . . . . . . . 467
9.40 Change in relevant directional parameters (severely damage buildings) vs
discontinuity ID. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 468
9.41 Change in controlling directional parameter (severely damaged buildings) vs
discontinuity ID. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 469
A.1 PW1 crack pattern in lower two stories following three cycles at 0.10% drift. . 494
A.2 PW1 crack pattern in lower two stories following three cycles at 0.25% drift. . 495
A.3 PW1 damage during 0.35% and 0.50% drift cycles. . . . . . . . . . . . . . . . 496
A.4 PW1 vertical splitting at top of east splice at first ER+ peak at 0.75% drift. 497
A.5 PW1: Spalling on west side of wall during 0.75% and 1.0% drift cycles. . . . 499
A.6 PW1: Initial bar buckling on west side at first WL- peak at 1.0% drift (Step
620) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 499
A.7 PW1: Vertical crack on east edge of wall at first ER+ peak at 1.0% drift
(Step 600) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 500
A.8 PW1: Multiple bars buckled at final WL- peak at 1.0% drift (Step 620) . . . 501
A.9 PW1: East side of wall during 1.5% drift cycles . . . . . . . . . . . . . . . . . 502
A.10 PW1: Interface crack opening at drop in lateral load carrying capacity (−1.38%
drift, Step 739) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 506
A.11 PW1: Longitudinal bar fracture on back face of east boundary element at
drop in lateral load carrying capacity (−1.38% drift, Step 739) . . . . . . . . 506
A.12 PW1: Longitudinal bar fracture on front face of east boundary element at
drop in lateral load carrying capacity (−1.38% drift, Step 739) . . . . . . . . 507
A.13 PW1: Longitudinal bar fracture on front face of east boundary element at
final peak (−1.523% drift, Step 740) . . . . . . . . . . . . . . . . . . . . . . . 507
A.14 PW1: Longitudinal bar fracture on back face of east boundary element at
final peak (−1.523% drift, Step 740) . . . . . . . . . . . . . . . . . . . . . . . 508
A.15 PW2 crack pattern in lower two stories following three cycles at 0.14% drift. . 510
A.16 PW2 crack pattern in lower two stories following three cycles at 0.21% drift. . 511
A.17 PW2 crack pattern in lower two stories following three cycles at 0.35% drift. . 512
A.18 PW2 crack pattern in lower two stories following three cycles at 0.14% drift. . 513
A.19 PW2: West boundary element at first ER+ peak at 0.75% drift, with 0.1
inch wide crack above splice (Step 870) . . . . . . . . . . . . . . . . . . . . . 515

xxii
A.20 PW2: Vertical splitting above west boundary element splice at WL- peak at
−0.75% drift (Step 910) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 515
A.21 PW2: West toe of wall at WL- peak at −0.75% drift (Step 910) . . . . . . . . 516
A.22 PW2: Spalling along cracks above splice in west boundary element at WL-
peak at −0.75% drift (Step 910) . . . . . . . . . . . . . . . . . . . . . . . . . 516
A.23 PW2: East toe of wall following third ER+ peak at 0.75% drift (Step 1030) . 517
A.24 PW2: Vertical cracks at top of splice in east boundary element following
third ER+ peak at 0.75% drift (Step 1030) . . . . . . . . . . . . . . . . . . . 517
A.25 PW2: Bar buckling in west boundary element following third WL- peak at
−0.75% drift (Step 1070) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 519
A.26 PW2: Bar buckling above east boundary element splice at third ER+ peak
at 1.0% drift (Step 1190) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 520
A.27 PW2: Extensive bar buckling above west boundary element splice at first
Wl- peak at 1.0% drift (Step 1150) . . . . . . . . . . . . . . . . . . . . . . . . 521
A.28 PW2: Spalling in east boundary element at 1.5% drift (Step 1250) . . . . . . 523
A.29 PW2: Damage to west boundary element at −1.05% drift (Step 1304). One
step prior to loss of load carrying capacity. . . . . . . . . . . . . . . . . . . . . 523
A.30 PW2: Web crushing and bar buckling following loss of load carrying capacity
(Step 1305) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 524
A.31 PW2: West boundary element and web following loss of load carrying capac-
ity (Step 1305) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 524
A.32 PW2: Damage to back side of wall following loss of load carrying capacity
(Step 1305) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525
A.33 PW2: East boundary element following loss of load carrying capacity (Step
1305) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525
A.34 PW2: Damage to the wall in its final resting place . . . . . . . . . . . . . . . 526
A.35 PW3: Crack pattern following 3 cycles at ±0.14% drift (Step 250) . . . . . . 529
A.36 PW3: Crack pattern following 3 cycles at ±0.21% drift (Step 370) . . . . . . 530
A.37 PW3: Crack pattern following 3 cycles at ±0.35% drift (Step 610) . . . . . . 531
A.38 PW3: Crack pattern following 3 cycles at ±0.50% drift (Step 850) . . . . . . 533
A.39 PW3: West boundary element at the third WL- peak at 0.50% drift (Step 830)533
A.40 PW3: Spalling in region above splice in east boundary element at first ER+
peak at 0.75% drift (Step 870) . . . . . . . . . . . . . . . . . . . . . . . . . . 535
A.41 PW3: Damage above splice in west boundary element at first ER+ peak at
0.75% drift (Step 870) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 535
A.42 PW3: Spalling above splice on east side of wall at second WL- peak at 0.75%
drift (Step 990) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 537

xxiii
A.43 PW3: Crack pattern following 3 cycles at ±0.75% drift (Step 1090) . . . . . . 538
A.44 PW3: Bar buckling on west side of wall above splice at −0.743% drift on way
to first WL- peak at 1.0% drift (Step 1145). . . . . . . . . . . . . . . . . . . . 540
A.45 PW3: Bar buckling on west side of wall above splice at first WL- peak at
1.0% drift (Step 1150). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 540
A.46 PW3: Bar buckling in east web at second ER+ peak at 1.0% drift (Step 1190)541
A.47 PW3: Bar buckling on east side at second ER+ peak at 1.0% drift (Step 1190)542
A.48 PW3: Damage to first two floors at second ER+ peak at 1.0% drift (Step
1190) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 542
A.49 PW3: Damage to first two floors at second WL- peak at 1.0% drift (Step 1230)543
A.50 PW3: Back of wall at loss of axial load capacity (1.28% drift, Step 1267) . . . 544
A.51 PW4: Crack pattern following 3 cycles at ±0.14% drift (Step 250) . . . . . . 547
A.52 PW4: Crack pattern following 3 cycles at ±0.21% drift (Step 370) . . . . . . 548
A.53 PW4: Crack pattern following 3 cycles at ±0.35% drift (Step 610) . . . . . . 549
A.54 PW4: Spalling at east toe of wall (Step 630) . . . . . . . . . . . . . . . . . . . 551
A.55 PW4: Spalling at east toe of wall (Step 750) . . . . . . . . . . . . . . . . . . . 551
A.56 PW4: Spalling at west side of wall following ±0.50% drift cycles (Step 830) . 552
A.57 PW4: Spalling of patched region of east boundary element following ±0.50%
drift cycles (Step 830) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 552
A.58 PW4: Extension of spalling in east boundary element at first ER+ peak at
0.75% drift (Step 870). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 554
A.59 PW4: Vertical crack in west boundary element at 0.602% drift (Step 866). . . 554
A.60 PW4: Exposed longitudinal reinforcement in east boundary element at first
ER+ peak at 0.75% drift (Step 870). . . . . . . . . . . . . . . . . . . . . . . . 555
A.61 PW4: Increase in vertical cracks on west toe of wall at −0.521% drift (Step
904). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 555
A.62 PW4: Bar buckling in east boundary element at second ER+ peak at 0.75%
drift (Step 950). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 556
A.63 PW4: Exposed longitudinal reinforcement in west boundary element at third
ER+ peak at 0.75% drift (Step 1030). . . . . . . . . . . . . . . . . . . . . . . 556
A.64 PW4: Damage to east boundary element core concrete at third WL- peak at
0.75% drift (Step 1070). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
A.65 PW4: Damage to west boundary element at third WL- peak at 0.75% drift
(Step 1070). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 557
A.66 PW4: Bar buckling and core damage in east boundary element at first ER+
peak at 1.0% drift (Step 1110). . . . . . . . . . . . . . . . . . . . . . . . . . . 559

xxiv
A.67 PW4: Face of east boundary element at the first ER+ peak at 1.0% drift
(Step 1110). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 559
A.68 PW4: Spalling in west boundary at the first WL- peak at 1.0% drift (Step
1150). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560
A.69 PW4: East Bottom camera at loss of axial load capacity (Step 1187). . . . . 562
A.70 PW4: West Bottom camera at loss of axial load capacity (Step 1187). . . . . 563
A.71 PW4: East boundary element damage at loss of axial capacity (Step 1187). . 563
A.72 PW4: Loss of axial capacity: bar buckling at the east side of the wall (Step
1187). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 564
A.73 PW4: Loss of axial capacity: spalling and bar buckling on the back side of
web next to the east boundary element (Step 1187). . . . . . . . . . . . . . . 564
A.74 PW4: Base of the west boundary element at the second ER+ peak at 1.0%
drift (Step 1190). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 565
A.75 PW4: East half of first floor at second ER+ peak at 1.0% drift (Step 1190). . 565
A.76 PW4: Out-of-plane displacement of wall as seen from the east side view of
the first floor at the second ER+ peak at 1.0% drift (Step 1190). . . . . . . . 566
B.1 Specimen PW1 displacement profiles at first peaks for each drift level. . . . . 568
B.2 Specimen PW1 rotation profiles at first peaks for each drift level. . . . . . . . 568
B.3 Specimen PW2 displacement profiles at first peaks for each drift level. . . . . 569
B.4 Specimen PW2 rotation profiles at first peaks for each drift level. . . . . . . . 569
B.5 Specimen PW3 displacement profiles at first peaks for each drift level. . . . . 570
B.6 Specimen PW3 rotation profiles at first peaks for each drift level. . . . . . . . 570
B.7 Specimen PW4 displacement profiles at first peaks for each drift level. . . . . 571
B.8 Specimen PW4 rotation profiles at first peaks for each drift level. . . . . . . . 571
B.9 Comparison of effective flexural stiffness values for the first floor (base fixed
and allowed to rotate) and the second floor. Values calculated using calcu-
lated using Timoshenko beam theory. . . . . . . . . . . . . . . . . . . . . . . 572
B.10 Comparison of effective shears stiffness values for the first floor (base fixed and
allowed to rotate) and the second floor. Values calculated using Timoshenko
beam theory. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 573
B.11 Specimen PW1 deformation contributions. . . . . . . . . . . . . . . . . . . . . 574
B.12 Specimen PW2 deformation contributions. . . . . . . . . . . . . . . . . . . . . 575
B.13 Specimen PW3 deformation contributions. . . . . . . . . . . . . . . . . . . . . 576
B.14 Specimen PW4 deformation contributions. . . . . . . . . . . . . . . . . . . . . 577
B.15 Vertical strain fields at first and last peaks at 0.50% drift. . . . . . . . . . . . 578
B.16 Vertical strain fields at first and last peaks at 0.75% drift. . . . . . . . . . . . 579

xxv
B.17 Vertical strain fields at first and last peaks at 1.00% drift. . . . . . . . . . . . 580
B.18 Vertical strain fields at first and last peaks at 1.50% drift. . . . . . . . . . . . 581
B.19 Shear strain fields at first and last peaks at 0.50% drift. . . . . . . . . . . . . 582
B.20 Shear strain fields at first and last peaks at 0.75% drift. . . . . . . . . . . . . 583
B.21 Shear strain fields at first and last peaks at 1.00% drift. . . . . . . . . . . . . 584
B.22 Shear strain fields at first and last peaks at 1.50% drift. . . . . . . . . . . . . 585
B.23 2nd principal strain fields at first and last peaks at 0.50% drift. . . . . . . . . 586
B.24 2nd principal strain fields at first and last peaks at 0.75% drift. . . . . . . . . 587
B.25 2nd principal strain fields at first and last peaks at 1.00% drift. . . . . . . . . 588
B.26 2nd principal strain fields at first and last peaks at 1.50% drift. . . . . . . . . 589
B.27 1st principal strain fields at first and last peaks at 0.50% drift. . . . . . . . . 590
B.28 1st principal strain fields at first and last peaks at 0.75% drift. . . . . . . . . 591
B.29 1st principal strain fields at first and last peaks at 1.00% drift. . . . . . . . . 592
B.30 1st principal strain fields at first and last peaks at 1.50% drift. . . . . . . . . 593
C.1 Envelope for PilaSW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 611
C.2 PilaSW4: DS1a - cracks at MDL-2 (peak after initial cracking) . . . . . . . . 611
C.3 PilaSW4: DS2d - cracks at MDL-16 (larger shear crack widths) . . . . . . . . 612
C.4 PilaSW4: DS4a - cracks at MDL-22 (core crushing) . . . . . . . . . . . . . . . 612
C.5 PilaSW4: DS4d - cracks at MDL-22 (core crushing) . . . . . . . . . . . . . . 613
C.6 PilaSW4: DSr1 - cracks at MDL-8 (peak of cycle following yield cycle) . . . . 613
C.7 PilaSW4: DSr4 - cracks at MDL-16 (larger shear crack widths) . . . . . . . . 614
C.8 Envelope for PilaSW5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 616
C.9 PilaSW5: DS1a - cracks at MDL-2 (peak after initial cracking) and MDL-4
(wall fully cracked) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 616
C.10 PilaSW5: DS1f - cracks at MDL-8 (shear cracks extend length of wall) and
MDL-10 (shear reinf. yield) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 617
C.11 PilaSW5: DS4i - cracks at MDL-8 (shear cracks extend length of wall) and
MDL-10 (shear reinf. yield) . . . . . . . . . . . . . . . . . . . . . . . . . . . . 617
C.12 Envelope for PilaSW6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 619
C.13 PilaSW6: DS1a - cracks at MDL-2 (peak after initial cracking) and MDL-4
(fully cracked) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 619
C.14 PilaSW6: DS2a - cracks at MDL-8 (yield) and MDL-16 (spalling/maximum
load) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 620
C.15 PilaSW6: DS4q - cracks at MDL-18 (open stirrup) and Failure . . . . . . . . 620
C.16 Envelope for PilaSW7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 621

xxvi
C.17 PilaSW7: DS1a - cracks at MDL-2 (peak after initial cracking) and MDL-4
(fully cracked) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 622
C.18 PilaSW7: DS4e - cracks at MDL-22 (failure due to bar fracture) . . . . . . . 622
C.19 PilaSW7: DSr5 - cracks at MDL-8 and MDL-4 (shear cracks in b.e.) . . . . . 623
C.20 Envelope for PilaSW8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
C.21 PilaSW8: DS1a - cracks at MDL-2 (peak after initial cracking) and MDL-4
(fully cracked) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
C.22 PilaSW8: DS2a - cracks at MDL-18 (spalling) and Failure . . . . . . . . . . . 626
C.23 PilaSW8: DSr4 - cracks at MDL-6 and MDL-12 . . . . . . . . . . . . . . . . . 626
C.24 Envelope for PilaSW9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 627
C.25 PilaSW9: DS1a - cracks at MDL-2 (peak after initial cracking) and MDL-4
(fully cracked) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 628
C.26 PilaSW9: DS2a - cracks at MDL-18 (spalling) and Failure . . . . . . . . . . . 628
C.27 PilaSW9: DSr4 - cracks at MDL-6 and MDL-14 . . . . . . . . . . . . . . . . . 629
C.28 Envelope for TasSHW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 630
C.29 Envelope for TasSHW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 631
C.30 Envelope for TasSHW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 632
C.31 Envelope for TasSHW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 633
C.32 Envelope for ThomRW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 635
C.33 ThomRW1: DS4b - Slight buckling of RW1 bars at 2% drift . . . . . . . . . . 635
C.34 Envelope for ThomRW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 637
C.35 ThomRW2: DS4f - RW2 damage at 2.5% drift . . . . . . . . . . . . . . . . . 637
C.36 Envelope for ThomTW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
C.37 ThomTW1: DS4f - TW1 damage at 1.25% drift . . . . . . . . . . . . . . . . . 639
C.38 Envelope for ThomTW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 640
C.39 ThomTW2: DS4f - TW2 damage at 2.5% drift . . . . . . . . . . . . . . . . . 641
C.40 Envelope for DazioWSH1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 643
C.41 Envelope for DazioWSH2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 645
C.42 Envelope for DazioWSH3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 647
C.43 Envelope for DazioWSH4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 649
C.44 Envelope for DazioWSH5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 651
C.45 Envelope for DazioWSH6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 653
C.46 Envelope for LefasSW21 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 655
C.47 Envelope for LefasSW22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 656
C.48 Envelope for LefasSW23 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 657

xxvii
C.49 Envelope for LefasSW24 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 659
C.50 Envelope for LefasSW25 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 660
C.51 Envelope for LefasSW26 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 662
C.52 Envelope for LefasSW30 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664
C.53 Envelope for LefasSW31 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 666
C.54 Envelope for LefasSW32 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 668
C.55 Envelope for LefasSW33 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 670
C.56 Envelope for OestR1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 672
C.57 Envelope for OestR2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 674
C.58 Envelope for OestB1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 676
C.59 OestB1: DS4b - Buckling of reinforcing bars . . . . . . . . . . . . . . . . . . . 676
C.60 Envelope for OestB2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 678
C.61 Envelope for OestB3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 680
C.62 Envelope for OestB4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 682
C.63 Envelope for OestB5 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 684
C.64 Envelope for OestB6 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 686
C.65 Envelope for OestB7 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 688
C.66 Envelope for OestB8 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 690
C.67 Envelope for OestB9 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 692
C.68 Envelope for OestB10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 694
C.69 Envelope for OestF1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 696
C.70 Envelope for OestF2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 698
C.71 Envelope for Morgan . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 700
C.72 Morgan: DS1a - crack map after steps 19/22 . . . . . . . . . . . . . . . . . . 700
C.73 Morgan: DS2a - crack map after steps 108/111 . . . . . . . . . . . . . . . . . 701
C.74 Morgan: DS2d - crack map after steps 88/91 . . . . . . . . . . . . . . . . . . 701
C.75 Morgan: DS3a - crack map after steps 128/131 . . . . . . . . . . . . . . . . . 702
C.76 Morgan: DSr1 - crack map after steps 65/68 . . . . . . . . . . . . . . . . . . . 702
C.77 Envelope for LiuW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 704
C.78 LiuW1: DS1b - photo after initial cracking . . . . . . . . . . . . . . . . . . . 705
C.79 LiuW1: DS1c - photo after 2.0Mcr cycles . . . . . . . . . . . . . . . . . . . . 705
C.80 LiuW1: DS1e - cracking of first yield . . . . . . . . . . . . . . . . . . . . . . . 706
C.81 LiuW1: DS2c - cracking at first cycle of 3.80 Delta y; spalling in web . . . . . 706
C.82 LiuW1: DS4b - cracking after final cycle of 3.80 Delta y; longitudinal buckling707
C.83 LiuW1: DS4f - photo at failure . . . . . . . . . . . . . . . . . . . . . . . . . . 707

xxviii
C.84 Envelope for LiuW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 709
C.85 LiuW2: DS1b - photo after initial cracking . . . . . . . . . . . . . . . . . . . 710
C.86 LiuW2: DS1c - photo after 2.0Mcr cycles . . . . . . . . . . . . . . . . . . . . 710
C.87 LiuW2: DS1f - cracking of 3.38 Deltay . . . . . . . . . . . . . . . . . . . . . . 711
C.88 LiuW2: DS2a - cracking of 2.25*Deltay . . . . . . . . . . . . . . . . . . . . . 711
C.89 LiuW2: DS2c - cracking at first cycle of 3.80 Delta y; spalling in web . . . . . 712
C.90 LiuW2: DS4b - cracking after cycle 40a . . . . . . . . . . . . . . . . . . . . . 712
C.91 LiuW2: DS4f - photo at failure . . . . . . . . . . . . . . . . . . . . . . . . . . 713
C.92 LiuW2: DSr1 - cracking of first yield . . . . . . . . . . . . . . . . . . . . . . . 713
C.93 Envelope for AliW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 715
C.94 AliW1: DS2a - photo of cycle 9 damage . . . . . . . . . . . . . . . . . . . . . 715
C.95 Envelope for TupperW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 717
C.96 TupperW3: DS2a - deflection of approximately 1.75 Deltay . . . . . . . . . . 717
C.97 TupperW3: DS4f - final damage of wall . . . . . . . . . . . . . . . . . . . . . 718
C.98 Envelope for MobeenW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 720
C.99 MobeenW1: DSr11 - Final damage of wall (close up of boundary element) . . 720
C.100Envelope for RivaW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 722
C.101Envelope for ShiuC1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723
C.102ShiuC1: DS1a - crack pattern following first phase of testing . . . . . . . . . 724
C.103ShiuC1: DS4b - final damage of wall . . . . . . . . . . . . . . . . . . . . . . . 724
C.104ShiuC1: DS4i - final damage of wall . . . . . . . . . . . . . . . . . . . . . . . 725
C.105Envelope for KhaC1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 726
C.106Envelope for ElnCW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 727
C.107Envelope for ElnCW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 728
C.108Envelope for IleX . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 729
C.109Envelope for IleY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 730
C.110Envelope for IleXY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 731
C.111IleXY: DS4e - Shear failure of flange . . . . . . . . . . . . . . . . . . . . . . . 732
C.112Envelope for TUA . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733
C.113Envelope for TUB . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 734
C.114Envelope for NTW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 737
C.115Envelope for NTW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 740
C.116Envelope for PW1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 741
C.117Envelope for PW2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742
C.118Envelope for PW3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 743

xxix
C.119Envelope for PW4 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 744
D.1 Presentation providing overview of Plaza del Rio Buildings A & B. . . . . . . 746
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 747
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 748
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 749
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 750
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 751
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 752
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 753
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 754
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 755
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 756
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 757
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 758
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 759
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 760
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 761
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 762
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 763
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 764
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 765
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 766
D.1 Presentation providing overview of Plaza del Rio Buildings A & B (con’t). . . 767
D.2 Presentation providing overview of Centro Mayor. . . . . . . . . . . . . . . . 768
D.2 Presentation providing overview of Centro Mayor. . . . . . . . . . . . . . . . 769
D.2 Presentation providing overview of Centro Mayor. . . . . . . . . . . . . . . . 770
D.2 Presentation providing overview of Centro Mayor. . . . . . . . . . . . . . . . 771
D.2 Presentation providing overview of Centro Mayor. . . . . . . . . . . . . . . . 772
D.2 Presentation providing overview of Centro Mayor. . . . . . . . . . . . . . . . 773
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 774
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 775
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 776
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 777
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 778
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 779

xxx
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 780
D.3 Presentation providing overview of Concepto Urbano. . . . . . . . . . . . . . 781
F.1 PR-A Floor 1: # of damage state occurances in walls achieving various
performance objectives. Color indicates the observed damage. . . . . . . . . . 785
F.2 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under ASCE 7 response spectrum. . . . . . . . . . . . . . . . . . . . . . 786
F.3 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under Chilean response spectrum. . . . . . . . . . . . . . . . . . . . . . . 787
F.4 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under proposed Chilean response spectrum. . . . . . . . . . . . . . . . . 788
F.5 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under Concepcion response spectrum. . . . . . . . . . . . . . . . . . . . 789
F.6 PR-A Floor 1: Wall damage for walls achieving various performance objec-
tives under San Pedro de la Paz response spectrum. . . . . . . . . . . . . . . 790
F.7 PR-B Floor 1: # of damage state occurances in walls achieving various per-
formance objectives. Color indicates the observed damage. . . . . . . . . . . . 791
F.8 PR-B Floor 1: Wall damage for walls achieving various performance objec-
tives under ASCE 7 response spectrum. . . . . . . . . . . . . . . . . . . . . . 792
F.9 PR-B Floor 1: Wall damage for walls achieving various performance objec-
tives under Chilean response spectrum. . . . . . . . . . . . . . . . . . . . . . . 793
F.10 PR-B Floor 1: Wall damage for walls achieving various performance objec-
tives under proposed Chilean response spectrum. . . . . . . . . . . . . . . . . 794
F.11 PR-B Floor 1: Wall damage for walls achieving various performance objec-
tives under Concepcion response spectrum. . . . . . . . . . . . . . . . . . . . 795
F.12 PR-B Floor 1: Wall damage for walls achieving various performance objec-
tives under San Pedro de la Paz response spectrum. . . . . . . . . . . . . . . 796
F.13 CM Floor 2: # of damage state occurances in walls achieving various perfor-
mance objectives. Color indicates the observed damage. . . . . . . . . . . . . 797
F.14 CM Floor 2: Wall damage for walls achieving various performance objectives
under ASCE 7 response spectrum. . . . . . . . . . . . . . . . . . . . . . . . . 798
F.15 CM Floor 2: Wall damage for walls achieving various performance objectives
under Chilean response spectrum. . . . . . . . . . . . . . . . . . . . . . . . . . 799
F.16 CM Floor 2: Wall damage for walls achieving various performance objectives
under proposed Chilean response spectrum. . . . . . . . . . . . . . . . . . . . 800
F.17 CM Floor 2: Wall damage for walls achieving various performance objectives
under Concepcion response spectrum. . . . . . . . . . . . . . . . . . . . . . . 801
F.18 CM Floor 2: Wall damage for walls achieving various performance objectives
under San Pedro de la Paz response spectrum. . . . . . . . . . . . . . . . . . 802

xxxi
F.19 PR-A Floor 1: Unfactored demand capacity ratios (DCRu ). . . . . . . . . . . 804
F.20 PR-A Floor 1: Action resulting in largest DCRu value. Blue: Moment; Red:
Shear; Black: Axial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 804
F.21 PR-A Floor 1: Factored demand capacity ratios (DCRf ) for immediate oc-
cupancy performance level. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 805
F.22 PR-A Floor 1: Action resulting in largest DCRf value for immediate occu-
pancy performance level. Blue: Moment; Red: Shear; Black: Axial . . . . . . 805
F.23 PR-B Floor 1: Unfactored demand capacity ratios (DCRu ). . . . . . . . . . . 806
F.24 PR-B Floor 1: Action resulting in largest DCRu value. Blue: Moment; Red:
Shear; Black: Axial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 807
F.25 PR-B Floor 1: Factored demand capacity ratios (DCRf ) for immediate oc-
cupancy performance level. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 808
F.26 PR-B Floor 1: Action resulting in largest DCRf value for immediate occu-
pancy performance level. Blue: Moment; Red: Shear; Black: Axial . . . . . . 809
F.27 CM Floor 2: Unfactored demand capacity ratios (DCRu ). . . . . . . . . . . . 810
F.28 CM Floor 2: Action resulting in largest DCRu value. Blue: Moment; Red:
Shear; Black: Axial . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 811
F.29 CM Floor 2: Factored demand capacity ratios (DCRf ) for immediate occu-
pancy performance level. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 812
F.30 CM Floor 2: Action resulting in largest DCRf value for immediate occupancy
performance level. Blue: Moment; Red: Shear; Black: Axial . . . . . . . . . . 813
F.31 PR-A Floor 1: Ratio of maximum axial load due to earthquake loading to
axial load due to gravity loading from different response spectra. . . . . . . . 814
F.32 PR-A Floor 1: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial
capacity of the wall (Ag fc0 ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 815
F.33 PR-B Floor 1: Ratio of maximum axial load due to earthquake loading to
axial load due to gravity loading from different response spectra. . . . . . . . 816
F.34 PR-B Floor 1: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial
capacity of the wall (Ag fc0 ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 817
F.35 CM Floor 2: Ratio of maximum axial load due to earthquake loading to axial
load due to gravity loading from different response spectra. . . . . . . . . . . 818
F.36 CM Floor 2: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial
capacity of the wall (Ag fc0 ). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 819
G.1 Summary of walls exceeding acceptance criteria (Concepcion ground motion). 821
G.2 Walls exceeding ASCE 41 rotation limits (Concepcion ground motion). . . . . 822

xxxii
G.3 Walls exceeding fragility function probabilities of 0.50 (Concepcion ground
motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 823
G.4 Walls exceeding strain limits (Concepcion ground motion). . . . . . . . . . . . 824
G.5 Summary of walls exceeding acceptance criteria (Concepcion ground motion
scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . 828
G.6 Walls exceeding ASCE 41 rotation limits (Concepcion ground motion scaled
to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . 829
G.7 Walls exceeding fragility function probabilities of 0.50 (Concepcion ground
motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . 830
G.8 Walls exceeding strain limits (Concepcion ground motion scaled to ASCE 7
design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 831
G.9 Summary of walls exceeding acceptance criteria (Concepcion ground motion
scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . 835
G.10 Walls exceeding ASCE 41 rotation limits (Concepcion ground motion scaled
to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 836
G.11 Walls exceeding fragility function probabilities of 0.50 (Concepcion ground
motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . 837
G.12 Walls exceeding strain limits (Concepcion ground motion scaled to ASCE 7
MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 838
G.13 Summary of walls exceeding acceptance criteria (San Pedro de la Paz ground
motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 842
G.14 Walls exceeding ASCE 41 rotation limits (San Pedro de la Paz ground motion).843
G.15 Walls exceeding fragility function probabilities of 0.50 (San Pedro de la Paz
ground motion). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 844
G.16 Walls exceeding strain limits (San Pedro de la Paz ground motion). . . . . . . 845
G.17 Summary of walls exceeding acceptance criteria (San Pedro de la Paz ground
motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . 849
G.18 Walls exceeding ASCE 41 rotation limits (San Pedro de la Paz ground motion
scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . 850
G.19 Walls exceeding fragility function probabilities of 0.50 (San Pedro de la Paz
ground motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . 851
G.20 Walls exceeding strain limits (San Pedro de la Paz ground motion scaled to
ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 852
G.21 Summary of walls exceeding acceptance criteria (San Pedro de la Paz ground
motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . 856
G.22 Walls exceeding ASCE 41 rotation limits (San Pedro de la Paz ground motion
scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . 857

xxxiii
G.23 Walls exceeding fragility function probabilities of 0.50 (San Pedro de la Paz
ground motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . 858
G.24 Walls exceeding strain limits (San Pedro de la Paz ground motion scaled to
ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 859
G.25 Summary of walls exceeding acceptance criteria (Llolleo ground motion scaled
to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . 863
G.26 Walls exceeding ASCE 41 rotation limits (Llolleo ground motion scaled to
ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 864
G.27 Walls exceeding fragility function probabilities of 0.50 (Llolleo ground motion
scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . 865
G.28 Walls exceeding strain limits (Llolleo ground motion scaled to ASCE 7 design
spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 866
G.29 Summary of walls exceeding acceptance criteria (Llolleo ground motion scaled
to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 870
G.30 Walls exceeding ASCE 41 rotation limits (Llolleo ground motion scaled to
ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 871
G.31 Walls exceeding fragility function probabilities of 0.50 (Llolleo ground motion
scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . 872
G.32 Walls exceeding strain limits (Llolleo ground motion scaled to ASCE 7 MCE
spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 873
G.33 Summary of walls exceeding acceptance criteria (Matanzas ground motion
scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . 877
G.34 Walls exceeding ASCE 41 rotation limits (Matanzas ground motion scaled to
ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 878
G.35 Walls exceeding fragility function probabilities of 0.50 (Matanzas ground mo-
tion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . 879
G.36 Walls exceeding strain limits (Matanzas ground motion scaled to ASCE 7
design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 880
G.37 Summary of walls exceeding acceptance criteria (Matanzas ground motion
scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . 884
G.38 Walls exceeding ASCE 41 rotation limits (Matanzas ground motion scaled to
ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 885
G.39 Walls exceeding fragility function probabilities of 0.50 (Matanzas ground mo-
tion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . 886
G.40 Walls exceeding strain limits (Matanzas ground motion scaled to ASCE 7
MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 887
G.41 Summary of walls exceeding acceptance criteria (Valparaiso ground motion
scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . 891

xxxiv
G.42 Walls exceeding ASCE 41 rotation limits (Valparaiso ground motion scaled
to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . 892
G.43 Walls exceeding fragility function probabilities of 0.50 (Valparaiso ground
motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . 893
G.44 Walls exceeding strain limits (Valparaiso ground motion scaled to ASCE 7
design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 894
G.45 Summary of walls exceeding acceptance criteria (Valparaiso Almendral ground
motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . 898
G.46 Walls exceeding ASCE 41 rotation limits (Valparaiso Almendral ground mo-
tion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . 899
G.47 Walls exceeding fragility function probabilities of 0.50 (Valparaiso Almendral
ground motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . 900
G.48 Walls exceeding strain limits (Valparaiso Almendral ground motion scaled to
ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 901
G.49 Summary of walls exceeding acceptance criteria (Vina del Mar - Centro
ground motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . 905
G.50 Walls exceeding ASCE 41 rotation limits (Vina del Mar - Centro ground
motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . 906
G.51 Walls exceeding fragility function probabilities of 0.50 (Vina del Mar - Centro
ground motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . 907
G.52 Walls exceeding strain limits (Vina del Mar - Centro ground motion scaled
to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . 908
G.53 Summary of walls exceeding acceptance criteria (Vina del Mar - Centro
ground motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . 912
G.54 Walls exceeding ASCE 41 rotation limits (Vina del Mar - Centro ground
motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . 913
G.55 Walls exceeding fragility function probabilities of 0.50 (Vina del Mar - Centro
ground motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . 914
G.56 Walls exceeding strain limits (Vina del Mar - Centro ground motion scaled
to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 915
G.57 Summary of walls exceeding acceptance criteria (Vina del Mar - El Salto
ground motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . 919
G.58 Walls exceeding ASCE 41 rotation limits (Vina del Mar - El Salto ground
motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . 920
G.59 Walls exceeding fragility function probabilities of 0.50 (Vina del Mar - El
Salto ground motion scaled to ASCE 7 design spectrum). . . . . . . . . . . . 921
G.60 Walls exceeding strain limits (Vina del Mar - El Salto ground motion scaled
to ASCE 7 design spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . 922

xxxv
G.61 Summary of walls exceeding acceptance criteria (Vina del Mar - El Salto
ground motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . 926
G.62 Walls exceeding ASCE 41 rotation limits (Vina del Mar - El Salto ground
motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . 927
G.63 Walls exceeding fragility function probabilities of 0.50 (Vina del Mar - El
Salto ground motion scaled to ASCE 7 MCE spectrum). . . . . . . . . . . . . 928
G.64 Walls exceeding strain limits (Vina del Mar - El Salto ground motion scaled
to ASCE 7 MCE spectrum). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 929

xxxvi
LIST OF TABLES

Table Number Page


2.1 Damage states and methods of repair for structural walls by Gulec et. al.
[41][40] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.2 Damage states and methods of repair for structural walls by Brown[12] . . . . 17
2.3 Planar wall design characteristics for building inventory by Brown et al. [13] 19
2.4 Planar wall shear demands and capacities for building inventory by Brown
et al. [13] . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5 Summary of earthquakes with reinforced concrete structural wall damage. . . 21
2.6 Building damage types by region . . . . . . . . . . . . . . . . . . . . . . . . . 62
2.7 Summary of variables considered in wall test programs. Geometric, reinforce-
ment, material property and loading data for each individual specimen are
provided in Tables 2.10 - 2.12. . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
2.8 Correlation of drift capacity and design parameters. . . . . . . . . . . . . . . 138
2.9 Drift capacity, failure and code compliance for experimental test programs . . 150
2.10 Geometric and material parameters for experimental test programs . . . . . . 152
2.11 Reinforcement ratios for experimental test programs . . . . . . . . . . . . . . 154
2.12 Load and capacity parameters for experimental test programs . . . . . . . . . 156
3.1 Test matrix of planar wall experimental test program . . . . . . . . . . . . . . 161
3.2 Average material properties for self-consolidating concrete (SCC) used for
wall specimens. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
3.3 Steel Material Properties (Summary) . . . . . . . . . . . . . . . . . . . . . . . 176
3.4 Design and Expected Nominal Shear Strengths . . . . . . . . . . . . . . . . . 177
3.5 Expected moment strengths . . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
3.6 Expected moment strengths in the spliced region . . . . . . . . . . . . . . . . 179
3.7 Prescribed Displacement History . . . . . . . . . . . . . . . . . . . . . . . . . 194
3.8 PW1 Actual Displacement History . . . . . . . . . . . . . . . . . . . . . . . . 194
3.9 Summary of intended and actual loads applied to wall specimens . . . . . . . 196
4.1 Specimen PW1 applied displacement history. . . . . . . . . . . . . . . . . . . 202
4.2 Specimen PW2 applied displacement history. . . . . . . . . . . . . . . . . . . 210
4.3 Specimen PW3 applied displacement history. . . . . . . . . . . . . . . . . . . 217

xxxvii
4.4 Specimen PW4 applied displacement history. . . . . . . . . . . . . . . . . . . 224
4.5 Summary of moment demands. . . . . . . . . . . . . . . . . . . . . . . . . . . 232
4.6 Summary of critical moment demands. . . . . . . . . . . . . . . . . . . . . . . 235
4.7 Summary of shear demands. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 237
4.8 Summary of response at crack damage states. Where two step numbers are
presented, the first is the step at which response values are reported and the
number in ( ) is the step the damage occurred. . . . . . . . . . . . . . . . . . 239
4.9 Summary of response at concrete damage states. Where two step numbers
are presented, the first is the step at which response values are reported and
the number in ( ) is the step the damage occurred. . . . . . . . . . . . . . . . 240
4.10 Summary of response at rebar damage states. Where two step numbers are
presented, the first is the step at which response values are reported and the
number in ( ) is the step the damage occurred. . . . . . . . . . . . . . . . . . 241
4.11 Initial Yield of Longitudinal Reinforcement . . . . . . . . . . . . . . . . . . . 242
4.12 Initial Yield of Longitudinal Reinforcement in Foundation . . . . . . . . . . . 243
6.1 Summary of Damage States . . . . . . . . . . . . . . . . . . . . . . . . . . . . 277
6.2 Critical section extreme fiber strains when nominal compressive strain first
reached. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318
6.3 Maximum tensile strains at wall-foundation interface. . . . . . . . . . . . . . 319
6.4 Critical compressive strains at failure. Gray columns indicate the strains at
failure. White columns indicate the maximum historic strains on the opposite
side of the wall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
6.5 Shear and axial demands assuming load carried only in the compressive region.323
7.1 Summary of Database Properties . . . . . . . . . . . . . . . . . . . . . . . . . 328
7.2 Damage States and Methods of Repair . . . . . . . . . . . . . . . . . . . . . . 343
7.3 Median EDP values of damage states using full database. Outliers not removed.353
7.4 Median EDP values of damage states using the full database, with outliers
removed. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 355
7.5 Median EDP values of damage states (walls sorted by load history) . . . . . . 356
7.6 Median EDP values of damage states for reduced data set (walls sorted by
scale) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 357
7.7 Median EDP values of damage states (walls sorted by shape) . . . . . . . . . 358
7.8 Median EDP values of damage states for reduced data set (walls sorted by
shear span) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 359
7.9 Median EDP values of damage states for reduced data set (walls sorted by
axial load ratio) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360

xxxviii
7.10 Median EDP values of damage states for reduced data set (walls sorted by
shear demand) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 360
7.11 Median EDP values of damage states for reduced data set (walls sorted by
shear demand ratio) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 361
7.12 Lognormal distribution parameters for cyclically loaded walls . . . . . . . . . 364
7.13 Recommended fragility function parameters . . . . . . . . . . . . . . . . . . . 366
8.1 Effective stiffness values for elastic building models. . . . . . . . . . . . . . . . 374
8.2 m-factors for reinforced concrete shear walls (ASCE/SEI 41-06 [8, 9] Tables
6-20 and 6-21). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 379
8.3 Gravity loads used in creating the elastic building models . . . . . . . . . . . 381
8.4 Comparison of code based spectral accelerations at fundamental periods. . . . 387
8.5 Location and soil type of the ground motions used to generate site-specific
response spectra. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 388
8.6 Comparison of spectral accelerations from ground motion records at funda-
mental periods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 389
8.7 Summary of ground motions studied. . . . . . . . . . . . . . . . . . . . . . . . 393
8.8 Factors to scale ground motions to ASCE 7 DBE response spectrum . . . . . 393
8.9 Acceptance criteria for reinforced concrete shear walls (ASCE/SEI 41-06 Ta-
bles 6-18 and 6-19). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
8.10 Summary of damage levels used to describe damage to buildings. . . . . . . . 401
9.1 Summary of building drifts from Tier 3 analyses. . . . . . . . . . . . . . . . . 429
9.2 Summary of story drifts from Tier 3 analyses. . . . . . . . . . . . . . . . . . . 430
9.3 Summary of effective height ratios from Tier 3 analyses. . . . . . . . . . . . . 431
9.4 Relationship between damage levels and target performance levels. . . . . . . 434
9.5 ASCE/SEI 41-06 relationship between performance level and observed damage435
9.6 Number of walls exceeding ASCE 41 rotation limits. . . . . . . . . . . . . . . 452
9.7 Number of walls exceeding fragility function probability of 0.50. . . . . . . . . 453
9.8 Number of walls exceeding strain limits. . . . . . . . . . . . . . . . . . . . . . 454
9.9 Comparison of local and global vertical discontinuites . . . . . . . . . . . . . 465
C.1 Geometric properties for slender walls included in fragility function database 596
C.2 Loading and reinforcement properties for slender walls included in fragility
function database . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 598
C.3 Rebar Properties (* denotes assumed value) - Part A . . . . . . . . . . . . . . 600
C.4 Rebar Properties (* denotes assumed value) - Part B . . . . . . . . . . . . . . 604
C.5 Damage States and Methods of Repair . . . . . . . . . . . . . . . . . . . . . . 608

xxxix
C.6 PilaSW4 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 610
C.7 PilaSW5 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 615
C.8 PilaSW6 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 618
C.9 PilaSW7 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 621
C.10 PilaSW8 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 624
C.11 PilaSW9 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 627
C.12 TasSHW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 630
C.13 TasSHW2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 631
C.14 TasSHW3 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 632
C.15 TasSHW4 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 633
C.16 ThomRW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 634
C.17 ThomRW1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . 634
C.18 ThomRW2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 636
C.19 ThomRW2 crack width information . . . . . . . . . . . . . . . . . . . . . . . . 636
C.20 ThomTW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 638
C.21 ThomTW2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 640
C.22 DazioWSH1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . 642
C.23 DazioWSH1 crack width information . . . . . . . . . . . . . . . . . . . . . . . 643
C.24 DazioWSH2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . 644
C.25 DazioWSH2 crack width information . . . . . . . . . . . . . . . . . . . . . . . 645
C.26 DazioWSH3 damage information . . . . . . . . . . . . . . . . . . . . . . . . . 646
C.27 DazioWSH3 crack width information . . . . . . . . . . . . . . . . . . . . . . . 647
C.28 DazioWSH4 damage information . . . . . . . . . . . . . . . . . . . . . . . . . 648
C.29 DazioWSH4 crack width information . . . . . . . . . . . . . . . . . . . . . . . 649
C.30 DazioWSH5 damage information . . . . . . . . . . . . . . . . . . . . . . . . . 650
C.31 DazioWSH5 crack width information . . . . . . . . . . . . . . . . . . . . . . . 651
C.32 DazioWSH6 damage information . . . . . . . . . . . . . . . . . . . . . . . . . 652
C.33 DazioWSH6 crack width information . . . . . . . . . . . . . . . . . . . . . . . 652
C.34 LefasSW21 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 654
C.35 LefasSW21 crack width information . . . . . . . . . . . . . . . . . . . . . . . 654
C.36 LefasSW22 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 656
C.37 LefasSW23 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 657
C.38 LefasSW24 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 658
C.39 LefasSW24 crack width information . . . . . . . . . . . . . . . . . . . . . . . 658
C.40 LefasSW25 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 660

xl
C.41 LefasSW26 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 661
C.42 LefasSW26 crack width information . . . . . . . . . . . . . . . . . . . . . . . 661
C.43 LefasSW30 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 663
C.44 LefasSW30 crack width information . . . . . . . . . . . . . . . . . . . . . . . 663
C.45 LefasSW31 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 665
C.46 LefasSW31 crack width information . . . . . . . . . . . . . . . . . . . . . . . 665
C.47 LefasSW32 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 667
C.48 LefasSW32 crack width information . . . . . . . . . . . . . . . . . . . . . . . 667
C.49 LefasSW33 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 669
C.50 LefasSW33 crack width information . . . . . . . . . . . . . . . . . . . . . . . 669
C.51 OestR1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 671
C.52 OestR1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . 671
C.53 OestR2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 673
C.54 OestR2 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . 673
C.55 OestB1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 675
C.56 OestB1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 675
C.57 OestB2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 677
C.58 OestB2 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 677
C.59 OestB3 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 679
C.60 OestB3 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 679
C.61 OestB4 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 681
C.62 OestB4 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 681
C.63 OestB5 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 683
C.64 OestB5 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 683
C.65 OestB6 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 685
C.66 OestB6 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 685
C.67 OestB7 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 687
C.68 OestB7 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 687
C.69 OestB8 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 689
C.70 OestB8 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 689
C.71 OestB9 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 691
C.72 OestB9 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 691
C.73 OestB10 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 693
C.74 OestB10 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . 693
C.75 OestF1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 695

xli
C.76 OestF1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 695
C.77 OestF2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 697
C.78 OestF2 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 697
C.79 Morgan damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 699
C.80 LiuW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 703
C.81 LiuW1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 704
C.82 LiuW2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 708
C.83 LiuW2 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 709
C.84 AliW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
C.85 AliW1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 714
C.86 TupperW3 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 716
C.87 TupperW3 crack width information . . . . . . . . . . . . . . . . . . . . . . . . 716
C.88 MobeenW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . 719
C.89 MobeenW1 crack width information . . . . . . . . . . . . . . . . . . . . . . . 719
C.90 RivaW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 721
C.91 RivaW1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . 721
C.92 ShiuC1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723
C.93 KhaC1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 726
C.94 ElnCW2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 727
C.95 ElnCW3 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . 728
C.96 IleX damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 729
C.97 IleY damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 730
C.98 IleXY damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 731
C.99 TUA damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 733
C.100TUB damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 734
C.101NTW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 735
C.102NTW1 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 736
C.103NTW2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . 738
C.104NTW2 crack width information . . . . . . . . . . . . . . . . . . . . . . . . . . 739
C.105PW1 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 741
C.106PW2 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 742
C.107PW3 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 743
C.108PW4 damage information . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 744
E.1 List of wall and building characteristics resulting in structural damage. . . . . 782

xlii
G.1 Sequence of exceedence of ASCE 41 acceptance criteria (Concepcion ground
motion). Shading indicates that the wall also exceeded the acceptance crite-
rion for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 825
G.2 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Concepcion ground motion). Shading indicates that the wall also
exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . . . 826
G.3 Sequence of exceedence of steel and concrete strain limits (Concepcion ground
motion). Shading indicates that the wall also exceeded the acceptance crite-
rion for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 827
G.4 Sequence of exceedence of ASCE 41 acceptance criteria (Concepcion ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 832
G.5 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Concepcion ground motion scaled to ASCE 7 design spectrum).
Shading indicates that the wall also exceeded the acceptance criterion for shear.833
G.6 Sequence of exceedence of steel and concrete strain limits (Concepcion ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 834
G.7 Sequence of exceedence of ASCE 41 acceptance criteria (Concepcion ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 839
G.8 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Concepcion ground motion scaled to ASCE 7 MCE spectrum).
Shading indicates that the wall also exceeded the acceptance criterion for
shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 840
G.9 Sequence of exceedence of steel and concrete strain limits (Concepcion ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 841
G.10 Sequence of exceedence of ASCE 41 acceptance criteria (San Pedro de la Paz
ground motion). Shading indicates that the wall also exceeded the acceptance
criterion for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 846
G.11 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (San Pedro de la Paz ground motion). Shading indicates that the
wall also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . 847
G.12 Sequence of exceedence of steel and concrete strain limits (San Pedro de
la Paz ground motion). Shading indicates that the wall also exceeded the
acceptance criterion for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . 848

xliii
G.13 Sequence of exceedence of ASCE 41 acceptance criteria (San Pedro de la Paz
ground motion scaled to ASCE 7 design spectrum). Shading indicates that
the wall also exceeded the acceptance criterion for shear. . . . . . . . . . . . . 853
G.14 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (San Pedro de la Paz ground motion scaled to ASCE 7 design spec-
trum). Shading indicates that the wall also exceeded the acceptance criterion
for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 854
G.15 Sequence of exceedence of steel and concrete strain limits (San Pedro de la
Paz ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 855
G.16 Sequence of exceedence of ASCE 41 acceptance criteria (San Pedro de la Paz
ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that
the wall also exceeded the acceptance criterion for shear. . . . . . . . . . . . . 860
G.17 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (San Pedro de la Paz ground motion scaled to ASCE 7 MCE spec-
trum). Shading indicates that the wall also exceeded the acceptance criterion
for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 861
G.18 Sequence of exceedence of steel and concrete strain limits (San Pedro de la
Paz ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 862
G.19 Sequence of exceedence of ASCE 41 acceptance criteria (Llolleo ground mo-
tion scaled to ASCE 7 design spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . . . 867
G.20 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Llolleo ground motion scaled to ASCE 7 design spectrum). Shad-
ing indicates that the wall also exceeded the acceptance criterion for shear. . 868
G.21 Sequence of exceedence of steel and concrete strain limits (Llolleo ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 869
G.22 Sequence of exceedence of ASCE 41 acceptance criteria (Llolleo ground mo-
tion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . . . 874
G.23 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Llolleo ground motion scaled to ASCE 7 MCE spectrum). Shading
indicates that the wall also exceeded the acceptance criterion for shear. . . . 875
G.24 Sequence of exceedence of steel and concrete strain limits (Llolleo ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 876

xliv
G.25 Sequence of exceedence of ASCE 41 acceptance criteria (Matanzas ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 881
G.26 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Matanzas ground motion scaled to ASCE 7 design spectrum).
Shading indicates that the wall also exceeded the acceptance criterion for
shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 882
G.27 Sequence of exceedence of steel and concrete strain limits (Matanzas ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 883
G.28 Sequence of exceedence of ASCE 41 acceptance criteria (Matanzas ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 888
G.29 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Matanzas ground motion scaled to ASCE 7 MCE spectrum). Shad-
ing indicates that the wall also exceeded the acceptance criterion for shear. . 889
G.30 Sequence of exceedence of steel and concrete strain limits (Matanzas ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 890
G.31 Sequence of exceedence of ASCE 41 acceptance criteria (Valparaiso ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 895
G.32 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Valparaiso ground motion scaled to ASCE 7 design spectrum).
Shading indicates that the wall also exceeded the acceptance criterion for
shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 896
G.33 Sequence of exceedence of steel and concrete strain limits (Valparaiso ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear. . . . . . . . . . . . . . . . . . 897
G.34 Sequence of exceedence of ASCE 41 acceptance criteria (Valparaiso Almen-
dral ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 902
G.35 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Valparaiso Almendral ground motion scaled to ASCE 7 MCE spec-
trum). Shading indicates that the wall also exceeded the acceptance criterion
for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 903
G.36 Sequence of exceedence of steel and concrete strain limits (Valparaiso Almen-
dral ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 904

xlv
G.37 Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - Cen-
tro ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 909
G.38 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - Centro ground motion scaled to ASCE 7 design
spectrum). Shading indicates that the wall also exceeded the acceptance
criterion for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 910
G.39 Sequence of exceedence of steel and concrete strain limits (Vina del Mar -
Centro ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 911
G.40 Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - Cen-
tro ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that
the wall also exceeded the acceptance criterion for shear. . . . . . . . . . . . . 916
G.41 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - Centro ground motion scaled to ASCE 7 MCE spec-
trum). Shading indicates that the wall also exceeded the acceptance criterion
for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 917
G.42 Sequence of exceedence of steel and concrete strain limits (Vina del Mar -
Centro ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 918
G.43 Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - El
Salto ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 923
G.44 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - El Salto ground motion scaled to ASCE 7 design
spectrum). Shading indicates that the wall also exceeded the acceptance
criterion for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 924
G.45 Sequence of exceedence of steel and concrete strain limits (Vina del Mar - El
Salto ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 925
G.46 Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - El
Salto ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 930
G.47 Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - El Salto ground motion scaled to ASCE 7 MCE
spectrum). Shading indicates that the wall also exceeded the acceptance
criterion for shear. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 931
G.48 Sequence of exceedence of steel and concrete strain limits (Vina del Mar - El
Salto ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear. . . . . . . . . . 932

xlvi
ACKNOWLEDGMENTS

The research presented in Chapters 3-6 was funded by the National Science Foundation
through the Network for Earthquake Engineering Simulation Research Program, Grant
CMMI-042157 and CMMI-0927178, Joy Pauschke, program manager. Any opinions, find-
ings, and conclusions or recommendations expressed in this material are those of the authors
and do not necessarily reflect the views of the National Science Foundation.
Funding for the research presented in Chapters 8 and 9 was provided by the NEHRP Con-
sultants Joint Venture (a partnership of the Applied Technology Council and Consortium
of Universities for Research in Earthquake Engineering), under Contract SB134107CQ0019,
Earthquake Structural and Engineering Research, issued by the National Institute of Stan-
dards and Technology.
Thank you to all of my collaborators on various projects throughout the years. It was
a pleasure working with you all. At the University of Washington, Josh Pugh and Jake
Turgeon. At the University of Illinois, Chris Hart, Dr. Dan Kuchma, and Ken Marley.
From Simpson Gumpertz and Heger, Ady Aviram and Dominic Kelly.
Thank you to all my friends, office mates and fellow graduate students throughout the
six plus years at the University of Washington graduate school for making my time in Seattle
so wonderful.
Thank you to my reading committee, Marc Eberhard and Greg Miller, for taking the time
to read and provide feedback. Thank you also to Ken-Yu Lin and Joyce Cooper for serving
as the graduate school representative for my general exam and final exam, respectively.
Thank you to my advisors, Laura Lowes and Dawn Lehman. I could not have done
this without either of you. The opportunities you gave me and your advice, guidance, and
encouragement have allowed me to achieve something I never dreamed of. I will be forever
grateful.

xlvii
And above all, to my parents, thank you for your never ending support and encourage-
ment. I love you.

xlviii
1

Chapter 1

INTRODUCTION

The 2010 Maule earthquake in Chile resulted in severe damage to concrete walls in
numerous buildings; some walled buildings collapsed partially or completely. This wide-
spread damage to concrete walls has left some structural engineers questioning the design
of concrete walls for earthquake loading in the United States as well as the earthquake
performance that can be expected from walls designed in compliance with codes in the
United States.
Reinforced concrete structural walls are one of the most commonly employed lateral-load
resisting systems for mid- and high-rise buildings. They are relatively stiff and strong; thus,
their use results in relatively less floor area of the building being reserved for lateral-load
resisting structural elements. Additionally, they are easily incorporated into the architec-
tural layout of mid- and high-rise buildings by placing them in the interior of the buildings
around elevators. Finally, it has been the consensus of the earthquake engineering commu-
nity that well-designed and correctly detailed reinforced concrete walls provide earthquake
performance that is superior to many other lateral load-resisting systems.
Given the high reliance by engineers on reinforced concrete walls and, perhaps, because
of the earthquake engineering community’s confidence in their superior earthquake per-
formance, there have been relatively few experimental tests investigating the earthquake
performance of modern, code-compliant structural walls. As such there are relatively few
experimental data sets available in the literature that provide, for modern walls, compre-
hensive response data for walls subjected to cyclic loading or comprehensive performance
data. Response data are required to validate numerical models for use in design and re-
search; performance data are required to develop damage-prediction models, often referred
to as fragility functions, for performance-based earthquake engineering of walled buildings.
To address the need for improved understanding of the earthquake response of modern
2

walls as well as advance seismic design of these systems, a research program was proposed
by faculty at the Universities of Washington (UW), Illinois Urbana-Champaign (UIUC),
and California, Los Angeles (UCLA) and funded by the National Science Foundation (NSF)
through the Network for Earthquake Engineering Simulation (NEES) Research program and
by the Charles Pankow Foundation (CPF). This research program includes experimental
testing of planar rectangular, planar coupled, and C-shaped wall subassemblages; evaluation,
development, and validation of numerical response models for walls, and development of
recommendations and the modeling tools required for traditional and performance-based
seismic design of structural walls.
A major component of the research presented here represents a part of the larger effort
by researchers at the Universities of Washington, Illinois, and California. The research
presented here primarily addresses the earthquake response of slender planar rectangular
walls. It seeks to use data from the laboratory tests conducted as part of the larger effort to
improve understanding of the earthquake response and performance of planar rectangular
walls, develop the data required to advance response and performance-prediction models
for these walls, develop recommendations and fragility functions for use in performance-
based earthquake engineering of walls, and validate these through design and assessment of
concrete-walled buildings.
In addition to the need for appropriate experimental data and performance-based de-
sign tools, engineers require tools for the evaluation of existing structures to determine the
seismic adequacy of the full building and/or specific building components. For this pur-
posed, the American Society of Engineers (ASCE) and the Structural Engineering Institute
(SEI) developed the standard “Seismic Evaluation of Existing Structures”, referred to more
commonly as ASCE/SEI 31-03 [7]. This standard offers three levels, or tiers, of evaluation,
ranging from quick check requiring minimal (or no) calculations to nonlinear analysis of a
full building. A related standard is ASCE/SEI 41-06 [8, 9]. The guidelines of ASCE/SEI
41-06 overlap with and expand on those of ASCE/SEI 31-03. Damage to buildings in the
2010 Maule earthquake provided a unique opportunity to study buildings subjected to major
earthquake demands and to evaluate the adequacy of the ASCE/SEI 31-03 and ASCE/SEI
41-06 guidelines. Consequently, an additional component of the work presented in this dis-
3

sertation focuses on the study of four buildings damaged in the Maule earthquake. The
project was sponsored by the NEHRP (National Earthquake Hazards Reduction Program)
Consultants Joint Venture (a partnership between the Applied Technology Council (ATC)
and the Consortium of Universities for Research in Earthquake Engineering (CUREE)) and
was conducted in collaboration with practicing engineers.

1.1 The Earthquake Response and Design of Reinforced Concrete Walls

Under earthquake loading, reinforced concrete walls carry flexural, shear and axial loads. If
walls have an aspect ratio (AR), defined as the ratio of the wall height to wall length, that is
small (less than 2), they are typically referred to as squat. Squat walls typically carry rela-
tively higher shear stress demands and exhibit shear-dominated response. Shear-dominated
response would include shear deformation accounting for a significant portion of the total
deformation, development of diagonal cracks through the web and boundary regions of the
wall, development of diagonal or horizontal zones of concrete crushing within the web that
may extend into the boundary regions of the wall, and possibly not achieving the nominal
flexural strength of the wall. Squat walls typically exhibit lower ductility capacity. If walls
have an AR that is large (greater than 2), they are typically referred to as slender. Slender
walls typically carry relatively lower shear stress demands and exhibit flexure-dominated
response. Flexure-dominated response would include flexural deformation accounting for a
relatively large portion of the total deformation, development of horizontal flexural cracks
in the boundary regions of the wall, development of diagonal shear cracks within the web
of the wall, loss of lateral load carrying capacity due to compressive failure of a boundary
region (crushing of core concrete and buckling of longitudinal steel) or due to fracture of
longitudinal steel. Typically, the axial load in a wall due to gravity loading is small (less
than 10% of the gross capacity); however is a coupled-wall system in which earthquake
overturning moment is resisted by flexure in individual wall piers as well as through tension
and compression forces in the wall piers, axial loads may increase significantly (typically up
to 40% of the gross capacity).
The earthquake response of slender concrete walls has been studied in the laboratory
and in the field. The typical crack patterns indicate boundary regions on the edges of
4

respond in a flexural manner (horizontal cracks) and the web of the wall responds in shear
(inclined cracks). The severe damage observed in the field following earthquakes is primarily
flexure, in the form of core crushing, bar buckling, and sometimes bar fracture; however
large diagonal cracks/damage planes and web crushing have been observed. Observed failure
mechanisms in the lab generally reflect that seen in the field, with compressive failures of the
boundary elements occurring most frequently. Observations from experimental research has
led to the current practice of providing well confined boundary elements to ensure ductile
behavior of walls.
Flexure, shear and axial demands are used to design the geometry of the wall as well
as the layout of longitudinal reinforcing steel, horizontal reinforcing steel, and transverse,
confining steel placed in the boundary regions of the wall. In the US, walls design is typically
done in accordance with the American Concrete Institute Building Code Requirements for
Structural Concrete and Commentary (ACI 318 2008), referred to as the ACI Code. In the
US, shear demand on the wall typically determines the gross area of the wall, as the ACI
p
Code requires that the shear strength of a wall not exceed 8 fc0 Acv where fc0 is the concrete
compressive strength in psi and Acv is the area, in square inches, of the wall that is expected
to carry shear. Horizontal reinforcement is designed (bar size and spacing) on the basis of
the shear demand. Longitudinal reinforced is designed (bar size and spacing) on the basis of
the flexural and axial load demands. Often longitudinal reinforcement is concentrated at the
ends of a planar wall (corners of a c-shaped wall), with the ACI Code minimum longitudinal
reinforcement distributed within the interior, or web, of the wall; this reinforcement layout
is more efficient for planar walls. Finally, for walls in regions of high seismicity, transverse
reinforcement is designed to confine concrete and delay buckling of longitudinal steel at the
ends of a planar wall (regions expected to experience high compressive strain demands for
non-planar walls). To reduce the likelihood of walls exhibiting less ductile, shear-dominated
failure; significantly lower resistance factors are defined for shear strength than for flexural
strength. Additional ACI Code requirements address minimum longitudinal and horizontal
reinforcement ratios as well as specifications for detailing of confined boundary regions, with
the intent of ensuring that walls exhibit acceptable performance under service-level loads
and in the event of design-level earthquake loading.
5

Current ACI Code design requirements for walls in regions of high seismicity are in-
tended to ensure that walls exhibit ductile, flexure-dominated response under earthquake
loading. However, as previously stated, relatively few experimental tests of modern ACI
Code-compliant walls have been conducted to quantify the earthquake performance of mod-
ern walls and to verify that modern design procedures produced the intended earthquake
response.

1.2 Objectives

To advance the knowledge of the earthquake performance and design of slender reinforced
concrete walls, the following objectives were established for this research effort.

1. Establish the current state of knowledge and design by reviewing i) the state of practice
of wall design, ii) damage observed to structural walls following earthquakes, and iii)
experimental tests of wall specimens.

2. For the four planar walls tested as part of the UW-UIUC-UCLA NEESR Project i)
document the damage progression, and ii) analyze experimental data and iii) establish
the design parameters that determined performance and failure mechanism.

3. Use experimental data from laboratory tests to develop fragility functions for slender
walls.

4. Study buildings damaged during the 2010 Maule earthquake through evaluation of
fragility functions and the provisions of ASCE/SEI 31-03 and ASCE/SEI 41-06.

5. Provide recommendations for improvement to future versions of the ASCE/SEI 41-06


standard.

1.3 Outline of Document

Chapter 2 presents information establishing the current state of knowledge of the earthquake
performance and design of walls. The current state of practice for the design of slender
6

reinforced concrete structural walls is reviewed. A comprehensive review of earthquake


damage to structural walls in buildings is presented. Previous research addressing slender
structural walls is summarized. A discussion on the failure modes and impact of wall design
parameters on wall performance follows.
Chapter 3 presents a brief overview of the planar wall test program for the UW-UIUC-
UCLA NEESR project. Details are presented in a document co-authored by all members
of the project team [63]. Included is a summary of the load conditions and data collection,
a presentation of the load-displacement response of the specimens and a summary of the
progression of observed damage.
Chapter 4 presents an overview of the planar wall specimen tests, including load-
deformation response and damage progression. Each test is summarized individually, fol-
lowed by a comparison of the four specimens and a discussion on the impact of the test
program variables.
Chapter 5 summarizes the methods used to analyze the data collected from instruments
during the planar wall tests. This includes evaluation of the yield behavior of the spec-
imens, calculation of strain fields from measured displacements, and calculation of shear
and deformation components of the test. The results of the data analysis are presented in
Chapter 6.
Chapter 7 presents fragility functions developed for ATC-58. Fragility functions were
developed using experimental data from test programs summarized in Chapter 2 and the
test program summarized in Chapters 3-6.
Chapters 8 and 9 presents results of a study of buildings damaged in the 2010 Maule
earthquake. The study involved performing the three tiered evaluation procedures of
ASCE/SEI 31-03 and comparing to the observed performance of the buildings.
Chapter 10 provides a summary of the dissertation.
7

Chapter 2

LITERATURE REVIEW

This chapter presents a review of the seismic design and performance of slender rein-
forced concrete walls. Section 2.1 presents an overview of the design provisions of ACI
318-08. Section 2.2 provides an overview of performance-based design methods, with an
emphasis on the development of fragility functions. In Section 2.3, construction drawings
for a collection of modern structural walls in buildings on the west coast are reviewed.
In Section 2.4 a review of observed damage to walls following earthquakes is presented.
Section 2.5 presents a review of previous experimental tests on slender reinforced-concrete
walls, with an emphasis on those tests for which significant damage data are available. In
Section 2.6, the previous research is collectively summarized by considering compliance with
the ACI 318 building code, similarity to the building inventory, similarity to damage ob-
served following earthquakes, and the impact of test variables on wall performance. Finally,
Section 2.7 presents a summary of the literature review.

2.1 Code-Based Design

The American Concrete Institute (ACI) Building Code Requirements for Structural Con-
crete (ACI 318-08) and Commentary (ACI 318R-08)[1], subsequently referred to as ACI
318-08, requires structural walls to resist shear, moment and axial loads. ACI 318-08 re-
quires that structures be assigned a Seismic Design Category (SDC) according to ASCE/SEI
7-2005. For structures with SDC D, E, or F, structural walls must be designed with seismic
detailing in accordance with Chapter 21.
The intent of Chapter 21 provisions is to ensure that structures subject to earthquake
induced forces are capable of sustaining multiple oscillations in the inelastic range of re-
sponse. When members are properly detailed, they will be able to maintain strength during
these oscillations but will have decreased stiffness and increased energy dissipation.
8

The following sections summarize the ACI 318-08 requirements for the design of RC
walls in SDC D, E, and F.

2.1.1 Design Loads and Strength Reduction Factors

Walls must be designed for the strength requirements provided in ACI 318-08 Chapter 9.
The design strength is equal to the nominal strength multiplied by a strength reduction
factor, φ, and must be equal to or greater than the required strength.
The required strength, U , of is determined from factored load combinations1 . For seismic
design, the load combination U = 1.2D + 1.0E + 1.0L + 0.2S must be considered, where D is
the dead load, L is the live load, S is the snow load, and E is the load effects of earthquakes.
The value of E can be determined by i) elastic analysis, ii) ASCE 7 [6] Equivalent Lateral
Force (ELF) procedure, or iii) response spectrum analysis.
The strength reduction factor, φ, is intended to account for uncertainties in the material
properties, inaccuracy of design equations, the importance of the member, and the duc-
tility/reliability of the member under the force in question2 . For special structural walls,
the strength reduction factor for shear is equal to 0.603 . The strength reduction factor for
flexure ranges from 0.65 for compression-controlled members to 0.90 for tension-controlled
members4 . The member is compression-controlled if the extreme tensile steel strain, εt , is
less than or equal to 0.002 at the nominal strength of the member. The member is tension-
controlled if εt is greater than or equal to 0.005 at nominal strength. The nominal strength
of the member is defined as that when the extreme compression fiber strain is equal to
−0.003.
1
9.2.1
2
R9.3.1
3
9.3.4
4
9.3.2
9

2.1.2 Material properties

Concrete compressive strength, fc0 , must be greater than or equal to 3000 psi (20.7 MPa) for
normal weight concrete5 and 5000 psi (34.5 MPa) for lighweight concrete6 . Reinforcement
is required to satisfy ASTM A706, with exceptions permitted for ASTM A615 Grades 40
and 60 if the actual yield strength does not exceed the specified yield strength by more than
18 ksi (124 MPa) and the ratio of actual tensile strength to the actual yield strength is at
least 1.257 . The value used in designing shear and confining reinforcement cannot exceed
60 ksi (413 MPa) and 100 ksi (689 MPa), respectively.8

2.1.3 Shear Strength and Distributed Web Reinforcement

The ratio of distributed horizontal, ρh , and longitudinal steel in the web, ρweb , of the wall
to the area of concrete must be at least 0.25% with spacing not exceeding 18 inches (45.7
cm) 9 and must be provided in at least two curtains10 .
The nominal shear strength of a wall is defined by Equation (2.1)11 , where αc is equal
to 2.0 for slender walls with aspect ratio (AR) greater than or equal to 2.0; λ is a factor
accounting for lightweight concrete (1 for normal weight concrete); and Acv is the area of
p
wall resisting the shear force12 . The nominal shear strength cannot exceed 8Acv fc0 13 .

 p  p
Vn = Acv αc λ fc0 + ρh fy ≤ 8Acv fc0 (2.1)

5
21.1.4.2
6
21.1.4.3
7
21.1.5.2
8
21.1.5.4,21.1.5.5,11.4.2
9
21.9.2.1
10
21.9.2.2
11
21.9.4.1
12
Equal to the gross area for rectangular walls and the area bound by the length and thickness of the
web/flange resisting shear in flanged walls.
13
21.9.4.4
10

2.1.4 Boundary Elements

The need for special boundary elements is determined using one of the following approaches:

1. Special boundary elements must be provided if the neutral axis depth, c, is equal to or
exceeds `w /(600δu /hw ) where δu is the design displacement of the wall with height hw ,
and the ratio of δu /hw is greater than or equal to 0.007.14 Boundary element confining
reinforcement must be extended vertically from the critical section a distance not less
than the larger of `w (wall length) and Mu /(4Vu ), Mu and Vu are the required moment
and shear strengths of the wall, respectively. This procedure is often referred to as
displacement-based design.

2. Special boundary elements must be provided where extreme fiber compressive stresses
exceed 0.2fc0 . The special boundary element detailing (or confining reinforcement) may
be discontinued where compressive stresses are less than 0.15fc0 . This criterion15 is
intended for walls that have a single critical section and are continuous and is believed
to provide a conservative design over that of displacement-based design.

The in-plane length of special boundary elements must extend at least c − 0.1`w and
c/2 for the load conditions associated with the calculation of c.16 For flanged walls, the
boundary element must contain at least 12 inches (30.5 cm) into the web and the effective
flange width. The specimen boundary element transverse reinforcement in the region must
meet the following requirements17 :

1. Extend into the support a minimum of the length of the development length of the
largest bar confined by the transverse reinforcement.

2. Provide adequate length to develop the yield strength of horizontal reinforcing bars
anchored in the boundary element.
14
21.9.6.2
15
21.9.6.3
16
21.9.6.4(a)
17
21.9.4.(b)-(e), 21.6.4.2-21.6.4.4
11

3. Transverse reinforcement must be spiral, circular or rectilinear hoops. Cross-ties are


not required and can be a smaller diameter than the hoops, but must engage a pe-
ripheral longitudinal bar. The in-plane spacing of hoops and cross-ties, hx , cannot
exceed 14 inches (35.6 cm).

4. The vertical spacing of the transverse reinforcement, s, must satisfy the following:

(a) Less than 1/3 the smallest dimension of the boundary element

(b) Less than 6 times the diameter of the smallest longitudinal bar

(c) Less than so = 4 in. + (14 in. − hx )/3 where 4 in. ≤ so ≤ 6 in. and hx (horizontal
spacing between cross-ties) is in inches.

5. The volumetric ratio of transverse steel to concrete must be at least ρs = 0.12fc0 /fyt
for circular and spiral reinforcement.

6. The cross-sectional area of rectangular hoop reinforcement, Ash must be at least


0.09sbc fc0 /fyt .

If special boundary elements are not required but the longitudinal reinforcement ratio of
the boundary, ρbe , is greater than 400/fy (fy in ksi), transverse reinforcement is required at
a spacing of no more than 8 inches (20.3 cm) and must extend the same length as specified
for special boundary elements.18

2.1.5 Splices

Lap splices are permitted in structural walls if i) the length is sufficient to develop the yield
strength of the steel in tension19 , and ii) the yield strength is multipled by a factor 1.25
if the splice is located in a region of the wall where yielding is expected.20 Welded splices
are not permitted within a distance equal to two times the member depth where yielding is

18
21.9.6.5
19
21.9.2.3
20
21.9.2.3(c)
12

likely to occur21 . Mechanical splices are permitted in special structural walls under certain
conditions22 .

2.1.6 Considerations for Bi-Directionally Loaded Walls

If walls are designed to resist loading in two orthogonal directions, effective flange widths
extend from the face of the web to the smaller of one-half the distance to another wall web
or 1/4 the wall height23 .

2.2 Performance-Based Seismic Design of Walls

The benefits of Performance-Based Seismic Design (PBSD) are indicated by the name of
the concept itself – an engineer is able to design a structure to meet specific performance
objectives given an expected level of earthquake demand. Performance objectives may
be damage related (“cracking” or “severe structural damage”) or functional (“immediate
occupancy” or “life safety”) or economic (“losses less than than 20% of replacement value”).
Because earthquake demands as well as the expected performance given a specific demand
are uncertain, PBSD is typically done in a probabilistic manner such that a structure is
designed to achieve low annualized risk of exceeding the damage or loss associated with
the performance objective (e.g. less than a 10% probability of damage exceeding cracking).
Consequently, PBSD can be used to provide a structure with an improved performance, as
evaluated by the specific desires of the buildings owners/occupants, over that of a building
designed strictly per the provisions of a building code such as ACI 318-08.
The development of PBSD occurred in the 1990’s as a result of interaction between
building owners and engineers to improve the seismic performance of existing structures.
Procedures for PBSD are documented by such publications as NEHRP Guidelines for the
Seismic Rehabilitation of Buildings (FEMA Publication 273)[35] and FEMA 356: Pre-
standard and Commentary for the Seismic Rehabilitation of Buildings[36], precursors to
the current ASCE/SEI Standard 41 Seismic Rehabilitation of Existing Buildings.

21
21.1.7
22
21.1.6
23
21.9.5.2
13

FEMA 445 outlines concerns with the methods of FEMA 356 and ASCE 41. Among
these concerns are i) how well models simulate actual building response, ii) ability of design
engineers to easily apply design procedures, iii) questions about how conservative accep-
tance criteria are, and iv) ability to communicate PBSD beyond the engineering commu-
nity. To address these concerns, the ATC 58 project titled Development of Next-Generation
Performance-Based Seismic Design Guidelines for New and Existing Buildings [5] was
funded by FEMA through NEHRP. A key component of the ATC 58 project is the de-
velopment of fragility functions.
In the following section, two levels of PBSD are discussed. Section 2.2.1 discusses capac-
ity based design recommendations for structural walls provided by the IBC (International
Building Code) 2006 Seismic Design Manual [47]. Section 2.2.2 discusses fragility functions
that relate the probability of achieving specific performance states to engineering demand
parameters.

2.2.1 IBC Seismic Design Recommendations

The IBC 2006 Seismic Design Manual [47] provides guidance for the design of structures
subject to earthquake loading. For the design of structural walls to resist shear load, the
manual recommends that the ACI 318 requirements be considered a minimum design re-
quirement as it will not satisfy the requirement that “the design shear strength φVn shall
not be less than the shear associated with the development of the nominal flexural strength
of the wall.” [47]. As an alternative design method, the manual recommends that an am-
plified shear demand be considered. The shear demand can be considered as either i) that
corresponding to the nominal flexural strength or ii) that corresponding to the probable
flexural strength, Mpr , which is the nominal moment calculated using a steel yield strength
equal to 1.25fy . By designing the wall to have shear strength greater than or equal to that
of the flexural strength, performance based seismic design is achieved in that the controlling
strength of the wall is flexural. This will result in a more ductile response and will ensure
that a shear failure is prevented.
14

2.2.2 Fragility Functions

Fragility functions are functional relationships that define the likelihood that, given a spe-
cific level of earthquake demand as defined by an engineering demand parameter (EDP),
a specific method of repair (MOR) will be required to restore a structural component to
pre-earthquake condition. The required MOR is determined by association with a damage
states (DS) that would trigger the need for a specific level of repair.
Fragility functions for reinforced concrete structural walls have been the focus of two
studies. Gulec et al. [40, 41] developed fragility functions for squat walls. Brown [12]
developed fragility functions for squat and slender walls. A summary of the results of these
studies are provided in the following sections.

2.2.2.1 Fragilities for Squat Walls

Gulec et al. [40, 41] developed fragilities for squat reinforced concrete walls with aspect
ratio less than or equal to 2.0, where aspect ratio is defined at the ratio of the wall height,
hw , to the wall length, `w . Fifty-one specimens had a rectangular cross-section, 32 had a
barbell cross section, and 28 had a flanged cross-section.
Table 2.1 lists the damage states and methods of repair used in the study. Four methods
of repair were specified. The first, cosmetic repair, is associated with initial cracking. The
second, epoxy-resin injection of cracks, is associated with initial yielding and cracks less
than 0.12 inches (3.0 mm) but greater than 0.02 inches (0.5 mm). The third, partial wall
replacement, is associated with concrete crushing (toe or web) and bar buckling. The forth,
full wall replacement, is associated with sliding shear, widespread concrete crushing, and
bar fracture.
The engineering demand parameters considered were i) drift, ii) dissipated hysteretic
energy, and iii) a functional form of maximum story drift and number of load cycles [14],[82];
ultimately drift was used to develop the fragility functions. The impact of wall geometry,
aspect ratio, reinforcement ratios, and axial load on damage progression and the fragilities
were considered, but only wall geometry was found to significantly affect the progression of
damage. Separate sets of fragilities were developed for walls with rectangular, barbell, and
15

Table 2.1: Damage states and methods of repair for structural walls by Gulec et. al. [41][40]

Damage States Method of Repair


DS1 - Initial cracking and crack MOR-1 - Cosmetic repair of sur-
widths less than 0.02 inches (0.5 mm) face finishes to restore aesthetic ap-
pearance, maintain first resistance and
prevent water infiltration into wall
DS2 - Initiation of yielding in rein- MOR-2 - Epoxy-resin injection of
forcement (boundary element, web, or cracks to restore component stiffness
horizontal) and cracks less than 0.12 and strength
inches (3.0 mm)
DS3 - Concrete crushing at toe or in MOR-3 - Partial wall replacement
web; Vertical cracking in toe; Buckling requiring removal and replacement of
of longitudinal reinforcement; Flex- damaged concrete
ural cracks exceeding 0.12 inches (3
mm)
DS4 - Sliding at base of wall; Wide MOR-4 - Wall replacement
diagonal cracks; Widespread con-
crete crushing; Reinforcement frac-
ture; Shear cracks exceeding 0.12
inches (3 mm)

flanged cross sections.


For wall tests with multiple occurrences of a damage state, all were recorded. Fragilities
were developed twice, once using all occurrences of a damage state and once using only
the first occurrence of a damage state. Minimal differences were observed in the fragilities.
The fragilities were created using the lognormal distribution and the Method of Maximum
Likelihood; they were evaluated using the Lilliefors test. Figure 2.1 shows the fragility
functions developed for rectangular squat walls by Gulec et al. [40, 41].

2.2.2.2 Fragilities for Squat and Slender Walls

Brown [12] developed fragilities for walls without consideration for slenderness. Only walls
with rectangular- and barbell-shaped cross-sections and that represented modern construc-
tion for zones of high seismicity were used. This resulted in a data set of 45 tests from 10
2a 0.167 0.500 A 0.110
Flanged 2b 0.324 0.194 A 0.270
3 0.176 0.500 A 0.109
4 0.197 0.500 A 0.084
Table 16 Lilliefors results for the lognormal distribution computed using Method 2
16

Figure 25 Method 2 fragility functions for rectangular walls


Figure 2.1: Fragility functions developed by Gulec et al. [40, 41] for rectangular squat walls.

54
research programs.
Table 2.2 provides a summary of the damage states and methods of repair used. Five
methods of repair were specified. The first cosmetic repair, is associated with initial hor-
izontal cracks. The second, epoxy injection of cracks, is associated with initial diagonal
cracks and initial yield of reinforcement. The third method of repair, concrete patching, is
associated with initial spalling of the cover concrete. The forth, replacement of concrete, is
associated with crushing of the concrete in the web. The fifth method of repair, replacement
of steel, is associated with bar buckling and damage resulting in loss of lateral strength of
at least 20%.
Brown considered the impact of the following engineering demand parameters (EDPs):
i) story drift, ii) number of load cycles, iii) displacement ductility, and iv) plastic rotation.
Drift was identified as the best predictor of damage progression and was used to develop
17

Table 2.2: Damage states and methods of repair for structural walls by Brown[12]

Damage States Method of Repair Repair


DS.0 - First recorded MOR.0 - Cosmetic Repair Replace and repair finishes
horizontal crack
DS.1 - First recorded Inject cracks with epoxy and
MOR.1 - Epoxy injection
diagonal crack replace finishes
DS.2 - Recorded
and/or measured yield
of extreme reinforce-
ment
DS.3 - Initial spalling MOR.2 - Patching Patch spalled concrete, epoxy
of concrete cover inject cracks and replace fin-
ishes
DS.4 - Crushing of web MOR.3 - Replace concrete Remove and replace damaged
concrete concrete, replaces finishes
DS.5 - Extreme dam- MOR.4 - Replace steel Replace damaged reinforcing
age, including buck- steel, remove and replace con-
ling of longitudinal re- crete, and replace finishes
inforcement or damage
resulting in reduction
of lateral strength by
20%

fragility functions. The impact on damage progression of design parameters such as shape,
aspect ratio, and shear demand-capacity ratio were evaluated. Of the parameters considered,
only shape was found to impact damage progression, and in particular only the median
drifts at which web crushing and extreme damage occurred. Separate fragilities were not
developed due to insufficient data.
Fragility functions were developed using a lognormal distribution and the Method of
Maximum Likelihood. Goodness-of-fit testing was done using the Lilliefors test. Figure 2.2
shows the final fragility functions developed by Brown [12].
..g
(\) , / / ' - _ - - ----- -----

18
118

Shear Walls

J -----

0 0. It-J
/ I / ,J

:~ 0. , I

~ 0.

p:: 0. c-I

Drift (%)

Figure 2.2: FragilityFigure


functions
5.4 developed by Brownfor
Fragility functions [12] for reinforced
structural walls. concrete walls.

Table 5. 6 shows the results of the hypothesis test. From the data in this table , it
2.3 Building Inventory Review
may be concluded that the lognormal distribution is acceptable for use in modeling.

Brown et al. [13] assembled a building inventory to determine characteristics of modern


According to the Lilliefors test the lognormal distribution is acceptable for all but
mid-rise buildings with structural walls as the primary lateral load resisting system. The
MORinventory
building 0; damage
wasdata
doneassociated
in the earlywith MOR
stages 0 experimental
of the was difficult testing
to consistently
program record due
discussed
in Chapter 3. of
to the lack Building plans information
crack width for 12 mid-rise structural
presented wall
in the buildings
wall reports. on the West Coast
of the United States were obtained from four design firms. Building heights ranged from
5 to 30 stories. All buildings were designed after 1991 and followed the provisions of the
1991 Uniform Building Code, the 1994 Uniform Building Code, the 1997 Uniform Building
Code or the 1998 Canadian Building Code. All wall configurations were considered, and
47 individual structural walls were identified, including 18 rectangular, 3 barbell, 10 C-
shaped, 8 L-shaped, 2 H-shaped, and 6 “box-shaped” walls. The geometry, reinforcement
distribution, and material properties were recorded for each wall.
Table 2.3 presents statistics for wall geometric characteristics that could be expected to
affect earthquake response of walls. These include wall thickness, tw ; length, `w ; boundary
19

element length, `be as a ratio of total wall length; longitudinal boundary region reinforce-
ment ratio, ρbe ; longitudinal web reinforcement ratio, ρweb ; total longitudinal reinforcement
ratio ρv ; and horizontal reinforcement ratio, ρh . For walls with a uniform distribution of
reinforcement, boundary element reinforcement ratios were not provided and the web and
total longitudinal reinforcement ratios were equal.

Table 2.3: Planar wall design characteristics for building inventory by Brown et al. [13]

Characteristic Min. Value Avg. Value Max. Value


tw , in 12 23.3 30
`w , ft 4.3 24.3 44.5
`be /`w , % 9.6 20.7 34.0
ρbe , % 1.54 2.98 3.97
ρweb , % 0.24 0.46 0.99
ρv , % 0.31 0.91 1.81
ρh , % 0.24 0.43 1.38

Table 2.4 presents statistics for shear capacity, Vn , and estimated shear demand, Vmn . Vn
was calculated using the ACI 318 equation for nominal shear strength. Vmn was estimated
as the shear demand at the nominal moment strength of the cross section assuming effective
heights, hef f of 1/2hw , 2/3hw , and hef f for an ASCE 7 lateral load distribution, where hw
is the wall height. Data for hef f = 2/3hw are presented in Table 2.4.
Results of the building inventory region by Brown et al. are presented in Section 2.6.2,
where the characteristics of the walls from the building inventory review are compared with
the characteristics of experimental test specimens.

2.4 Wall Damage Observed Following Earthquakes

Wall damage observed following earthquakes is valuable for comparison to experimentally


observed damage. To provided this comparison a review of the literature was done to
establish a qualitative understanding of earthquake damage. Earthquake reconnaissance
reports and journal articles published after earthquakes typically provide summaries of the
performance of all structure types following a specific earthquake. As such, the discussion
20

Table 2.4: Planar wall shear demands and capacities for building inventory by Brown et al.
[13]
p p
Vn /Acv fc0 Vmn /Acv fc0 Vmn /Vn
Min. 3.81 1.33 0.25
Avg. 5.68 2.77 0.50
Max. 13.71 4.67 1.00

of wall damage is often limited to a few sentences. A few documents specifically address
reinforced concrete systems [29, 52, 66, 67, 94, 95, 98, 69] or structural walls [91, 117, 116].
All of these documents provided a limited understanding of earthquake performance of
structural walls, in that they described the performance of walls designed and constructed in
accordance with local building codes and standard construction practices and were subjected
to specific earthquake demands. However, no documentation was found that provided a
global overview of the performance of structural walls.
In reviewing the literature, the following information was recorded: i) earthquake name,
date, and location, ii) number of buildings for which damaged structural walls were reported,
iii) building height and/or number of floors (if available), iv) type of damage observed, and
v) classification of failure mode. A summary of the types of damage observed will provide a
useful comparison to experimentally observed damage when developing performance based
design tools.
This section is organized as follows: Section 2.4.1 provides an overview of the earthquakes
for which damage to structural walls was observed, including the i) magnitude, location,
and date of the earthquake, ii) sources from which damage information was collected, and
iii) summary of reported damage to structural walls. Section 2.4.2 provides an overview
of the types of damage observed in structural walls, accompanied by photos collected from
the literature. Section 2.4.3 provides a discussion of the observed failure mode in structural
walls, including correlation with building height and region.
21

2.4.1 Earthquakes

Damage data from previous earthquakes were reviewed to evaluate the damaged to re-
inforced concrete structural walls. The following paragraphs provide an overview of the
earthquakes for which damaged walls were found. Table 2.5 provides a summary of the
earthquakes for which reinforced concrete structural wall damage was found in the litera-
ture, including the number of buildings with observed wall damage.

Table 2.5: Summary of earthquakes with reinforced concrete structural wall damage.

Earthquake Location Date # Buildings1 Sources


Mexico 1957 Mexico City, Mexico 7-28-1957 2 (0) [74, 76]
Chile 1960 Valdivia, Chile 5-22-1960 5 (4) [74]
Hollister Hollister, California, USA 4-8-1961 1 (1) [74]
Alaska Anchorage, Alaska, USA 3-27-1964 10 (7) [10, 74, 112]
Fremont Fremont, California, USA 9-18-1965 1 (0) [74]
San Fernando Los Angeles, California, USA 2-9-1971 5 (5) [57, 74]
Nicaragua Managua, Nicaragua 12-23-1972 1 (1) [74]
Japan 1978 Sendai City, Japan 6-12-1978 7 (7) [31]
Chile 1985 Vina del Mar, Chile 3-3-1985 3 (3) [74, 117, 91]
Mexico 1985 Mexico City, Mexico 9-19-1985 4 (4) [16]
Armenia Armenia, Soviet Union 12-7-1988 1 (1) [118]
Loma Prieta San Francisco, California, USA 10-17-1989 3 (2) [33]
Turkey 1992 Erzincan, Turkey 3-13-1992 2 (2) [93, 99]
Guam Guam 8-8-1993 3 (3) [21]
Northridge California, USA 1-17-1994 6 (6) [71, 105, 67, 74]
Kobe Kobe, Japan 1-17-1995 12 (8) [96, 20, 19, 66, 119, 74]
Turkey 1999 Kocaeli (Izmit), Turkey 8-17-1999 10 (0) [94, 98, 74, 95, 97]
Chi-Chi Chi-Chi, Taiwan 9-21-1999 1 (0) [60]
Bingöl Bingöl, Turkey 5-1-2003 1 (0) [52, 29]
Nisqually Olympia, Washington, USA 2-28-2001 2 (0) [30, 111]
Japan 2007 Niigataken Chuetsu-oki, Japan 7-15-2007 1 (1) [119]
Chile 2010 Chile 2-27-2010 14 (13) [74, 69, 46]
Unknown Unknown Unknown 2 (0) [51]
1
# Buildings Total (# Buildings of Known Height)

2.4.1.1 Chile 1960

The 9.5 magnitude (Richter scale) Valdivia, Chile earthquake occurred May 22, 1960. No
reports were found documenting damage to buildings with structural walls; however, some
images of damage and associated information were obtained from the NISEE image database
[74].
22

2.4.1.2 Alaska 1964

The 8.4 magnitude (Richter scale) Anchorage, Alaska earthquake occurred on March 27,
1964; its epicenter was located approximately 80 miles east of Anchorage. Damage data
for 10 walled buildings were available from i) a reconnaissance report by the American
Iron and Steel Institute (AISI) [10], ii) the NISEE image database [74], and iii) the USGS
Photographic Library [112].
The primary causes of damage to structural wall buildings were i) poor construction
quality, including construction joint quality, and ii) lack of seismic detailing, including in-
sufficent confinement and inadequately designed and detailed lap splices in critical locations.
Walled buildings for which damage was documented included i) the identical 14-story Mt.
McKinely and 1200 L Street apartment buildings that had exterior rectangular walls (12
inches thick at the base) and interior core walls, ii) the 14-story steel Anchorage Westward
Hotel with interior core walls, iii) the 5-story J.C. Penney Building with exterior struc-
tural walls, iv) the 6-story Cordova building with interior core walls, and v) the unoccupied
6-story Four Seasons Apartment Building that collapsed due to failure of the core walls.

2.4.1.3 San Fernando 1971

The 6.6 magnitude (Richter scale) San Fernando, California earthquake occurred on Febru-
ary 9, 1971. Damage 5 wall buildings was documented in a U.S. Department of Commerce
report [57] and the NISEE image database [74].
Of particular interest is the 6-story Indian Hills Medical Center, designed per the 1966
building code [57], the floor plan for which is shown in Figure 2.3. The wall damage ob-
served in this building included: i) diagonal cracking, ii) horizontal cracking at construction
joints, iii) spalling and bar buckling at the wall edges, iv) bar fracture and v) crushing at
the base of the wall. This building was repaired and damaged again in the 1994 Northridge
earthquake. Damage to structural walls was also documented for five other mid-rise struc-
tures and included diagonal and horizontal crackings, especially at construction joints, and
spalling/crushing at the toe of walls.
23

Figure 2.3: Indian Hills Medical Center plan. Building was damaged in both the 1971 San
Fernando and 1994 Northridge earthquakes.

2.4.1.4 Japan 1978

The 7.4 magnitude (Richter scale) Miyagi-ken-oki (Sendai City, Japan) earthquake occurred
on June 12, 1978. The damage to 7 walled buildings was documented in a U.S. Commerce
Department report[31].
Most reinforced concrete structural wall damage was limited to diagonal and horizontal
cracking and spalling of cover, but web crushing was also observed. A three-story building
with walls on only one side of the building experienced a complete shear failure of the wall
that caused the first story to collapse.

2.4.1.5 Chile 1985

The 8.0 magnitude (moment magnitude scale) Vina del Mar, Chile earthquake occurred on
March 3, 1985. Damage to 3 walled buildings was documented in i) technical reports by
Wood et al. [117] and Rafael et al. [91], and ii) the NISEE image database [74].
These reports indicate that structural walls were the typical load-resisting system in a
Chilean building, and the ratio between wall and floor areas were much higher than for U.S.
buildings. The detailing of Chilean walls typically did not include confining reinforcement
and did not meet the ductile detailing requirements of the ACI Building Code at the time.
Shear reinforcement was typically provided by two layers of welded wire fabric. Figure 2.4
24

shows a typical detail. The typical thickness of walls was reported to range from 8 inches
in low-rise buildings to 28 inches at the base of a 22-story wall. Lap splices were typically
placed above the floor slabs. Tall buildings, having 16 or more stories, were reported to
have sustained light or no damage. The most severe damage was observed in buildings of
12-15 stories. These buildings had an initial period for which there was a high amplification
factor in the response spectra for the earthquakes ground motion.

Figure 2.4: Typical detail of Chilean wall [91].

2.4.1.6 Mexico City 1985

The 8.1 magnitude (Richter scale) Mexico City, Mexico earthquake occurred on September
19, 1985. Damage to 4 walled buildings was reported by a New Zealand reconnaissance
team [16].
Most buildings in Mexico City were RC frames with masonry in-fill; although some
buildings were constructed with structural walls to resist lateral forces. Of detailed case
studies of 22 buildings, four had structural walls. These buildings were reported to have
preformed remarkably well, with little or no damage occurring to the walls and requiring
only cosmetic repair. No images were available for these buildings and thus they are not
included in the damage statistics provided in Section 2.4.3.

2.4.1.7 Armenia 1988

The 6.8 magnitude (Richter scale) Spitak, Armenia earthquake occurred on December 7,
1988. Damage to 2 walled buildings was documented by Wyllie [118].
25

Wyllie reported that building construction of the affected areas was primarily stone-
masonry or precast concrete; however, two cast-in-place reinforced concrete lift-slab build-
ings with structural wall lateral systems were present in the affected area. One of these,
a 10-story building collapsed. A 16-story building experienced significant damage and re-
quired demolition following the earthquake. Failure of this building was due to damage to
the first floor core walls.

2.4.1.8 Loma Prieta 1989

The 6.9 magnitude (Richter scale) Loma Prieta, California earthquake occurred on Octo-
ber 17, 1989. Damage to walled 3 walled buildings was reported by an EQE Engineering
reconnaissance team [33]. The reconnaissance team reported that many mid-rise reinforced
concrete structures were extensively damaged; damage to structural walls was typically
restricted to diagonal cracking and concrete spalling.

2.4.1.9 Turkey 1992

The 6.8 magnitude (Richter scale) Erzincan, Turkey earthquake occurred on March 13,
1992. Damage data for 2 walled buildings was provided by i) an EERI reconnaissance team
[99], and ii) Saatcioglu and Bruneau [93].
Saatcioglu and Bruneau [93] reported that approximately 300 reinforced concrete build-
ings collapsed or were damaged beyond repair. The use of structural walls in buildings
was not common, but some dual frame-wall lateral systems were found in newer, mid-rise
buildings. These buildings were reported to have performed well, with wall damage limited
to cracking.

2.4.1.10 Guam 1993

The 7.8 magnitude (Richter scale) Guam earthquake occurred on August 8, 1993. An
Earthquake Engineering Research Institute (EERI) reconnaissance team reported on dam-
age following the earthquake [21]. Damage was concentrated in buildings over four stories;
many hotels were damaged. Structural walls were the typical lateral load resisting system
26

in mid-rise structures. The local building code was based on the Uniform Building Code
(UBC). However, boundary elements were typically detailed with little or no confinement.
The extent of damage in the walls was typically cracking and spalling. For the 16-story Reef
Hotel, cracking in exterior walls were repaired by epoxy injection (approximately 15,000 lin-
ear feet of cracks were injected). In the 7-story Guam Hilton hotel, built in 1974, cracking
and spalling occurred at the construction joints, splitting and spalling were observed at the
base of most walls where splices were located, and, in a few locations, bar buckling was
observed.

2.4.1.11 Northridge 1994

The 6.8 magnitude (Richter scale) Northridge, California earthquake occurred on January
17, 1994; the region affected was the as that same affected by the 1971 San Fernando
earthquake. Damage to walled buildings resulting from this earthquake is documented in
reports written by teams from EERI [43], NIST [105], and EERC (Earthquake Engineering
Research Center) [71] as well as by Mitchell et al. [67].
The EERI reconnaissance team reported that reinforced concrete structural wall build-
ings performed well with respect to life safety and collapse prevention. Most observed dam-
age was diagonal flexure-shear cracking and sliding along horizontal construction joints.
The team also reported that ten healthcare facilities were red-tagged due to severe diagonal
cracks in structural walls, which often extended through the entire thickness of the wall.
Mitchell et al. noted that structural wall buildings performed well, with crack widths that
did not appear sufficient to yield longitudinal reinforcing steel and were easily repairable, if
the number and distribution of walls was adequate. The NIST reconnaissance team reported
that walls in the Kaiser Permanente garage and the Northridge Fashion Center did not show
crack patterns typical of structural walls and, thus, were likely not loaded as intended.
Damage to the Indian Hills Medical Center was documented in the NISEE image database
[74]. The structural walls of this 6-story structure are located on the exterior of the building
(Figure 2.3), making damage observation possible. The walls in this building were damaged
in the 1971 San Fernando earthquake and repaired afterwards; details on the repair were
27

not found in the literature.

2.4.1.12 Kobe 1995

The 7.2 magnitude (Richter scale) Kobe, Japan earthquake occurred on January 17, 1995.
Shear failure was reported in walls with insufficient vertical and horizontal reinforcement
[66]. A reconnaissance team from the Structural Engineers Association of Washington
(SEAW) [96] also noted that damaged structural wall buildings typically had low percentages
of reinforcement and the presence of many door and windows openings that resulted in
‘X’-crack patterns in spandrels of walls. A NIST report [19] stated that no structural wall
related collapses occurred in tall buildings, for which structural walls were a common lateral
system, but some short buildings with walls did collapse. Damage was limited to diagonal
cracking in buildings that did not collapse or exhibit shear failure of walls.

2.4.1.13 Turkey 1999

The 7.4 magnitude (Richter scale) Izmit, Turkey earthquake occurred on August 17, 1999.
Damage to 10 walled buildings was provided from the i) a PEER reconnaissance team
[97], and ii) journal articles by Saatcioglu et al. [94, 95] and Sezen et al. [98]. These
reports indicated that structural walls were found primarily in newer buildings; some of
these buildings had walls with length to width ratios that were so small that they could
be considered to be wide rectangular columns. Damage to walls was typically observed at
stiffness discontinuities.

2.4.1.14 Nisqually 2001

The 6.8 magnitude (Richter scale) Nisqually, Washington earthquake occurred on February
28, 2001. An EERI preliminary reconnaissance report[30] indicates that the damage to
structural walls was limited to diagonal cracking and damage at the construction joints.
28

2.4.1.15 Turkey 2003

The 6.4 magnitude (Richter scale) Bingöl, Turkey earthquake occurred on May 1, 2003.
Kaplan et al.[52] reported that buildings built using the 1998 Turkish Earthquake Code
either collapsed or were badly damaged due to poor detailing, poor construction practice,
or poor material properties. Poor confinement and large spacing of transverse reinforcement
were common in reinforced concrete structures. Doğangün[29] reported that buildings with
structural walls as the lateral load resisting system generally experienced little or no damage
but one building with a dual frame-wall lateral-load resisting system had extensive web
crushing and diagonal cracking in the walls.

2.4.1.16 Chile 2010

The 8.8 magnitude (moment magnitude scale) Maule, Chile earthquake occurred on Febru-
ary 27, 2010; its epicenter was located off the Chilean coast. Numerous reconnaissance
teams visited affected areas. The majority of the damage data summarized in this thesis
was collected by an EERI reconnaissance team and provided by Jack Moehle [69], the EERI
reconnaissance team leader; additional pictures of damaged walls were provided by John
Hooper [46], Director of Earthquake Engineering at Magnusson Klemencic Associates, Seat-
tle. Structural walls are the most common type of lateral load resisting system in mid- to
high-rise reinforced concrete buildings; some dual frame-wall systems exist as well. Newer
construction was designed using the 1985 Chilean Building Code; this code was developed
based on the satisfactory performance of buildings during the 1985 Chile earthquake. The
EERI team reported that failure in structural walls was typically related to: i) compressive
failures due to high axial stresses or compressive-flexural stresses in flanged walls, ii) insuf-
ficient transverse reinforcement, iii) vertical irregularities, and/or iv) lap splices in critical
locations.

2.4.2 Types of Damage

The development of performance-based seismic design tools, specifically fragility functions,


requires establishing a relationships between observed damage and engineering demand pa-
29

rameters. Although demands are not known for the buildings documented in the literature,
images and descriptions of the observed damage provide insight into the extent of damage
in buildings. From the earthquake reconnaissance reports discussed in Section 2.4.1, images
of damage were collected and sorted by damage type. This section discusses the types of
damage observed and provides images representative of these damage states.
First, damage indicating moderate demand (cracking and vertical splitting) is presented.
Second, damage indicating a flexural response (bar fracture) is discussed. Then, compressive
damage to boundary regions (spalling, bar buckling, and core crushing) is discusssed. Next,
damage indicating shear failure (diagonal-tension and diagonal-compression) is discussed,
followed by discussions damage related to discontinuities (construction joint damage, dam-
age at splice and horizontal failure planes at geometric discontinuities). Finally, structural
wall buildings that collapsed are discussed.

2.4.2.1 Horizontal and Diagonal Cracking

The most common type of damage seen in structural walls in the field is horizontal and
vertical cracking; this type of cracking is the damage observed first in lab tests. Horizontal
cracks occur as a result of flexural loading and diagonal cracks occur as a result of shear
loading. Due to the wide-spread occurrance of cracking following an earthquake, only a
few images are included here. Figure 2.5 and 2.6 shows example of diagonal and horizontal
cracks in walls, respectively. Figure 2.7 shows the presence of both horizontal and diagonal
cracks in one wall, with horizontal cracks occurring at the edges of the wall and becoming
inclined in the web of the wall. Figure 2.8 shows the crack map for the core wall of the Mt.
McKinley building [74].
A map of cracks on the elevation of the Mt. McKinley building core walls is shown in
Figure 2.8.

2.4.2.2 Splitting

In the laboratory, vertical splitting cracks are typically observed as an indication bond slip.
Vertical splitting of the concrete was observed at the base and beneath the 4th construction
30

(a) San Fernando: 3-story Pacoima Memorial (b) Guam: 16-story ho-
Lutheran Hospital [57] tel with epoxy injected
cracks [21]

(c) Northridge: 15-story Koll building in (d) Northridge: In-


Santa Monica [67] dian Hills Medical Cen-
ter [74]
H. Sezen et al. / Engineering Structures 25 (2003) 103–114 113

10. Conclusions

Much has been written in the aftermath of the August


17, 1999, earthquake about the poor quality of residential
and commercial construction in the epicentral region.
The detailing and quality of the residential construction,
perhaps most of it not rigorously engineered, was poor
by modern US and Turkish standards. The reconnais-
sance team also documented many collapses of commer-
cial reinforced concrete moment frame buildings that
were likely carefully engineered. Both poor construction
practices and the continued use of nonductile seismic
detailing were the primary reasons for most of the build-
ing collapses. Shear reinforcement was lacking in most
damaged columns observed by the authors. In contrast
with the code design provisions, common use of 90-
degree hooks for transverse reinforcement reduced the
lateral strength and confinement of columns. Short col-
(e)wallTurkey
Fig. 16. Damaged 1999
frame building due [98]
to ground failure and umns,
(f) poor detailing
Japan in beam-column joints, strong-
1978: 3-story
wall rotation. beam weak-columns, and use of inconsistent unre-
Izumi High School [31]were among other reasons
inforced masonry infill walls
for the widespread destruction in the region affected by
the earthquake. For the most part, buildings with shear
Figure 2.5: Examples of diagonal cracks in structural walls.
walls survived with limited or no damage.

Acknowledgements

The work described in this paper made use of the


Pacific Earthquake Engineering Research Center shared
facilities supported by the Earthquake Engineering Cen-
ters Program of the National Science Foundation under
Award Number EEC-9701568. This support is grate-
31

joints in an exterior wall of the Indian Hills Medical Center following the 1971 San Fernando
earthquake and is shown in Figure 2.9.

2.4.2.3 Bar Fracture

In the lab, bar fracture typically results in loss of lateral load carrying capacity. Figure 2.10
shows a fractured bar in the 4th floor of an exterior wall in the Indian Hills Medical Center
following the 1971 San Fernando earthquake. Moehle[69] reported that bar fracture was not
uncommon in high-rise structures damaged during the 2010 Chile earthquake. Figure 2.11
shows fracture of boundary element steel in a 16-story building in Santiago, Chile. Bars
in both the boundary element and web are buckled, indicating significant cyclic demands
prior to bar fracture.
32

(a) Northridge: 15-story apartment building [74] (b) Kobe: 10-story apartment
building [96]

Figure 2.6: Examples of diagonal cracks in structural walls.

Figure 2.7: Horizontal and diagonal cracks in wall of 13-story building in Concepcion fol-
lowing the 2010 Chile earthquake [69]
33

Figure 2.8: Crack map of Mt. McKinley core walls following the 1964 earthquake [74]

(a) Splitting below 4th floor construction (b) Splitting at base of corner structural wall
joint

Figure 2.9: Indian Hills Medical Center following the 1971 San Fernando earthquake [57].
34

Figure 2.10: Fractured longitudinal reinforcement in 16-story Indian Hills Medical Center
following the 1971 San Fernando earthquake [74].

Figure 2.11: Fractured longitudinal reinforcement in 16-story Santiago building following


the 2010 Chile earthquake [69].
35

2.4.2.4 Compression Boundary Element Damage

In the lab and the field, compressive damage to the boundary elements or at the edges of a
wall includes cover spalling, bar buckling, and core crushing. Because these types of damage
are frequently observed together, they are presented together. The most minor damage seen
in boundary regions is spalling of the cover concrete; examples are shown in Figure 2.12.
More severe damage to boundary regions includes core crushing and buckling of extreme
longitudinal reinforcement; examples are shown in Figures 2.13 and 2.14.
In reviewing images of earthquake damage to boundary regions of reinforced concrete
structural walls, it was observed that many boundary regions did not contain confining
reinforcement. For those walls with confining reinforcement, 90 degree ties rather than 135
degree hooks were often observed and the volumetric ratio of confining steel appeared to be
significantly less than that required by ACI 318-08 for special structural walls. No image
was found showing a damaged boundary region that had confining reinforcement that was
thought to satisfy ACI 318-08 requirements for special structural walls.

2.4.2.5 Shear Failure

Diagonal-tension shear failure is indicated by large diagonal shear cracks spanning the length
of the wall. Fracture of transverse reinforcement in the web may also be present. The
following paragraphs discuss shear failures observed in specific earthquakes.
Figure 2.15 shows shear failure in two buildings, built prior to 1950, following the 1960
Chile earthquake. No information was available on the design and construction of the
structures. Figure 2.16 shows shear failure of a wall following the 1978 Japan earthquake;
in this building, walls were only located at one end of the 3-story structure. Figure 2.17
shows two examples of shear failure of walls during the Kobe earthquake. Figure 2.17a is a 9-
story building, with damage to an exterior wall occurring at a location where there is a plan
discontinuity. Figure 2.17b shows damage to an interior wall in the 4-story Uegahara Junior
High School. The wall is 5 m long and 100 mm thick, with 6 mm diameter smooth wire
reinforcement, spaced at 250 mm in both the vertical and horizontal directions. Boundary
elements were 500 mm square with ties spaced at 100 mm. Figure 2.18 shows the failure of
36

(a) Spalling at construction (b) Spalling and bar buckling at corner of 3-


joint. Mt. McKinley Bldg story API (Alaska 1964) [10]
(Alaska 1964) [10]

(c) 4th floor of Indian Hills (d) First floor core wall of 10-story Kaiser
Medical Center (San Fer- Foundation Hospital (San Fernando 1971)
nando 1971) [57] [57]

(e) Splice in 6-story St. John’s Hospital (f) Spalling at corner of exterior wall of un-
(Northridge 1994) [74] known building (Nisqually 2001) [30]

Figure 2.12: Boundary region cover spalling and minor bar buckling
37

(a) Anchorage Westward (b) Interior of


Hotel (Alaska 1964) [10] ENALUF(Nicaragua
1972) [74]

(c) 15-story Acapulco Building (Chile 1985) (d) 15-story Acapulco Building(Chile 1985)
[74] [74]

(e) Base of core wall (Armenia 1988) [118]

Figure 2.13: Boundary region core crushing and bar buckling


38

(a) 26-story building in Santi- (b) 16-story building in Santiago [69]


ago [69]

(c) 16-story building in Santiago [69] (d) 15-story building in Chillan [69]

Figure 2.14: Boundary region core crushing and bar buckling following the 2010 Chile
earthquake
39

a structural wall following the 1999 Turkey earthquake. Diagonal shear failure of a building
under construction was observed following the 1999 Chi-Chi, Taiwan earthquake, shown in
Figure 2.19. Two examples of shear failure in walls following the 2010 Chile earthquake are
shown in Figure 2.20.

(a) Valdivia Region Hospital, built 1932-1935 (b) Unknown building, Osorno, Chile, built
[74] 1950 [74]

Figure 2.15: Shear failure of walls following the 1960 Chile earthquake.

2.4.2.6 Web Crushing

Web crushing in a structural wall can occur when high shear stress demands cause diagonal
shear struts to crush.
Figure 2.21 shows web crushing during the 1960 Chile earthquake. Figure 2.22 shows
additional examples of walls that exhibited web crushing during other earthquakes. Walls
that had a horizontal failure plane are not included in this section. For walls in which the
web concrete was crushed, buckled bars were typically observed as well.
40

Figure 2.16: Shear failure of 3-story Paloma building in Sendai, Japan following the 1978
Japan earthquake. Structural walls were only located on one end of the building [31].

(a) 9-story [74] (b) 4-story Uegahara Junior High School [66]

Figure 2.17: Shear failure of walls in Kobe, Japan, 1995.


41

Figure 2.18: Shear failure of wall in an unknown building following the 1999 Turkey earth-
quake [74].

Figure 2.19: Shear failure in a building under construction following the 1999 Chi-Chi
earthquake [60].
42

(a) 5-story building in Santiago (b) 24-story building in Concepcion

Figure 2.20: Shear failure of walls following the 2010 Chile earthquake [69].

(a) Valdivia Region Hospital, built 1932-1935 (b) Unknown building, Osorno, Chile, built
1950

Figure 2.21: Web crushing/uniform failure plane in walls following the 1960 Chile earth-
quake [74].
43

(a) 14-story Anchorage Westward Hotel interior (b) First floor elevator tower of 6-story Cordova
wall (Alaska 1964) [10] Building (Alaska 1964) [10]

(c) 3-story Taiyo Fisheries Plant (Japan 1978) [31] (d) 15-story Acapulco condominium (Chile 1985)
[74]
A. Doǧangün / Engineering Structures 26 (2004) 841–856 847

(e) Interior wall


Fig. 7. Heavily of Bingol
damaged High
structural wall School
of school building(Turkey
[13]. (f) 14-story building in Concepcion (Chile 2010)
2003) [29] [69]
Lateral rigidity of the infill walls are ignored in the flat slab and bare or hollow block floor and waffle slab
design. But, these walls contributed significantly to the Figure
system 2.22: Web crushing
whose columns innot
and beams do structural
satisfy the walls
lateral stiffness of buildings during the earthquake. requirements given for high ductility level.
Firstly, they tried to reduce lateral displacement of Only one frame-structural wall system into which a
building. Once the brick infills failed, the lateral strength structural wall was incorporated inadequately was
and stiffness had to be provided by the frames alone, damaged in Bingöl. This building was not the com-
which then experienced significant inelasticity in the pleted high school. The structural wall of this building
critical region. At this stage, the ability of reinforced was damaged as shown in Fig. 7. Similar structural
concrete columns, beams, and beam-column joints to wall damages were observed for the January 17, 1995
sustain deformation demands depended on how well the Kobe earthquake [19]. The building might have been
seismic design and detailing requirements were followed saved from collapse due to the large rigidity of a struc-
in both the design and construction stage. tural wall. In general, buildings with structural walls
Soft stories should have been protected by adding an survived with limited or no damage during the earth-
appropriate amount of structural walls. Structural quakes that occurred in Turkey in the last years [14].
44

2.4.2.7 Horizontal Failure Plane

A type of damage pattern that is observed in some structural walls but is difficult to
categorize is referred to here as a horizontal failure plane. In such walls, there is a horizontal
plane of spalling, crushing and buckling that spans the entire length of the wall. This can
be viewed as a combination of web crushing and boundary element damage, although it
is not clear which type of damage was ultimately responsible for failure of the wall. The
damage may also indicate a sliding shear failure. A continuous plane of damage across the
length of the wall is not commonly considered when wall failure modes are discussed; thus,
it is of particular interest to note here. The following paragraph describes instances of an
observed horizontal failure plane.
Following the 1964 Alaska earthquake, failure planes occurred at construction joints in
the 14-story Mt. McKinley building. In the 1200 L Street building of the same design, the
damage at the construction joints was not as extensive; however, a uniform failure plane was
observed just below the construction joint (and just above a change in stiffness) in the west
wall of the south face. This damage is shown in Figure 2.23 and was attributed to lack of
adequate axial-flexural capacity[74]. Figure 2.23 shows the damaged second floor structural
wall in the West Anchorage High School. Note the lack of confining reinforcement in the
boundary regions and minimal transverse reinforcement in the web. Figure 2.25 shows the
damage to a light-weight concrete core wall in a 15-story building in Oakland, CA following
the Loma Prieta earthquake. Horizontal failure planes were observed in many buildings
following the 2010 Chile earthquake; examples are shown in Figure 2.26.
45

Figure 2.23: South face of the 1200 L Street building following the 1964 Alaska earthquake
[74].

Figure 2.24: Second story of West Anchorage High School following the 1964 Alaska earth-
quake. Note the lack of confining reinforcement in the boundary regions [10].
46

Figure 2.25: 15-story building in Oakland, CA following the 1989 Loma Prieta earthquake
[33].
47

(a) 16-story building in Santiago. Wall was (b) 12-story building in Santiago. Multiple
below grade on the side of the building that walls crushed in the first floor (Below grade
experienced compressive damage. level).

(c) 17-story building in Concepcion. Failure (d) 24-story building in Concepcion.


plan occurred in wall on far left side of the
picture.

(e) 13-story building in Con- (f) Emerald Building in Santiago


cepcion.

Figure 2.26: Horizontal failure plane in walls following 2010 Chile earthquake [69].
48

2.4.2.8 Construction Joint Damage

Construction joints are present in most walls and often the location at which damage is
concentrated. Of the 183 images of damage collected, 36 (20%) show damage at construction
joints. The severity of the damage ranges from minor horizontal cracking to crushing of
the concrete through the thickness of the wall and along the entire length (or width) of the
wall. In the following list, the construction joints that are discussed were identified as such
by the authors of the reconnaissance reports.

• Alaska 1964

– Exterior walls in the 14-story Mt. McKinley and 1200 L Buildings were damaged
at construction joints. In Figure 2.27a, damage is seen at the second floor of the
Mt. McKinley building. Figure 2.27b shows a close up of this damage. On one
side of the building, construction joint damage was seen in the east exterior wall
at the third floor. A smooth surface can be observed at the damage section of
the wall in Figure 2.27c.
– A neighboring building of nearly identical design, 1200 L Street saw only minimal
damage at the construction joint, with a horizontal plane of damage developing
just below the construction joint near the base of the wall, as shown in Fig-
ure 2.27d.
– The 14-story tower of Anchorage Westward Hotel was constructed in two phases,
six years apart. The construction joint between these two phases was located at
the 8th floor and was the location of the principal failure of the structure, shown
in Figure 2.28a.
– Damage at construction joints observed in other buildings are shown in Fig-
ures 2.28b-2.28d.

• San Fernando 1971

– Damage at the construction joints in the 7-story Holy Cross Hospital in the
exterior and interior structural walls was observed, as shown in Figure 2.29.
49

(a) Mt. McKinley south face. Damage at 2nd (b) Close up of construction joint damage to the Mt.
floor construction joints [10]. McKinley building south face [10].

(c) North face of Mt. McKinley building [10]. (d) 1200 L Street Building south face. Note
the damage occurred below the construction joint
here unlike in the identical Mt. McKinley build-
ing [10].

Figure 2.27: Construction joints in the Mt. McKinley and 1200 L Street buildings (1964
Alaska).
50

(a) Damage at 8th floor construction joint between (b) 2-story West Anchorage High School [10]
old and new construction in the Anchorage Westward
Hotel [10]

(c) 6-story Cordova building [74] (d) J.C. Penney Department Store [112]

Figure 2.28: Damage at construction joints following the 1964 Alaska earthquake.
51

(a) 4th floor east elevation of ele- (b) 3rd floor west elevation of el-
vator tower [57] evator tower [57]

(c) Interior stair well [57] (d) Interior stair well [74]

Figure 2.29: Holy Cross Hospital construction joint damage following the 1971 San Fernando
earthquake
52

– Similar damage was observed on the exterior of the 6-story Indian Hills Medical
Center (Figure 2.30) and the Nethercutt Museum (Figure 2.31).

• Armenia 1988 - The failure of the core wall in a 16-story lift-slab building, shown in
Figure 2.32, was reported to have occurred at the first-floor construction joint [118].
Welded lap splices at this location were observed to have remained intact.

• Loma Prieta 1989 - Damage at a construction joint is shown in Figure 2.33.

• Northridge 1994 - The 15-story Indian Hills Medical Center, damaged during the San
Fernando earthquake, was damaged again during the Northridge earthquake, as shown
in Figure 2.34, with damage occurring at a construction joint on the exterior of the
elevator tower.

• Kobe 1995 - Figure 2.35 shows damage to construction joints during the 1995 Kobe
earthquake.

• Turkey 1999 - Figure 2.36 shows web crushing to an interior wall at the location of a
construction joint that occurred during the 1999 Turkey earthquake.

• Nisqually 2001 - Little damage to structural walls beyond horizontal and diagonal
cracking was observed; however, spalling of concrete was observed at some construction
joints [30], as shown in Figure 2.37.
53

Figure 2.30: Damage to construction joints of exterior walls of the Indian Hills Medical
Center following the 1971 San Fernando earthquake [57]

(a) (b)

(c)

Figure 2.31: Nethercutt construction joint damage following the 1971 San Fernando earth-
quake [57]
54

Figure 2.32: Crushing and bar buckling at construction joint of core-wall in 16-story lift
slab building following the 1988 Armenia earthquake [118].

Figure 2.33: Maryknoll retirement home in Mountain View, CA following the 1989 Loma
Prieta earthquake. Horizontal cracks at construction joints can be seen in multiple locations
[33].
55

Figure 2.34: Damage to the Indian Hills Medical Center in the Northridge earthquake was
located at the construction joint of the elevator tower [74].

(a) (b)

Figure 2.35: Construction joint damage following the 1995 Kobe earthquake [66].
56

(b)

Fig. 14: A frame-wall building in Izmit without any significant


Figure 2.36: Web damage:
crushing(a)atexterior
construction
view; andjoint followingwall
(b) a structural thewith
1999 Turkey earthquake [94].
concrete
placement problem

Turkish seismic code requirements


Many reinforced concrete buildings that suffered
damage during the earthquake were designed in a period
when the 1975 edition of the “Specifications for
Structures to be Built in Disaster Areas,”1 that had been
issued by the Ministry of Reconstruction and
Resettlement of the Government of Turkey, was in effect.
Therefore, this edition of the specifications, referred to as
the Turkish Code, is highlighted in this article, though a
more recent edition was issued in 1998. The emphasis is
(a) placed on reinforced concrete frame buildings with
masonry infills, because this type of structural system
dominated not only the concrete construction, but the
entire building inventory, of the earthquake-stricken areas.
Design base shear (F) in the Turkish Code is ex-
pressed below and resembles that specified in North
ges American building codes of the era.
e collapsed during the earthquake.
ced concrete, multiple single-span F = CW (1)
ver the Ankara-Istanbul Tollway, near
Figure
ated in an area that suffered 2.37: Damage at construction joint
signifi- C = following
C0 KSI the 2001 Nisqually(2) earthquake [30].
ement. It was also very close to the
showed a horizontal offset of about where W is total structural weight, and C is seismic
fault offset illustrated in Figure 2(b) coefficient as defined below:
km (0.6 mi) away from the collapsed
e, shown in Fig. 16, was attributed to ■ C0 is seismic zone coefficient, and is equal to 0.1 for
irders. It was clear that the excessive seismic Zone 1, which is the zone of the area that was
t in the area caused the separation of affected by the earthquake;
lumns that resulted in unseating of ■ K is a coefficient related to structural type, and is
equal to 0.8 for ductile concrete frames with
eavy damage sustained in the area, unreinforced masonry infills, and equal to 1.5 for
57

2.4.2.9 Damage at a Splice in the Longitudinal Reinforcement

Damage located at spliced longitudinal reinforcement is of particular interest when examin-


ing as few experimental data exist to indicate potential damage modes, yet most structural
wall buildings are of sufficient height to require the use of splices in construction.
The following describes images found indicating damage at longitudinal reinforcement
lap splices. Figure 2.38 shows spalling that exposed the splice at the 4th story of the Indian
Hills Medical Center following the 1994 Northridge earthquake. Figure 2.39 shows similar
damage at a 1st floor lap splice in a 16-story building in Santiago following the 2010 Chile
earthquake. The side of the building was reported to be in tension. Moehle[69] reported
that lap splices were located at some failure locations in walls. Figure 2.40 shows lap splices
of smooth longitudinal bars in walls with wire mesh transverse reinforcement.
In addition to the images of splice damage discussed above, some researchers described
the observed damage. Walls in buildings affected by the 1988 Armenia earthquake were
constructed with welded lap splices. No failures of these splices were observed [118]. Wood
et al [117], in a report on observed building damage following the 1985 earthquake reported
that lap splices were commonly located immediately above the floor slabs, but did not report
any damage explicitly connected with the performance of splices. Spalling and splitting was
reported at splices in the base of the walls in the 7-story Guam Hilton hotel following the
Guam 1993 earthquake [21]. Detailed information about the splice or images were not
available.

2.4.2.10 Collapse

In the preceding paragraphs, damage to buildings that remained standing following an


earthquake was discussed. For structural wall buildings that collapsed during earthquakes, it
is difficult to determine the mechanism that caused collapse and little information is available
in the literature addressing the cause of collapse. Collapsed buildings with structural walls
are shown in Figure 2.41.
The El Faro building (Figure 2.41a) was designed in 1979, failed during the 1985 Chile
earthquake, and was destroyed shorty after. Wood et al. [117] reported that the discrep-
58

Figure 2.38: Spalling exposing splice at 4th story level of the Indian Hills Medical Center
following the 1994 Northridge earthquake [74].

Figure 2.39: Spalling at 1st floor in 16-story Santiago building following the 2010 Chile
earthquake [69].
59

(a)

(b)

Figure 2.40: Damaged core concrete at lap splice locations following the 2010 Chile earth-
quake [46].
60

(a) 8-story El Faro


building in Viña del
Mar (Chile 1985) [117].

(b) 6-story unoccupied apartment building (c) 9-story building (Kobe 1995) [20].
(Alaska 1964) [74].

(d) Apartment building (Kobe 1995) [19]. (e) 12-story building in Concepcion (Chile
2010 [69]).

Figure 2.41: Buildings with structural walls that collapsed.


61

ancies existed between the architectural and structural drawings and the structure more
closely followed the architectural drawings. Honeycombing in the concrete and poor con-
struction joints in the foundation were also reported. The damage to the walls was crushing
below the first floor slab. A detailed analysis of the failure is provided by Wood et al[116].
Figure 2.41b shows the Four Seasons apartment building that collapsed during the 1964
Alaska earthquake. The lateral system for the 6-story building was two central core walls,
which failed in the first story.
Details of the other collapsed buildings shown in Figure 2.41 were not available.

2.4.3 Building Damage Summary

The review of the literature addressing earthquake damaged buildings resulted in 91 build-
ings, 62 with known height in stories, for which damage information was available. The
buildings were located in one of the following countries/regions: a) United States, b) Chile,
c) Asia (Japan, Taiwan) or d) other (Mexico, Nicaragua, Turkey, Guam, Soviet Union).
For each building with significant structural wall damage that could be expected to require
partial or full replacement of the wall, the dominant type of damage was classified as follows:

• Compressive boundary element damage: The predominant damage was spalling,


bar buckling, and/or core crushing in the boundary regions.

• Diagonal: Damage was predominately large diagonal cracks.

• Web crushing: The predominant damage was a region of horizontal and/or diagonal
crushing in the wall web.

• Horizontal failure plane: Damage characterized by a horizontal region (i.e. plane)


of crushed concrete including the boundary region and web. It is difficult to know if
web crushing or boundary region damage occurred first or if both occurred simulta-
neously. As is discussed in Section 2.5, experimental tests have been done in which
web crushing occurred first and loss of lateral load carrying capacity due to boundary
62

element crushing; in other laboratory tests, web crushing occurred after significant
damage had accumulated in the boundary region.

• Collapse: Partial or complete collapse of the structure. For structural wall build-
ings that collapse, the dominant type of damage, which caused failure, is difficult to
determine.

For each significantly damaged wall in a building, the type of damage was classified.
Thus, for some buildings, multiple damage types were assigned. An example is the building
shown in Figure 2.26c, in which the wall on the far left of the picture has a horizontal failure
plane and the wall in the center has significant web crushing and compressive damage to
the boundary region.
Table 2.6 presents the number of buildings in each region in which walls that were
classified as having the damage types listed above. This is illustrated graphically by the
histograms in Figure 2.42, where the number of buildings with a particular damage type are
presented as a percentage of total damaged buildings in the region. For buildings of known
heights, the percentage of buildings in which each damage type was observed is are shown
as a function of building height in Figure 2.43 (all regions) and Figures 2.44 - 2.46 (Asia,
the United States, and Chile).

Table 2.6: Building damage types by region

# of Damage Type
1
Region # Buildings B.E. Diagonal Web Crush Horizontal Collapse
All 91 (62) 28 (26) 21 (13) 13 (10) 18 (16) 6 (5)
United States 27 (22) 14 (13) 3 (2) 2 (2) 6 (6) 1 (1)
Chile 23 (21) 12 (12) 6 (6) 6 (6) 9 (8) 2 (2)
Asia 21 (15) 1 (1) 5 (4) 2 (2) 2 (1) 3 (2)
Other 20 (4) 1 (0) 7 (1) 3 (0) 1 (1) 0 (0)
1
# Total (# Known Height)

Comparison of the damage types by region (Figure 2.42) reveals that boundary element
damage is the dominant behavior in approximately 50% of the damaged walls in U.S. and
63

(a) All regions (b) United States (c) Chile

(d) Asia (e) Other

Figure 2.42: Percent of buildings in which a particular type of damage was observed, by
region.

Chilean buildings. In U.S. buildings, few structural walls are observed to have web crushing
or diagonal failure planes. For Chilean construction, diagonal cracking and crushing, web
crushing, and horizontal damage planes were observed more than in U.S. buildings, and
may be a result of the lack of confined boundary regions typically used in Chilean walls.
Figures 2.45 and 2.46 show the damage for the U.S. and Chile as a function of building
height. In the United States, approximately 40% of buildings with boundary element dam-
age were between 2 and 6 stories, while in Chile, approximately 40% of the buildings with
boundary element damage were between 12 and 16 stories.
The high concentration of damage in 12-16 story buildings in Chile may be due to
specific earthquake characteristics [91]; damage for Chilean buildings where available for
3 earthquakes, 2 of which reported damage to buildings of this height range. Damage to
buildings in the U.S. were available for 7 earthquakes, but the majority of wall damage was
reported for buildings of 2-6 stories.
This is consistent with the findings of Rafael et al. [91] that buildings of this height
range had periods that were more highly excited by the 1985 Chile earthquake, although it
64

(a) B.E. compressive failure (b) Diagonal shear failure (c) Web crushing failure

(d) Horizontal failure plane (e) Collapse

Figure 2.43: Primary damage type for buildings of known height (All regions)

(a) B.E. compressive failure (b) Diagonal shear failure (c) Web crushing failure

(d) Horizontal failure plane (e) Collapse

Figure 2.44: Primary damage type for buildings of known height (Asia)
65

(a) B.E. compressive failure (b) Diagonal shear failure (c) Web crushing failure

(d) Horizontal failure plane (e) Collapse

Figure 2.45: Primary damage type for buildings of known height (United States)

(a) B.E. compressive failure (b) Diagonal shear failure (c) Web crushing failure

(d) Horizontal failure plane (e) Collapse

Figure 2.46: Primary damage type for buildings of known height (Chile)
66

should be noted that the number of Chilean buildings used are dominated by the damage
data from the more recent 2010 Chile earthquake.
The relative occurrence of large diagonal cracks (such as that shown in Figures 2.15 -
2.20) was much greater in Asia and other regions than in the United States and Chile. The
few instances of diagonally dominant behavior in U.S. buildings occurred in small wall piers
between window/door openings.

2.5 Experimental Research on Performance of Walls

This section summarizes previous experimental research investigating the performance of


slender walls. Sections 2.5.1 through 2.5.18 present summaries of experimental research
programs in which conventionally reinforced slender walls (shear span ratio greater than
or equal to 2.0) without openings were tested in the laboratory. The research programs
are presented in chronological order. Rectangular, barbell, C-shape, I-shape, and T-shape
cross-sections are considered. Walls subject to dynamic loading are not considered. The ex-
perimental research were used to develop fragility functions defining the likelihood of an RC
wall exhibiting a given damage state as a function of maximum deformation demand (see
Chapter 7). The following aspects of each experimental research program are summarized:
i) research team/program and purpose of the experimental investigation, ii) design of spec-
imens, iii) details of specimen construction and material properties, iv) experimental test
set-up, including the shear span ratio M/(V `w ) (where M/V is the ratio of base moment to
base shear and `w is the length of the wall) and instrumentation used to monitor specimen
behavior, v) loading protocol, vi) failure of specimens, and vii) observations/conclusions
made by the research team. Table 2.7 lists the research programs included in this review
and the parameters considered by each.
Section 2.5.19 provides brief summaries of experimental programs that were reviewed
but not used to develop fragility functions.
Section 2.5.20 provides a review of previous research comparing the results of multiple
experimental test programs.
Table 2.7: Summary of variables considered in wall test programs. Geometric, reinforcement, material property and loading
data for each individual specimen are provided in Tables 2.10 - 2.12.

Geometry Reinforcement Mat’l Prop Loading


Test Program # Walls1 Shape Openings tw hw Confine Vertical Horizontal fc0 fy M/(V `w ) P History Repair
Ali 3(1) x
Beyer et al 2 x x
Brueggen 2 x x
Dazio et al 6 x x
Elnady 3(2) x
Ile et al 2 x
Khalil et 3(1) x
Lefas et al 16(10)? x x x x x x
Liu 2 x
Mobeen 3(1) x
Morgan et al 1
Oesterle et al 14(10) x x x x x x x x
Pilakoutas et al 9(6) ? x
Riva et al 1
Shiu et al 2(1) x
Tasnimi 4 x
Thomsen et al 4 x x
Tupper 3(1) x

a
Number of walls used to develop fragility functions is provided in parentheses.
67
68

2.5.1 Portland Cement Association

Fourteen wall speciens were tested by Oesterle et al. [77, 78, 80] at Portland Cement
Association Construction Technology Laboratories as part of a three-phase program to
develop design criteria for walls in earthquake resistant buildings. The objective of the
study was to develop design recommendations to increasethe energy dissipation of walls.
Phase I of the program included tests of eight walls. Phase II includeded tests of six walls.
These two phases are summarized in detail here. Phase III included testing of additional
walls and is referenced in several papers [38, 79]; however, communications with the PCA
library [37] indicated that a final report was not written for Phase III. Thus, Phase III is
not discussed due to lack of information. The following sections provide details of Phase I
and Phase II followed by conclusions made for the full test program.

2.5.1.1 Phase I

The variables considered for the eight wall specimens were i) shape (rectangular (2), barbell
(6) and H-shaped (1) walls were tested), ii) loading (one wall was tested monotonically,
all others cyclically), iii) repair method (one wall was repaired and re-tested; this repaired
test is not discussed here), iv) longitudinal boundary element reinforcement ratio, and v)
boundary element confining reinforcement ratio.
Design of the wall specimens was not based on a prototype building. Wall designs were
approximately 1/3-scale. The 1971 ACI building code was used. Longitudinal reinforcement
in the wall webs was kept at the minimum required by the ACI Code. The shear reinforce-
ment was designed to meet the ACI Code requirements and to allow a shear demand defined
by the nominal expected flexural strength of the wall. Walls were designed using a nominal
steel yield strength of 60 ksi (414 MPa), and a nominal concrete compressive strength of 6.0
ksi (41 MPa). Two methods of detailing the confining reinforcement were used i) ordinary
column ties and ii) rectangular hoops and cross-ties.
The walls were cast vertically, with each floor cast separately. Construction joints were
made “following standard practice” [80]; this is not described in the report. Longitudi-
nal reinforcement was continuous from the foundation block to the slab at the top of the
69

(a) R1 (b) R2 (c) F1

(d) B1 (e) B2 (f) B3, B4 (g) B5

Figure 2.47: Cross-sections of Oesterle Phase I wall specimens [80].


70

wall. The maximum concrete aggregate dimension was 3/8 inches (9.5 mm). Compressive
strength and modulus of rupture tests were done. The actual concrete strength ranged from
5.6 − 7.8 ksi (38 − 54 MPa). Tension testing of steel coupons was done; yield strengths of
the reinforcement were approximately the specified value of 60 ksi (414 MPa).
The walls were tested vertically between two reaction frames (shown in Figure 2.48). The
foundation block of the wall was bolted to the laboratory strong floor. Lateral loads were
applied to the slab at the top of the wall, resulting in a shear span ratio of M/(V `w ) = 2.40.
Lateral displacements were measured on both sides of the wall in the lower half of the wall
and on one side in the top half of the wall. Diagonal displacement measurements were also
taken at the points at which lateral displacement was measured in the lower half of the wall
and used to calculate wall rotations. Diagonal relative displacement measurements were
taken in the two lowest levels of the wall to measure shear distortion. Strain gauges were
placed on vertical and horizontal reinforcement in all walls and on hoops and cross-ties of
well confined walls. Dial gauges were used to measure slip at the construction joints.

Figure 2.48: Test set-up for walls tested by Oesterle et al. [80].

No axial loads were applied to the walls. Reverse-cyclic loading was applied to all but
one wall. Tests were force-controlled prior to yield (approximately three levels of maximum
71

force demand for each test) and displacement-controlled following yield. Three cycles were
done for each peak force/displacement. Free vibration tests were done at some points during
the tests to determine the frequency and damping characteristics of the walls.

2.5.1.2 Phase II

The variables considered for the Phase II tests were i) shape (barbell (6) and H-shaped
(1) walls were tested), ii) loading (two walls tested with only cycles at large drift levels),
iii) repair method (one wall was repaired and re-tested; this repaired tests in not discussed
here), iv) longitudinal boundary element reinforcement ratio, v) boundary element confining
reinforcement ratio, vi) horizontal reinforcement ratio, vii) concrete compressive strength
and viii) axial load ratio.
Design, construction, material, and instrumentation were the same as for Phase I. Fig-
ure 2.49 shows the cross-sections for the Phase II walls. The test configuration was modified
to allow application of axial loads, shown in Figure 2.50.
Axial loads of 0.1Ag fc0 were applied to the walls. Lateral load was applied at height
that created a shear span ratio of M/(V `w ) = 2.40. Tests were force-controlled prior
to yield and displacement-controlled thereafter. Three cycles were done for each peak
force/displacement. The displacement history for Specimens B9 and B10 differed from
that used for the other specimens in that an initial small displacement cycle to crack the
specimen was followed by a cycle to a large ductility demand, with subsequent cycles at a
lesser ductility demand. Free vibration tests were done at some points during the test to
determine the frequency and damping characteristics of the walls

2.5.1.3 Observations and Conclusions

The two phases of experimental tests conducted by Oesterle et al. [77, 78, 80] resulted in
the observations and conclusions discussed here.
Typical crack patterns for walls with low and high shear stress demands are shown in
Figure 2.51. In walls with low shear stress demands, horizontal cracks extended a significant
distance into the wall web before becoming inclined cracks. In walls with high shear stress
72

(a) F2 (b) B6

(c) B7 and B9 (d) B8 (e) B10

Figure 2.49: Cross-sections of Oesterle et al. [80] Phase II wall.


73

Figure 2.50: Test setup axial load modification for Phase II tests by Oesterle et al. [78].

demands, horizontal cracks were present only in the boundary elements and shear was
transferred by truss action, controlling the failure of the wall. The only exception to this
was walls with insufficient boundary element confinement.
Oesterle et al. made the following additional observations and conclusions:

1. Walls designed using the 1971 ACI Building Code without detailing for seismic load
will reach design shear and flexure strengths and possess significant inelastic deforma-
tion capacity.
2. Wall ductility increased with decreasing levels of nominal shear stress demand.
3. For walls with low shear stress, bar buckling limits the drift capacity of the wall but
increasing teh confining reinforcement ratios delayed the onset of bar buckling.
4. For walls with high shear stress, web shear distress limits wall performance and axial
load increases wall ductility.
5. Confinement in boundary regions improved inelastic performance by: i) delaying bar
74

(a) Low shear stress

(b) High shear stress

Figure 2.51: Typical crack patterns in Oesterle et al. walls [80].


75

buckling, ii) increasing stiffness, and iii) increasing shear capacity of boundary ele-
ments.
6. Displacements due to shear distortions were a significant part of total lateral inelastic
displacement.
7. Yielding of reinforcing steel extended approximately lw to 1.5lw up the height of the
wall. Damage was limited to a height of 0.5lw .
8. Stiffness decreased with accumulated number of load reversals.

2.5.2 Shiu et al

Shiu et al. [100] tested two rectangular walls at the Construction Technology Laboratories
of the Portland Cement Association to evaluate the impact of openings on the seismic
behavior of structural walls. One wall without openings was tested as a control specimen
and is discussed in detail here; wall specimens with openings are not discussed.
The test specimen was 1/3-scale of a prototype wall elements of a coupled wall system,
for which a prototype wall was designed using the 1971 ACI Building Code and the 1976
Uniform Building Code (Earthquake Zone 4) and checked by dynamic analyses using ac-
tual earthquake ground motions. The final design of the 6-story solid wall specimen had
boundary elements where most of the flexural reinforcement was concentrated and is shown
in Figure 2.52a.
The specimen was cast vertically, one floor at a time. Slabs (2.5 inches (63.5 mm) thick)
were provided at each floor. Concrete strength of 3000 psi (20.7 MPa) and Grade 60 steel
were specified. Boundary element reinforcement was #4 bars, confining reinforcement was
D-3 deformed wire. Web and horizontal reinforcement was 6 mm diameter bars.
The test setup is shown in Figure 2.52b. The walls were post-tensioned to the laboratory
floor and tested between two reaction walls. Lateral stability was provided to prevent out-
of-plane movement of the wall. No axial load was applied to the walls. Lateral load was
applied at the top of the wall, creating a shear span ratio of M/(V `w ) = 2.88. Approximately
150 steel strain gauges were used on the vertical, horizontal, and confining steel. Lateral
deflections were measured at the first, second, third and sixth floors on both sides of the
76

(a) Cross-section

(b) Test setup

(c) Load history (d) Damage to lower two floors

Figure 2.52: Shiu [100] solid wall.


77

walls. Vertical measurements were taken at the base, first and second floors to determine
rotation. Shear deformations were measured using diagonally mounted instruments in the
first and second floors.
A displacement-controlled reverse-cyclic history, broken up into three phases (Figure 2.52c),
was used to test the walls. The first two phases were designed to represent severe earth-
quakes, with a total of ten load cycles completed, five of them beyond elastic loading. The
first cycle was to a specified rotation ductility; subsequent cycles were displacement con-
trolled. The third phase consisted of cycles of incrementally increase displacements and
continued until a loss of lateral load carrying capacity occurred.
The solid wall, C1, had a sliding shear failure at 2.86% drift. The final damage to the
lower two floors is shown in Figure 2.52d. Although Shiu et al reported that construction
joints were only present at the top of the floor slabs, the sliding shear failure occurred along
a horizontal crack that opened very early in the test, suggesting a construction joint was
also present in the middle of the first floor of the wall. The pierced wall performed similar
to the solid wall in terms of shear strength, displacement capacity and energy dissipation,
but loss of lateral strength was due to a shear-compression failure of the boundary elements.
Shiue et al concluded that i) design practice of placing discontinuous reinforcement on each
side of openings functioned well, and ii) stress concentrations at openings did not effect
behavior of the pierced wall.

2.5.3 Morgan et al

One isolated barbell-shaped wall was tested by Morgan et al. [72, 73] at the Construction
Technology Laboratories of the Portland Cement Association as part of a joint program
between the U.S. and Japan to improve seismic safety by improving the understanding of
the relationship between meduim- and small-scale tests, component tests, and analytical
studies. The program also tested a wall-frame assembly and small scale walls; reports on
these tests were not applicable to this study or unavailable and thus are not discussed in
further detail.
The wall design matched that of the wall in the wall-frame assembly also tested as a
78

part of the program; details on the design procedure were not available. The specimen
represented the bottom five stories of a seven story building and was constructed at a scale
of 1 : 3.5 of the prototype building. Figure 2.53 shows the cross-section of the wall. At the
floor levels, additional horizontal reinforcement for the beams in the wall-frame subassembly
was provided to create similarity between the two tests.

Figure 2.53: Cross-section of wall tested by Morgan et al. [72].

The specimen was cast horizontally, with the foundation block cast after the wall. Lon-
gitudinal reinforcement was continuous. Boundary element longitudinal reinforcement was
deformed 6 mm rolled bars with properties similar to Grade 60 steel. All other steel was
deformed wire that was heat treated to create a stress-strain response similar to Grade 60
steel. Small aggregrate concrete was used, with a compressive strength of 3.84 ksi (26.5
MPa).
The wall was tested between two reaction walls with supports for out-of-plane stabil-
ity provided at the second and fifth stories. Axial loads were applied at the top of the
boundary elements and distributed to the wall through a concrete beam at the top of the
wall. A single actuator applied lateral load at the top of the wall specimen, thus creating
a shear span ratio of M/(V `w ) = 2.75, equivalent to that of an inverted triangular load
distribution on the prototype seven story wall. Strain gauges were used on i) corner longi-
tudinal bars in each boundary element at three different levels, ii) three different elevations
of web longitudinal reinforcing bars, iii) confining reinforcement at three elevations, and iv)
horizontal reinforcement at three elevations. Absolute vertical and horizontal displacement
were measured on both sides of the walls, but more densely on one side.
79

A constant axial load of 0.05Ag fc0 was applied to the wall prior to beginning the dis-
placement controlled reverse cyclic load history. Three cycles were completed at each dis-
placement level.
The test reached an ultimate drift of 1.79%. Failure of the wall was not reported by the
researchers, but a drop in lateral load carrying capacity of more than 20% had occurred.
A sketch of the final damage state of the wall is shown in Figure 2.54. The following
observations/conclusions were reported by the researchers: i) the maximum shear stress
p
reached was 3.1 fc0 , which was 1.04 times the analytically determined capacity of the wall,
ii) vertical deformations of the tensile boundary element were found to be concentrated in the
first story, iii) the boundary element confining reinforcement experienced significant strain
in the first story of the wall, but did not reach yield strains, and iii) at large displacements,
compressive deformations of the boundary element were relatively small relative to the
deformations in the tensile boundary element, indicating a neutral axis depth near the edge
of the compression boundary element.

Figure 2.54: Crack map after testing by Morgan et al. [72].


80

••n••n•Er

2.5.4 Lefas, Kotsovos, and Ambraseys

Thirteen walls were tested by Lefas, Kotsovos, and Ambraseys to investigate the the fol-
lowing parameters: i) aspect ratio, ii) axial load, iii) concrete strength, and iv) amount of
shear reinforcement. Seven walls had an aspect ratio of 1.0 and are not discussed here. The
remaining 6 walls, with an aspect ratio equal to 2.0, are discussed. [56, 54]
Design of the specimens followed the requirements of the 1983 ACI Building Code at a
• • • • • . • .•
scale of 1:2.4. The reinforcement layout of all 2.0 aspect ratio walls were the same except• for

one, which had 1/2 the shear reinforcement of the other walls and a specified concrete com-
pressive strength (3.5 ksi or 24 MPa, versus approximately 5.0 ksi or 34 MPa). Figure 2.55a
1 00000000000 • • • • •
shows the cross-section of the walls. High-tensile strength deformed bars were used for the
• 1 •

-A
vertical reinforcement (8 mm) and horizontal reinforcement (6 mm). Confining steel was
250 650 foundation blocks 250 67.5 65 67.5
4 mm mild steel bars. The walls had both and wall
04 caps to anchor the
1150 200
reinforcement. The foundation block and cap were cast separately from the wall.

SECTION

SECTION IFF=1H_ 165
A-A
• • • •
B-B

140 370 140


4. .4
650
(a) Cross-section (dimensions in mm)

All dimensions in mm

Fig. 7.1. Geometry and reinforcement details of type II walls.

(b) Test set-up

Figure 2.55: Monotonic wall tests by Lefas et al. [56].


81

The walls were tested vertically between a steel reaction frame and concrete reaction
blocks. Axial loads were applied in the center of the wall to a beam that spread the load
out to the wall cap. Lateral forces were applied to the wall cap to create a shear span ratio
of M/(V `w ) = 2.0. LVDTs measured the lateral displacement in five locations along the
height of the wall. Vertical deformations were measured on both sides of the wall at the wall-
foundation interface, to measure rotation, and at the top of the wall, to measure vertical
elongation of the wall. Out-of-plane displacements were also measured but found not to
exceed 0.2% of the wall height. Steel strain gauges were placed on the corner longitudinal
bars near the foundation.
Application of the lateral forces was force-controlled and monotonic. Axial loads were
applied to the walls prior to application of the axial loads. The magnitude of the force was
varied between the walls and was either 0.0 (three walls), 0.1Ag fc0 (one wall) or P = 0.2Ag fc0
(two walls). Specimen SW25, with a load of 0.2Ag fc0 , was accidentally loaded eccentrically
(both axial and lateral forces) partially through the test.
All walls failed due to vertical splitting in the compression zone. Figure 2.56 shows
examples of the damage at the base of the walls for the three axial load ratios. The slender
walls with no axial load exhibited a more ductile response, with drifts as much as 3.8%,
than did those with axial loads (2.54% drift for 0.2Ag fc0 ). The wall with reduced shear
reinforcement and concrete compressive strength had a lateral load capacity approximately
7% of the comparison walls. The following observations/conclusions were made by the
researchers: i) axial compression increases the lateral load capacity and secant stiffness of a
wall and reduces the lateral displacements of a wall, ii) horizontal web reinforcement does
not appear to have significant effect on shear capacity, iii) concrete compressive strength
did not affect the strength and deformational response of walls, iv) damage to compressive
zone was greater for walls with lower aspect ratios and increasing axial loads, v) increase
in axial load did not affect the maximum compressive strain but reduced the maximum
recorded tensile steel strain, and vi) shear resistance of the walls is associated with triaxial
compressive stress at the compressive toe of the wall.
82

(a) No axial load (b) 0.1Ag fc0 (c) 0.2Ag fc0

Figure 2.56: Typical damage during monotonic wall tests by Lefas et al. [56].

2.5.5 Lefas and Kotsovos

Four walls were tested by Lefas and Kotsovos [55, 54] to investigate repair methods for
walls. Specimens were tested, repaired, and re-tested. Only the original tests are discussed
here. The tests discussed here were a continuation of testing done by Lefas, Kotsovos and
Ambraseys.
The design of the walls was done using the 1983 ACI Building Code. The amount of
vertical reinforcement was designed for a specific moment capacity at the base of the wall.
The amount of horizontal reinforcement was based on the results of the previous portion of
the research program. The boundary element longitudinal reinforcement is the same as the
monotonically tested specimens but the web vertical reinforcement ratio is lower, resulting
in lower nominal moment demands. The construction, test setup and instrumentation of
the walls is the same and not repeated here.
No axial loads were applied to the walls. Lateral loading of the tests was force-controlled.
Of the four walls tested in this study, one was monotonically loaded and three were cyclically
loaded. The first cyclic tests applied four cycles with equal peak forces of a magnitude less
than yield than pushed the wall to failure. The second cyclic test applied four cycles with
equal peak forces of a magnitude greater than yield than pushed the wall to failure. The
third cyclic test applied cycles of increasing peak forces, with two cycles at each force.
Failure of the compressive zone of the walls was reported for all specimens, occurring
just after the formation of vertical cracks and spalling of the cover concrete. Loss of lateral
83

load carrying capacity occurred at drifts of 1.61 − 1.92%; the monotonically loaded wall
had the lowest drift capacity. The typical crack pattern and final damage of the walls
is shown in Figure 2.57. The researchers concluded that the strength and deformational
response of walls is independent of the cyclic loading regime used and that the wall capacity
was associated with the strength of concrete in the compression region where the bending
moment is maximum, not the common belief that the strength of the inclined compression
strut controlling the strength of the wall.

(a) Typical crack pattern (b) Typical damage at failure

Figure 2.57: Damage to cyclic wall tests by Lefas et al. [55].

2.5.6 Ali and Wight

Four barbell walls were tested at the University of Michigan by Ali and Wight [3] to in-
vestigate the impact of the number and location of openings on the performance of walls.
A reference specimen, W1, was tested without openings. Walls without openings are not
discussed in detail here.
The design of the wall specimens was based on the Almendral building, built in 1972 and
damaged during the March 3, 1985 Chilean earthquake. The prototype wall was U-shaped
84

with an aspect ratio of 6.0. The five-story specimen was constructed as a barbell wall, due
to limitations in loading capacity, at 1/5-scale and had an aspect ratio of 2.9. The walls
were designed to represent the characteristics of Chilean design and construction practices.
The 1988 Uniform Building code was followed, although some criteria were not considered
if inappropriate for a scaled model. The final design had a lower reinforcement ratio than
that of the prototype to prevent a shear failure of the specimen. The cross-section of the
wall is shown in Figure 2.58a

(a) Cross-section and strain gauge layout. (b) Damage at 1.5% drift.

Figure 2.58: Ali [3] specimen W1.

The wall was cast vertically in five lifts, one for each floor. The foundation block was cast
separately from the first floor. The concrete design strength was 5.0 ksi, with a maximum
aggregate size of 3/8 inches (9.5 mm). Longitudinal reinforcement was continuous from the
foundation block to the top of the wall. Grade 60 steel was specified for all reinforcement,
85

but was only actually achieved for the #4 bars used for the boundary element longitudinal
reinforcement. Plain #2 bars were used for the vertical and horizontal web reinforcement.
Confining reinforcement was 3/16 in (4.8 mm) wire.
The foundation block was bolted to the laboratory floor and lateral load was applied
to a transfer beam that transferred load to the wall at four points. At the height of the
load transfer, supports were provided to prevent out-of-plane deformation of the wall. Axial
load was applied using pre-stressing strands that were anchored to the base block. Strain
gauges were provided on boundary element longitudinal reinforcement, web longitudinal
reinforcement and horizontal reinforcement. Layout of the gauges is shown in Figure 2.58a.
Displacement transducers were used to measure the lateral displacement at each floor and
the vertical displacement of the first floor.
Prior to testing, vibration tests were done to determine the frequency and damping
characteristics of the walls. A constant axial load of 0.08Ag fc0 was applied to the wall. The
lateral displacement of the test was a displacement-controlled reverse cyclic history with
two cycles were completed at each displacement level. Cycle increments were 0.25% drift
up to 1.0% and 0.5% thereafter. The tests terminated at 3.0% drift due to actuator capacity
limitations.
In the reference specimen, W1, the development of diagonal struts was clearly visible in
the upper stories of the walls. The construction joints tended to disrupt the progression of
the cracks and re-direct them along the joint. Sliding was observed along the construction
joint between the first and second stories but the overall impact on the displacement of the
wall was minimal. Loss of lateral load of the specimen did not occur due to the loading
limitations mentioned previously. Figure 2.58b shows the damage to the wall at 1.5% drift.
Conclusions/observations reported by the researchers: i) location of door openings did not
affect the flexural behavior of walls, ii) staggered openings are a viable alternative to aligned
door openings, iii) walls with openings had compressive zone failures that were likely the
result of the close proximity of wall openings, iv) strain gauges in wall without openings
indicated truss action, and v) the plastic hinge length extended just higher than the first
floor (≈ 0.6lw ).
86

2.5.7 Pilakoutas and Elnashi

Nine rectangular walls were tested by Pilakoutas and Elnashi [86, 88, 87] to quantify the
ductility and energy absorption of reinforced concrete walls. Three walls were small scale
and not discussed in this literature review. Six walls 1/2.5 scale walls were tested under
cyclic loading and are discussed here. The primary variable tested was the amount of shear
reinforcement in the wall.
Design of the six 1/2.5 scale specimens was subdivided to three pairs of walls (SW4 and
SW6; SW5 and SW7; SW8 and SW9). Each pair of walls had the same flexural reinforcement

but different shear reinforcement. The selection of the flexural
I 6F reinforcement for each pair
5HT6EF


I 6F 5HT6EF 5R6ES
of walls was done by concentrating
I 6F the steel in boundaryI elements.
5R6ES
5HT6EF
-I
I 4HT]2ES
The confinement of the
2 HT16 I
I -I 5R6ES 5 liT 6 EF 9
2 HT16 I
boundary regions varied as I Ia 4HT]2ES
result
-I of the different shear
I 4HT]2ES
5 liT 6 EF
reinforcement.
2 HT16 I
Figure 2.59 shows
9 R4 EF
5 liT 6 EF 9 R4 EF
the reinforcement layout of the six walls. 5 liT 6 ES 5
5 liT 6 ES 5R4ES I
5 liT 6 ES 5R4ES I

Ii.. 11

I 6F
5HT6EF
I
11
6F 5HT6EF 2HT6EF 5HT6EFfl

I 6F

I
I
-I
4HT]2ES I
I
-I
5HT6EF
4HT]2ES
5R6ES

5 liT 6 EF
5R6ES 2HT6EF
2HT6EF 2 HT16 I

5 liT 6 EF
4 HT 12 ES
2 HT16 I
5HT6EFfl
9 R4 EF 4 HT 12 ES
9 R4 EF
5HT6EFfl

Ii.. 11
Ii. .
9R4EF
2HT16EF}I
I_I 'I I WI5H
t HI-
5R6ES 2HT16EF}I t HI---
H
I
4 HT 12 ES
I
I 4HT]2ES
-I
2 HT16 I
9R4EF 2HT16EF}I
I_I
I_I t HI---
'I I WI5HT6EF
9R4EF ' I WI5HT6EF

H H
I

5 liT 6 EF
5 liT 6 ES 5 liTEF
9 R4 6 ES 5R4ES 5R4ES I I
I I
5 R4 ES i II1OH
5 R4 ES i II1OHT6ES
5 R4 ES i II1OHT6ES
5 liT 6 ES 5R4ES I
(a) SW4 (b) SW5 (c) SW6

2HT6EF 2HT6EF 5HT6EFfl


5HT6EFfl

Ii.. 11
Ii. . 11
Ii.. 11
6 HT 10 ES 6HTIGEPI
6 HT 10 ES 2HT16EF}I
2HT6EF
4 HT 12 ES
45HT6EFfl
HT 12 ES 6 HT 10 ES2HT16EF}I 2 HT 6 EF 6HTIGEPI
t HI---
6HTIGEPI 2 HT 6 EF
9R4EF
I_I
I_I t HI---
'I I WI5HT6EF
4R
9R4EF 2 HT 6 EF ' I WI5HT6EF 2 HT 6 EF 2 HT 6 EF

H H
I
2 HT 6 EF 4R4EF 16 R 4 EF @
I I
4R4EF
12 R4 ES
4 HT 12 ES @70
2HT16EF}I
I_I
t HI--- 12 R4 ES 12i R4 ES 16 R 4 EF 16 R 4 EF @100
@140 @140
9R4EF ' I 5 R4 ES
WI5HT6EF II1OHT6ES
8R4EF@70 @70
5 R4 ES @100

H
I I
i @100
II1OHT6ES @100
8R4EF 8R4EF 9R4ES
@100 @100 @70
5 R4 ES i II1OHT6ES 9R4ES 9R4ES
@70 @70
9
9 9

3.10 Reinforcement details for walls SW4 SW9
6 HT 10 ES (d) SW7 (e) SW8 (f) SW9
3.10 3.10 Reinforcement
6HTIGEPI
Reinforcement details fordetails for walls
walls SW4 SW4SW9 SW9
6 HT 10 ES 6HTIGEPI
2 HT 6 EF 2 HT 6 EF
4R4EF
6 HT 10 ES 2 HT 6 EF
Figure 2.59: Reinforcement layouts of Pilakoutas and
6HTIGEPI
Elnashai [88] wall tests.
2 HT 6 EF 16 R 4 EF 4R4EF @140
12 R4 ES
2 HT 6 EF @70
2 HT 6 EF @100
4R4EF 16 R 4 EF @140 81 81
8R4EF 12 R4 ES
@140 @70
12 R4 ES
@100
16 R 4 EF @100
9R4ES
8R4EF @70 @70
@100
8R4EF @100
9R4ES
@100 @70 9
9R4ES
@70
9
3.10 Reinforcement
9 details for walls SW4 SW9

87

The walls were monolithically cast, along with the foundation block and cap beam, in a
horizontal position. The concrete compressive strength ranged from 4.6 − 6.6 ksi (32 − 46
MPa) with a maximum aggregate size of 0.4 inches (10 mm). Longitudinal reinforcement
was continuous from the foundation block to the wall cap with yield strength of 62 − 78 ksi
(430 − 535 MPa).
The walls were tested vertically against a reaction wall (Figure 2.60). The foundation
block was pre-stressed to the laboratory floor. A single lateral force applied to the cap beam
resulted in a shear span ratio of M/(V `w ) = 2.0. LVDTs were used to measure vertical and
horizontal displacement on both sides of the wall at the base, wall cap, 1/4 height and 1/2
height points. Strains were measured at 30 locations. No axial loads were applied to the
walls. Lateral load application was displacement-controlled, with two cycles at each 2 mm
increase in the maximum displacement.

Figure 2.60: Pilakoutas and Elnashai [88] test setup.

Figure 2.61 shows the damage to the six walls at failure of the walls. Specimen SW5
had a diagonal-tension shear failure at 0.83% drift. SW7, which differed from SW5 only in
that it had a greater percentange of shear reinforcement (at greater spacing), failed due to
bar fracture of the shear reinforcement at a drift of 1.83%. The SW4 and SW6 pair of walls
failed due to crushing in the boundary elements at drifts of 2.0% and 1.83%, respectively.
Both SW8 and SW9 reached a drift of approximately 2.2%. Failure of the two walls was
reported by the researchers when the lateral load was 25% of the maximum; however, the
damage causing failure was not specified and had to be inferred from the provided damage
maps. For SW8, failure appeas to have been due to loss of cover at the wall toes. For SW9,
88
4.(4).3 Crack pattern of wall SW4 at MDL-8 and MDL-16

Crack pattern of wall SW8 at MDLI and MDL-12

(a) SW4 (b) SW5 (c) SW6


Figure 4.(4).4 3 Crack pattern
Crack pattern of wall SW4 at MDL-22 of wall
and4.(6).4
Figure failure SW5 at MDL-8 and MDL-1O
Crack pattern of wall SW6 at MDL-18 and MDL-22
Following failure in one direction, it was decided to avoid cycling the
wall twice at the same MDL and failure in the other direction was
immediate after exceeding MDL-10
4.6.2 Load-displacement 108 KM. The last MDL
at a load of about curves
imposed in both directions was MDL-12. For the sake of demonstrating the
effect of cycling on a wall that failed in shear, displacements
The load-displacement were
curves are given in figures A.(6).1 - A.(6).1O
increased by 2 mm in each fordirection as shown
displacements in 12.
3 to
103 figure
Base4.(5).1, untiland
vertical theout-of-plane displacements are
limits of the control transducers
shown inwere reached.
figures The state
A.(6).11 of the wallrespectively.
and A.(6).12, at The wall extensions at
MDL-14 is shown in figure top,4.(5).4,
mid anddemonstrating failure
quarter height in the forward
locations is shown in figures A.(6).13 - A.(6).15,
direction which was caused by the crack originating at the top left corner.
respectively.
Again two main cracks have opened considerably. After failure, several
hoops and stirrups opened up and considerable opening was noticed in the
middle sections of the wall. Additionally, it could be seen that the wall was
displacing in a rigid body4.6.3
modeStrain
abovegauge results
the main cracks. The considerable
degradation of the wall(d)atSW7
MDL. 26 is shown in figure(e) SW8
4.(5).4. (f) SW9
4.(7).4 Strainofgauge
Crack pattern readings
wall SW7 are plotted against force in figures 4.(6).16 to
at MDL-22
4.(8).4 Figure
Crack2.61:
pattern of wall
Pilakoutas
4.(6).38. SW8 at
and
The value MDL-18 and
Elnashai
of strain failure
[88]
is given wall test damage
in microstrain. atstrain
Most of the failure.
Crack pattern
gauges were distributed along the heightof wallofSW9
the at MDL-18
wall on theand failure
flexural
4.7.2 Load-displacement curves
reinforcement. Three strain gauges were located on the stirrups and two on
the hoop reinforcement near the bottom.
The load-displacement curves are given in figures A.(7).1 - A.(7).1O
for displacements
failure appears to have 4.9.2
3 to beenLoad-disDiacement
12. Base vertical
due to and
damage curves
out-of-plane displacements
to the are
web concrete. Pilakoutas and Elnashai
shown in figures A.(7).11 and A.(7).12, respectively. The wall extensions at
heightThe load-displacement 119 in figures A.(9).1 - A.(9).1O
made thetop,
following observations/conclusions:
mid and quarter locations i)curves
is shown in figures the are- given
amount
A.(7).13 of shear reinforcement impacted
A.(7).15,
respectively. for displacements 3 to 12. Base vertical and out-of-plane displacements are
107
the failure mechanism shown
of theinwall but
figures not the
A.(9).11 strength
and A.(9).12 and deformation
respectively. characteristics,
The wall extensions at ii)
top, half and quarter height locations is shown in figures A.(9).13 - A.(9).15,
failure of4.7.3
theStrain
specimens occurred when the shear resistance of the concrete in compression
gauge results
respectively. 112
was exceeded,Strain
iii)gauge
concrete
readingsdilation
are plottedcaused extension
against force of thetowall in longitudinal and lateral
in figures A.(7).16
A.(7).41. The value of strain is given in microstrain. Almost all the strain
directions, and
gauges iv)wall
in this vertical
lasted extension
4.9.3 until
Shear
ofplastic
considerable
and flexural
the wall were
strains
deformation
occurred
achieved.after yield and was due to the
comDonents
Seven strain gauges were located on stirrups and one on the bottom LHS
accumulation
hoop. of plastic strains in the plastic hinge region.
As discussed in section 3.6, shear deformations can be separated
from the total horizontal deformations after calculating the flexural
contribution. The flexural contribution in scale 1:5 walls was calculated by
using area 'al' of figure 3.12 since only top displacements were available.
The shear deformation diagrams as calculated by this method are given in
sections 4.2 and 4.3. For scale 1:2.5 walls, the flexural contribution could be
116
obtained by using the improved approximation of area 'cz2'. The results of
wall SW9 can be used as an example to demonstrate the differences between

123
89

2.5.8 Thomsen and Wallace

Four walls (two rectangular and two T-shaped) were tested by Thomsen and Wallace
[104, 103] to verify displacement based design approaches proposed by the Wallace and
Moehle [113]. In addition to verifying the design approach, the study looked at the type of
confinement used.
The specimens were approximately one-quarter scale versions of walls in a 6-story pro-
totype office building in aRW1—design
Fig. 3. Specimen regionstrain
ofprofile
high seismicity designed per the equivalent static load
Fig. 5. Specimens RW1 and RW2—computed moment–curvature
relations

procedure verse
in UBC-91. Detailing
reinforcement for the ofwasthe
boundary element about walls
the same was done for a design drift of 1.5% drift using
for the two specimens 共0.0116 for RW1 and 0.0132 for RW2兲, the 共0.000 56/in.兲. The moment capacities of the walls at the target
tighter spacing used for RW2 was expected to delay the onset of curvature are approximately 4,700 in. kips 共531 kN m兲; therefore,
the proposed procedure which determines region requiring special confinement based on the
buckling for the two longitudinal bars closest to the wall bound- the predicted failure load for the relatively slender walls
ary, and result in a slightly greater displacement capacity. For (h w /ᐉ w ⫽3), given that a single lateral load was applied at the
practical reasons, the hoops used on RW1 and RW2 extended top of the wall during testing, is approximately 33 kips 共147 kN兲.
strain distribution. In the rectangular walls, RW1 and RW2, shown in Figure 2.62, the
around all eight vertical bars at the wall boundary, with the cen- The shear stress normalized to the square root of the concrete
terline of the hoop approximately 7 in. 共178 mm兲 from the edge of strength associated with this failure load is
the wall; therefore, the transverse reinforcement extended beyond 2.72冑 f c⬘ psi (0.23冑 f c⬘ MPa). According to ACI 318-99, equation
volume of confinement in the boundary regions was kept consistent but spacing and use of
the length required 共4.4 in. or 112 mm, as noted above兲, to confine 21-10, adequate shear strength was provided to avoid shear fail-
the boundary region with compression strains 共for 1.5% drift兲 ure prior to reaching the ultimate curvature (V n ⫽62 kips or 275
between 0.0015 and 0.0055 共Fig. 3兲.
cross-ties was varied. The detailing of one T-wall, TW1 was the same as the rectangular
Moment–curvature relations for RW1 and RW2 were com-
kN兲.
Strain distributions for the design displacement are presented
puted using BIAX 共Wallace 1992兲 and are plotted in Fig. 5. in Fig. 7 for the T-shaped wall. For this analysis, the effective
Stress–strain relations were developed for 4 ksi 共27.6 MPa兲 con-
walls while the other, TW2, was detailed per the proposed design procedure. Figure 2.63 flange width for both tension and compression was assumed to
crete and for Grade 60 reinforcement 共Fig. 6兲. Confined concrete include the entire flange based on ACI 318-99 requirements. For
was incorporated using the model proposed by Saatcioglu and the case with the flange in tension, it was necessary to include the
shows the cross-section of the T-walls.
Razvi 共1992兲. The plots reveal that the rectangular walls have influence of confinement on the stress–strain model for the con-
sufficient curvature capacity to reach the ultimate curvature crete in compression given the relatively large compressive

Fig. 4. Specimens RW1 and RW2—boundary reinforcing details


Figure 2.62: Rectangular specimens tested by Thomsen and Wallace [104].
620 / JOURNAL OF STRUCTURAL ENGINEERING © ASCE / APRIL 2004

Downloaded 17 Apr 2010 to 128.95.204.63. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyright

Walls were cast vertically with lifts for the foundation and each of the four floors. Slabs
were constructed at each level for the T-walls. Concrete compressive strength ranged from
4.2 − 6.3 ksi (28.6 − 43.6 MPa) Reinforcement was continuous from the foundation block
to the top of the wall and was specified as Grade 60 steel; the 3/16 diameter wire used for
90

Fig. 8. Reinforcing details—共a兲 Specimen TW1 and 共b兲 Specimen TW2


Figure 2.63: T-wall specimens tested by Thomsen and Wallace [104].
longitudinal reinforcement, 共2兲 deformed No. 2 共6.4 mm兲 bars for
uniformly distributed horizontal and vertical web reinforcement,
and 共3兲 3/16 in. 共4.75 mm兲 diameter smooth wire for boundary
transverse reinforcement. The boundary transverse reinforcement
was heat treated to produce material properties similar to those of
Grade 60 共414 MPa兲 reinforcing steel. Fig. 10共b兲 plots the mea-
confinement was heat treated to have properties similar to Grade 60 steel.
sured stress–strain relations for each type of reinforcement.

The walls were tested vertically against a strong wall as shown in Figure 2.64. The foun-
Testing and Instrumentation
The wall specimens were tested in an upright position 共Fig. 11兲. A
specially fabricated steel load transfer assembly was used to
dation block was
transfer secured
both axial to tothe
and lateral loads laboratory
the wall specimen. An axial floor and out-of-plane supports were provided.
load of approximately 0.10A g f ⬘c was applied at the top of the wall
Axial load by
The
hydraulic jacks mounted on top of the load transfer assembly.
wasaxialapplied usingconstant
stress was maintained post-tensioning
throughout the duration cables. Lateral
Fig. 9. Specimens TW1 loads were applied
and TW2—computed to the top of
moment–curvature
relations
of each test. Cyclic lateral displacements were applied to the
the wall through a pair of steel channels, creating a shear span ratio of M/(V `w ) = 3.0. A
622 / JOURNAL OF STRUCTURAL ENGINEERING © ASCE / APRIL 2004

set of LVDTs measured vertical displacements at the base of the wall to produce curvature
Downloaded 17 Apr 2010 to 128.95.204.63. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyright

profiles. Sets of diagonal wire potentiometers measured shear deformation in the two two
floors of the wall. Steel strain gauges and embedded concrete gauges were also used.
Constant axial loads of approximately 0.1Ag fc0 were applied to the walls prior to applica-
tion of the displacement controlled lateral load history. Two cycles were performed at each
drift increments of 0.1%, 0.25%, 0.50%, 0.75%, 1.0%, and increasing increments of 0.5%.
91

In the rectangular walls, yielding, vertical splitting and minor crushing were observed
at the same drifts in each test. Specimen RW1 failed due to bar buckling at 2.5% drift.
Specimen RW2 had a delayed onset of bar buckling and also failed at 2.5% drift, but
during the second cycle to this drift level. Specimen TW1 (poor detailing) failed due to bar
buckling at a drift of 1.25%. Failure due to bar buckling also occurred in specimen TW2,
but only after multiple cycles at 2.5%. Figure 2.65 shows the walls at failure. Thomsen
and Wallace made the following observations/conclusions: i) smaller spacing of transverse
reinforcement delayed the onset of bar buckling, ii) at 1.5% drift (used for design), there was
good agreement between the experimental and analytical strain profiles for walls designed
using the proposed procedure, and iii) detailing per the proposed design procedure was
found to be conservative for rectangular walls and adequate for T-shaped walls.

2.5.9 Tupper

Three rectangular walls were tested by Tupper [106] at McGill University (Montreal, Canada)
to evaluate the performance of walls with structural steel boundary elements. Two walls
were tested with structural steel boundary elements and one wall was tested, as a reference
specimen, using conventional boundary element reinforcement. Only the latter wall (W3)
is discussed in detail here.
Design of the specimen followed the requirements of the CSA Standard A23.3 (1994)
for ductile flexural wall systems with a force modification factor of R = 3.5. A prototype
wall was designed and the specimen was constructed as approximately a 1/2 scale version
of the prototype. The axial load used in design was approximated as the total gravity load
of a 12-story structure. Design of the individual specimens were done to achieve the same
moment demand. The spacing of distributed reinforcement in the web was based on the
minimum code requirements. In the conventionally reinforced wall (Figure 2.66), 8 20M
bars in each boundary element were confined by 6 mm diameter ties spaced at one-half the
wall thickness in the plastic hinge region of the wall. Shear reinforcement was provided by
pairs of 10M bars.
Details of specimen construction were not specified by Tupper. Specified concrete
Fig. 12. Applied displacement history

walls by a 125 kip 共556 kN兲 hydraulic actuator mounted horizon-


tally to a reaction wall 12 ft 共3.66 m兲 above the base of the wall.
Out-of-plane support was provided to prevent twisting of the wall
Specimen during testing. The displacement histories applied to
Specimens RW2 and TW2 are shown in Fig. 12. The displace-
ment history for RW1 was similar to RW2, except that the four
additional cycles at 1% and 1.5% drift were not applied after
applying the first two cycles at 1.5% drift. All measurements were
read by a computer data acquisition system at 32 points through-
Fig. 10. Material stress–strain relations: 共a兲 concrete 共b兲 reinforce- out each cycle 共Fig. 12兲. Displacement histories for TW1 and
ment TW2 are identical to RW2. 92

Fig. 11. Specimen test setup—RW1 and RW2

Figure 2.64: Experimental test setup JOURNAL


of Thomsen and Wallace walls [103].
OF STRUCTURAL ENGINEERING © ASCE / APRIL 2004 / 623

Downloaded 17 Apr 2010 to 128.95.204.63. Redistribution subject to ASCE license or copyright; see http://pubs.asce.org/copyright

Fig. 15. Flange reinforcement tensile strains—Specimen TW2

maximum hoop/crosstie spacing of 6d b is used in ACI 318-99 to


suppress buckling of vertical bars. Spacing of special boundary
element transverse reinforcement exceeded this limit for RW1
and TW1 (8d b ), was slightly less than this for RW2 (5.33d b ),
and considerably less than this for TW2 (4d b ).
For Specimens TW1 and TW2, the observed behavior depends
on the direction of the applied loads as noted earlier 共Figs. 7 and
9兲. For the case with the flange in compression, the behavior of
Specimens TW1 and TW2 is very similar; the moment capacity of
the walls is slightly greater than that for RW1 and RW2, and
eilestrains—Specimen
strains—Specimen substantial
TW2 TW2 curvature(a) RW1 capacity
ductility (prior toexists. fail- As a result (b)ofRW2 the (c) TW1
poor detailing provided ure)at the boundary of the web opposite the
flange Fig. 16. Specimen 共a兲 RW1 damage—2% drift; 共b兲 RW2 damage—
sile strains—Specimen TW2for Specimen TW1, the lateral load capacity dropped sud- 2.5% drift; 共c兲 TW1 damage—1.25% drift; and 共d兲 TW2 damage—
denlytoattoan applied lateral drift of approximately 1.25%. The lat-
6db isb isused
used in in
ACI ACI 318-99
318-99 2.5% drift
pacing eral load at this drift level, approximately 65 kips 共289 kN兲, is
Spacingofofspecial specialboundary
boundary
ceeded this roughly equal to the nominal wall shear strength 共62 kips, 276
6d b is used
xceeded thislimitACIfor318-99
inlimit forRW1 RW1 to
an this kN兲, but considerably less than the shear strength computed using
than
Spacing thisforofforRW2 RW2(5.33d
special (5.33d b ),b ),
boundary
W2 (4d measured material properties. Therefore, the loss in lateral load ture during the first cycle to 2.5% lateral drift. As expected, the
TW2
exceeded (4d b ).bthis
). limit for RW1
observed capacity is attributed to the large spacing of transverse reinforce- behavior of specimen RW2 was slightly better then specimen
hethan this behavior
observed RW2depends
forbehavior depends
(5.33d b ), at the wall boundary, which was inadequate to suppress
ment used RW1 due to the closer hoop spacing. Fig. 16共b兲 reveals that, at
sas
TW2 noted (4dearlier
noted ). 共Figs.
bearlier 共Figs. 7 and
7 and
mpression, the behaviorbucking of of the vertical web 共No. 2兲 and boundary 共No. 3兲 rein- failure, all eight longitudinal bars of Specimen RW2 buckled si-
compression,
he observed behavior the behavior depends of
forcement given the deep neutral axis depth and associated high multaneously. The brittle failure of Specimen TW1 is shown in
slar;
as the
milar; the
noted moment
moment 共Figs.
earliercapacity
capacity 7ofandof
tcompression,
forforRW1 compressive strains 共Fig. 7兲. By applying displacement-based de- Fig. 16共c兲, where buckling of all boundary bars, as well as some
hat RW1and theRW2,
and RW2,and
behavior and of
ymilar; sign, transverse reinforcement at the boundary opposite the flange web vertical bars, is observed. The relatively large spacing of
ityexists. theAs
exists. Asa aresult
moment result ofofthethe
capacity of
ry offor was placed at a closer spacing and over an increased depth of the hoops and crossties at the boundary, and the large spacing of the
dary
hat ofthethe
RW1 webweb opposite
and RW2,theand
opposite the
load capacity dropped cross
sud- section Fig. forFig.
Specimen
16.16.Specimen TW2共a兲compared
Specimen 共a兲
RW1RW1 with Specimen
damage—2%
damage—2% drift; 共b兲
TW1
drift;
(d) 共b兲
RW2
TW2RW2 horizontal web bars, allowed buckling to occur as expected. The
damage—
damage—
al
city load capacity
exists. As adropped
result ofsud- the
roximately 关Fig.lat-8共b兲兴. Fig. 2.5% 14共d兲
2.5% drift;
drift;共c兲共c兲
indicates
TW1TW1 that Specimen TW2
damage—1.25%
damage—1.25% drift;did
and
drift; not
and共d兲共d兲
ex-
TW2TW2 effectiveness of the transverse boundary reinforcement for speci-
damage—
damage—
dary of the1.25%.
pproximately 1.25%.
web TheThe lat-
opposite the
perience a loss2.5%of
2.5% lateral
drift
drift load capacity共a兲 until the second and third
共b兲 menbyTW2 is apparent
andinWallace
Fig. 16共d兲, where extensive spalling is
ately
load65capacity
mately
al 65kipskips共289 共289kN兲,
dropped is is
kN兲,sud- Fig. 16. Specimen RW1 damage—2%
Figure 2.65: Damage at failure of walls tested drift; RW2 damage— Thomsen [103].
cycles at a lateral drift level of
共c兲 approximately 2.5%, at which time
共d兲 observed. Lateral load capacity for TW2 was limited by out-of-
strength共62
ear strength
shear
pproximately 共62kips,
1.25%. kips,
The 276276
lat- 2.5% drift; TW1 damage—1.25% drift; and TW2 damage—
the web boundary 2.5%compression
drift zone began to experience an out- plane buckling 共perpendicular to the direction of the applied
ar strength
hear
imately 65 computed
strength 共289 using
computed
kips usingis
kN兲,
of-plane stability failure 共buckling over several hoop spacings兲. It load兲, likely exacerbated by the geometry of the test specimen,
ore,
efore, the
shear the lossloss in共62
strength in lateral
lateral load
kips, load276 ture during the first cycle to 2.5% lateral drift.
ture during the first cycle to 2.5% lateral drift. As expected, the As expected, the
ing is noted that the peak moment strength of TW2 with the flange in with a very narrow confined core once the cover concrete spalled
hearofstrength
acing oftransverse
transverse reinforce-
computed reinforce-
using behavior
behaviorofofspecimenspecimenRW2 RW2was wasslightly
slightlybetter
betterthen thenspecimen
specimen
was inadequate tension does not develop until the drift cycles of 2.5% were im- off.
ch was
efore, the loss into to
inadequate suppress
suppress
lateral load RW1 RW1
ture duedueto tothe
during thecloser
the closer
first hoop
hoop
cycle tospacing.
spacing.
2.5% Fig.
Fig.16共b兲
lateral 16共b兲reveals
drift. Asreveals that,
expected,that,atthe
at
posed. This is a result of the gradual yielding of the tension rein- The
兲acing
and andboundary
boundary
of transverse 共No.
共No.3兲 3兲rein-
rein-
reinforce- failure,
failure,
behaviorallalleight
ofeightlongitudinal
longitudinal
specimen RW2 barsbars
was ofofSpecimen
Specimen
slightly RW2
betterRW2 buckled
then buckled si-si-test results reported in Figs. 14 and 15, as well as the
specimen
forcement within the flange, with reinforcement closest to the photos
xis
chdepth
was andandassociated
depthinadequate associated highhigh
to suppress multaneously.
multaneously.
RW1 dueyielding toTheThebrittle
the brittlefailure
closer failure
hoop ofofSpecimen
spacing.Specimen
Fig. TW1
16共b兲TW1 is isshown
reveals shown inshown
that, in
at
in Fig. 16 reveal that the use of displacement-based
web–flange intersection first, and subsequently progress- design was an effective tool in assessing detailing requirements to
93

Figure 2.66: Cross-section of conventionally reinforced wall tested by Tupper [106].

strength for the walls was 5.1 ksi (35 MPa) and steel properties were per CSA Standard
G30.18. Compressive strength, splitting strength, modulus of rupture and shrinkage tests
were done for the concrete.
The walls were tested horizontally, with the foundation block post-tensioned to the
laboratory reaction floor using high-strength threaded rods. Axial load was applied to the
wall using four prestressing strands. A lateral force was applied to the top of the specimen
at a height that produced a shear span ratio of M/(V `w ) = 3.75. A steel frame was used
to prevent out-of-plane movement of the walls. Lateral displacements were measured at the
base and top of the wall. Vertical displacements were measured at four locations along the
edges of the wall to determine curvature of the walls. Vertical and diagonal displacements
were measured in the plastic hinge region to determine shear contributions. Movement of the
base block relative to the reaction floor was measured and found to have an insignificant
impact on the wall displacement. Steel strains gauges were placed on i) the outer most
longitudinal reinforcing bars, ii) horizontal reinforcement in the plastic hinge region, and
iii) confining hoops in the plastic hinge region. Concrete strain gauge rosettes were also
used in the plastic hinge region.
A constant axial load of 0.1Ag fc0 was applied prior to application of a reverse cyclic
lateral load history. The initial cycles were force controlled to i) half the cracking moment,
ii) the cracking moment, iii) first yield of longitudinal reinforcement and iv) general yield,
defined as a significant drop in specimen stiffness. The remaining cycles were displacement
controlled as multiples of the displacement at general yield (1.25, 1.5, 2, 2.5, 3, and 3.5).
94

Only one cycle at each force/displacement level were completed.


For the conventionally reinforced wall (W3 ), loss of lateral load carrying capacity oc-
curred at 2.9% drift due to core crushing and bar buckling in the boundary element. The
damage immediately following failure is shown in Figure 2.67. Tupper concluded that walls
with structural steel shape boundary element reinforcement had ductility and energy dissi-
plation comparable to a conventionally reinforced wall.

Figure 2.67: Failure of Tupper conventionally reinforced wall [106].

2.5.10 Tasnimi

Four rectangular walls were tested by Tasnimi [102] at Tarbiat Modarres University to
evaluate the Iranian seismic design code. All four walls had the same design but were
tested with different load histories.
The strength and dimensions of the walls were designed at 1/8-scale based on the capac-
ity of the testing facilities, with reinforcing details designed in accordance with the ACI 318
building code. The boundary elements of the walls did not have confining reinforcement.
The reinforcement layout of the walls is shown in Figure 2.68a.
95

314 A.A. Tasnimi / Engineering Structures 22 (2000) 311–322

316 A.A. Tasnimi / Engineering Structures 22 (2000) 311–322

Fig. 1.
(a)Geometry and reinforcement details of wall specimens.
Reinforcement layout
Table 1 crete properties are summarized in Tables 2 and 3,
Properties of reinforcing bars respectively.
Type Yield strength fsy Ultimate strength fsu
(Mpa) (Mpa) 3.2. Loading history and testing procedures

3 mm diameter bar 216 317 To simulate loading sequence that might be expected
6 mm diameter bar 276 475 to occur during an earthquake, four simplified types of
horizontal cyclic loading history were adopted. Since no
standard cyclic test procedures has been introduced for
The concrete for all specimens was made from type testing reinforced concrete structures, the horizontal load
(I) Portland cement, river sand and 10 mm maximum
side crushed gravel. Measured slumps ranged from 60 Table 2
to 180 mm for all specimens. From the six 150 ⫻ 300 Concrete mix proportion by weight
mm concrete cylinders, three were tested in compression Properties by weight Specimens SHW1 Specimens SHW3
at 28 days and the remaining three used for the tensile and SHW2 and SHW4
splitting test. Also compression tests were carried out on (kg) (kg)
150 ⫻ 150 mm cubes. The modulus of elasticity of con-
crete was calculated on the basis of data obtained from 10 mm aggregate 825 827
River sand 950 954
cylinder compression tests. The concrete strain corre- Cement 320 325
sponding to its strength was measured for each speci- Free water 250 254
mens. Full details of the concrete mix used and the con-

(b) with
Fig. 3. Schematic view of test setup Testthe setup
positions of strain gauges and transducers.

Table 4 Figure 2.68: Wall tests by Tasnimi [102].


Experimental and calculated principal results in successive loading cycle for specimen SHW1

Cycle no. Horizontal load (kN) and direction Average top Average stiffness Average moment Average applied Moment ratio
deflection of resistance moment

⇒ ⇐ (mm) (kN/m) (1) (2) (1)/(2)

1 1.57 1.77 0.520 3296.5 1.69 1.72 0.97


2 2.95 2.55 0.870 3112.7 2.67 2.79 0.96
3 5.99 6.00 2.390 2508.5 5.92 6.18 0.96
4 10.11 10.11 4.480 2258.3 9.79 10.41 0.95
5 14.34 14.63 13.46 1190.9 14.03 14.92 0.94
6 15.42 14.53 16.45 937.4 14.87 15.88 0.95

these tables, the variation of calculated average secant 5. Discussion


stiffness in successive load cycles are presented. The
applied and resisting moments at the base and their ratio It should be emphasized that all the walls were essen-
indicating good agreement are given in the same tables. tially subjected to the intended in-plane action where the
Fig. 1 illustrates the lateral load versus top displacement out-of-plane displacement was prevented by special
96

Each wall was cast monolithically, including the foundation block. Longitudinal rein-
forcement was continuous from the foundation block to the top of the wall. Compressive
strengths and splitting strength tests of the concrete were done. Concrete strengths ranged
from 3.1 − 3.4 ksi (21.6 − 23.5 MPa). Figure 2.68b shows the test set-up. The lateral loads
were applied below the top of the wall, creating a shear span ratio of M/(V `w ) = 2.20.
No axial loads were applied to the walls. The walls were restrained against out-of-plane
displacement at the point of lateral load application. Lateral displacements were measured
at the height of the lateral load and approximately 5% up the height of the wall. Vertical
displacements of the foundation block were measured but foundation movement was found
to contribute insignificantly to the overall movement of the wall. Strain gauges were in-
cluded at the base of the wall on the boundary element reinforcing bars. Concrete strain
gauges were included on one specimen.
The application of lateral loads was force-controlled, and each wall was tested with a
slightly different load history. The loads increased monotonically, with no repeated cycles.
The variation in the loading was the rate at which the load increased between cycles24 .
The tests were completed when a loss in load carrying capacity was reached. For all
walls, the damage at this point, shown in Figure 2.69 consisted of initiation of spalling and
vertical cracking in the compressive boundary zones of the wall. The ultimate drift capacities
ranged from 0.86%-1.70%, with the lowest drift occurring in SHW2, which was loaded with
the slowest increase in force and thus more cycles were completed prior to reaching the
maximum load capacity. Crack patterns indicated flexurally dominated behavior. The
researchers concluded that strength and deformation response of the walls were independent
of the load histories used.

2.5.11 Mobeen

Mobeen [68] tested barbell walls at the University of Alberta to evaluate the impact of
double-headed studs as a replacement for conventional cross-tie confinement. Two walls
were tested with double-headed studs. A third wall, W1, was tested using conventional

24
The difference between the load history for the four walls was very little and can be considered as
essentially the same for all four specimens
97

Fig. 8. Typical mode of failure for specimens SHW1 and SHW4.


Figure 2.69: Typical damage at base of Tasnimi wall specimens [102].

Table 8
Curvature and displacement ductility of the specimens

Specimen Ductility Calculated ductility based on test results Ratio of calculated


reinforcement and based
is discussed
on ACI here. ductility to ACI

Based on wall curvature Based on wall displacement


The specimen was designed as a full-scale wall from a prototype 6-story, 5 bay residential
␮␾ ␾ *y10−6 ␾ *u10−6 ␮*␾10−6 ⌬y ⌬u ⌬u/⌬y (4)
(1) (2) (3) (4) (5) (6) (7) (1)
building. The
SHW1
design
3.98
was 4.49
completed
8.980
using
2.00
the static
4.48
load
16.45
procedure
3.67
from
0.51
the 1997 Uniform
SHW2 3.45 3.60 8.140 2.26 3.39 11.26 3.32 0.66
Building Code
SHW3 (Seismic
SHW4
4.36
4.69
Zone
4.29
4.89
IV) and detailing
8.980
15.15
2.10
2.10
5.05requirements
4.76
15.50
18.20
of ACI0.46
3.10
3.82 0.45
318-2002. The wall
AVG 4.12 4.32 9.060 2.12 4.42 15.35 3.48 0.45
cross-section is shown in Figure 2.70a.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

(a) Cross Section (b) Test set-up

Figure 2.70: Conventional wall specimen tested by Mobeen [68].

The wall specimens were cast horizontally, with the foundation block cast separately.
Longitudinal reinforcement was continuous from the foundation block to the top of the wall.
Normal weight concrete with compressive strength of 3.48 ksi (24 MPa) was used. 15M bars
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
98

were used for the boundary element longitudinal reinforcement had a yield strength of 55.7
ksi (383.5 MPa). All other reinforcement was 10M bar with a yield strength of 57.3 ksi
(394.5 MPa).
Figure 2.70b shows the test-set up of the wall. The foundation block was attached to the
laboratory strong floor using prestressing rods. Axial loads were applied to a W-shape steel
beam at the top of the wall using four hydraulic jacks. Lateral forces, applied using two dual
action jacks, were resisted by a steel reaction frame. The height at which the lateral force
was applied created a shear span ratio of M/(V `w ) = 2.74. Lateral bracing was provided
near the top of the specimen to prevent out-of-plane movement of the wall. Strain gauges
were provided on the longitudinal reinforcing bars at the base of the wall and on transverse
reinforcement to study the confining behavior of the steel. Vertical displacement of the wall
was measured with respect to the strong floor. Horizontal displacements were measured
with respect to the steel reaction frame. Rotation of the loading assembly was measured.
A constant axial load of 0.15Ag fc0 was applied to the wall. Application of shear force was
displacement controlled in a reverse cyclic loading pattern. Three cycles were completed at
top displacement increments of 3 mm up to 30 mm, followed by cycles of 42, 54, 66, 78, 125
and 135 mm.
The test of the conventionally reinforced wall had reached a drift of 3.9% when the test
was concluded due to concern for damage to the out-of-plane supports of the test set-up.
At a drift of 2.2%, an 8% drop in lateral load occurred when a longitudinal reinforcing
bar fractured in the corner. Figure 2.71 shows damage to the ends of the wall following
completion of the test. Conclusions/observations of the test program were i) double-headed
studs improved the displacement ductility and energy absorption of walls when compared
to a similar wall with conventional confining reinforcement, ii) compared to walls with same
shape, axial load ratio, and shear force-capacity ratio, walls with increasing amounts of
confining reinforcement have increased displacement ductility, iii) yield of the conventional
confining reinforcement occurred at lower displacement ductility than did double-headed
studs, and iv) prior to yield of the longitudinal reinforcement, energy absorption is greater
in the walls with double-headed studs, with energy absorption dropping in both after yield.
99

(a) Fractured bar in south boundary (b) Buckled bars in north boundary
element. element.

Figure 2.71: Damage to wall tested by Mobeen [68].

2.5.12 Riva, Meda, and Giuriani

Riva et al. [92] tested a full-scale rectangular wall at the University of Brescia, Italy to
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

address the lack of large-scale wall tests found in the literature.


The wall, shown in Figure 2.72 was a four-story building with an underground floor,
designed per the European Seismic Code (Eurocode 8) for moderate seismic actions, soil
type B, and medium ductility. The ground floor was designed as a box foundation system
in which the horizontal forces applied to the wall were resisted by diaphragm actions of the
ground and basement level slabs, allowing for reduction in the dimensions of the foundation.
The critical section was at the ground level. The foundation level of the wall exhibited shear
panel behavior and was designed to have a greater thickness than the rest of the wall. The
flexural reinforcement was concentrated in confined boundary regions.
Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Details on the construction of the wall were not provided. The concrete cube compressive
strength was 5.9 ksi (40.7 MPa). Simple supports were provided at the foundation and
ground levels to create the effect of the box foundation system. Due to the size of the
specimen, the wall was tested horizontally and axial load was not applied. Lateral loads
were applied at two points along the height of the wall. The relationship of the forces was
controlled to provide a shear span ratio of M/(V `w ) = 3.17, creating a relationship between
Theoretical and experimental bending moment and shear strength
horizontal displacement.
values: design (MRd, VRd,s); nominal (Mn, Vn); structural yield (My, Vy); The applied load was measured by means of a full
collapse (Mu, Vu) bridge resistive pressure transducer placed on the pump
manifold. The displacements were measured using 17
MRd=4015 kN m VRd,s=715 kN potentiometric transducers as shown in Fig. 6: two wire
Mn=5910 kN m Vn=1165 kN transducers (16, 17) measured the vertical displacement
My=5300 kN m Vy=615 kN
Mu=6200 kN m Vu=720 kN
of the wall; 11 linear transducers (1–8, 14, 15) measured
100 close
the displacements in the upper and lower chords
to the critical section; two linear transducers (11, 12)

Fig. 4. Test set-up: wall supports (a); ribs simulating the ground and basement floor diaphragms (b); post tensioned strands and bars (c); steel
Figure
frame to avoid 2.72:
lateral Wall(d);tested
instability byframes
additional Rivaforetimproving
al. [92]:safety
(a)inwall supports,
the test set-up (e). (b) ribs simulating ground
and basement floor diaphragms, (c) post tensioned strands and bars, (d) steel frame to
avoid lateral instability, and (e) additional frames for improving safety in the test set-up.

the shear and bending forces at the critical section that was equivalent to an inverted
triangular distribution of lateral forces for a 4-story wall. Openings in the wall web were
provided at the loading elevations to accommodate the actuators applying load to the walls.
A steel frame was provided near the load points to provide lateral stability to the walls.
Vertical displacements were measured at the critical section to determine rotation of the
wall. Lateral displacements were measured near the points of load application. The test was
force-controlled up to the theoretical yield load (Figure 2.73a) and displacement controlled
thereafter (Figure 2.73b).
Shear failure of the wall occurred when a longitudinal web reinforcing bar fractured while
unloading from 3.14% drift. Figure 2.74a shows the fractured bar and Figure 2.74b shows
the critical section at the time of failure. The damage to the boundary elemect is shown in
Figure 2.74c. The following conclusions/observations were made following the test: i) after
yield, damage progressively increased with increased displacement, with damage localized at
the critical section, ii) there was no significant strength and/or stiffness degradation in the
cycles following yield, iii) minimum web reinforcement requirements of Eurocode 8 were not
adequate to prevent a shear failure, iv) web reinforcement should be increased to prevent
sliding shear failure, and v) ductility of 3 relative to observed yield.
101

5. Action diagram: comparison between true scheme and the two jacks scheme. 840 P. Riva et al. / Engineering Structures 25 (2003) 835

앫 in order to prevent slidingtility


3

be obtained by either reducing


reinforcement amount should
when axial force in the wall

friction resistance might be overestimated;


conservative design for sliding shear effects, as shear
presented, the adopted approach might lead to non-
normal strength concrete. Based on the results herein
with a slight increase of the friction contribution for
ure, the same formulation adopted in [10] is proposed,
version [10] is given. Concerning sliding shear fail-
same overstrength factor adopted by the 1994 EC 8
tility class M walls, while for class H structures the
2

δ/ δ y
0
24 25 26 27 28 29 30 31 32
-1

-2

-3

-4

cycles

bution, or increasing the overstrength factor when

Research) within the program COFIN-99 “Safety of RC


The research was co-financed by MURST (Italian

criteria of ultimate strength and damage limitation given


structures under seismic actions with reference to design

cestruzzi s.p.a., Ferriera Valsabbia s.p.a., Italcables s.p.a.


Ministry of University, Scientific and Technological

by EC8”. The contribution of UNIECO s.c.r.l., Cal-


앫 in order to prevent sliding shear failure, particularly

Fig. 10. Force F v

Fig. 16.
be obtained by either reducing the friction contri-
reinforcement amount should be increased. This could
when axial force in the wall is negligible, the web

friction resistance might be overestimated;


conservative design for sliding shear effects, as shear
presented, the adopted approach might lead to non-
normal strength concrete. Based on the results herein
with a slight increase of the friction contribution for
ure, the the
version
sameisoverstrength

(a) Force
Fig. 7. Loading control
history up prior to
to the theoretical yield
yield load: M/Myt versus (b) Displacement
Fig. 8. Loading history after thecontrol
structural after yieldd/dy versus
yield load:
shearclass

cycle number. cycle number.


be increased.

load are equal


Figure 2.73: Load history for wall tested by Riva et al. [92]
respectively, wh

Cracking development around critical section from dy to collapse.


negligible, the
failure,

3. Experimental results
samefriction
[10] is given.

displacement is
M walls,

at the critical section), corresponding to an experimental Fig. 10 show


Fig. 9 shows the load versus end displacement curves
formulation

maximum displacement at the wall end (dyt) equal to From this figure

determining the design shear force.


(F–d) for the whole loading history, where F is the net
particularly

cks (a) and downward jacks (b)),


This could

90 mm, was reached. After three cycles with maximum transverse load, equal to the difference between the total
displacement equal to dyt, the load was further increased 앫 due to crac
factor

jack load and the jack load necessary to equilibrate the


contri-adoptedcestruzzi

decreases for
while for class

in both directions until the displacements corresponding wall weight, i.e. the load which annuls the bending
앫 cycles with c
Concerning
webadoptedcriteria

he deformation of the panel to the structural yield in both directions were detected moment in the critical section. The results are discussed

Cracking development around critical section from dy to collapse.


Fig. 16. Ministry

linear transducers (9, 10) (+dy⬇120 mm and ⫺dy⬇⫺120 mm). These displace- in the following, by analysing separately the load cycles
stant dissipa
between the wall and the ments were determined by intersecting two lines tangent damage after
before yielding, typical of service conditions, and after
der to monitor any potential to the load–displacement experimental curve in the II 앫 the wall cyc
yielding, typical of design seismic loads. The results of
by sliding

structures
Research) within the program COFIN-99 “Safety of RC

Acknowledgements

Acknowledgements
the signals were conditioned stage (after cracking) and in the III stage (after yielding).
Cracking development

the experimental test not reported and discussed herein


in [10] iss.p.a.,

by theof1994

The research was co-financed by MURST (Italian

on system (Mod. UPM 100 The following loading history was defined by imposing
determining the design shear force.
bution, or increasing the overstrength factor when

may be found in [12].


EC8”. shear

Table 2
H structures

a PC. cycles of increasing displacement amplitude until failure Dissipated energy


rformed by applying cyclic was reached (Fig. 8). The collapse occurred during the
ude (Fig. 7), until the theor- second cycle at a displacement equal to three times the 3.1. Behaviour under service loads
d +–
of University,
under seismic

Cycle
ultimate
proposed,
The contribution

eoretical load for which the (a) Fractured web bar (b) Critical
structural yield displacement, also equal to four times the section at fail- (c) B.E. damage
the outermost rebar occurs theoretical first yield displacement (3dy⬇4dyture⬇360 mm). The behaviour in service conditions was analysed by
EC strength

applying load cycles lower or equal to the first theoreti-


Ferriera Valsabbia s.p.a., Italcables s.p.a.
fail-

the

cal yield load, defined as the load for which the stress mm
Figure 2.74: Damage to large scaleinwall tested reinforcement
by Riva et at al.the[92]
8

around critical

the external critical section is


equal to Re=560 MPa. In the present case, the theoretical 6 26.
actions

yield bending moment at the critical section and applied 7 26.


tility class M walls, while for class H structures the

reinforcement amount should be increased. This could


version [10] is given. Concerning sliding shear fail-

when axial force in the wall is negligible, the web


ure, the same formulation adopted in [10] is proposed,

conservative design for sliding shear effects, as shear


same overstrength factor adopted by the 1994 EC 8

with a slight increase of the friction contribution for


normal strength concrete. Based on the results herein
presented, the adopted approach might lead to non-

be obtained by either reducing the friction contri-


Scientific

앫 in order to prevent sliding shear failure, particularly


cestruzzi s.p.a., Ferriera Valsabbia s.p.a., Italcables s.p.a.
by EC8”. The contribution of UNIECO s.c.r.l., Cal-
criteria of ultimate strength and damage limitation given
structures under seismic actions with reference to design
Research) within the program COFIN-99 “Safety of RC
Ministry of University, Scientific and Technological

Acknowledgements

8 26.
and
The research was co-financed by MURST (Italian

determining
bution,
of UNIECO s.c.r.l., Cal-

9 26.
sectionand
damage limitation

2.5.13 Liu
with or

10 25.
11 28.
friction resistance might be overestimated;
reference

from Technological

Two walls were tested by Liu [59] at McGill University (Montreal, Canada) to investigate 12 30.
increasing
the design

dy to collapse.

13 61.
the influence of high strength concrete on structural wall ductility.
14 71.
to design

Specimens were designed per a draft version of CSA Standard A23.3-04 Clause 21.6
Fig. 16.

15 77.
given

16 77.
the overstrength factor when

for ductile walls. Dimensions of the walls were selected based on the capacity of the test- 17 78.
shear force.

18 100
ing facility. Longitudinal reinforcement was the same in both specimens. Specimen W1 19 100

was designed with normal strength concrete of 5.1 ksi (35 MPa). Specimen W2 had high 20 179
21 198
22 199
strength concrete of 10.9 ksi (75 MPa). Due to Fig.
the9. increase in end
Force F versus moment capacity,
displacement the
d curve for all shear
the cycles.

reinforcement in W2 was one-quarter higher than in W1. The cross-section of the walls is
102

Figure 2.75: Cross-section of walls tested by Liu [59].

shown in Figure 2.75.


The walls were cast horizontally, with the foundation block cast separately from the
walls. Longitudinal reinforcement was continuous from the foundation block to the top of
the wall. All rebar was met the requirements for CSA Standard G30.18, weldable Grade
400 steel. Compressive strength, splitting strength, modulus of rupture and shrinkage tests
were done for the concrete. The actual compressive strengths were 4.8 ksi (33.1 MPa) for
W1 and 10.3 ksi (70.6 MPa) for W2.
The walls were tested horizontally with the lateral displacement of the wall occurring
perpendicular to the laboratory floor. The foundation block was connected to the reaction
floor using high-strength threaded rods. Axial load was applied using four pre-stressing
strands. The lateral load was applied below the top of the wall specimen, creating a shear
span ratio of M/(V `w ) = 3.125. A steel frame was used to prevent out-of-plane movement
of the wall. Lateral displacement of the wall was measured at the height of the lateral load.
Strains were measured on the outermost longitudinal reinforcing bars and on the confining
hoops and horizontal reinforcement in the plastic hinge region. Vertical LVDTs were used
to measure curvature at two elevations in the plastic hinge region of the wall. Diagonally
mounted LVDT’s measured relative displacements in the plastic hinge region for use in
determining shear deformations.
A constant axial load force of 135 kips (600 kN) was applied to each wall, resulting in
103

axial load ratios of 0.04Ag fc0 and 0.08Ag fc0 for W1 and W2, respectively. The lateral load
was applied in a reverse cyclic pattern that was force-controlled prior to general yield of the
walls, with cycles to i) half the cracking moment, ii) the cracking moment, iii) two times the
cracking moment, iv) first yield of longitudinal reinforcement and v) general yield. The tests
were displacement controlled beyond yield at increments of the general yield displacement.
Three cycles were done for each load/displacement level.
Lateral load carrying capacity was lost at 2.9% drift for both walls. In W1, this was
due to extreme bar buckling and core crushing. In W2, buckling and crushing had also
occurred, but the drop in load was ultimately a result of low-cycle fatigue of the bars. The
final damage to the walls were shown in Figure 2.76. Liu concluded that with increasing
concrete compressive strength: i) displacement and curvature ductilities are increased, ii)
plastic hinge length is decreased, iii) cumulative energy dissipation is increased, iv) initial
loading response is stiffer, and v) cracks initiate at higher loads, develop quicker, and have
smaller crack widths.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

(a) W1 (b) W2

Figure 2.76: Damage to walls tested by Liu [59].


104

2.5.14 Khaliil and Ghobarah

Three walls were tested by Khalil and Ghobarah [53] to investigate failure methods and
repair techniques for walls under earthquake loading. Walls were large-scale models of the
plastic hinge region of a wall and were loaded with a moment and shear at the top to
simulate base reactions for the full wall. Two walls were tested after retrofitting had been
applied and are not discussed here. The third wall was a control wall and is discussed here.
Wall dimensions were designed at one-third scale, with the height of the specimen based
on the 1994 CSA code requirements for wall plastic hinge length and the thickness selected
such that the ratio of the wall length to the wall thickness was less than or equal to 10. The
reinforcement used in the wall complied with the ACI 1968 and CSA 1977 design provisions;
boundary regions were not confined. Figure 2.77 shows the crosss-section of the control wall.

Figure 2.77: Cross-section of wall tested by Khalil and Ghobarah[53].

Details on the construction of the wall were not provided by the researchers. Concrete
compressive strength was 5.5 ksi (38 MPa). Vertical reinforcing steel had a yield strength of
68 ksi (470 MPa) and horizontal reinforcing steel had a yield strength of 87 ksi (600 MPa).
Figure 2.78a shows the experimental test setup. The foundation block was secured to
the laboratory floor and a wall cap was use to apply the loads to the specimen. One actuator
applied the lateral force to the wall while two vertical actuators applied axial and moment
forces. The relationships between the shear and moment forces was controlled such that the
effective shear span ratio was M/(V `w ) = 2.25. Lateral deformations were measured at the
top of the wall. LVDTs measured the elongation and shear deformations for the full wall
height, the bottom half of the wall, and the top half of the wall. Twenty-six strain gauges
were used on the vertical and horizontal steel.
The wall was loaded with an axial load of 0.07Ag fc0 prior to application of the lateral
105

(a) Experimental setup (b) Final damage

Figure 2.78: Walls tested by Khalil and Ghobarah [53].

load and held constant throughout the test. The application of the lateral load was force
controlled prior to yield and displacement-controlled by multiples of the yield displacement
after yield. Throughout the entire test, the moment applied to the wall was a function of
the applied lateral load such that the shear span ratio was kept constant.
The wall reached a maximum lateral drift of 2.5% before a sudden diagonal shear failure
occurred, accompanied by concrete crushing and bar buckling at the toe of the wall. Hor-
izontal reinforcement had yielded but the vertical reinforcement had not yet yielded. The
final damage to the wall is shown in Figure 2.78b. Khalil and Ghobarah concluded from the
full test program that rehabilitated walls sustained 60% more lateral drift and dissipated
significantly more energy than did the control wall.

2.5.15 Ile and Reynouard

Three identical U-shaped walls were tested by Ile and Reynouard [49] at the European
Laboratory for Structural Assessment (ELSA) to investigate the impact of load direction
and bi-directional loading.
The walls were designed by Reynouard and Fardis (2001) to represent full scale walls in
106

an elevator shaft. The walls were designed per the requirements of Eurocode 8 with confined
boundary elements in the flange tips and corners. Figure 2.79 shows the wall cross-section.

Figure 2.79: Cross section of walls tested by Ile and Reynouard [49].

Walls were cast monolithically with a foundation block and wall cap. Compressive
strength of the concrete ranged from 3.0 − 3.5 ksi (21 − 24 MPa). Reinforcing steel was
continuous from the wall foundation to the wall cap. Walls were tested against two reactions
walls with two actuators loading the wall in each orthogonal direction. Axial load was
applied using post-tensioning bars. No details were provided on the instrumentation used
to measure the response of the wall.
All three walls were loaded with axial load equal to 0.10Ag fc0 . Lateral loading was
quasi-static. For the uni-directional tests, the lateral movement was displacement-controlled
and a zero-torsion target between the two actuators was force-controlled. Three cycles
at displacement levels of 1, 2 (weak-axis bending only), 4, 8 and 12 cm were applied.
The bidirectional test had three displacement-controlled actuators, with the forth force-
controlled. The load was applied in a butterfly path at 4 and 8 cm.
The wall tested under strong-axis bending, X, reached a maximum drift of 3.08% and
failed when previously buckled vertical reinforcing bars fractured in the flange tips. Speci-
107

men Y, tested under weak-axis bending, also reached a maximum drift of 3.08% and failed
due to bar buckling in both the web and flange tips. The bi-directional test maximum drift
was 2.03%; failure was due to the fracture of previously buckled reinforcing bars. The dam-
age to the bi-directional test, XY, is shown in Figure 2.80. A drop in strength was noted for
walls that had been previously loaded in an orthogonal direction. Despite the different drift
capacities for uni- and bi-directional loading, Ile and Reynouard calculated a displacement
ductility of 6 for all three walls. The researchers concluded that the design approach for
walls in the newest version of the building code were satisfactory. The experimental results
were used to validate the accuracy of numerical models created as a part of the study.

Figure 2.80: Flanged after bi-directional loading of U-shape wall tested by Ile and Reynouard
[49].

2.5.16 Beyer, Dazio and Priestley

Two U-shaped walls with different thicknesses were tested under bi-directional loading by
Beyer, Dazio, and Priestley [11] at ETH Zurich to provide experimental data on an infre-
quently tested wall configuration.
Specimen were constructed at one-half scale of walls from a 6-story prototype building
containing a single U-shaped wall and a bi-directional moment-frame. The walls were de-
signed for high ductility but detailing did not follow the guidelines of any particular code.
The horizontal reinforcement was the same for both walls and the total area of vertical
FIGURE 4 TUA and TUB: Cross section, shear keys, and elevation of TUA (a-
108

6 × shear keys
(100 × 60 mm, 40 mm deep)

12 × shear keys
(65 × 60 mm, 30 mm deep)
wall

wall
reinforcement was similar between the two walls; the fundamental difference between the

(b)

(e)
footing
(without blocks)

footing
(without blocks)
walls was the thickness. Figure 2.81 shows the wall cross-sections.

1050 1050

25

25
25
100 150 150 150 125 125 100 100 100 50 150 150 150 150 75 75 100
25
25

25
100 50
100
150

shear reinforcement

175

shear reinforcement
650

650

TUB
TUA
150

Ø6 s = 125

150

Ø6 s = 125
ties Ø6 s = 50

ties Ø6 s = 50
shear reinforcement

shear reinforcement
150

tie Ø6 s = 50

150

tie Ø6 s = 50
75

Ø6 s = 125

Ø6 s = 125
tie Ø6 s = 50
12 Ø 6

14 Ø 6

tie Ø6 s = 50
4 Ø 12

8Ø6

7 Ø 12

4 Ø 12

10 Ø 6

6 Ø 12
(d)
(a)

Figure 2.81: Cross-section of walls tested by Beyer et al[11].

The walls
1030were cast separately from the foundation block and shear keys were provided
to prevent sliding shear failure at the wall-foundation interface. A thick concrete beam
Downloaded By: [University of Washington Libraries] At: 19:21 2 July 2010
was provided at the top of the wall to apply loads to the walls. The maximum aggregate
size was 0.63 inches (16 mm). Specified concrete strength was 6.5 ksi (45 MPa), but was
significantly higher for the thicker wall, 11.3 ksi (78 MPa), resulting in higher than intended
shear capacity of the wall. Reinforcing steel was continous from the foundation block to the
cap beam and met the requirements for Eurocode class “C” steel.
The walls were tested vertically against a strong wall and steel reaction frame as shown in
Figure 2.82a. Two actuators were used to apply strong-axis bending to the wall and one to
apply weak-axis bending. The loads were applied at different heights, resulting in shear span
ratios of M/(V `w ) = 2.58 and M/(V `w ) = 3.2, respectively. Axial loads were applied by pre-
tensioned tendons. Lateral displacements were measured at the height of load application
and sliding shear displacements were measured at the wall base. The height of the wall was
divided into thirds, with vertical and diagonal displacements measured. Strain gauges were
109

placed on horizontal reinforcing bars and Demac or Whitmore gauge measurements were
taken to measure strain in the lower portion of the wall.
1032 K. Beyer, A. Dazio, and M. J. N. Priestley

NS-W actuator
(Capacity: 1000 kN, 1200 mm)

Transfer beam
Reaction wall
Reference column
for EW direction
NS-E actuator
(Capacity: 1000 kN,
Tendon for axial load
1200 mm)
1500 kN hollow core jack Strong floor
for constant axial load
Foundation blocks
post-tensioned to
the strong floor
Downloaded By: [University of Washington Libraries] At: 19:21 2 July 2010

Test unit

EW actuator
(Capacity: 1000 kN, 1200 mm) Reaction frame

Quasi-Static Cyclic Tests of Two RC U-Shaped Walls 1033


FIGURE 6 Isometric view of the test setup.
(a) Test set-up

(b) D
on of Line of action of μ = 8.0
tor NS-E actuator
μ = 6.0
G F
Web Line of action μ = 4.0
of EW actuator μ = 3.0
East corner μ = 2.0
μ = 1.0
North A B
O
East flange
West East

South E H
East flange end
C
FIGURE 7 Photo of the test setup.
(b) Load pattern
dinal points and labelling of different wall sections (a) and target
The most important measurements were the actuator forces, the displacements of the
tern (b). Figure 2.82: Beyer
wall head, the sliding displacements at the et
wallal.
base,wall tests [11].
the elongation of the wall edges,
and the shear displacements (Fig. 9). The elongation of the edges was measured by four
chains of linear variable differential transformers (LVDTs). Each chain consisted of

The same axial force was applied to both walls, resulting in load ratios of 0.02Ag fc0
LVDTs for
ormeasuring
0.04Ag fc0 . Lateral load application was force-controlled at 25, 50, 75 and 100% of the
global displacement NS-W actuator (b)
against reference yield load and displacement-controlled thereafter. The loading pattern, shown
predicted
columns
in Figure 2.82b, consisted of displacements i) parallel to the web (strong-axis bending), ii)
NS-E actuator
parallel to flange (weak-axis bending), iii) diagonal with one flange end in compression, iv)
EW actuator
Demec measurement
points on the
1000

inner faces of the LVDT chains


web and flanges along the edges
110

diagonal with one corner in compression, and v) clover leaf sweeping cycle.
The thicker wall, TUA, reached a maximum drift of 3.1% and failed due to bar buckling.
The thinner wall had a maximum drift of 2.7% and failed due to crushing in the unconfined
region of the wall web. Beyer et al concluded that the most critical direction of loading was
the diagonal direction because the moment and displacement capacity under this loading
were less than anticipated.

2.5.17 Brueggen

Two large-scale T-shaped walls were tested by Brueggen [15] at the University of Minnesota
NEES MAST facility to investigate the effects of multi-directional loading. The walls were
different heights with the same effective shear span ratio. One wall was constructed with
lap splices in the second story.
The wall specimens were one-half scale of a 6-story prototype building, designed following
the guidelines of ACI 318-02 and IBC 2003 (seismic category D). Shear reinforcement of
the walls was designed to allow development of the plastic moment-capacity of the wall
with an inverted triangular load distribution. For the scaled specimen, the clear cover
and confinement spacing requirements of ACI 318 were also scaled. Figure 2.83 shows the
cross-sections of the two walls. NTW1 (4-stories) was based on the prototype wall, with
flexural reinforcement in the flange concentrated in boundary elements. NTW2 (2-stories)
was modified, based on the experimental results of NTW1, to have a uniform distribution
of flexural reinforcement in the flange. This resulted in a decreased weak-axis bending
moment-capacity.
The foundation block, floors, and wall cap were cast separately in the vertical position.
Slabs were constructed at each floor level. Maximum aggregate size was 3/8 inches (9.5
mm). Concrete compressive strengths in the first floor of NTW1 and NTW2 were 7.3 and
6.6 ksi (50 and 45 MPa), respectively. Longitudinal reinforcement was continuous in NTW1
and spliced in the second floor base in NTW2. For reinforcing bars size No. 3 and higher
A615 Grade 60 steel was used. No. 2 bars were deformed wires meeting the requirements
of ASTM 496.
111

72" 1 long hoop, 2 square hoops


Flange Longitudinal Flange Horizontal Steel:
#3@ 14'' O/C
HORIZONTAL Steel: (28) #4 + (12) #5 #3@ 9'' O/C

5@312"

6" 385"

#2@ 2 1/2'' O/C Web Horizontal Steel:


2 hoops + 2 crossties #3 @ 6'' O/C

#3 @ 6'' O/C
HORIZONTAL
Clear Cover 0.75" except
at boundary elements

Web Distributed Steel:


#3@ 12.5'' O/C #3@ 12'' O/C
VERTICAL EACH 90"
FACE

#2@ 2'' O/C


1 long hoop, 3 square hoops

crossties @ 2" oc
3 hoops + 1 crosstie
312"
#2 hoops &

6@312"
Web Tip Longitudinal Steel:
(8) #6 + (4) #5 + (4) #3

Figure 3.6. Section view of first story of NTW1 showing detailing.


(a) NTW1 (4-stories) Figure 3.9. Section view of first story and lower portion of second story of NTW2 showing
(b) NTW2 (2-stories) detailing.

Figure 2.83: Cross-sections of walls tested by Brueggen [15].

95
92
112

The walls were tested in the MAST (Multi-Axial Subassemblage Testing) facility allowed
application of shear, moment and axial forces from a 6-degree of freedom ‘cross-head’. Steel
strain gauges were provided on vertical, horizontal and confining reinforcement. Embedded
concrete gauges at 45 degrees were also provided. Relative displacements (vertical and
diagonal) of the wall were measured in the first floor using an optical measuring system
(Krypton) and LVDTs. LVDTs were also used in the upper floors.
flange-in-tension
Constant axial loads and flange-in-compression
of 0.03A 0 loading directions, the initial cycle to 25
g fc were applied to the walls prior to application of lateral

percent
forces. The of yieldingbetween
relationship in the flange
thedirection, and all of
base moment the base
and non-orthogonal
shear weredirection cycles, such that
controlled
which
the effective were span
shear only applied
of theonce.
wallTypically,
was M/(V a slightly
`w ) =larger force
3.47 is required
under to take a test
strong-axis bending and
M/(V `w ) specimen
= 4.33 under weak-axis
to a given bending.
displacement level theApplication
first time than of
thelateral loads
second and were displacement-
subsequent

controlled.times,
Figure 2.84 inshows
as shown Figure the
3.18.displacement
The first loadinghistory of the
causes some walls,
damage which
to the included cycles
specimen,
of i) strong-axis bendingof only,
such as cracking ii) which
concrete, weak-axis
reducesbending
its stiffnessonly, iii) displacement
for future at 45 degrees,
loadings. In general,
iv) 100% displacement in one direction and 30% in the other, v) figure eight displacement
unless the specimen is very near to failure, these future loadings should not cause
pattern, and vi) displacement along the analytical yield envelope shape.
significant additional damage to the specimen.

15

10

5
Applied displacement (in.)

0
0 20 40 60 80 100

-5
Web direction displacement

Flange direction displacement

-10

-15
Ramp number

Figure 3.17. Displacements applied to NTW1 in each ramp.


Figure 2.84: Displacement history of walls tested by Brueggen [15].

The four-story wall (NTW1) exhibited lateral strength loss at a drift of 2.0% under
strong-axis bending, exhibiting crushing of core concrete and bar buckling in the bound-
ary region in the web tip. The two-story wall (NTW2) lost lateral strength at a drift of

115
113

2.5%; bars buckled under compression and confining hoops fractured under tensile loading.
Figure 2.85 shows the first floors of the wall after strong-axis bending failure. The follow-
ing observations/conclusions were made by Brueggen: i) changing specimen height while
keeping ratio of base reactions constant had little impact on the wall behavior, ii) the wall
with flexural reinforcement concentrated in the boundary regions had increased shear lag
and moment strength, iii) weak axis drift capacities between the two walls were similar,
and iv) lap splice in the second floor prevented extension of the plastic hinge and decreased
the flexural and shear deformations in hte spliced region, while only slightly chnaging the
stiffness and displaced shape of the specimen.

Figure 5.13. First story web of NTW1 after web failure.


(a) NTW1 (4-stories)

Figure 5.14. Second story web of NTW1 after web failure.


Figure 6.7. First story web of NTW2 after web failure.
(b) NTW2 (2-stories)
153
Figure 2.85: First floor of Brueggen wall tests following strong-axis bending failure [15].
114

2.5.18 Dazio et al

Dazio et al. [26, 25] tested six rectangular walls at ETH Zurich to investigate the effect of
i) vertical longitudinal reinforcement ratios, ii) amount and detailing of confining reinforce-
ment, iii) ductility of reinforcement and iv) axial load ratio.
The walls were designed as the lateral system for a 6-story prototype building, with
the test specimens constructed as a 1/2 scale representation of the bottom 6-stories of the
prototype walls. The specimen height was 45% of the prototype height to account for the
reduction in shear span due to higher mode effects. All walls had the same geometry and
the same shear reinforcement, which satisified capacity design of the walls. The selection
of the longitudinal reinforcement layouts was not presented; the spacing of longitudinal
reinforcement in the web and boundary regions was approximately consistent with larger
bars and slightly smaller spacing in the boundary regions. Confinement of the walls was i)
none, ii) multiple hoops or iii) a single hoop with ties. In the walls with confinement every
longitudinal bar was engaged by a hoop or cross-tie. The cross-sections of the six specimens
are shown in Figure 2.86.
A. Dazio et al. / Engineering Structures 31 (2009) 1556–1571 1557

Fig. 1. Reinforcement layout in the plastic zone of the test units. All dimensions in [mm].
Figure 2.86: Cross-sections of walls tested by Dazio et al. [25]
Table 1
Summary of the test unit properties.
Test unit Sectional forces at the base Reinforcement ratios Stabilising reinf.
N (kN) N /Ag fc0 (–) V /0.8lw bw b (MPa) αN = 0.45lw N /M (–) ρbound (%) ρweb (%) ρtot (%) ρh (%) s (mm) s/Dnom (–)
a
WSH1 689±3 0.051 1.40 0.40 1.32 0.30 0.54 0.25 75 7.50
The
WSH2 foundation
691±4 block 1.50
0.057 and bottom0.38half of the 1.32
walls were
0.30 cast
0.54 vertically
0.25 75 while
7.50 the top
WSH3 686±7 0.058 1.89 0.30 1.54 0.54 0.82 0.25 75 6.25
WSH4 695±6 0.057 1.85 0.31 1.54 0.54 0.82 0.25 No ties
half ofWSH5
the wall
1474 ± 29and the wall
0.128 1.83 cap were0.66
cast horizontally.
0.67 The
0.27 concrete
0.39 compressive
0.25 50 6.25strength
WSH6 1476 ± 6 0.108 2.49 0.49 1.54 0.54 0.82 0.25 50 4.17
a

ranged Standard
from
b
Ratios
deviation of the applied axial load over the duration of the test.
5.5-6.7
computed with largestksi (38
sectional forces− 46 MPa).
measured during the test. One of the objectives of the program was to test

walls 2.with non-ductile steel; Figure 2.87 shows the typical stress-strain response of the steel
Experimental investigation

The test series comprised six large-scale RC cantilever walls


used. tested under quasi-static cyclic loading. In the following, the
geometry of the test units, the material properties, the test setup,
the instrumentation, and the loading history are described.

2.1. Test units

The test units were half-scale models of the lower part of a RC


wall in a six-storey reference building with a total height of 20.4
m [7]. The height of the test units was determined considering
569.2 ± 4.0 700.2 ± 3.3 1.23 ± 0.01 7.34 ± 0.29 H– M B (∼C)
489.0 ± 4.3 552.2 ± 3.3 1.13 ± 0.00 6.45 ± 0.33 M– M B
562.2 ± 1.8 615.0 ± 3.0 1.09 ± 0.00 3.06 ± 0.66 – –

576.0 ± 2.6 674.9 ± 1.8 1.17 ± 0.01 7.29 ± 0.61 H–M B (∼C)
583.7 ± 5.5 714.4 ± 5.1 1.22 ± 0.01 7.85 ± 0.66 H–H C
518.9 ± 13.8 558.7 ± 6.7 1.08 ± 0.02 5.45 ± 0.41 M–M B
562.2 ± 1.8 615.0 ± 3.0 1.09 ± 0.00 3.06 ± 0.66 – –
mples tested. 115
05 < Low < 1.08 < Medium < 1.15 < High.
Low < 5% < Medium < 7.5% < High.
Annex C1 [9].

cing bar is displayed in Fig. 3. All values are


ined from uniaxial tension tests on reinforcing
r length of 750 mm. The ductility properties
by the hardening ratio Rm /Rp02 and the total
ximum force Agt , i.e. the ultimate strain. The
m and the proof stress at 0.2% non-proportional
were obtained from the material tests dividing
testing machine by the nominal area of the
as an approximation of the yield strength fy
bars used for the construction of the test units
unced yield plateau [8]. The ultimate strain Agt
s the stroke of the testing machine at maximum
Figure while
he clear length of the test specimen 2.87: also
Typical Fig.stress-strain
3. Typical stress–strain relationship
response of and ductility
steel used properties
in wallsfor D12 bars usedby Dazio et al. [25]
tested
within the tests.
elastic deformation of the testing machine. Note
considerably between different test methods.
Table 3
employed here aimed at deriving an ultimate Mechanical properties of the concrete.
hich is comparable to the strain capacity obtained
Test unit ρc (kg/m3 ) fc w (MPa) fc0 (MPa) Ec (GPa)
ethod according to EC2 [9]. In this method
Walls were the tested vertically with axial load applied by post-tensioned tendons. The
a reinforcing bar is determined as the plastic WSH1 2397 ± 28 55.3 ± 1.8 45.0 ± 2.1 44.4 ± 5.1
WSH2 2421 ± 23 55.0 ± 2.1 40.5 ± 2.8 37.1 ± 0.1
m gauge length – not including the part ofblocks
foundation the of the walls
WSH3 2381were
± 18 post-tensioned
56.8 ± 1.6 to.2 ±
39 the
2.1 laboratory
35.2 ± 1.5 strong floor. The
ctured – plus the elastic strain at maximum load. WSH4 2378 ± 15 58.8 ± 1.7 40.9 ± 1.8 38.5 ± 2.0
rain as average strain over the rather
wall capslong cleartapered,
were WSH5 with lateral2404 ± 19 lateral 56.0 load
± 2.3 applied 38.3 ± at1.4an elevation
36.1 ± 1.5 to produce a shear
serves as an approximation of the strain capacity WSH6 2383 ± 22 59.2 ± 2.9 45.6 ± 0.3 36.9 ± 0.7
span ratio
their region of rupture. In addition of M/(V `w ) = 2.28 for five walls and M/(V `w ) = 2.26 for the final wall. Lateral
to the stroke
xtensometer measurements with a gauge length
casting. The density ρc and the cube strength fc w were obtained
carried out. Measurements displacement of thefrom
wallscubewerespecimens
measured at afive200vertical elevations andtherelative deformations
on bars where the
with mm edge length while
outside the extensometer gauge length showed 0
cylinder strength fc and the modulus of elasticity Ec were obtained
from the extensometer readings werewas measured
about 8% vertically along the chords of the walls and in two sets of diagonals on the
from cylinder specimens with a diameter of 150 mm and a height
btained from the stroke. In the last two columns
of 300 mm. Further details on the material tests are given in [6].
lower half
nforcing bars are classified according of [9]
to EC2 the walls. Stains gauges were placed on the corner longitudinal reinforcing
ening ratio Rm /Rp02 and Agt . The mean hardening
barsvaried
udinal and shear reinforcing bars in the foundation
between block.
2.3. Test setup,Strains in the and
instrumentation, lower third
loading of the wall were measured using a
history
ile their strain capacity ranged between 2.34%
grid of Demac or Whitmore A photo ofgauges.
the test setup is shown in Fig. 4. The test units were
was cast in two phases: In the first phase the clamped onto the strong floor by large diameter post-tensioning
0
and the wall up to a height ofAxial 1.5 mloads bars.−
above of 0.05 At0.13A
the wall g fhead
c were applied
the test to the
units were walls toprior
subjected to beginning the reverse
a horizontal
lock were cast upright. For the second phase displacement history applied by two servo-controlled hydraulic
laid down horizontally in ordercyclic tests. The
to simplify the first two cycles
actuators of the
with a force tests of
capacity were
±500 force
kN and controlled to three-quarters of the
a displacement
mechanical properties of the concrete given in capacity of ±100 mm each. The two actuators were connected in
the samples taken during the nominal
first phaseyield
of theforce.series
The displacement
since the upper bound atestimate
this point
of the was used tocapacity
displacement determine displacement
ductilities, which were used in the displacement-controlled cycles for the remainder of the
tests. Two cycles were performed at each displacement ductility increment (1, 2, 3,...).
Four of the wall tests failed due to fracture of the longitudinal reinforcement in the web
or boundary region. The remaining two walls failed due to concrete crushing. The final
damage to the walls is shown in Figure 2.88. The wall with the least ductile steel, WSH1,
exhibited the least ductile response at an ultimate drift of 0.68%. A wall with similar
layout but more ductile steel, WSH2, also failed due to boundary element bar fracture,
but at a drift of 1.80% and only after the bars had previously buckled. In both cases, the
web reinforcement had fractured prior to spalling of the cover concrete (but did not cause
116

failure). A third wall, SHW3, with a similar cross-section but steel meeting the European
standard of ductile avoided this premature fracture of the bars and failed at 2.04% drift
due to fracture of a previously buckled longitudinal reinforcing bar. This cross-section was
tested again without confining reinforcement (WSH4). Spalling occurred at the same drift
in these two walls but in WSH4 the bar buckling occurred simultaneously with the spalling
and the wall failed due to core crushing at a drift of 1.63%. The final two wall tests had
increased axial loads and thus the longitudinal reinforcement was reduced to keep moment
strength the same. WSH5 was similar to WSH3 in layout, strength and ductility properties
of steel, but axial load was increased and confining hoops were used instead of cross-ties.
Failure occurred due to fracture of boundary element reinforcement at 1.35% drift. The
final wall had slightly higher shear span ratio than the others, the same axial load ratio as
WSH5 and a higher moment strength. The maximum drift reached as 2.11%, but failure
occurred at 1.93% in the opposite direction due to crushing of the core concrete, which was
allowed by fracture of confining reinforcement.
Dazio et al concluded/observed that i) walls with low longitudinal reinforcement ratios
had reduced flexure-shear cracking, ii) ductility of web reinforcement is more critical that
usually assumed, and iii) at three-quarters yield strength, crack patterns less developed in
walls with greater moment strength due to axial loads.

2.5.19 Other Sources

The preceding sections provide detailed summaries of the research programs that tested
walls that were used in this thesis to develop a database of slender wall tests. The walls
included in the database had shear span ratios greater than or equal to M/(V `w ) = 2.0 and
because they were used to develop fragility, had damage information available. This section
presents a brief summary of other reinforced concrete wall research that was not used in the
database but provided a background for the work contained in this thesis.
117

1562 A. Dazio et al. / Engineering Structures 31 (2009) 1556–1571

Fig. 8. Crack patterns of the six test units after failure.

Figure 2.88: Final damage to walls tested by Dazio et al. [25].


displacement ∆ = 6.2 mm used in the loading history had been
Fig. 8(d) shows that the damage to the compression zone of WSH4 y
was significantly more severe than the damage to the compression determined according to the 3/4-rule which was later found not to
zone of WSH3. be the most suitable method for determining yield displacements
WSH5: Test Unit WSH5 was designed for a similar moment (Section 2.3). During the cycles with displacement ductilities µ∆ =
capacity as WSH3 but since the axial load was about 2.14 times 2–4 it was noticed that increasing the displacement amplitudes by
larger, its longitudinal reinforcement ratio had to be reduced. The just 6.2 mm would require an unrealistic number of cycles to reach
spacing of the confining reinforcement of WSH5 was reduced failure. From LS44 onwards the increment of the displacement
to 50 mm because of the smaller diameter of the longitudinal amplitudes was therefore doubled. Onset of spalling was observed
reinforcing bars in the boundary regions (s/Dnom = 50/8 = 6.25). for δ = 0.55% (LS36). Shortly before reaching LS46 (δ = 0.82%)
2.5.19.1 Zhang and Wang
The web reinforcing bars of WSH5 had poorer ductility properties problems with the load-follower occurred yielding relatively large
compared to those of WSH3 while the ductility properties of the oscillations in the axial load (the axial load history is included in
boundary reinforcement were similar despite the fact that D8 mm the recorded data [6]). However, the first reinforcing bars only
fractured at LS56 (δ = 1.01%) after these problems had been
Zhang and Wang [120] tested rectangular walls with high axial loads as large 0.35Ag fc0 .
instead of D12 mm bars were employed for WSH5. The yield

The test configurations suggested that the walls had a shear span ratio of M/(V `w ) = 2.14
but the reported shear span ratio was 1.80. Base moment vs curvature and base shear vs
displacement are both provided and the relationship between the base reactions suggests a
shear span of 1.80, thus, the walls were considered squat for this study and not included in
the damage database.
The study concluded that axial load ratio had an important effect on stiffness, ductility
and failure mode, but that more research was needed to make definitive conclusions. The
researchers also looked at the shear compression ratios of walls (the maximum shear force
118

normalized by the axial capacity of the walls) and determined that it had an effect on the
post-yield behavior of walls.

2.5.19.2 Cardenas and Magura

Cardenas and Magura [18] tested six walls, four with shear span ratios of M/(V `w ) = 3.36.
Monotonic load histories were applied and axial loads ranging from 0.056 − 0.072Ag fc0 were
applied to the walls. Walls had either a uniform distribution of longitudinal reinforcement
or two-thirds of the steel concentrated in boundary regions; confining reinforcement was not
provided. These wall tests were not included in the damage database because i) material
properties reinforcement ratios were provided but not the layout of the reinforcement, ii)
damage progression was not reported, and iii) lateral deformations of the wall were recorded
but only moment-curvature was reported by the researchers.
Failure modes of the walls was reported as i) fracture of reinforcement for wall with 0.27%
gross longitudinal reinforcement, ii) crushing of concrete in compression zone after yielding
of reinforcement in tension for walls with 1.0% and 2.3% gross longitudinal reinforcement or
iii) fracture of shear reinforcement in the wall with 3.0% gross longitudinal reinforcement.
p
Walls reached a shear stress of 1.7 − 5.3 fc0 . Cardenas and Magura made the following
conclusions/observations: i) concentrating reinforcement near the ends of structural walls
may be advantageous and ii) axial load decreases the ultimate curvature of a wall.

2.5.19.3 Oh, Han, and Lee

Oh, Han and Lee [81] tested four full scale walls designed for a Korean building, where wall
area to floor area ratios and detailing requirements are similar to Chilean buildings. Three
walls had a rectangular cross-sections and the forth had a barbell cross-section. Walls had
a shear span ratio of M/(V `w ) and axial loads of 0.1Ag fc0 .
Walls with confinement in the boundary element failed due to crushing of the concrete
and bar buckling at drifts in excess of 2.5% and the wall without confinement failed due to
concrete crushing at a drift of approximately 2.0%. The researchers made the following con-
clusions: i) open hoops and cross-ties provided confinement to concrete in the compression
119

zone that enhanced the drift, ductility and energy dissipations of the wall, ii) comparable
barbell and rectangular walls produced nearly identical results, and iii) walls without con-
fining reinforcement may be adequate for buildings with wall to floor areas similiar to that
of Chilean buildings.

2.5.19.4 Paterson

Paterson and Mitchell [85, 84] tested rectangular walls to investigate a proposed seismic
retrofit method for an existing building. Two walls were tested without retrofit methods
and were considered here. The walls lacked adequate shear strength and boundary element
confinement and both contained lap splices; one with the splice located at the base of
the wall (W1) and one with the splice located lw /2 above the base of the wall (W2).
Figure 2.89 shows the cross section of the un-retrofitted specimens. Walls were tested
horizontally without axial loads. Lateral loads were applied at a height producing a shear
ear wall and test setup: 共a兲 Typical core wall
span ratios of M/(V `w ) = 2.83 and M/(V `w ) = 3.25 for specimens W1 and W2 respectively.
core wall elevation showing lap splices above
The共d兲reverse-cyclic
setup, lap splice at wall base; and test setup, test was force-controlled prior to general yielding of the specimen and
om wall base
displacement-controlled thereafter.

eversed cyclic loading was applied 150 mm


wall by hydraulic jacks using the strong floor
c and d兲兴. Each cycle of the loading consisted
up on the tip of the wall兲 and negative load-
d cycles were done at each load/deflection
ns were cycled to selected load levels, at a
ut 15 kN/min, up until general yielding oc-
l yielding the specimens were cycled to Figure
mul- 2.89: Cross-section of walls tested by Paterson and Mitchell [85].
eflection, at a deflection rate of about 4 mm/ Fig. 2. Specimen details: 共a兲 Specimen W1 and 共b兲 Specimen W2

nd negative peaks of each cycle, testing was As-Built Specimens W1 and W2


tographs and measure crack widths. During
Specimen Specimens W1 and W2 were tested in the as-built condition. They
ollected by load cells, linear voltage differen- W1 failed at a drift of 0.43% due to failure of the lap splices at the edges of
were detailed to simulate the critical region of the existing core
VDTs兲, and strain gauges on the reinforcing
the walls
cement, and carbon fiber wrap. A potentiom-
(Figure 2.90a);
wall thatyielding
required had occurred
retrofit. The twoonly at test
as-built thespecimens
base of wall
modelon the bar extending
the presence of lap splices at various floor levels. Specimen W1
asure the tip deflection of the specimen.
from For
the foundation block. Loss
was designed andofconstructed
lateral load withcapacity also
a lap splice at occurred
the base of due
the to the bond failure
and W2R, horizontal LVDTs were placed at
wall 关Fig. 1共c兲兴. This is intended to model the situation where the
l parallel to the longitudinalof reinforcement.
the lap splice in lap specimen W2 during theofsecond cycleSpecimen
to a drift
splice occurs at the base the structure. W2ofwas1.76% (Figure 2.90b).
ured displacements over a length of 50 mm,
designed and constructed with a lap splice beginning d/2 共or 600
ure to be determined at this This
criticalfailure occurred
location. once yielding was observed in the lap splice. Both walls were able to
mm兲 from the base of the wall 关Fig. 1共d兲兴. This specimen simu-
d from the LVDTs were supplemented with
reach W1R,the expected lates the situation whereofa lap
moment-strength thesplice is present just above a region
cross-section.
m the strain gauges. For specimen the
of flexural hinging, as would be the case at the second or third
rs at the top of the reinforced concrete collar
zontal LVDTs at this location 共with Duea to lack of sufficient
gauge damage
floor level of information,
the structure the Paterson and Mitchell wall tests were
after retrofit.
were used to determine the curvature. In Fig.
Design and Construction of Specimens W1 and W2
sponses of the specimens, the locations of the
ermine the curvatures and the locations of the The cross section of the as-built specimens was detailed to have
dicated. similar dimensions and properties to the critical section of the

JOURNAL OF STRUCTURAL ENGINEERING © ASCE / MAY 2003 / 607


120

not included in the damage database. However, these tests presented data on the perfor-
mance of lap splices in walls, a consideration made in this dissertation but for which there
is little information available. Figure 2.91 present strain profiles for specimens W1 and W2.

2.5.19.5 Ghorbani-Renani

Ghorbani-Renani et al. [39] tested four rectangular walls to assess the effect of scaling and
loading. Two tests were a full-scale wall and two were 1:2.37 scaled versions. One monotonic
and one cyclic test was complete for each size of wall. The walls had a shear span ratio of
M/(V `w ) = 2.08 and axial loads of 0.01Ag fc0 .
Tests were ended prior to failure of the monotonically loaded specimens. The large
scale specimen had a significant drop in lateral load carrying capacity at approximately
3.0% drift while the small-scale specimen had reached 80% of the maximum lateral load at
approximately 2.0%. Failure modes of the cyclically loaded walls was not reported by the
researchers. These wall tests were not included in the damage database due to a lack of
damage information.

2.5.19.6 Wolschlag

Wolschlag tested 1/8 scale H-shaped walls under both static (2 specimens) and dynamic
loading (4 specimens). The static tests were conducted by placing lateral loads to slabs at
the first, second and third floors of a three story wall specimen. The test was displacement-
controlled in at the first floor level and the load necessary to produce the desired displace-
ment was also applied to the upper floors. The reported displacements at the first and third
floors were essentially the same, and because of this, the small scale of the specimens and
the lack of detailed damage information led to the exclusion of these tests from the damage
database.

2.5.19.7 Su and Wong

Su and Wong[101] tested three one-quarter scale walls to investigate the impact of axial load
ratio and confining reinforcement ratios. Walls had a shear span ratio of M/(V `w ) = 4.0
121

(a) W1 (b) W2

Figure 2.90: Final damage of walls tested by Paterson and Mitchell [84].

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

(a) W1 (b) W2

Figure 2.91: Strain profiles at splice in walls tested by Paterson and Mitchell [84]. Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
122

and had an initial axial load of either 0.25Ag fc0 or 0.50Ag fc0 which decreased throughout
the test. Two confinement ratios were considered for the walls loaded with the higher axial
loads.
The wall with 0.25Ag fc0 axial load failed due to buckling of longitudinal reinforcement
at a drift of 2.2%. The walls with higher axial loads experienced less cracking and failed
at approximately 1.3% drift due to out-of-plane displacement of the walls along cracks
penetrating the full thickness of the wall at the edges of the walls. The walls with higher axial
load exhibited lower energy dissipation, drift capacity and ductility. When the confining
reinforcement ratio was doubled, the strength and ductility of the specimens was minimally
impacted, although it should be noted that the increased reinforcement ratio was an increase
in the area of steel, not an increase in the spacing of the steel.
Although the tests by Su and Wong provided valuable results for the behavior of walls
under large axial loads, they were not included the database used for this research due to a
lack of detailed damage information and because force-displacement relationships were not
reported, only force-rotation.

2.5.19.8 Johnson

Johnson tested rectangular walls with un-symmetric reinforcement to simulate the response
of T-shaped walls. The longitudinal reinforcement was spliced in various methods to eval-
uated the impact of the splice on the performance of the walls. Due to the unsymmetric
reinforcement, the walls were not included in the damage database. However, the behavior
of the splice and the determination of deformation components are of interest to the research
presented in this thesis.

2.5.20 Failure modes

This section considers the relationship between failure mode of experimental wall subassem-
blage tests and wall characteristics. Section 2.5.20.1 and Section 2.5.20.2 summarize research
found in the literature that specifically addresses this idea.
123

2.5.20.1 Failure Classification by Wood

Wood [115] examined 27 wall tests to investigate failure modes of walls. The data set
consisted of wall specimens that were i) barbell and rectangular, ii) had squat and slender
aspect ratios, iii) subjected to cyclic and monotonic loading, iv) had axial loads less than
0.15Ag fc0 , and v) had drift capacities in excess of 1.0%. Failure mechanism was classified
as either flexure or shear on the basis of the crack patterns developed in the walls. Wood
determined the nominal shear and flexural strengths of the walls and noted that the lower
of these two values did not predict the failure mode but did provide a good indication of the
experimental strength of the wall. Shear failure occurred in walls with shear stress in excess
p p
of 4 fc0 and flexural failure in walls with shear stresses less than 4 fc0 . Wood also noted
that ductility decreased with increased boundary element reinforcement ratio and that the
number of cycles prior to yield impacted the deformation capacity of the walls, but did not
note an impact of axial load or web reinforcement. Deformation capacity was not identified
to be dependent on the failure mode.

(a) (b)

Figure 2.92: Failure mode of walls studied by Wood [115].

2.5.20.2 Failure Classification by Fang

Fang [34] examined approximately 45 experimental tests of moderate and tall walls to in-
vestigate the failure modes of walls. Specific characteristics of the wall specimens were not
available. Failure modes were classified as i) flexure (bar fracture; bar buckling; concrete
124

crushing), ii) flexure-shear (sliding shear or failure along inclined crack after yield of lon-
gitudinal reinforcement), iii) out-of-plane failure and iv) shear (failure prior to yielding of
longitudinal reinforcement). The failure mode was evaluated as a function of i) shear span
ratio, ii) relationship between shear and flexural strength, iii) nominal shear stress, iv) wall
shape, and v) wall detailing.
Fang observed that shear failure prior to yield occurred in walls with shear span ratios
as high as 3.0 if the shear strength was less than the shear force associated with the flexural
strength. Flexural failure occurred in walls with low nominal shear strengths and greater
flexural strength than shear strength. Fang also observed that shear-flexure failures occurred
in walls with higher nominal shear strengths, with the flexural strength having been reached
but the high shear demand limiting the ductility of the walls. Out-of-plane failure of walls
occurred in walls with thin cross-sections (relative to wall length) and Fang suggested that
no amount of confinement could prevent premature failure in this mode.
Fang recommended that i) shear failure can be avoided by designing walls with greater
flexural strength than shear strength, ii) shear span ratios equal to or greater than 1.5
should be used to ensure adequate ductility, and iii) special attention should be paid to
prevent out-of-plane buckling of a wall.

2.6 Summary of Experimental Research

This section provides a summary of the experimental tests summarized in Section 2.5.
Section 2.6.1 compares compliance of the experimental wall specimens with the ACI 318-08
building code. Section 2.6.2 evaluates the characteristics of the experimental wall specimens
with those of the building inventory discussed in Section 2.3. Section 2.6.3 summarizes the
failure modes of the experimental wall specimens. Section 2.6.4 compares the observed fail-
ure mode of experimental wall specimens with the observed damage states of walls following
earthquakes. Section 2.6.5 discusses the impact of design parameters on the performance of
experimental wall specimens.
125

2.6.1 Evaluation of Code Compliance

The ACI 318-08 Building Code [1] requirements for structural walls are summarized in
Section 2.1. These requirements were used to determine whether experimental specimens
from the test programs summarized in Section 2.5 meet the modern design requirements for
structural walls. In evaluating compliance with ACI 318-08, the following assumptions and
omissions are made:

1. Criteria for material properties are ignored. This includes concrete compressive strength
and reinforcing bar conformance with ASTM standards.

2. Actual rather than nominal/design material properties are used in all calculations.

3. Criteria affected by scaling of specimens are ignored. This includes absolute spacing
limits (i.e. requirement that spacing of horizontal reinforcement be less than 18 inches)
and spacing limits based on specimen thickness (i.e. requirement that spacing of
confining hoops not exceed 1/3 the minimum dimension of the wall). This does not
included spacing requirements that are a function of member dimensions or reinforcing
bar sizes.

4. Code compliance of splices lengths and/or locations of splice are not considered.

5. The availability of adequate boundary element length to develop the yield strength of
horizontal reinforcing bars is not considered.

6. The need for special boundary elements is assessed using the displacement-based de-
sign approach.

7. A design drift of δu /hw = 0.02 is used to assess the need for special boundary elements.
Neutral axis depth of the cross-section is determined from a moment-curvature analysis
of the wall section at a strain of −0.003 in/in in the extreme compression fiber.25

25
All walls tested by Tasnimi[102] are classified as non-compliant based on the need for a special boundary
element. Evaluation of neutral-axis depth values are very close to the design limit and are sensitive to the
126

8. Criteria for spacing and reinforcement areas/ratios are considered to be met if the
wall designss were within 5% of the specified values.

Table 2.926 indicates whether each specimen meets the ACI 318 Building Code require-
ments and where applicable, the provision(s) not met. Of the 72 walls considered, 23 are
code compliant. The following paragraphs discuss for each cross-sectional shape the number
of compliant and non-compliant walls, as well as the reasons for non-compliance.
Ten of 43 (23%) rectangular walls, meet the ACI 318-08 requirements. Of the 33 non-
compliant walls, the reasons for non-compliance is i) lack of special boundary element where
one is required for 10 walls (30%), ii) inadequate boundary element length for 2 walls (6%),
iii) inadequate transverse reinforcement spacing for 22 walls (67%), and iv) inadequate
vertical or horizontal steel in the web for 3 walls (9%). The percentanges listed total to
greater than 100% because some walls are non-compliant for multiple reasons.
Twelve of 13 (92%) barbell walls meet the ACI 318-08 requirements. The one non-
compliant wall, W1, did not provide the required two curtains of horizontal steel.
Three of the four (75%) T-shape walls meet the ACI 318-08 requirements. Only TW2
tested by Thomsen and Wallace [104] is non-compliant as it does not meet the provisions for
special boundary element length and transverse reinforcement spacing; it was specifically
designed by the researchers to have poor detailing. Both H-shaped walls (100%) meet the
ACI 318-08 requirements. Of the five C-shaped walls, only one wall (20%) (TUA tested by
Beyer et al. [11]) is code compliant. The four non-compliant specimens fail to meet the
requirements for spacing of transverse reinforcement.

2.6.2 Comparison to Building Inventory

This section presents a comparison of the experimental tests discussed in Section 2.5 with
the building inventory discussed in Section 2.3. Geometric, material, and capacity/demand

varied concrete strength between the specimens. Because the specimen design was identical, it is desired
that they all be classified the same; a non-compliant classification is selected due to the specimens falling
at the limit for requiring special boundary elements and none was provided. Note that no other specimens
are border line in determining if the need for special boundary elements is satisfied.
26
Provided at end of Section 2.6 due to length.
127

parameters for the experimental test specimens are presented in Tables 2.10 - 2.1227 . Only
walls with rectangular cross-sections are considered in the comparison of the test specimen
and building inventory wall properties. Wall characteristics for the building inventory and
experimental tests are compared by plotting them against the cross-section aspect ratio
of the walls (lw /tw ). The cross-section aspect ratio was used because the two groups of
walls have similar ranges for this parameter; for other parameters, such as wall aspect ratio,
hw /lw , the range of values for the lab tests is significantly different than for buildings. In
plotting the wall characteristics for experimental tests, different markers are used for walls
tested with and without axial load, as the former provide a more accurate representation of
true wall behavior.
Figure 2.93 shows the reinforcement ratios of the walls. For the walls in the building
inventory, walls with a uniform distribution of reinforcement are not included in the plot of
boundary element reinforcement ratio (Figure 2.93a); for experimental test specimens, with
uniformly distributed steel, the same refinrocement ratio is plotted in Figures 2.93a and2.93c
Figure 2.94 shows the wall aspect ratio, hef f /lw ; the boundary element length, `b , rela-
p
tive to the wall length, `w ; the nominal shear stress demand Vn /( fc0 Acv ) (fc0 in psi); and
the shear demand-capacity ratio, Vmn /Vn . The aspect ratio, hef f /lw , is calculated using
hef f = 2/3hw for the building inventory and hef f = Mb /Vb for the experimental test speci-
mens, where Mb /Vb is the ratio of base moment to base shear. The shear demand-capacity
ratio is the shear at nominal moment strength, Vmn , divided by the nominal shear strength,
Vn , as calculated per ACI 318-08 [1]. For the building inventory, the shear at nominal mo-
ment strength was determined using an effective height of 2/3hw , where hw is the height of
the wall.
In Figure 2.93 and 2.94 the characteristics for all experimental test specimens are com-
pared to the characteristics of walls in the building inventory. These figures are reproduced
in Figures 2.95 and 2.96 to show the characteristics of only the ACI 318-08 compliant test
specimens compared to the characteristics of walls in the building inventory.
Comparison of Figure 2.93 to Figure 2.95 and Figure 2.94 to Figure 2.96 indicates that

27
Provided at end of Section 2.6 due to length.
128

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
12 2.5

10
2

8
1.5

ρweb
ρbe

6
1
4

0.5
2

0 0
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(a) Longitudinal boundary element steel (b) Longitudinal web steel

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
5 1.4

1.2
4
1
3 0.8
ρh
ρv

2 0.6

0.4
1
0.2

0 0
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(c) Total longitudinal steel (d) Horizontal steel

Figure 2.93: Reinforcement ratios for building inventory and experimental test programs.
129

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
20 35

30
15
25

ℓb /ℓw (%)
heff /lw

10 20

15
5
10

0 5
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(a) Aspect ratio (b) Boundary element length (percent of wall
length)

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
14 2

12
1.5
fc′ ) (psi)

10
Vmn /Vn
p

8 1
Vn /(Acv

6
0.5
4

2 0
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(c) Nominal shear stress (d) Shear demand-capacity ratio

Figure 2.94: Geometry and demand characteristics for building inventory and experimental
test programs.
130

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
6 1

5
0.8
4

ρweb
ρbe

3 0.6

2
0.4
1

0 0.2
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(a) Longitudinal boundary element steel (b) Longitudinal web steel

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
3 1.6

2.5 1.4

1.2
2
1
ρh
ρv

1.5
0.8
1
0.6
0.5 0.4

0 0.2
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(c) Total longitudinal steel (d) Horizontal steel

Figure 2.95: Reinforcement ratios for building inventory and experimental test programs
(ACI 318-08 compliant walls only).
131

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
20 35

30
15
25

ℓb /ℓw (%)
heff /lw

10 20

15
5
10

0 5
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(a) Aspect ratio (b) Boundary element length (percent of wall
length)

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
14 1.2

12
1
fc′ ) (psi)

10
0.8
Vmn /Vn
p

8
Vn /(Acv

0.6
6

0.4
4

2 0.2
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(c) Nominal shear stress (d) Shear demand-capacity ratio

Figure 2.96: Geometry and demand characteristics for building inventory and experimental
test programs (ACI 318-08 compliant walls only).
132

i) ACI 318-08 compliant experimental test specimens are generally representative of walls
in the building inventory, ii) experimental test specimens with characteristics outside the
range of characteristics of the building inventory are generally non-compliant with ACI 318-
08. Although walls in the building inventory are generally well represented by ACI 318-08
compliant test specimens, Figures 2.95 and 2.96 result in the following observations:

1. Experimental test specimens with cross section aspect ratios, `w /tw , at the upper
bound of values for the building inventory have been tested only without axial load.
2. Experimental test specimens fall at the lower limits of shear span ratios, hef f /`w , in
the building inventory (Figure 2.96a).
3. For a given cross section aspect ratio, `w /tw , the boundary element length, `b , relative
to the wall length, `w , is lower for the experimental test specimens than for the building
inventory walls (Figure 2.96b).

These observations can be reasonably attributed to the limitations typically encountered


when testing specimens in a laboratory. Higher shear spans, longer boundary elements, and
axial load are all factors that increase the flexural demand on specimens. The load capacity
of equipment at a testing facility is often a limiting factor in specimen design.

2.6.3 Evaluation of Failure Mechanism

Failure of the experimental test specimens refers to the mechanism that causes a loss in
lateral load carrying capacity. Each experimental wall specimen is labeled as having one of
the following failure mechanisms, as determined from the failure descriptions provided by
the researchers:

1. Shear failure, including sliding shear failure, fracture of shear reinforcement, and ten-
sion failure along a large diagonal crack.
2. Web crushing failure, where loss of strength occurs as a result of crushing of diagonal
compression struts in the wall web.
3. Flexural failure, including fracture of longitudinal reinforcing bars and failure due to
bond slip
133

4. Shear-compression failure, including bar buckling and/or core crushing in boundary


regions and unspecified compressive failures reported by researchers.

Table 2.928 presents the failure mode for each specimen, along with the maximum drift
capacity and information related to code compliance of (see Section 2.6.1). Figure 2.97
shows the number of occurrences of each failure mode for all specimens, as well as the
number of occurrences of each failure mode based on cross-section shape. The following
sections provide a brief discussion of specimens in each failure mode category.

50 60
45
40 50
% Specimens

% Specimens
35 40
30
25 30
20
15 20
10 10
5
0 0
Flex. Comp. Shear Web Flex. Comp. Shear Web
(a) All (b) Rectangular

60 60
50 50
% Specimens

% Specimens

40 40
30 30
20 20
10 10
0 0
Flex. Comp. Shear Web Flex. Comp. Shear Web
(c) Barbell (d) Flanged

Figure 2.97: Distribution of failure modes for experimental test specimens.

28
Provided at end of Section 2.6 due to length.
134

2.6.3.1 Shear Failure

Five specimens, all rectangular, failed due to a shear mechanism:

• Specimen C1 tested by Khalil and Ghobarah [53], a non-compliant wall tested as a


control wall for a retrofit study, had a diagonal-tension failure at 2.6% drift.
• Specimen C1 tested by Shiu et al. [100] had a sliding shear failure that occurred at
2.9% drift at an apparent construction weakness in the first floor.
• The large-scale wall tested by Riva et al. [92] failed at 3.2% drift due to fracture
of longitudinal web reinforcement at along a large diagonal crack in the web; the
specimen does not meet the ACI 318-08 minimum requirements for longitudinal and
horizontal steel in the web.
• Pilakoutas and Elnashai [88] specimens SW5 and SW7 failed at drifts of 0.8% and
1.8% due to diagonal-tension failure and fracture of the shear reinforcement, respec-
tively. Specimen SW7 had a greater amount of flexural reinforcement than did speci-
men SW5, but both had the same amount of flexural reinforcement (ρbe ≈ 11%) that
resulted in shear demands well in excess of the nominal shear strength of the walls.

2.6.3.2 Web Crushing Failure

Nine specimens (barbell or flanged only) failed due to a web crushing mechanism:

• Six barbell wall specimens tested by Oesterle et al. [78, 80] failed due to web crushing,
four of which were tested with axial loads applied29 . The barbell wall specimens
without axial load that failed due to web crushing had shear-demand capacity ratios
greater than the other barbell wall specimens without axial load. These higher shear-
demands were the result of increased moment capacity due to higher boundary element
reinforcement ratios.
• Both H-shaped walls tested by Oesterle et al. failed due to web crushing.

29
One barbell wall with axial load (B10 ) tested by Oesterle et al. [78] did have web crushing, although
this was not the ultimate cause of failure; instead failure was due to compressive failure in the boundary
element at a region where the concrete was significantly damage during construction,
135

• Specimen TUB, the thinner of the two C-shaped walls tested by Beyer et al. [11], also
failed due to crushing of diagonal compression struts in the web.

2.6.3.3 Flexural Failure

Nineteen specimens, including eleven rectangular walls, failed due to a flexural mechanism.
Four walls tested as control walls in programs investigating retrofit methods failed due to
bond slip:

• The two walls tested by Elnady [32] failed due to a bond slip at drifts of 0.2% and
0.3%.
• The two walls with splices tested by Paterson [84] failed due to bond slip at the splice
locations at drifts of 0.5% and 1.6%.

The remaining 15 flexural failures were a result of fracture of the longitudinal reinforce-
ment and occurred after bar buckling and concrete crushing had occurred:

• Four specimens tested by Dazio et al. [25] with non-ductile steel failed due to fractures
of the longitudinal reinforcement30 .
• Oesterle et al. [78, 80] tested two rectangular walls that failed due to fracture of the
longitudinal reinforcement. Both walls were tested without axial load.
• Liu [59] tested a higher strength concrete wall that failed due to fracture of the lon-
gitudinal reinforcement in the boundary element.

2.6.3.4 Shear-Compression Failure

Thirty-five wall specimens, including twenty-eight rectangular walls, failed due to shear-
compression failure (bar buckling or core crushing in the boundary region).

• The three rectangular walls tested by Oh, Han and Lee failed due to core crush-
ing at drifts of 2.2 − 2.9%, with greater drifts obtained in specimens with confining
reinforcement.
30
The two Dazio et al. walls that did not fail due to bar fracture had either no confining reinforcement or
a high shear demand-capacity ratio; both had shear-compression failure modes
136

• The wall tested by Tupper [106] failed due to core crushing at a drift of 2.9%.
• The specimen tested by Liu [59] with normal weight concrete failed due to bar buckling
and core crushing at a drift of 2.9%.
• The ten specimens tested by Lefas et al. [56, 55] were reported to have loss of strength
at 0.8 − 1.9% drift due to compressive failure of the compression zones; these tests
were force controlled and essentially were push-over tests.
• The four specimens tested by Tasnimi [102] were also force controlled and reached ul-
timate drift of 1.0−1.7%, where compressive boundary element damage was observed.
• Four of six specimens tested by Pilakoutas and Elnashi [88] failed due to crushing or
unspecified compressive damage at 1.8 − 2.2% drift.
• Two specimens tested by Dazio et al. [25] failed due to concrete crushing in the
boundary region at drifts of 1.4 − 1.6%; one did not have confining reinforcement and
the other had a high shear demand relative to the others in the test program.
• The two rectangular walls tested by Thomsen and Wallace [104] both failed due to
bar buckling at 2.5% drift.

2.6.4 Comparison of Laboratory and Field Damage

In this section, the failure mode of the experimental test specimens (see Section 2.6.3) are
compared with the post-earthquake damage types presented in Section 2.4.3.
Figure 2.98 shows the percentage of failure modes as a function of the shear span ratio,
M/(V `w ). Consideration of the shear span ratio allows comparison to the assumed failure
mode of buildings presented in Figure 2.43. The following observations are made:

• Shear failures appear independent of wall height.


• Web crushing failures generally occur at higher heights than does flexure/compressive
failures.
• Compressive boundary element failure is less likely to occur in slender walls.
137

Figure 2.98: Failure mode of experimental test specimens as a function of shear span ratio,
M/(V `w ).

2.6.5 Impact of Design Parameters

This section explores the impact of design parameters on wall response. For each design
parameter considered, the following is presented i) summary of observations from research
programs that varied a specific design, ii) consideration of effect of parameter on failure
modes, and iii) consideration of effect of parameter on the drift capacity.
When considering the failure modes and drift capacities, only rectangular specimens are
considered. The drift capacity of for each specimen is plotted against the design parameters
considered, with failure mode indicated by different markers31 . Table 2.8 provides the
correlation coefficients (R) and coefficients of determination (R2 ) indicating the strength of
relationships between drift capacity and the design parameters. Tables 2.10 - 2.1232 provide
the geometric, material, and capacity/demand parameters for each test.

31
Web crushing is not shown as it was only observed in experimental tests of barbell walls. Different
markers are used to indicate those tests which were tested under monotonically increasing lateral force.
32
Provided at end of Section 2.6 due to length.
138

Table 2.8: Coefficients of determination, R2 , for design parameters and drift capacities by
failure mode (rectangular walls only). R2 values greater than or equal to 0.20 are considered
significant and are indicated by gray cells. Positive (+) or negative (−) correlation is
determined from the correlation coefficient, R.

All1 Flexure1 Shear-Compression1,2 Shear1


Parameter R2 +/− R2 +/− R2 +/− R2 +/−
fc0 0.07 + 0.41 + 0.05 (0.17) + (+) 0.02 −
fy 0.01 + 0.01 + 0.02 (0.13) + (+) 0.05 −
εu 3 0.07 + 0.02 + 0.55 (0.55) + (+) 0.82 +
tw /12, (in) 0.05 + 0.05 − 0.42 (0.37) + (+) 0.50 +
`w /tw 0.04 + 0.29 + 0.25 (0.34) − (−) 0.07 +
`b /`w 0.02 − 0.09 − 0.10 (0.07) − (+) 0.30 +
`b /tw 0.00 − 0.20 + 0.29 (0.09) − (−) 0.28 +
ρbe 0.01 − 0.13 − 0.01 (0.00) + (+) 0.67 −
ρcon 0.36 + 0.50 + 0.42 (0.46) + (+) 0.02 +
ρweb 0.21 − 0.40 − 0.29 (0.00) − (−) 0.51 −
ρh 0.04 − 0.47 − 0.00 (0.24) − (+) 0.04 +
M/(V `w ) 0.01 + 0.10 − 0.29 (0.22) + (+) 0.71 +
λN 0.04 − 0.47 − 0.00 (0.35) − (+) 0.04 +
p
Vn /(Acv fc0 ) 0.01 − 0.10 + 0.09 (0.31) − (+) 0.03 −
Vmn /Vp n 0.04 − 0.43 − 0.00 (0.01) + (−) 0.43 −
Ven /(Acvp fc0 ) 0.07 − 0.36 − 0.10 (0.03) − (+) 0.43 −
Vu /(Acv fc0 ) 0.03 − 0.05 − 0.14 (0.04) − (+) 0.65 −
Vu /Vn 0.01 − 0.09 − 0.02 (0.01) − (−) 0.79 −
Vu /Ven 0.06 + 0.28 + 0.06 (0.00) − (−) 0.00 +
1 # Walls: All (43), Flexure (11), Shear-Compression (27 (17)), Shear (5)
2 Values in ( ) excluded monotonically loaded tests.
3 ε not available for all test specimens. All (22), Flexure (9), Shear-Compression (10), Shear (3)
u

2.6.5.1 Material Properties

Several research studies investigated the impact of concrete compressive strength and steel
stress-stain response on wall performance. Liu [59], Lefas et al. [56], and Oesterle [78, 80]
investigated the impact of concrete compressive strength:

• Liu [59] tested two walls that were identical with the exception of concrete compres-
sive strength and shear reinforcement. Both failed at the same drift and due to the
139

same mechanism: bar buckling of longitudinal reinforcement in the boundary element.


However, the wall with normal strength concrete exhibited concrete crushing over a
much larger area of the wall than did the wall with high strength concrete, for which
concrete crushing was limited to the boundary regions.
• Lefas et al. [56] varied the concrete strength in one specimen (SW26 )and concluded
that concrete compressive strength did not impact the wall strength or deformational
capacity of the wall.
• Oesterle et al. [78, 80] investigated the impact of concrete compressive strength on
the performance of one barbell wall design, testing two identically configured walls
constructed with different strength concrete; however one wall was subjected to axial
load while the other was not, therefore the results are not conclusive with respect to
concrete compressive strength.

Only one previous study investigated the impact of steel ductility on the performance of
walls (Dazio et al. [25]). Dazio et at. found that specimens constructed using low ductility
steel failed due to bar fracture at drifts lower than specimens constructed using more ductile
steel.
Figure 2.99 shows the drift capacity as a function of the material properties (concrete
compressive strength and steel ultimate strain, εu ) for rectangular walls. Table 2.8 indicates
that i) walls for which failure is due to a flexure mechanism have an increased drift capacity
for higher concrete compressive strengths, and ii) walls for which failure is due to a shear-
compression mechanism have an increased drift capacity for larger ultimate steel strains.

2.6.5.2 Geometry

None of the experimental test programs summarized in Section 2.5 varied the geometry (e.g.
length `w or thickness, tw ) of rectangular cross-sections. Figure 2.100 plots drift capacity
as a function of a few selection geometric properties of rectangular walls, with failure mode
indicated by marker shape. The following paragraphs discuss these geometric properties
individually.
140

Shear Shear
Flexure Flexure
Shear-Compression (Cyclic) Shear-Compression (Cyclic)
Shear-Compression (Monotonic) Shear-Compression (Monotonic)

4 4

3 3
% Drift Capacity

% Drift Capacity
2 2

1 1

0 0
2 4 6 8 10 12 0 50 100 150 200
fc′ εu (mµ)

(a) Concrete compressive strength (b) Steel ultimate strain

Figure 2.99: Drift capacity for rectangular walls as a function of material strengths.

Scale: Different researchers define test specimen scale using different measures and many
researchers do not report scale. Thus, the repeated scale is an arbitrary quantity for in
comparing test specimens. In Figure 2.100a, drift is plotted as a function of the calculated
scale of the walls, where the calculated scale is defined as the thickness of the wall, in
inches, divided by 12 inches; thus a 12 inch wall is considered full-scale. As is discussed in
Section 2.3, 12 inches represents a lower bound on the thickness of walls in the West Coast
building inventory. Figure 2.100a and Table 2.8 indicate a positive correlation between
computed scale and drift capacity (increased drift at larger scales) for shear-compression
and shear failures. The data in Figure 2.100a do show that few large scale tests (scale ≥ 0.5)
have been done.

Cross-section aspect ratio: The cross-section aspect ratio is the length of the wall,
lw , divided by the thickness of the wall, tw , and is a useful parameter for considering the
geometry of the specimens as it is independent of scale. As is discussed in Section 2.6.2,
many tests have been done with cross-section aspect ratios less than 14.0 but few walls
with very slender cross-sections (`w /tw > 14) have been tested, despite the indication of
141

Shear Shear
Flexure Flexure
Shear-Compression (Cyclic) Shear-Compression (Cyclic)
Shear-Compression (Monotonic) Shear-Compression (Monotonic)

4 4

3 3

% Drift Capacity
% Drift Capacity

2 2

1 1

0 0
0 0.2 0.4 0.6 0.8 1 0 5 10 15 20
Shear Scale Shear ℓ w /tw
Flexure Flexure
Shear-Compression
(a) Scale (Cyclic)
(tw /12 inches) Shear-Compression
(b) Cross-section (Cyclic)
aspect ratio (lw /tw )
Shear-Compression (Monotonic) Shear-Compression (Monotonic)

4 4

3 3
% Drift Capacity

% Drift Capacity

2 2

1 1

0 0
0.05 0.1 0.15 0.2 0.25 0.5 1 1.5 2 2.5 3
ℓ be /ℓ w ℓ be /tw

(c) Boundary element length as percentage of wall (d) Boundary element aspect ratio
length

Figure 2.100: Drift capacity as a function of wall geometry.

the building inventory that such walls are used commonly in mid-rise structures. The data
in Figure 2.100b and Table 2.8 indicate a positive correlation between `w /tw and drift
capacity for flexural failures (increased drift capacity with increased slenderness) and a
negative correlation for shear-compression failures (decreased drift capacity with increased
slenderness).
142

Boundary element geometry: The drift capacity of rectangular walls is plotted versus
the boundary element length normalized by the wall length, `b /`w , in Figure 2.100c and ver-
sus the boundary element aspect ratio, `b /tw , in Figure 2.100d. These figures and Table 2.8
indicate for walls for which failure is due to a shear mechanism that walls with longer, more
slender boundary elements have higher drift capacities. When considering flexure and shear
compression failure mechanisms, there is a positive and negative correlation, respectively
for boundary element slenderness, `b /tw .

2.6.5.3 Reinforcement Ratios

Figure 2.101 shows drift capacity plotted versus reinforcement ratios for rectangular walls.
The impact on wall response of the boundary element longitudinal reinforcement ratio, ρbe ,
boundary element confining reinforcement ratio, ρcon , web vertical reinforcement ratio, ρweb ,
and web horizontal reinforcement ratio, ρh are discussed in the following paragraphs.

Boundary Element Reinforcement: The boundary element longitudinal reinforcement


ratio, ρbe , was a test parameter in the studies by Oesterle et al. [78, 80], Lefas et al. [56],
Pilakoutas and Elnashai [88], and Dazio et al. [25]. The walls tested by Pilakoutas and El-
nashai have reinforcement, ρbe , in excess of 6.9% and are not considered to be representative
of modern walls. Dazio et al. concluded that walls with larger longitudinal reinforcement
ratios have reduced flexure-shear cracking; however, horizontal reinforcement ratios were
held constant for these tests, and the difference in crack patterns may have been affected
by the shear demand-capacity ratio.
Figure 2.101a shows drift capacity as a function of ρbe . Most wall specimens have
reinforcement ratios between 1.5% and 4.0% and, thus, are representative of modern con-
struction (see Figure 2.93 presenting building inventory). Those tests with greater than
4.0% reinforcement included i) non-modern walls tested by Elnady [32], ii) four of the five
specimens that failed in shear33 iii) the wall tested by Tupper [106] and iv) the walls tested
by Pilakoutas and Elnashai [88]. Table 2.8 indicates a negative correlation between drift
capacity and boundary element reinforcement for walls for which failure is due to shear;
33
The wall tested by Riva et al. [92] failed in shear but had ρbe = 2.3%.
143

Shear Shear
Flexure Flexure
Shear-Compression (Cyclic) Shear-Compression (Cyclic)
Shear-Compression (Monotonic) Shear-Compression (Monotonic)

4 4

3 3
% Drift Capacity

% Drift Capacity
2 2

1 1

0 0
0 2 4 6 8 10 12 0 1 2 3 4
Shear ρ be Shear ρ con
Flexure Flexure
Shear-Compression
(a) B.E. (Cyclic) Shear-Compression
(b) Confining (Cyclic)
Shear-Compression (Monotonic) Shear-Compression (Monotonic)

4 4

3 3
% Drift Capacity

% Drift Capacity

2 2

1 1

0 0
0 0.5 1 1.5 2 2.5 0 0.2 0.4 0.6 0.8 1
ρ w eb ρh

(c) Web (d) Horizontal

Figure 2.101: Drift capacity as a function of reinforcement ratios.

however, if only the walls with less than ρbe = 4.0% are considered, there is no indication
of correlation between drift capacity and boundary element reinforcement ratios.

Confining Reinforcement: Studies by Oesterle et al. [78, 80] and Thomsen and Wallace
[104] address the impact of the boundary element confining reinforcement ratio,rhocon , on
the performance of walls. Oesterle et al. [80] tested two specimens, one with conventionally
detailed column confinement and one with hoop and cross-tie confinement, and concluded
that confinement should be detailed to delay bar buckling and to confine the core concrete,
144

but not to increase the usable strain of the core concrete. Thomsen and Wallace [104] tested
two rectangular walls with the same confining reinforcement ratio, ρcon , but different spac-
ing and configuration of the confinement and found that smaller spacing of reinforcement
delayed the drift at which bar buckling first occurred but did not affect the drift capacity
of the walls.
Figure 2.101b shows drift capacity as a function of the confining reinforcement ratio,
ρcon , for rectangular walls. Table 2.8 indicates strong correlation (R2 ≥ 0.46) between drift
capacity and confining reinforcement ratio, ρc , for walls for which failure is due to a flexure
or shear-compression mechanism. No correlation is indicated for walls for which failure is
due to a shear mechanism.

Web Longitudinal Reinforcement: No previous research study has investigated the


impact of web longitudinal reinforcement ratio, ρweb , on wall performance. Figure 2.101c
shows drift capacity as a function of ρweb . Table 2.8 indicates that regardless of mech-
anism associated with failure (flexure, shear-compression, or shear), there is a negative
correlation between web longitudinal reinforcement and drift capacity (decreased capacity
with increased ρweb ). However, if the walls loaded under monotonically increasing force-
controlled loading are neglected, this correlation is no longer present for walls exhibiting
shear-compression failures.

Horizontal Reinforcement Ratios: The impact of the web horizontal reinforcement


ratio, ρh , was investigated by Oesterle et al. [78, 80] and Pilakoutas and Elnashai [88].
Pilakoutas and Elnashai concluded that the volume of horizontal reinforcement determines
the failure mechanism but not the strength and deformation characteristics of the wall.
Figure 2.101d shows drift capacity as a function of ρh for rectangular walls. Table 2.8
indicates that walls for which failure is due to a flexural mechanism, there is a decrease
in drift capacity for increased web horizontal reinforcement ratio. For cyclically loaded
walls for which failure is due to a shear-compression mechanism, there is a increase in drift
capacity for increased web horizontal reinforcement.
145

2.6.5.4 Shear Span Ratio

The shear span ratio, M/(V `w ), is the ratio of the effective loading height to the wall
length. Lefas et al. [56] investigated shear span ratio but looked at ratios for squat walls
(M/(V `w ) < 2.0) and slender walls (M/(V `w ) = 2.0). Dazio et al. [25] and Paterson
and Mitchel [85] varied the shear span ratio for test specimens; however, the variations
were small. Brueggen [15] did not investigate the impact of shear span ratio for T-shape
walls, but did investigate the impact of specimen height for a fixed shear span ratio; the
results of the tests indicate that, if the specimen height exceeds the plastic hinge length and
specimens are loaded with appropriate boundary conditions, shorter specimens exhibit the
same performance as taller specimens with the same ratio of base reactions.

Shear
Flexure
Shear-Compression (Cyclic)
Shear-Compression (Monotonic)

3
% Drift Capacity

0
2 3 4 5 6
M /(V ℓ w )

Figure 2.102: Drift capacity as a function of shear span.

Figure 2.102 shows the drift capacity versus the shear span ratio, M/(V `w ). Most test
specimens have shear span ratios less than 2.5; thus it is difficult to assess the impact
of this parameter on wall response. As the building inventory summarized in Section 2.3
indicates, the shear span ratios of experimental tests found in the literature are lower bound
representations of modern construction. Additional tests of walls with higher aspect ratio
are needed to better understand the impact of shear span ratio on the failure mode and
146

drift capacity of walls. From the limited data available for larger shear span ratios, the
correlation coefficients in Table 2.8 suggests that for specimens for which failure is the
result of a shear-compression mechanisms, higher shear span ratios result may result in
higher drift capacity.

2.6.5.5 Axial Load Ratio

The axial load ratio, λN , is the ratio of the applied axial force to the gross axial capacity
of the cross-section, Ag fc0 . The impact on response of axial load ratio was investigated by
Oesterle et al. [78, 80] and Lefas et al. [56]. Oesterle et al. [80, 78] investigated the impact of
axial load on the behavior of barbell walls and found that, contrary to accepted knowledge,
axial load increased the ductility of walls; Oesterle et al. [78, 80] concluded this was due to
a reduction in shear deformations in walls with higher axial loads. Lefas et al. [56] tested
walls under monotonically increasing lateral loading and constant axial load ratios of 0.0,
0.1, and 0.2; they observed that axial load increased the stiffness and strength of the walls
but i) decreased the drift capacity, ii) increased compressive damage to boundary elements,
iii) reduced the maximum recorded tensile strain and iv) but did not affect the maximum
compressive strain.

Shear
Flexure
Shear-Compression (Cyclic)
Shear-Compression (Monotonic)

3
% Drift Capacity

0
0 0.05 0.1 0.15 0.2 0.25
λN

Figure 2.103: Drift capacity as a function of axial load.


147

Figure 2.103 shows the drift capacity versus the axial load ratio, λN , for rectangular
walls. Table 2.8 indicates a negative correlation between axial load ratio and drift capacity
(decreased drift capacity for increased axial load ratio) for walls for which the failure was
due to a flexural mechanism. No trend is observed for walls for which failure was due to a
shear-compression mechanism; however, given the sparsity of the data over a range of axial
load rations, it is not possible to identify relationships between drift capacity and axial load
ratio.

2.6.5.6 Shear Demand and Capacity

To evaluate the shear demand and capacitity of experimental tests of rectangular slender
walls, the following values were considered i) ACI 318-08 nominal shear strength, Vn , ii) ratio
of the shear demand at nominal moment, Vmn , to the nominal shear strength, Vmn /Vn , iii)
the expected shear demand, Ven , determined as the smaller of Vn and Vmn , iv) the maximum
experimental shear demand, Vu , v) the ratio of the maximum shear demand to nominal shear
strength, Vu /Vn , and vi) the ratio of the maximum to the expected shear demand, Vu /Ven .
Figure 2.104 shows the drift capacity versus each of these values; note that Vn , Ven , and Vu
p
are plotted normalized by Acv fc0 , where Vn , Ven , and Vu are in lbs, fc0 is in psi, and Acv is
in inches.
Wall specimens with flexural failure modes generally have expected shear stresses,
p p
Ven /(Acv fc0 ), less than 3.0 and experimental maximum shear stress demands, Vu /(Acv fc0 ),
less than 4.0. This is consistent with the observation by Wood [115], in her study inves-
tigating the failure mechanism of structural walls, that flexural failures occurred in walls
p
with shear stress demands less than 4 fc0 .
Shear-compression failures are observed for walls with a wide range of expected shear
stress and measured shear stress demands. Table 2.8 suggests there is no significant correla-
tion between drift capacity and shear demand for these walls; however, Figure 2.104c and 2.104d
p
suggest that for shear demands less than 4Acv fc0 , shear does not impact the drift capacity
p
of the walls, but for demands greater than 4Acv fc0 , increased shear demand is ocrrelated
with a decreased drift capacity.
148

Shear Shear
Flexure Flexure
Shear-Compression (Cyclic) Shear-Compression (Cyclic)
Shear-Compression (Monotonic) Shear-Compression (Monotonic)

4 4

3 3

% Drift Capacity
% Drift Capacity

2 2

1 1

0 0
2 4 6 8 10 12 0 0.5 1 1.5 2
ShearV /(A pf ′ ) Shear Vmn/Vn
n cv c
Flexure Flexure
Shear-Compression
(a) (Cyclic) (b)
Shear-Compression (Cyclic)
Shear-Compression (Monotonic) Shear-Compression (Monotonic)
4 4

3 3
% Drift Capacity

% Drift Capacity

2 2

1 1

0 0
1 2 3 4 5 6 7 0 2 4 6 8 10
ShearVen /(Acv pf ′ ) ShearV /(A pf ′ )
c u cv c
Flexure Flexure
Shear-Compression
(c) (Cyclic) Shear-Compression
(d) (Cyclic)
Shear-Compression (Monotonic) Shear-Compression (Monotonic)

4 4

3 3
% Drift Capacity

% Drift Capacity

2 2

1 1

0 0
0 0.5 1 1.5 0.5 1 1.5
Vu /Vn Vu /Ven

(e) (f)

Figure 2.104: Drift capacity as a function of shear demand and capacity values.
149

The most striking observation to be made from the plots of drift capacity versus the
shear demand and capacity parameters relates to the ratio of maximum experimental shear
demand to expected shear demand (Figure 2.104f). The majority of walls reached or ex-
ceeded the expected shear demand, Ven . The four flexure failures with ratios just larger
than 0.50 are older specimens tested as control walls for retrofit programs (Elnady [32] and
Paterson and Mitchell [84]) and failed due to bond slip of lap splices, a failure mode con-
sidered for in the calculation of the expected shear. Walls that failed in shear had Vu /Ven
ratios ranging between 1.14 and 1.49, indicating that the ACI 318-08 Code definition of
shear strength is conservative. The walls with non-bond slip failures and Vu /Ven < 1.0 are
i) both walls tested by Liu [59], ii) three walls tested by Pilakoutas and Elnashai [88], and
iii) one specimen tested by Tasnimi [102]; however, these are all greater than 0.90.
150

Table 2.9: Drift capacity, failure and code compliance for experimental test programs

ACI Compliance
1
Researcher Name Drift Capacity (%) Failure Mode Y/N Reasons2
Ali, Wight W1 3.1 F N H
TUA 3.1 F Y
Beyer et al.
TUB 2.7 WC N TS
NTW1 1.0 SC Y
Brueggen
NTW2 4.1 SC Y
WSH1 1.1 F N TS
WSH2 1.4 F N TS
WSH3 2.0 F Y
Dazio
WSH4 1.6 SC N SBE
WSH5 1.5 F Y
WSH6 1.4 SC Y
C1 2.6 S N SBE
Khalil et al. CW2 0.2 F N SBE
CW3 0.3 F N SBE
X 3.1 F N TS
Ile et al. Y 3.1 F N TS
XY 2.0 F N TS
SW21 1.6 SC N TS
SW22 1.2 SC N TS
SW23 1.0 SC N BEL, TS
SW24 1.4 SC N TS
SW25 0.8 SC N BEL, TS
Lefas
SW26 1.6 SC N TS
SW30 1.6 SC N TS
SW31 1.7 SC N TS
SW32 1.9 SC N TS
SW33 1.9 SC N TS
W1 3.0 SC Y
Liu
W2 3.0 F Y
Mobeen W1 3.8 F Y
Morgan et. al. W1 1.8 SC Y
1
F - Flexure, S - Shear, SC - Shear-compression, WC - Web crushing
2
TS - Inadequate transverse reinforcement spacing, SBE - No Special B.E.
BEL - Inadequate B.E. Length, H - Inadequate horizontal reinf., W - Inadequate web reinf.
Continued on next page. . .
151

Table 2.9: Drift capacity, failure and code compliance for experimental test programs (con’t)

ACI Compliance
1
Researcher Name Drift Capacity (%) Failure Mode Y/N Reasons2

R1 2.8 F Y
R2 3.4 F Y
B1 3.4 SC Y
B2 2.9 WC Y
B3 4.5 F Y
B4 7.4 F Y
B5 2.8 WC Y
Oesterle et al.
B6 1.4 WC Y
B7 2.7 WC Y
B8 3.2 WC Y
B9 3.0 WC Y
B10 3.4 SC Y
F1 2.4 WC Y
F2 2.6 WC Y
WR-20 2.7 SC N TS
WR-10 2.9 SC N TS
Oh et al
WR-0 2.2 SC N SBE
WB 2.8 SC N TS
W1 0.5 F N H, SBE
Paterson et al.
W2 1.6 F N H, SBE
SW4 2.0 SC N TS
SW5 0.8 S N TS
SW6 1.8 SC N TS
Pilakoutas
SW7 1.8 S N TS
SW8 2.2 SC N TS
SW9 2.2 SC N TS
Riva et al. W1 3.2 S N W, H
Shiu et al. C1 2.9 S Y
SHW1 1.6 SC N SBE
SHW2 1.0 SC N SBE
Tasnimi
SHW3 1.0 SC N SBE
SHW4 1.7 SC N TS
RW1 2.5 SC N TS
RW2 2.5 SC Y
Thomsen et al.
TW1 1.3 SC N BEL, TS
TW2 2.0 SC Y
Tupper W3 2.9 SC Y
1
F - Flexure, S - Shear, SC - Shear-compression, WC - Web crushing
2
TS - Inadequate transverse reinforcement spacing, SBE - No Special B.E.
BEL - Inadequate B.E. Length, H - Inadequate horizontal reinf., W - Inadequate web reinf.
152

Table 2.10: Geometric and material parameters for experimental test programs

fc0 fy
1 2
Researcher Name Shape lw /tw Scale ksi (MPa) ksi (MPa)
Ali, Wight W1 B 16.00 0.25 4.7 (32) 78 (540)
TUA C 8.67 0.49 11.3 (78) 71 (488)
Beyer et al.
TUB C 13.00 0.33 7.9 (55) 68 (471)
NTW1 T 15.00 0.50 7.3 (50) 64 (438)
Brueggen
NTW2 T 15.00 0.50 6.6 (45) 67 (460)
WSH1 R 13.33 0.49 6.5 (45) 79 (545)
WSH2 R 13.33 0.49 5.9 (41) 84 (580)
WSH3 R 13.33 0.49 5.7 (39) 87 (599)
Dazio
WSH4 R 13.33 0.49 5.9 (41) 83 (574)
WSH5 R 13.33 0.49 5.6 (38) 85 (583)
WSH6 R 13.33 0.49 6.6 (46) 83 (574)
C1 R 8.33 0.39 5.5 (38) 68 (470)
Khalil et al. CW2 R 8.33 0.39 5.4 (37) 65 (450)
CW3 R 8.33 0.39 5.4 (38) 65 (450)
X C 6.00 0.82 3.4 (24) 75 (516)
Ile et al. Y C 6.00 0.82 3.4 (24) 75 (516)
XY C 6.00 0.82 3.0 (21) 75 (516)
SW21 R 10.00 0.21 5.0 (34) 68 (470)
SW22 R 10.00 0.21 5.1 (35) 68 (470)
SW23 R 10.00 0.21 5.6 (39) 68 (470)
SW24 R 10.00 0.21 5.1 (35) 68 (470)
SW25 R 10.00 0.21 5.3 (37) 68 (470)
Lefas
SW26 R 10.00 0.21 3.5 (24) 68 (470)
SW30 R 10.00 0.21 3.5 (24) 68 (470)
SW31 R 10.00 0.21 4.1 (28) 68 (470)
SW32 R 10.00 0.21 6.0 (42) 68 (470)
SW33 R 10.00 0.21 5.4 (37) 68 (470)
W1 R 6.00 0.66 4.8 (33) 66 (458)
Liu
W2 R 6.00 0.66 10.3 (71) 66 (458)
Mobeen W1 B 6.00 0.66 3.5 (24) 57 (395)
Morgan et. al. W1 B 27.50 0.19 4.6 (32) 65 (448)
1
R - Rectangular, B - Barbell, C - C-shaped, H - H-shaped, T - T-shaped
2
Scale = tw /12 (inches)
Continued on next page. . .
153

Table 2.10: Geometric and material parameters for experimental test programs (con’t)

fc0 fy
1 2
Researcher Name Shape lw /tw Scale ksi (MPa) ksi (MPa)

R1 R 18.75 0.33 6.5 (45) 74 (511)


R2 R 18.75 0.33 6.7 (46) 65 (450)
B1 B 18.75 0.33 7.7 (53) 65 (449)
B2 B 18.75 0.33 7.8 (54) 60 (410)
B3 B 18.75 0.33 6.9 (47) 64 (438)
B4 B 18.75 0.33 6.5 (45) 65 (450)
B5 B 18.75 0.33 6.6 (45) 64 (444)
Oesterle et al.
B6 B 18.75 0.33 3.2 (22) 64 (440)
B7 B 18.75 0.33 7.2 (49) 66 (457)
B8 B 18.75 0.33 6.1 (42) 65 (447)
B9 B 18.75 0.33 6.4 (44) 62 (429)
B10 B 18.75 0.33 6.6 (46) 55 (378)
F1 H 18.75 0.33 5.6 (38) 65 (444)
F2 H 18.75 0.33 6.6 (46) 65 (444)
WR-20 R 7.50 0.66 5.0 (34) 65 (449)
WR-10 R 7.50 0.66 5.3 (36) 65 (449)
Oh et al
WR-0 R 7.50 0.66 4.8 (33) 65 (449)
WB B 12.00 0.41 5.3 (36) 65 (449)
W1 R 4.00 0.98 3.8 (26) 61 (423)
Paterson et al.
W2 R 4.00 0.98 4.8 (33) 61 (423)
SW4 R 10.00 0.20 5.4 (37) 68 (470)
SW5 R 10.00 0.20 4.6 (32) 78 (535)
SW6 R 10.00 0.20 5.6 (39) 68 (470)
Pilakoutas
SW7 R 10.00 0.20 4.6 (32) 78 (535)
SW8 R 10.00 0.20 6.6 (46) 62 (430)
SW9 R 10.00 0.20 5.6 (39) 62 (430)
Riva et al. W1 R 9.33 0.98 4.7 (33) 81 (560)
Shiu et al. C1 R 18.75 0.33 3.4 (23) 69 (476)
SHW1 R 10.00 0.16 3.1 (22) 40 (276)
SHW2 R 10.00 0.16 3.1 (22) 40 (276)
Tasnimi
SHW3 R 10.00 0.16 3.3 (22) 40 (276)
SHW4 R 10.00 0.16 3.4 (23) 40 (276)
RW1 R 12.00 0.33 4.6 (32) 63 (433)
RW2 R 12.00 0.33 4.9 (34) 63 (433)
Thomsen et al.
TW1 T 12.00 0.33 6.3 (44) 63 (433)
TW2 T 12.00 0.33 6.0 (42) 63 (433)
Tupper W3 R 6.58 0.50 5.6 (39) 65 (450)
1
R - Rectangular, B - Barbell, C - C-shaped, H - H-shaped, T - T-shaped
2
Scale = tw /12 (inches)
154

Table 2.11: Reinforcement ratios for experimental test programs

ρbe ρweb ρh ρc
Researcher Name % % % %
Ali, Wight W1 3.20 0.28 0.28 2.67
TUA 1.29 0.25 0.30 1.64
Beyer et al.
TUB 2.17 0.38 0.45 2.44
NTW1 3.78 0.28 0.29 1.13
Brueggen
NTW2 2.64 0.86 0.61 1.12
WSH1 1.57 0.30 0.25 1.08
WSH2 1.57 0.30 0.25 1.12
WSH3 1.74 0.54 0.25 1.04
Dazio
WSH4 1.74 0.54 0.25 0
WSH5 0.77 0.27 0.25 1.00
WSH6 1.74 0.54 0.25 1.68
C1 10.00 0.83 0.27 0
Khalil et al. CW2 6.41 1.67 0.26 0
CW3 6.41 1.67 0.26 0
X 0.98 0.25 0.54 1.09
Ile et al. Y 0.98 0.25 0.54 1.09
XY 0.98 0.25 0.54 1.09
SW21 2.99 2.49 0.82 0.56
SW22 2.99 2.49 0.82 0.56
SW23 2.99 2.49 0.82 0.56
SW24 2.99 2.49 0.82 0.56
SW25 2.99 2.49 0.82 0.56
Lefas
SW26 2.99 2.49 0.41 0.56
SW30 2.99 1.55 0.36 0.49
SW31 2.99 1.55 0.36 0.49
SW32 2.99 1.55 0.36 0.49
SW33 2.99 1.55 0.36 0.49
W1 3.00 0.33 0.40 2.32
Liu
W2 3.00 0.33 0.47 2.32
Mobeen W1 1.60 0.67 1.00 2.15
Morgan et. al. W1 1.11 0.42 0.44 1.20
Continued on next page. . .
155

Table 2.11: Reinforcement ratios for experimental test programs (con’t)

ρbe ρweb ρh ρc
Researcher Name % % % %

R1 1.47 0.28 0.31 0


R2 4.00 0.28 0.31 3.99
B1 1.11 0.28 0.31 0.48
B2 3.67 0.28 0.63 0.42
B3 1.11 0.28 0.63 4.75
B4 1.11 0.28 0.63 4.75
B5 3.67 0.28 0.63 5.91
Oesterle et al.
B6 3.67 0.28 0.63 3.57
B7 3.67 0.28 0.63 5.91
B8 3.67 0.28 1.38 5.91
B9 3.67 0.28 0.63 5.91
B10 1.97 0.28 0.63 5.93
F1 3.89 0.28 0.71 1.11
F2 5.01 0.24 0.55 3.15
WR-20 1.47 0.32 0.28 1.37
WR-10 1.47 0.32 0.32 2.74
Oh et al
WR-0 1.47 0.32 0.28 0
WB 0.99 0.36 0.28 3.20
W1 ) 2.26 0.38 0.19 0
Paterson et al.
W2 2.26 0.38 0.19 0
SW4 6.85 0.75 0.39 1.54
SW5 11.41 0.79 0.35 0.88
SW6 6.85 0.75 0.35 0.71
Pilakoutas
SW7 11.08 0.79 0.39 1.85
SW8 7.14 0.75 0.42 0.85
SW9 7.14 0.75 0.60 1.22
Riva et al. W1 2.30 0.17 0.17 0.96
Shiu et al. C1 5.58 0.24 0.37 2.78
SHW1 3.23 0.28 0.28 0
SHW2 3.23 0.28 0.28 0
Tasnimi
SHW3 3.23 0.28 0.28 0
SHW4 3.23 0.28 0.28 0
RW1 2.93 0.33 0.33 1.06
RW2 2.93 0.33 0.33 1.21
Thomsen et al.
TW1 3.27 0.33 0.33 1.06
TW2 2.51 0.45 0.33 0.89
Tupper W3 5.36 0.44 0.73 0.96
Continued on next page. . .
156

Table 2.12: Load and capacity parameters for experimental test programs

Researcher Name M/(V L) λn αVn αVmn αVen αVu Vmn /Vn Vu /Vn Vu /Ven
Ali, Wight W1 3.04 0.08 5.3 3.5 3.5 3.7 0.67 0.69 1.04
TUA 2.58 0.02 4.1 3.7 3.7 3.2 0.90 0.79 0.87
Beyer et al.
TUB 2.58 0.04 5.8 6.3 5.8 5.7 1.09 0.99 0.99
NTW1 3.47 0.03 4.5 3.0 3.0 4.9 0.67 1.09 1.62
Brueggen
NTW2 3.47 0.03 7.5 3.3 3.3 5.4 0.44 0.72 1.64
WSH1 2.28 0.05 4.6 1.9 1.9 2.0 0.41 0.44 1.06
WSH2 2.28 0.06 4.3 2.0 2.0 2.3 0.46 0.53 1.15
WSH3 2.28 0.06 4.4 2.7 2.7 2.9 0.61 0.67 1.09
Dazio
WSH4 2.28 0.06 4.5 2.6 2.6 2.8 0.59 0.63 1.07
WSH5 2.28 0.13 4.5 2.5 2.5 2.9 0.55 0.63 1.14
WSH6 2.26 0.11 4.3 3.2 3.2 3.5 0.75 0.81 1.09
C1 2.25 0.07 5.2 9.6 5.2 5.9 1.85 1.14 1.14
Khalil et al. CW2 5.00 0.08 4.9 2.6 2.6 1.6 0.53 0.33 0.61
CW3 2.25 0.08 4.9 5.9 4.9 3.5 1.19 0.71 0.71
X 2.40 0.10 9.4 5.8 5.8 6.0 0.62 0.64 1.03
Ile et al. Y 2.40 0.10 9.4 3.5 3.5 3.0 0.37 0.32 0.86
XY 2.40 0.11 9.9 6.2 6.2 6.0 0.62 0.60 0.97
SW21 2.00 0.00 10.8 5.0 5.0 6.2 0.46 0.57 1.24
SW22 2.00 0.12 10.7 5.8 5.8 7.2 0.54 0.68 1.24
SW23 2.00 0.21 10.3 6.2 6.2 8.3 0.61 0.80 1.33
SW24 2.00 0.00 10.7 4.9 4.9 5.8 0.46 0.54 1.17
SW25 2.00 0.21 10.5 6.2 6.2 7.1 0.59 0.67 1.14
Lefas
SW26 2.00 0.00 7.2 5.5 5.5 7.1 0.76 0.99 1.30
SW30 2.00 0.00 6.6 4.9 4.9 6.8 0.74 1.03 1.39
SW31 2.00 0.00 6.3 4.7 4.7 6.2 0.74 0.99 1.33
SW32 2.00 0.00 5.5 4.1 4.1 4.9 0.74 0.89 1.20
SW33 2.00 0.00 5.7 4.3 4.3 5.2 0.74 0.91 1.23
W1 3.13 0.08 5.9 2.3 2.3 2.3 0.40 0.39 0.99
Liu
W2 3.13 0.04 5.1 1.9 1.9 1.7 0.37 0.33 0.90
Mobeen W1 2.74 0.15 11.4 3.5 3.5 3.8 0.31 0.34 1.10
Morgan et. al. W1 2.79 0.05 6.8 2.0 2.0 2.5 0.30 0.37 1.24
Continued on next page. . .
157

Table 2.12: Load and capacity parameters for experimental test programs (con’t)

Researcher Name M/(V L) λn αVn αVmn αVen αVu Vmn /Vn Vu /Vn Vu /Ven

R1 2.40 0.00 4.9 1.1 1.1 1.1 0.22 0.22 1.01


R2 2.40 0.00 5.0 1.9 1.9 2.0 0.37 0.40 1.07
B1 2.40 0.00 4.7 2.6 2.6 2.3 0.55 0.49 0.90
B2 2.40 0.00 7.5 5.8 5.8 5.8 0.78 0.77 1.00
B3 2.40 0.00 7.2 2.6 2.6 2.5 0.36 0.34 0.95
B4 2.40 0.00 7.7 2.7 2.7 3.1 0.35 0.41 1.16
B5 2.40 0.00 7.6 6.5 6.5 7.0 0.85 0.93 1.09
Oesterle et al.
B6 2.40 0.13 10.2 10.3 10.2 11.0 1.01 1.07 1.07
B7 2.40 0.08 7.2 8.1 7.2 8.7 1.11 1.20 1.20
B8 2.40 0.09 14.3 8.4 8.4 9.4 0.59 0.66 1.11
B9 2.40 0.09 7.2 8.1 7.2 9.2 1.12 1.27 1.27
B10 2.40 0.08 7.3 5.5 5.5 9.0 0.75 1.24 1.65
F1 2.40 0.00 9.3 9.8 9.3 8.4 1.05 0.90 0.90
F2 2.40 0.07 6.5 8.5 6.5 8.2 1.30 1.25 1.25
WR-20 2.00 0.10 4.0 2.7 2.7 3.0 0.67 0.76 1.14
WR-10 2.00 0.10 4.2 2.7 2.7 2.8 0.64 0.67 1.06
Oh et al
WR-0 2.00 0.10 4.0 2.7 2.7 3.0 0.66 0.73 1.12
WB 2.00 0.10 3.9 3.9 3.9 3.8 0.98 0.95 0.97
W1 2.71 0.10 3.5 2.3 2.3 1.3 0.66 0.36 0.55
Paterson et al.
W2 3.13 0.10 3.4 2.0 2.0 1.1 0.60 0.33 0.54
SW4 2.00 0.00 6.2 6.0 6.0 5.7 0.97 0.92 0.95
SW5 2.00 0.00 5.0 8.2 5.0 7.0 1.64 1.40 1.40
SW6 2.00 0.00 4.7 5.9 4.7 5.8 1.25 1.24 1.24
Pilakoutas
SW7 2.00 0.00 6.6 8.2 6.6 7.5 1.25 1.15 1.15
SW8 2.00 0.00 5.0 5.5 5.0 4.7 1.11 0.94 0.94
SW9 2.00 0.00 6.6 5.9 5.9 5.3 0.88 0.79 0.90
Riva et al. W1 3.17 0.00 4.0 1.4 1.4 2.0 0.34 0.51 1.49
Shiu et al. C1 2.88 0.00 6.3 3.4 3.4 4.3 0.54 0.69 1.26
SHW1 2.20 0.00 3.6 1.7 1.7 1.6 0.47 0.45 0.94
SHW2 2.20 0.00 3.6 1.7 1.7 2.1 0.47 0.57 1.22
Tasnimi
SHW3 2.20 0.00 3.6 1.7 1.7 1.7 0.47 0.47 1.00
SHW4 2.20 0.00 3.5 1.6 1.6 2.3 0.47 0.66 1.42
RW1 3.00 0.10 5.1 2.4 2.4 2.6 0.48 0.50 1.05
RW2 3.00 0.07 5.0 2.2 2.2 2.6 0.44 0.53 1.21
Thomsen et al.
TW1 3.00 0.07 4.7 3.1 3.1 4.3 0.66 0.92 1.39
TW2 3.00 0.07 4.7 3.1 3.1 5.5 0.66 1.16 1.76
Tupper W3 3.75 0.10 8.9 3.3 3.3 4.3 0.37 0.48 1.29
158

2.7 Summary of Literature Review

This chapter presented a review of the seismic design and performance of slender reinforced

concrete walls, including i) a review of code-based and performance-based seismic design

procedures (Sections 2.1 and 2.2), ii) a review of drawings for West Coast walled buildings

(Section 2.3), iii) a review of observed damage to structural walls following earthquakes

(Section 2.4), and iv) a review of previous experimental research that investigated wall

performance. In Section 2.6, these individual reviews were considered together to provide

an understanding of how well wall design and performance of structural walls is replicated

in the laboratory. The following summarizes the observations made:

1. Only 10 rectangular specimens meeting ACI 318-08 [1] Code requirements have been

tested in the laboratory. These walls are generally representative of modern West

Coast construction, however, boundary element lengths relative to wall lengths are

shorter in specimens than in actual walls. Few of these tests have shear span ratios

greater than 3.0 and no tests have been tested for long, skinny walls with axial loads.

2. The major damage type observed most frequently (more than 20% of all observed dam-

age types) in U.S. structural walls is compressive damage (bar buckling and concrete

crushing) in the boundary element. By contrast, the major damage types observed

frequently (more than 20% of all observed damage types) in Chilean construction

include large diagonal shear cracks, web crushing, and horizontal planes of damage

that span the length of the wall, in addition to compressive damage in the boundary

element.

3. Damage in the lab is primarily compressive, reflecting that seen post-earthquake.

However, the damage seen at discontinuities in the field is a condition not observed

in the lab.

4. Previous experimental research has indicated that wall drift capacity is affected by

i) axial load, ii) shear demand, iii) confining reinforcement ratio, and iv) shear span

ratio. Evaluation of all previous experimental research indicates that drift capacity
159

may also be affected by i) cross-sectional aspect ratio, ii) web reinforcement ratios,

and iii) specimen scale.


160

Chapter 3

NEESR COMPLEX WALL PROJECT: PLANAR WALL TEST


PROGRAM

This chapter summarizes the planar wall experimental test program of the the project

“Behavior, Analysis, and Design of Complex Wall Systems”. The planar wall testing pro-

gram investigates the seismic response of large-scale wall reinforced concrete structural

walls to simulate modern mid-rise building construction on the West Coast. The parame-

ters considered by the test program were i) shear demand, ii) distribution of longitudinal

reinforcement, and iii) spliced longitudinal reinforcement. A detailed overview on the test

program can be found in the project summary document [63]. The information presented

in this chapter is limited to that needed to provide the background for the contents of this

dissertation (Chapter 4 summarizes the results of the tests and Chapters 5 and 6 present

analysis of experimentally collected data).

Section 3.1 provides a summary of the experimental test matrix and a brief overview

of the design procedures. Sections 3.2 and 3.3 provide material properties and expected

strengths, respectively. Section 3.4 provides a summary of specimen construction. Sec-

tion 3.6 provides an overview of the instrumentation used to monitor wall behavior. Sec-

tion 3.7 provides a summary of loading and displacement histories.

3.1 Test Matrix

A four-specimen test matrix was designed to investigate the impact of i) longitudinal re-

inforcement layout, ii) lateral load distribution and its impact on shear demand, and iii)

spliced longitudinal reinforcement at the base of the wall. Table 3.1 summarizes the test

program.

Specimen PW2 differs from each of the other walls in just one of the three test variables
161

Table 3.1: Test matrix of planar wall experimental test program

Specimen Reinforcement Lateral Load Distribution Splice?


PW1 B.E. ASCE 7 Yes

PW2 B.E. Uniform Yes

PW3 Uniform Uniform Yes

PW4 B.E. Uniform No

and is considered the reference specimen. Specimen PW2 had longitudinal reinforcement

concentrated in boundary elements, loading equivalent to a uniform lateral load distribution,

and spliced reinforcement at the base of the wall.

Specimen PW1 differs from Specimen PW2 in its lateral load layout. Specimen PW1

had longitudinal reinforcement concentrated in boundary elements, loading equivalent to

an ASCE 7 lateral load distribution, and spliced reinforcement at the base of the wall.

Specimen PW3 differs from Specimen PW2 in its reinforcement layout. Specimen PW3

had a uniform distribution of longitudinal reinforcement, loading equivalent to a uniform

lateral load distribution, and spliced reinforcement at the base of the wall.

Specimen PW4 differs from Specimen PW2 in the use of splices at the base of the wall.

Specimen PW4 had longitudinal reinforcement concentrated in boundary elements, loading

equivalent to a uniform lateral load distribution, and no splice at the base of the wall.

The following sections discuss the development of the test matrix. Section 3.1.1 com-

pares the test matrix to the building inventory presented in Section 2.6.2. Section 3.1.2

discusses the wall dimensions. Section 3.1.3 discusses the lateral load distributions con-
162

sidered. Sections 3.1.4 summarizes the reinforcement design. Section 3.1.5 discusses the

detailing of the lap splices.

3.1.1 Comparison of Design Parameters to Building Inventory and Other Tests

In Section 2.6.2, characteristics of experimental wall tests were compared to characteristics

of walls in the West Coast mid-rise structural wall building inventory. Figures 2.93 and 2.94

are reproduced here as Figures 3.1 and 3.2, and include the four specimens from this test

program1 . In each figure, the variables are plotted against the cross-sectional aspect ratio

(lw /tw ).

The wall specimens fall at the upper limit of cross-sectional aspect ratios found in the

building inventory. As was noted in Section 2.6.2, few experimental tests for walls of similar

cross-section aspect ratios (`w /tw ) can be found in the literature, and while these tests are

representative of modern construction in terms of reinforcement ratios, shear capacity, and

shear-demand capacity ratio, they were tested without axial load. Figure 3.1 shows that the

specimens in the test matrix presented in this chapter have reinforcement ratios represen-

tative of modern construction. Figure 3.2b shows the walls have boundary element lengths

representative of modern construction. Figure 3.2d shows the expected shear-demand ca-

pacity ratio of the specimens is on the upper end of demands for modern construction. Thus,

the specimens geometry reflects modern construction and the demands can be considered

as representative of worse-case scenario demands on modern walls.

3.1.2 Dimensions

A prototype wall was established based on a building inventory review (see Section 2.6.2)

and consultation with practicing engineers. The wall was 120 feet (36.6 m) tall (10 12

foot (3.66 m) tall stories), 18 inches (45.7 cm) thick, and 30 feet (9.1 m) in length. The

specimens tested in the laboratory were one-third scale of the prototype wall. Figure 3.3

1
Strength calculations were done using the actual material properties, which are found in Section 3.2
163

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
PW Tests PW Tests
6 1.6

5 1.4

1.2
4
1

ρweb
ρbe

3
0.8
2
0.6
1 0.4

0 0.2
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(a) Longitudinal boundary element steel (b) Longitudinal web steel

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
PW Tests PW Tests
3 1.6

2.5 1.4

1.2
2
1
ρh
ρv

1.5
0.8
1
0.6
0.5 0.4

0 0.2
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(c) Total longitudinal steel (d) Horizontal steel

Figure 3.1: Reinforcement ratios for building inventory and experimental test programs
(ACI 318-08 compliant walls only).
164

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
PW Tests PW Tests
20 35

30
15
25

ℓb /ℓw (%)
heff /lw

10 20

15
5
10

0 5
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(a) Aspect ratio (b) Boundary element length (percent of wall
length)

Bldg Inventory Bldg Inventory


Test - No Axial Test - No Axial
Test - Axial Test - Axial
PW Tests PW Tests
14 1.2

12
1
fc′ ) (psi)

10
0.8
Vmn /Vn
p

8
Vn /(Acv

0.6
6

4 0.4

2 0.2
0 5 10 15 20 25 0 5 10 15 20 25
ℓw /tw ℓw /tw
(c) Nominal shear stress (d) Shear demand-capacity ratio

Figure 3.2: Geometry and demand characteristics for building inventory and experimental
test programs (ACI 318-08 compliant walls only).
165

shows the geometry of the prototype and scaled walls. The test specimens were the lower 3

stories of the scaled wall and were 12 feet (12.2 m) tall (3 stories of 4 feet (1.22 m) tall), 6

inches (15.2 cm) thick, and 10 feet (3.05 m) in length.

`w = 10 ft
120 ft (36.6 m)

(3.05 m)
30 ft (9.1 m)

h10 = 40 ft (12.2 m)
18 in tw = 6 in
(45.7 cm) (15.2 cm)

hw = 12 ft (3.66 m)
Specimen

Full-scale Prototype 1
3 Scale

1
Figure 3.3: Dimensions of 10-story prototype wall and 3 scaled wall. The tested specimens
were the lower 3 floors of the scaled 10-story wall.

3.1.3 Lateral Load Distribution

To investigate the impact of shear demand, loads were applied to simulate two different

lateral load distributions on a 10-story wall. This section discusses the lateral load distri-

butions considered.

The first loading condition considered simulated the demands resulting from the ASCE

7-05 equivalent lateral force distribution. Shown in Figure 3.4, this load distribution acts

at an effective height of 0.71h10 , where h10 is the height of the 10-story scaled specimen.
p
At the design nominal moment strength, the expected shear demand was 2.75Ag fc0 , or
166

0.67Vn , where Vn is the design nominal shear strength of 210 kips.

lw

V10 tw

V9
V8
V7
Vb
V6
V5

hw
V4

hef f
V3
V2
V1

Mb Vb

Figure 3.4: ASCE 7 ELF lateral load distributions.

The second load distribution considered was a uniform distribution of lateral forces.

The second loading condition considered simulated the demands resulting from a uniform

distribution of lateral forces. Shown in Figure 3.5, this load distribution acts at an effective

height of 0.5h10 . At the design nominal moment strength, the expected shear demand was
p
3.90Ag fc0 , or 0.95Vn .

3.1.4 Reinforcement Design

The walls were reinforced to i) meet the minimum code specified horizontal reinforcement

ratio, ii) meet the capacity of the MUST-SIM equipment, iii) simulate average properties

(e.g. bar size, longitudinal reinforcement ratio) of modern construction, and iv) meet ACI

318-05 requirements2 . Two layouts of longitudinal reinforcement were designed: i) concen-

2
Specimens were design using the ACI 318-05 Building Code, but also meet the requirement of the ACI
318-08 Building Code.
167

V10
V9
P10
V8 Vapp = i=3 Vi

P10
V7 − 3))

hef f = 0.50h10 = 240 in (6.1 m)


Mapp = i=4 (Vi h(i

h10 = 40 ft (12.2 m)
V6 N= 0.1Ag fc0
V5
V4 N
Vapp Mapp
V3

hw = 144 in
(3.66 m)
V2 V2
=
V1 V1

Mb Vb Mb Vb
N N
10-Story Scaled Prototype 3-Story Specimen

Figure 3.5: Uniform latera load distributions.

tration of longitudinal steel in boundary elements with minimum code required steel in the

web of the wall, and ii) uniform distribution of longitudinal steel along the length of the

wall. The wall design is documented in detail in the project summary document [63]. The

specimens were designed using concrete compressive strength of 5000 psi (34.5 MPa) and

A706 reinforcing steel with a yield strength of 60 ksi (413 MPa). The following sections

provide an overview of the design.

3.1.4.1 Horizontal Reinforcement

The ACI 318 building code specifies a maximum spacing of horizontal reinforcement of 18

inches (45.7 cm) and a minimum reinforcement ratio of 0.25%. For the wall geometry and
p
design material properties, the nominal shear strength, Vn , was 210 kips, or 4.1 fc0 psi. The
p
shear demand, Vu = φVn (φ = 0.60 for shear), is 126 kips or 2.5 fc0 . The actual horizontal

reinforcement for the specimens is two curtains of #2 bars spaced at 6 inches, or 0.27%

horizontal reinforcement ratio.


168

3.1.4.2 Longitudinal Reinforcement

The longitudinal reinforcement was designed to meet the design shear demand, Vu = 126

kips with an ASCE 7 lateral load distribution (see Section 3.1.3). The effective height of

0.71H1 0 resulted in a moment demand of Mu = 42900 kip-in (4847 kN-m). Using a strength

reduction factor of φ = 0.90 for flexure, the required moment strength was Mn = 47700

kip-in. Two layouts of longitudinal reinforcement were designed. For both, an iterative

procedure was done to establish the layout of the reinforcement:

1. Choose trial boundary element length and reinforcement ratios.

2. Determine nominal moment strength and neutral axis depth from the results of a

moment-curvature analysis.

3. Adjust boundary element length to meet code-based requirements and adjust rein-

forcement ratios to increase or decrease flexure strength.

4. Repeat process until nominal moment strength was achieved and code requirements

were met.

For Specimens PW1, PW2, and PW4, the longitudinal reinforcement was concentrated

in the boundary elements and code-minimum longitudinal steel was provided in the web.

The boundary elements have three curtains of 7 − #4 bars spaced at 3 inches (7.62 cm) on

center (a reinforcement ratio of 3.5%). The vertical reinforcement in the web was #2 bars

spaced at 6 inches (15.24 cm) on center. The cross section details are shown in Figure 3.6.

This design is referred to as the boundary element reinforcement distribution.

In the second longitudinal reinforcement design, longitudinal reinforcement was uni-

formly distributed. Two curtains of #4 bars at 4.25 inches (10.80 cm) on center were

used. The 6 bars on either side of the wall length were confined. Figure 3.7 shows the

reinforcement layout. This design is referred to as the uniform reinforcement layout.


169

Figure 3.6: Cross-section of boundary element longitudinal steel distribution. Used for
Specimens PW1, PW2, and PW4. See Chapter 4 for cross-sections of individual walls as
cover dimensions of the constructed specimens may vary slightly. Dimensions shown are in
inches [44].
170

Figure 3.7: Cross-section of uniform longitudinal steel distribution. Used for specimen
PW3. Dimensions shown are in inches [44].
171

3.1.4.3 Confining Reinforcement

Confining reinforcement in the boundary elements was designed per the ACI 318-05 pro-

visions. Figures 3.6 and 3.7 show the confining reinforcement detailing. For the boundary

element reinforcement layout, the confining hoops and ties are spaced at 2.0 inches (5.08

cm) on center. For the uniform reinforcement distribution, the confining hoops and ties are

spaced at 1.75 inches (4.45 cm) on center. For both designs, every bar in the outer curtains

is engaged by either a hoop or cross-tie and the direction of cross-ties alternate.

3.1.5 Splice

The ACI 318-05 Building Code allows for the use of lap splices in the expected plastic hinge

region, yet little experimental data is available to understand the impact of splices on the

performance of structural walls. One of the objectives of the planar wall test program was

to investigate this impact. The length of the lap splices was determined using the provisions

for the full-scale equivalent of the actual bar sizes used. For #4 bars, a #11 bar was used for

the full-scale equivalent (an equivalent #12 bar does not exist). For #2 bars, a #6 bar was

used for the full-scale equivalent. Detailed calculations of the splice lengths are provided in

the project summary document [63]. The splice length for the #4 bars was 24 inches (61.0

cm). The splice length for the #2 bars was 9 inches (22.9 cm).

Figure 3.8 shows the the spliced bar configuration for the boundary element layout of

longitudinal reinforcement. The #4 bars that originate above the wall foundation interface

were bent at the top of the splice. This bend is a common practice (often referred to as a

“dog-leg”) that is used to facilitate construction.

3.2 Material Properties and Construction Details

This section presents a brief overview of the material properties used. More detailed infor-

mation is provided in the project summary document[63]. Section 3.2.1 presents concrete

material properties. Section 3.2.2 present reinforcing bar material properties.


172

Figure 3.8: Typical detail of splice in boundary regions. Dimensions are in inches [44].
173

3.2.1 Concrete Properties

Self-consolidating concrete was used to construct the wall specimens. Concrete compressive

strengths, fc0 , were determined from 4”x8” cylinder tests conducted on the first day of

testing. Three cylinders were tested for each specimen and the maximum compressive

strengths were averaged. Figure 3.9 shows the stess-strain response of the cylinder tests. The

average compressive strength and the corresponding strain, εc0 are provided in Table 3.2.

The compressive strengths were used to calculate the the modulus of rupture (fr ), the elastic

modulus (Ec ), and the shear modulus (G) using the formulas provided in Equations 3.1 - 3.3.

Values are provided in Table 3.2.

p
fr = 7.5 fc0 (psi) (3.1)
p
Ec = 57000 fc0 (psi) (3.2)
Ec
G= (psi); ν = 0.2 (3.3)
2(1 + ν)

Table 3.2: Average material properties for self-consolidating concrete (SCC) used for wall
specimens.

Wall fc0 (psi) εc0 (in/in) fr (psi) Ec (ksi) G (ksi)


PW1 5231 0.0023 542 4123 1718
PW2 5843 0.0025 573 4357 1815
PW3 4980 0.0028 529 4022 1676
PW4 4272 0.0021 490 3726 1552

3.2.2 Reinforcing Bar Properties

Stress-strain curves for the rebar used in each specimen were determined from standard

tension tests. For each batch of steel, between 2 and 6 bars were tested. Specimens PW1

and PW2 were constructed using the same batch of #4 bars. All specimens were constructed
174

6000 6000

5000 5000

4000 4000
Stress, psi

Stress, psi
3000 3000

2000 2000
Cylinder 1 (Avg.) Cylinder 1 (Avg.)
1000 Cylinder 2 (Avg.) 1000 Cylinder 2 (Avg.)
Cylinder 3 (Avg.) Cylinder 3 (Avg.)
Avg. fc′ and ǫc0 Avg. fc′ and ǫc0
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Strain, in/in −3 Strain, in/in −3
x 10 x 10
(a) PW1 (b) PW2

6000 6000

5000 5000

4000 4000
Stress, psi

Stress, psi

3000 3000

2000 2000
Cylinder 1 (Avg.) Cylinder 1 (Avg.)
1000 Cylinder 2 (Avg.) 1000 Cylinder 2 (Avg.)
Cylinder 3 (Avg.) Cylinder 3 (Avg.)
Avg. fc′ and ǫc0 Avg. fc′ and ǫc0
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Strain, in/in −3 Strain, in/in −3
x 10 x 10
(c) PW3 (d) PW4

Figure 3.9: Concrete cylinder test stress-strain response.

using the same batch of #2 bars.

For each bar tested, the values corresponding to yield, strain hardening, maximum stress

and ultimate strain were determined. The average values for each batch are provided in
175

Table 3.3. Figure 3.10 show the experimental stress-strain data and the key points described

below.

An elastic modulus of Es = 29, 000 ksi was assumed for all bars. The yield strength (fy )

was determined from the 0.2% offset method. The stress at which this line crosses the test

data is the yield stress, fy . The corresponding yield strain was calculated as εy = fy /Es . The

strain at which strain hardening initiated (εsh ) was reported. A stress, fsh is associated with

this point that is larger than the yield stress; this creates a slight slope in the yield plateau,

which elimates issues associated with a 0 slope when performing moment-curvature analysis

in OpenSees. The maximum stress and strain, fmax and εmax , respectively, correspond to

the maximum stress reported. The ultimate strain, εu , is recorded as the strain at a stress

of 0.9fmax .

3.3 Expected Strengths

The measured material properties of concrete and reinforcing steel were used to calculate

expected strengths of the wall specimens. Section 3.3.1 presents the expected nominal

shear strengths. Section 3.3.2 presents the expected cracking, yield, and nominal moment

strengths.

3.3.1 Shear Strength

ACI 318-08 defines the nominal shear strength of a wall as:

p
Vn = Acv (αc fc0 + ρt fy )

where Acv is the gross area of concrete section bounded by web thickness and length of

section in the direction of the shear force (the gross wall area for rectangular walls); αc is a

coefficient dependent on the wall dimensions ( αc = 2.0 for hw /lw ≥ 2.0 ); ρt (0.27%) is the

amount of horizontal reinforcement resisting the shear; fy is the yield stress of the horizontal
176

Table 3.3: Steel Material Properties (Summary)

fy εy fsh εsh fmax εmax fu εu


Wall (ksi) (in/in) (ksi) (in/in) (ksi) (in/in) (ksi) (in/in)
W1 (#4) 84.0 0.0029 84.9 0.015 100.8 0.086 91.2 0.12
W2 (#4) 84.0 0.0029 84.9 0.015 100.8 0.086 91.2 0.12
W3 (#4) 51.3 0.0018 52.0 0.0115 77.9 0.140 70.2 0.20
W4 (#4) 67.1 0.0023 68.1 0.0075 109.5 0.094 99.0 0.13
All (#2) 75.7 0.0026 77.0 0.015 84.6 0.050 76.3 0.058

120 120

100 100

80 80
Stress, ksi

Stress, ksi

60 60

40 40

20 20

0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Strain, in/in Strain, in/in
(a) #2 bars (b) #4 bars (PW1 & PW2)

120 120

100 100

80 80
Stress, ksi

Stress, ksi

60 60
Test
Yield (y)
40 40
Strain Hardening (sh)
20 20 Maximum (max)
Ultimate (u)
0 0
0 0.05 0.1 0.15 0.2 0 0.05 0.1 0.15 0.2
Strain, in/in Strain, in/in
(c) #4 bars (PW3) (d) #4 bars (PW4)

Figure 3.10: Reinforcement stress-strain response with critical values


177

reinforcement (#2 bars for all walls). Table 3.4 lists the ACI nominal shear stress using the

design material properties and the actual material properties presented in Section 3.2.

Table 3.4: Design and Expected Nominal Shear Strengths

Design Expected
p p
Specimen Vn (kips) α fc0 Vn (kips) α fc0
PW1 219 4.30 251 4.83
PW2 219 4.30 257 4.67
PW3 219 4.30 249 4.90
PW4 219 4.30 241 5.13

3.3.2 Moment Strength

This section discusses the theoretical moment strength of the cross-sections at cracking,

yield and nominal strength. Table 3.5 provides these values.

3.3.2.1 Cracking Moment

ACI defines the cracking moment in Section 9.5.2.3 (ACI 318-05) by Equation 3.4, where

Ig is the moment of the gross cross section and yt is the distance from the centroid of the

cross section to the extreme tension fiber (0.5`w for planar walls). Table 3.5 provides the

cracking moments of the wall specimens.

fr Ig
Mcr = (3.4)
yt

3.3.2.2 Yield and Nominal Moments

OpenSees [65] was used to perform moment-curvature analysis of fiber models of the cross

sections to calculate the yield and nominal moment strengths. The material properties

presented in Section 3.2 and the as-built layout of the reinforcing bars were used to construct
178

the fiber sections. Concrete was modeled using the modified Kent-Park [83] response curve

(Concrete01 uniaxial material model in OpenSees). The measured compressive strength of

the concrete was modeled with a corresponding compressive strain of −0.002 in/in. The

impact of confining reinforcement on the compressive response of the concrete was not

considered and the concrete tensile strength was neglected. The discretization of fibers was

50 (length) by 10 (thickness). The reinforcing bars were modeled using the ReinforcingSteel

uniaxial material model in OpenSees.

The yield moment is defined by the extreme tension bar reaching the yield strain. The

nominal moment strength is the moment corresponding to the extreme compression fiber

reaching a strain of −0.003 in/in. These values are listed in Table 3.5.

In specimens PW1, PW2, and PW3, the lap splice located at the base of the wall results

in double the amount of longitudinal reinforcement in the cross section and thus can be

expected to increase the strength of the wall in this region. A second moment-curvature

analysis was conducted for each Specimens PW1-PW3 to determine the strength of the

cross section assuming that both bars in the splice contributed fully to the strength of the

section. The yield and nominal moment strengths for the spliced sections are presented in

Table 3.6.

Table 3.5: Expected moment strengths

Mcr My Mn
Specimen (k-ft) (k-ft) (k-ft)
PW1 651 4761 4849
PW2 688 5455 5533
PW3 635 3710 3891
PW4 588 4073 4333
179

Table 3.6: Expected moment strengths in the spliced region

My,sp My,sp /My Mn,sp Mn,sp /Mn


Specimen k-ft (kN-m) k-ft (kN-m)
PW1 7862 (10659) 1.65 7967 (10802) 1.64
PW2 8577 (11629) 1.57 8645 (11721) 1.56
PW3 5657 (7670) 1.52 5849 (7930) 1.50
PW4 - - - -

3.4 Specimen Construction

The wall specimens were constructed by the UIUC project team and lab technicians at the

NEES MUST-SIM facility. To connect the wall specimens to the laboratory strong floor

and the LBCBs, a foundation block and wall cap (shown in Figure 3.11) were designed to

ensure appropriate dimensions and strength. The details of the foundation block and wall

cap are provided in the project summary document [63]. The sequence of construction was

as follows (additional details can be found in the project summary document [63]):

1. Assemble rebar cage for foundation block, including longitudinal reinforcement origi-

nating in the foundation.

2. Cast concrete for foundation.

3. Assemble wall reinforcement, including vertical and horizontal reinforcing bars and

confining reinforcement.

4. Cast wall concrete.

5. Assemble reinforcing cage for wall cap.

6. Cast concrete for wall cap.

When the form work was removed for Specimen PW4, a small region of concrete on the

front east bottom corner (south face, right side) was slightly damaged; at the deepest, the

damaged concrete exposed confining reinforcing steel. This region of concrete was patched

with aesthetic mortar; additional details are provided in the project summary document
180

[63]. It was estimated that the patch was approximately 1/4 inch (6.4 mm).

(a) Foundation block. (b) Wall cap.

Figure 3.11: Construction of foundation block and wall cap.

3.5 Experimental Test Setup

The four wall specimens were tested using two Load and Boundary Condition Boxes (LBCBs),

shown in Figure 3.12. Each LBCB consists of six-actuators that act as a system to impose

loads and displacements that are controlled in 6 degrees of freedom about a single load

point.

The specimens were located on a strong floor and in front of a strong wall as shown

in Figure 3.13. The LBCBs were supported on the strong wall and attached to the test

specimen cap via a steel transfer beam (see Figure 3.14) that evenly distributed the ap-

plied forces to the specimen. The LBCBs were synchronized to apply a total target shear,

overturning moment, and axial force to the wall. In addition to the loads applied to the

top of the wall, side-mounted actuators were used to apply lateral forces at the first and

second floor levels to simulate an accurate distribution of lateral forces in the 3-story test

specimen3 . Section 3.7 discusses the application of load to the walls. Figure 3.14 shows

3
The side-mounted actuators were not used for Specimen PW1 because the control software used was not
able to control them. Because the lateral floor loads at the lower floors are a very small percentage of the
181

Figure 3.12: Load and Boundary Condition Box (LBCB) at the UIUC MUST-SIM Facility
for control of load point in 6 degrees of freedom (http://nees.uiuc.edu/facilities/
LBCBs.html).

a front view of the test set-up. Detailed information on the test set-up is provided in the

project summary document [63].

3.6 Instrumentation

An instrumentation plan was developed to monitor global and local deformations. Dis-

tributed targets associated with non-contact instrumentation systems allowed measurement

of displacement and strain fields along the length and height of the wall. This section

provides an overview of the instrumentation layout. Section 3.6.1 discusses layout of the

non-contact systems. Section 3.6.2 discusses external contact instruments. Section 3.6.3

discusses internal contact instruments. The project summary document [63] provides addi-

tional information on the instrumentation and collection of data.

total base shear, the absence of these forces was considered to have a minimal impact on the specimen
behavior
182

Figure 3.13: Plan view of experimental test set-up for Specimen PW1 [44]. Dimensions are
in inches.
183

Figure 3.14: Front view of experimental test set-up [44]. Dimensions are in inches.
184

3.6.1 Non-Contact Instrumentation

3.6.1.1 Metris/Krypton

A Metris/Krypton non-contact measurement system was used to monitor the displacements

of the wall specimens. The system consists of a specialized set of three cameras that track

the movement of light emitting diodes (LEDs) attached to the surface of the specimens.

The distance of the cameras from the specimens was selected to ensure accurate readings

could be obtained at small displacements. Due to the limited measurement range of the

camera (see Figure 3.15), LEDs were provided only on the first two stories of the front

(south) face of the specimens. Figure 3.16 shows a typical layout of LED targets; targets

were provided in a grid that provided rectangular regions similar to a finite element mesh.

Additional information on the set-up, data collection and post-processing of the Krypton

measurement system can be found in the project summary document [63].

Figure 3.15: Range available for Krypton non-contact measurement system [44].
185

Figure 3.16: Krypton LED locations for specimen PW1 [44]. Units in inches.

3.6.1.2 High-resolution Cameras

Digital cameras were used to take high resolution still images of the wall specimens through-

out testing. Photogrammetric targets were provided on the surface of the walls with the

intent of tracking the movement of the specimens using the high-resolution images4 . Fig-

ure 3.17 shows the locations of the cameras with respect to the wall specimens. Six cameras

were provided to take close-up images of the of the front of the wall. Two cameras were used

to photograph the full specimens. For Specimen PW1, an additional camera was provided

to photograph the east side of the wall. In addition to the cameras in fixed locations, a

‘roaming’ camera was used to take pictures of specific areas of interest throughout testing.

3.6.2 External Instrumentation

Relative displacements within segments of the specimens were measured using a grid of

linear pots on the back side (north face) of the wall. The measured vertical, horizontal,

4
Photogrammetric data was unable to be used due to a lack of sufficient overlap between the cameras
186

Figure 3.17: Location of fixed cameras used to capture high-resolution images [44].

and diagonal extensions were used to monitor and shear and rotation for each segment.

Figure 3.18 shows the layout of the grid. For each floor, two large segments were used to

monitor deformations in the web. In the first floor of the wall, two segments (half the floor

height) monitored deformations in each boundary element. The linear pots were attached

to threaded rods embedded in the wall.

Figure 3.19 shows locations of distributed string pots used to monitor the horizontal

movement of the specimen. The string pots were mounted to a steel reference column

located to the west (left) of the wall specimens. The string pots were attached to embedded

rods located two inches from the west edge of the wall (see Figure 3.13). The total lateral

displacement of each specimen was measured using a string pot connected to the bottom

center of the wall cap. This movement was used as part of the control algorithm. A string

pot measured the lateral displacement of the foundation block.

Local wall and specimen vertical displacements were monitored at the base of the wall
187

Figure 3.18: Layout of instruments that measured relative displacements on the north (back)
face of the wall [44]. Units are in inches.

Figure 3.19: Locations of in-plane displacement measurements [44]. Units are in inches.
188

and at the base of the foundation block (instruments 1-4 in Figure 3.13). The total vertical

and diagonal displacements of the wall cap relative to the foundation block were measured

by string pots labeled beginning with ‘D’, and ‘S’ in Figure 3.19. Absolute out-of-plane dis-

placements were measured by instruments mounted on the strong wall behind the specimen;

see Figure 3.20 for locations.

Figure 3.20: Locations of absolute out-of-plane displacement measurements [44]. Units are
in inches.

Concrete surface gauges were provided on the north (back) and south (front) faces of the

wall at the locations shown in Figure 3.21. At the top and bottom of the wall, the gauges

were oriented vertically; in the middle of the wall, pairs of gauges were oriented at 45◦ and

135◦ from the vertical.

3.6.3 Steel Strain Gauges

Internal instrumentation of the wall specimens consisted of steel strain gauges on the longi-

tudinal and horizontal reinforcement. Figure 3.22 shows the location of strain gauges on the

horizontal reinforcement in the four walls. Gauges were provided in the unconfined region
189

Figure 3.21: Location of concrete surface gauges. Gauges are located on the front and back
faces of the wall [44]. Dimensions are in inches.

of the wall at four locations on three bars in the first floor and at three locations on two bars

in the second floor. Additional gauges were located at the ends of the bars in the boundary

elements.

Figure 3.23 shows the locations of strain gauges on the longitudinal wall bars5 . and

on the longitudinal foundation bars6 Gauges were provided in all three floors and in the

foundation on the outer and inner most bar in each boundary element and the outer most

web bars. Additional gauges were provided on boundary element and web bars in the first

floor. For the spliced specimens, gauges were provided on the foundation and wall bars at

the same location for two elevations7 .


5
For walls with splices at the base of the wall (PW1, PW2, and PW3), the term ‘wall bars’ refers to
the bars that begin just above the foundation and extend to the wall cap. For the wall specimen with
continuous reinforcement (PW4), the term ‘wall bars’ refers to the longitudinal reinforcement
6
For walls with splices at the base of the wall (PW1, PW2, and PW3), the term ‘foundation bars’ refers
to the bars that begin in the foundation and terminate in the lower portion of the wall. For the wall
specimen with continuous reinforcement (PW4), foundation bars are not provided.
7
In spliced specimens, the strain gauges located at 2 inches (5.08 cm) from the wall foundation interface
were located near the end of the wall bars.
190

Figure 3.22: Locations of strain gauges on horizontal reinforcement for PW1 [44]. Units
are in inches. Horizontal strain gauge locations for specimens PW2, PW3, and PW4 are
similar.

3.7 Loading

The loading scheme was developed to control i) the horizontal displacement to meet a target

history, ii) the axial load to be constant, and iii) the relationship between the moment and

shear force to be constant. Equipment and specially developed software has been developed

to allow for mixed-mode control (force and displacement) of experiments. Details of the

control for the wall specimen tests are not discussed here but details can be found in the

project summary document [63] and in the dissertations of Hart [44] and Marley [64]. An

overview of the applied loads and displacements is provided here. Section 3.7.1 discusses

the displacement histories of the tests. Section 3.7.2 discusses the applied axial loads.

Section 3.7.3 discusses the control of the applied forces and the relationships between them.
191

(a) PW1 (b) PW2

(c) PW3 (d) PW4

Figure 3.23: Locations of strain gauges on vertical reinforcement [44]. Units are in inches.
192

3.7.1 Displacement History

The primary test control was a prescribed lateral displacement history at the top of the

wall (3rd story; instrument xc in Figure 3.19). The top (3rd story) loads were applied to

achieve this displacement and the target ratios of moment to shear and story shear to top

shear (see Section 3.7.3).

A monotonically increasing cyclic displacement history was used. Table 3.7 provides the

peak drifts for each cycle. Figures 3.24 shows the applied history, including the relationship

between the applied displacement and step numbers.. Drifts at the top of the third floor

were calculated as the displacement divided by the three story height of the specimens, hw ,

and reported as a percentage. At low drift levels, three cycles were applied. At drifts of 1.0%

or larger, two cycles were applied. The first displacement level corresponded to the cracking

moment of the cross-section. The second displacment level was approximately twice the

cracking moment. The next three drift levels corresponded to approximately 50%, 75%,

and 125% of the theoretical yield displacement, where the theoretical yield was determined

from a numerical model of the walls[64]. The remaining drift levels were 0.50%, 0.75%,

1.0% and 1.5%. For Specimen PW1, the displacement history was different than that of

the other walls due to an error in the calibration factor for the control string pot. Table 3.8

provides the displacement history used for PW1.

Within each cycle, the history between peaks was applied in 5 to 20 small displacement

steps. These steps were necessary for successful implementation of the control algorithm.

At each of these steps, data was recorded from the instrumentation on the walls, the fixed

location cameras were triggered to take still images of the specimen, and notes on damage

progression and crack width measurements were made.

In this document, step numbers are only referred to when necessary, such as in docu-

menting the damage progression for the tests. Frequently, only the peaks of the cycles are

discussed. The peaks are specified by the magnitude of the cycle as a drift (e.g. 0.50%)

and the direction of the loading (ER+ indicates loading in the positive direction to the east
193

SimCor Step

10 70 130 230 350 470 590 670


1.5 2.16

1 1.44

0.5 0.72

Displacement, [in]
Drift, [%]

0 0

−0.5 −0.72

−1 −1.44

−1.5 −2.16

%
%

%
%

50
50

75

00
0

7
05

10

25

34

1.
0.

0.

1.
Drift Cycles
0.

0.

0.

0.

(a) PW1

SimCor Step

10 70 130 250 370 610 850 1090 1250


1.5 2.16

1 1.44

0.5 0.72
Displacement, [in]
Drift, [%]

0 0

−0.5 −0.72

−1 −1.44

−1.5 −2.16
%
%

%
0. 4%
0. 8%
9%

8%

7%

50
50

75

00
01
02
13

20

34

1.
0.

0.

1.

Drift Cycles
0.

0.

0.

(b) PW2 (last step = 1305), PW3 (last step = 1287) and PW4 (last step = 1250)

Figure 3.24: Applied displacement histories


194

Table 3.7: Prescribed Displacement History

Displacement Level, in (mm) Drift Level Approx. Disp. Level Cycle #’s Steps per Cycle
0.02 (0.51) 0.014% Cracking 1-3 10
0.04 (1.02) 0.028% 4-6 10
0.2 (5.08) 0.139% 50% Yield 7-9 20
0.3 (7.62) 0.208% 75% Yield 10-12 20
0.5 (12.7) 0.347% 125% Yield 13-15 40
0.72 (18.29) 0.50% 16-18 40
1.08 (27.43) 0.75% 19-21 40
1.44 (36.58) 1.0% 22-23 40
2.16 (54.86) 1.5% 24-25 40

Table 3.8: PW1 Actual Displacement History

Displacement Level, in (mm) Drift Level Approx. Disp. Level Cycle #’s Steps per Cycle
0.072 (1.83) 0.05% 1-3 10
0.145 (3.68) 0.101% 4-6 10
0.362 (9.19) 0.251% 7-9 20
0.5 (12.7) 0.347% 125% Yield 10-12 20
0.72 (18.29) 0.50% 13-15 20
1.08 (27.43) 0.75% 16-18 20
1.44 (36.58) 1.0% 19-20 20
2.16 (54.86) 1.5% 21-22 20

(right); WL- indicates loading in the negative direction to the west (left)). Positive and

negative forces are associated with positive and negative peak displacements, respectively.

3.7.2 Axial Load

The target axial load for the walls was 360 kips (1601 kN). The axial loads were applied

by the LBCBs and held at an approximately constant force throughout the tests. Table 3.9

specifies the target and applied axial forces (N ) and axial load ratios (λn ). Actual values

are the mean value of all load steps. For Specimen PW2, a load that was larger that the

target value was applied. This, and the variation in concrete strengths resulted in axial load

ratios from 0.095 to 0.13.


195

3.7.3 Lateral and Overturning Loads

As discussed in Section 3.1, the wall specimens were one-third scale of a ten-story prototype

building, with only the bottom three stories constructed. The LBCBs (see Section 3.5)

applied shear and moment to the top of the three story specimens to simulate the demand

originating at the upper stories and met the target drift history. In Specimens PW2-PW4,

ancillary actuators were used to apply the story shear forces. Figure 3.25 shows the notation

used to describe the applied loads and base reactions.

N
Vapp Mapp

V2
hw
h2

V1
h1

Mb Vb
N
Figure 3.25: Notation for loads applied to specimens.

The displacement at the top center of the wall specimen was controlled to meet the

displacement history described in Section 3.7.1. The control software determined the lateral

shear force necessary to achieve the necessary displacement and ensured that the applied

moment was a constant relationship to this force. The relationship between the moment

and shear at the top of the wall was determined based on the specimen geometry and

the assumed story force distribution for the upper 7 stories. For Specimens PW2, PW3,

and PW4, side-mounted actuators provided additional load at the floor levels; these forces

were a function of the top shear force to simulate a uniform load distribution. Table 3.9

summarizes the intended and actual loads applied to the specimens. The values for the
196

actual loads are the mean values from all loads steps.

Table 3.9: Summary of intended and actual loads applied to wall specimens

Value PW1 PW2 PW3 PW4


Load Distribution ASCE 7 Uniform Uniform Uniform
αef f 0.71 0.50 0.50 0.50
Mapp /Vapp , ft (m) 16.4 (5.00) 11.5 (3.51) 11.5 (3.51) 11.5 (3.51)
Target

V1 /Vapp , % - 12.5 12.5 12.5


V2 /Vapp , % - 12.5 12.5 12.5
N , kips (kN) 360 (1601) 360 (1601) 360 (1601) 360 (1601)
λN 0.10 0.10 0.10 0.10
αef f 0.71 0.54 0.51 0.52
Mapp /Vapp , ft (m) 16.5 (0.42) 11.9 (0.30) 11.8 (0.30) 11.2 (0.28)
V1 /Vapp , % - 8.7 12.4 12.2
Actual

V2 /Vapp , % - 8.6 12.5 12.2


N , kips (kN) 359 (1597) 546 (2429) 360 (1601) 360 (1601)
λN 0.095 0.130 0.100 0.117

3.8 Summary

This chapter provides a brief overview of the planar wall test program that is a component

of the NEESR project “Design, Behavior and Analysis of Complex Wall Systems”. Four

specimens make up the planar wall test program and were designed to investigate the impact

of i) lateral load distribution and the resulting shear demand, ii) distribution of longitudinal

reinforcement, and iii) the use of lap splices at the base of the walls. The walls designs

were compared to geometry and strength characteristics of walls on the West Coast of the

United States to verify that the test program is representative of modern construction. The

specimen designs can be summarized as follows:

• PW1 - Varies the lateral load distribution from that of the reference specimen (PW2).

Had a boundary element distribution of longitudinal reinforcement, an ASCE-7 lateral

load distribution, and a lap splice at the base of the wall.


197

• PW2 - The reference specimen; every other specimen had one characteristic changed

from this wall. Had a boundary element distribution of longitudinal reinforcement, a

uniform lateral load distribution, and a lap splice at the base of the wall.

• PW3 - Varies the longitudinal reinforcement distribution from that of the reference

specimen (PW2). Had a uniform distribution of longitudinal reinforcement, a uniform

lateral load distribution, and a splice at the base of the wall.

• PW4 - Varies the use of a lap splice at the base of the wall from that of the reference

specimen (PW2). Had a boundary element distribution of longitudinal reinforcement,

a uniform lateral load distribution, and continuous longitudinal reinforcement (no

splice).

The walls were tested at the University of Illinois, Urbana-Champaign NEES MUST-

SIM Facility to allow testing of large scale specimens (1/3 scale of a prototype based on the

West Coast building inventory) while representing the base reactions of a taller wall. The

specimens were constructed as the bottom three stories of the ten-story prototype and the

reactions for the desired lateral load distributions on the 10-story structure were achieved

using Load and Boundary Conditions Boxes (LBCBs) available at the MUST-SIM Facility.

High resolution-data was collected for each test, including traditional instrumentation such

as linear potentiometers, LVDTs, and steel strain gauges, as well as advanced systems

such as Krypton/Metris non-contact measurement system and high resolution images. The

walls were loaded with approximately 0.10Ag fc0 axial load and a reverse cyclic, displacement

history. The applied forces were determined based on the desired lateral load distribution

and the lateral force necessary to achieve the desired displacement.

A component of all NEESR funded projects is the archival of all experimental data

collected. This data is provided in the NEES Project Warehouse (http://nees.org). A

summary of the available data can be found in the project summary document [63]. Chap-

ter 3 provides an overview of the experimental results of the planar wall test program.
198

Chapter 5 - 6 discuss analysis of the experimental data collected by the instrumentation

outlined in this chapter.


199

Chapter 4

SUMMARY OF EXPERIMENTAL RESULTS

This chapter summarizes the observed results of four planar wall tests that were con-

ducted at the University of Illinois, Urbana-Champaign (UIUC) NEES (The George E.

Brown, Jr. Network for Earthquake Engineering Simulation) MUST-SIM (Multi-Axial Full-

Scale Sub-Structured Testing and Simulation) Facility (http://nees.uiuc.edu). The walls

simulated modern construction and were subjected to realistic loading and boundary con-

ditions, as afforded by the advanced equipment and control available at the facility. During

testing, observations of key damage states were noted, in part to support the development

of performance based tools for seismic design.

This chapter presents the observed damage for each specimen. Specifically, the following

is provided for each specimen: i) a summary of the specimen design and construction, ii)

applied axial load and the base moment to base shear ratio (lateral load distribution), iii)

summary of general response of the wall, including force-drift hysteresis plots, drift capacity

and failure mode, and iv) summary of damage progression. In summarizing the damage, the

aim is to a) show the crack pattern, b) identify when key damage states occurred, c) identify

the key locations of damage and provide images to illustrate the extent of the damage, d)

describe the failure ans show the final damage state of the wall, and e) identify when

and where initial yield occurred. Each wall specimen is presented in an individual section

(Sections 4.1 - 4.4). Appendix A provides detailed descriptions of damage progression,

including an expanded set of images.


200

4.1 Specimen PW1

Specimen PW1 had heavily reinforced boundary elements and a lap splice at the base of

the wall. Figure 4.1 and 4.2 show the cross section and elevation drawings of the as-built

specimen.

The notation used to describe the applied loads and reactions are shown in Figure 4.3.

An average axial load of N = 359 kips (1597 kN) or λN = 0.095, was applied to the

wall. A lateral force (Vapp ) and overturning moment (Mapp ) were applied to the top of the

specimen. Throughout the test, the ratio of the lateral force to the overturning moment was

held constant such that the base reactions (base shear Vb and base moment Mb ; measured at

the wall-foundation interface) were equivalent to those of a 10-story wall with a lateral load

distribution calculated from the Equivalent Lateral Force procedure in ASCE 7-05 [6]. The

applied load ratio (Mapp /Vapp ) was 16.5 feet (0.42 m) and the ratio of the base reactions

(Mb /Vb ) was 28.8 feet (0.73 m)1 . The effective height of the load was 0.71h10 , or 2.37hw .

Table 4.1 summarizes the applied displacement history.

The following sections summarize general response and damage progression of the test.

Section 4.1.1 provides a general overview of the test, including load-drift response, failure

mode, and comparison of the maximum strength to the expected strength. Section 4.1.2

provides an overview of the wall damage.

4.1.1 General Response

The load-displacement hystereses for Specimen PW1 are shown in Figure 4.4. Figure 4.4a

shows the base shear versus drift. Figure 4.4b shows the base moment versus drift. The

wall lost lateral load carrying capacity during the second cycle to 1.5% drift due to fracture

of longitudinal reinforcement at the wall-foundation interface in the east (right) boundary

element.
1
Ancillary actuators were not provided at the first and second floor levels as were in the other three
specimens.
201

Figure 4.1: Specimen PW1 cross-section. Drawing provided by Chris Hart [44].

Figure 4.2: Specimen PW1 elevation. Drawing provided by Chris Hart [44].
202

N
Vapp Mapp

V2

hw
h2
V1

h1
Mb Vb
N
Figure 4.3: Notation for loads applied to specimens.

Table 4.1: Specimen PW1 applied displacement history.

Displacement, in (mm) Drift Cycles SimCor Steps


0.072 (1.83) 0.05% 1-3 10-70
0.145 (3.68) 0.101% 4-6 70-130
0.362 (9.19) 0.251% 7-9 130-230
0.5 (12.70) 0.347% 10-12 230-350
0.72 (18.29) 0.50% 13-15 350-470
1.08 (27.43) 0.75% 16-18 470-590
1.44 (36.6) 1.0% 19-20 590-670
2.16 (54.9) 1.5% 21-22 670-750

At the base of the wall, the maximum shear demand (Vb ) was approximately 0.73Vn ,

where Vn is the expected shear strength of the cross section (see Section 3.3.1). The maxi-

mum moment demand (Mb ) was approximately 1.08Mn , where Mn is the maximum expected

moment strength of the cross section (see Section 3.3.2). The expected nominal moment

moment strength of the wall was first reached at approximately +0.53% and −0.62% drift;

for the purpose of data analysis, the nominal strength is considered to occur at Step 480 in

the ER+ direction and Step 500 in the WL- direction.

At the center of the splice (elevation of 12 inches (30.5 cm)), the maximum moment de-

mand (Msp,mid ) was approximately 0.62Mn,sp , where Mn,sp is the expected moment strength
203

300 6000 8000


1200
6000
200 900 4000

600 4000
100 2000
300 2000

M b , kN-m
M b , k-ft
Vb , kip

Vb , kN
0 0 0 0

−300 −2000
−100 −2000
−600 −4000
−200 Test −900 −4000 Test
−6000
Vn Mn
−1200
−300 −6000 −8000
−1.5 −1 −0.5 0 0.5 1 1.5 −1.5 −1 −0.5 0 0.5 1 1.5
% Drift % Drift
(a) Base shear (b) Base moment

Figure 4.4: Specimen PW1 load-drift response.

of the spliced cross section (see Section 3.3.2). At the top of the splice (elevation of 24 inches

(61.0 cm)), the maximum moment demand (Msp,top ) was approximately 1.0Mn . The nomi-

nal moment strength was reached only in the ER+ direction at approximately 1.28% drift.

4.1.2 Damage

A summary of the primary damage states and their progression is presented here. Ap-

pendix A.1 provides a detailed description of the damage progression for Specimen PW1,

including a large set of images depicting the damage. A time-lapse movie of damage in

the first floor (lower one-third of the specimen) can be viewed at http://www.youtube.

com/user/NEESRWallProject#p/u/1/KC7b-k3ZRjs. The key damage states and the corre-

sponding drift at which it occurred follows.

1. Horizontal cracking initiated at 0.06% drift.

2. Diagonal cracking initiated at 0.1% drift.

3. Vertical cracks initiated at 0.34% drift.


204

4. Tensile yield of extreme vertical reinforcing bars was indicated by strain gauges at

0.35% drift.

5. Compressive yield of extreme vertical reinforcing bars was indicated by strain gauges

at 0.64% drift.

6. Cover spalling initiated at the toe of the wall at 0.56% drift.

7. Cover spalling initiated above the splice during that second cycle to 0.75% drift.

8. Concrete damage extensive enough to expose the longitudinal reinforcement occurred

splice during the third cycle to 0.75% drift.

9. Bar buckling was observed above the splice during the second cycle to 1.0% drift.

10. Damage (crushing) of the confined core of the boundary element was observed above

the splice during cycles to 1.5% drift.

11. Bar fracture occurred during the second cycle to 1.5% drift.

Although cracking initiated in the 0.10% drift cycles, a stable crack pattern, shown in

Figure 4.5, was not developed until the 0.25% drift cycles. Damage other than horizontal and

diagonal cracking was spread from the wall-foundation interface to a height of approximately

36 inches (91.4 cm), or 12 inches (30.5 cm) above the top of the splice, but was restricted

to the edges of the wall and the wall-foundation interface.

Figure 4.6b shows, at 1.0% drift, a vertical splitting crack along the east (right) edge

of the wall and running the full height of the splice. The extent of the damage along the

edges of the wall was i) greater above the top of the splice than in the splice region and ii)

greater in the west (left) boundary element than in the east (right) boundary element.

Figure 4.7a shows damage to the west (left) edge of the wall at 1.0% drift. Above the top

of the splice, bars were buckled, however, within the spliced region, the cover was spalled

but the longitudinal reinforcement was not exposed. Figure 4.7b shows the damage to the

east (right) edge of the wall at 1.5% drift; the longitudinal reinforcement was exposed in

the spliced region and one bar was buckled at the top of the splice, but to a lesser extent

than on the west (right) side of the wall.


205

Figure 4.5: Specimen PW1: Crack pattern in lower two stories following three cycles to
0.25% drift [44].

Leading up to the final cycle, the major observed damage was concentrated above the

west boundary element splice, however, loss of lateral load capacity ultimately resulted from

bar fracture at the wall-foundation interface in the east boundary element in the second half

(WL- peak) of the second cycle to 1.5% drift, at −1.38% drift. The bars fractured along the

crack at the wall-foundation interface, which had been observed previously, but the width

of which increased significantly at this step. At the final WL- peak at 1.5% drift, all of

the bars on the front of the east boundary element, expect the extreme bar, had fractured.

Figure 4.8 shows the fractured bars on the front east side of the wall at the final peak.

Figure 4.9 shows a photograph and the final crack pattern of the wall at the end of the test.
206

(a) Spalling at top of west (left) (b) Vertical crack at corner of east
splice. (right) splice

Figure 4.6: Specime PW1: Damage during first ER+ half-cycle at 1.0% drift (Step 600).

(a) Bar buckling at final WL- peak (b) Bar buckling on way to first
at 1.0% drift (Step 620). ER+ peak at 1.5% drift (Step 680)

Figure 4.7: Specimen PW1 damage along wall edges. Damage was more severe above the
top of the splice and in the west boundary element.
207

(a) Close-up

(b) Front east (right) boundary element.

Figure 4.8: Specimen PW1: Longitudinal bar fracture on front face of east boundary element
at final peak (−1.52% drift, Step 740)
208

(a) (b)

Figure 4.9: Specimen PW1: Final damage state and crack pattern.

4.2 Specimen PW2

Specimen PW2 had heavily reinforced boundary elements and a lap splice at the base of

the wall. Figure 4.10 and 4.11 show the cross section and elevation drawings of the as-built

specimen.

An average axial load of N = 546 kips (2429 kN) or λN = 0.130, was applied to the

wall. A lateral force (Vapp ) and overturning moment (Mapp ) were applied to the top of the

specimen. Throughout the test, the ratio of the lateral force to the overturning moment was

held constant such that the base reactions (base shear Vb and base moment Mb ; measured

at the wall-foundation interface) were equivalent to those of a 10-story wall with a uniform

lateral load distribution. The applied load ratio (Mapp /Vapp ) was 11.9 feet (0.30 m), ancillary

actuators provided first and second floor lateral loads of 8.7% and 8.6% of Vapp , respectively,

and the ratio of the base reactions (Mb /Vb ) was 21.4 feet (0.54 m). The effective height of
209

Figure 4.10: Specimen PW2 cross-section. Drawing provided by Chris Hart [44].

Figure 4.11: Specimen PW2 elevation. Drawing provided by Chris Hart [44].
210

Table 4.2: Specimen PW2 applied displacement history.

Displacement, in (mm) Drift Cycles SimCor Steps


0.02 (0.5) 0.014% 1-3 10-70
0.04 (1.0) 0.028% 4-6 70-130
0.2 (5.1) 0.139% 7-9 130-250
0.3 (7.6) 0.208% 10-12 250-370
0.5 (12.7) 0.347% 13-15 370-610
0.72 (18.3) 0.50% 16-18 610-850
1.08 (27.4) 0.75% 19-21 850-1090
1.44 (36.6) 1.0% 22-23 1090-1250
2.16 (54.9) 1.5% 24-25 1250-1305

the load was 0.54h10 , or 1.80hw . Table 4.2 summarizes the applied displacement history.

The following sections summarize general response and damage progression of the test.

Section 4.2.1 provides a general overview of the test, including load-drift response, failure

mode, and comparison of the maximum strength to the expected strength. Section 4.2.2

provides an overview of the wall damage.

4.2.1 General Response

The load-displacement hystereses for Specimen PW2 are shown in Figure 4.12. Figure 4.12a

shows the base shear versus drift. Figure 4.12b shows the base moment versus drift. The

wall lost lateral load carrying capacity during the first cycle to 1.5% drift in the WL-

direction due to extensive damage to the concrete and severe bar buckling of the steel above

the splice in the west (left) boundary element.

At the base of the wall, the maximum shear demand (Vb ) was approximately 1.06Vn ,

where Vn is the expected shear strength of the cross section (see Section 3.3.1). The maxi-

mum moment demand (Mb ) was approximately 1.05Mn , where Mn is the maximum expected

moment strength of the cross section (see Section 3.3.2). The expected nominal moment

moment strength of the wall was first reached at approximately +0.72% and −0.64% drift;

for the purpose of data analysis, the nominal strength is considered to occur at Step 870 in

the ER+ direction and Step 910 in the WL- direction.


211

300 6000 8000


1200
6000
200 900 4000

600 4000
100 2000
300 2000

M b , kN-m
M b , k-ft
Vb , kip

Vb , kN
0 0 0 0

−300 −2000
−100 −2000
−600 −4000
−200 Test −900 −4000 Test
−6000
Vn Mn
−1200
−300 −6000 −8000
−1.5 −1 −0.5 0 0.5 1 1.5 −1.5 −1 −0.5 0 0.5 1 1.5
% Drift % Drift
(a) Base shear (b) Base moment

Figure 4.12: Specimen PW2 load-drift response.

At the center of the splice (elevation of 12 inches (30.5 cm)), the maximum moment de-

mand (Msp,mid ) was approximately 0.63Mn,sp , where Mn,sp is the expected moment strength

of the spliced cross section (see Section 3.3.2). At the top of the splice (elevation of 24 inches

(61.0 cm)), the maximum moment demand (Msp,top ) was approximately 0.95Mn .

4.2.2 Damage

A summary of the primary damage states and their progression is presented here. Ap-

pendix A.2 provides a detailed description of the damage progression in Specimen PW2,

including a large set of images depicting the damage. A time-lapse movie of damage in

the first floor (lower one-third of the specimen) can be viewed at http://www.youtube.

com/user/NEESRWallProject#p/u/2/Miepayt10Vk. The key damage states and the corre-

sponding drift at which it occurred follows.

1. Horizontal cracking initiated at 0.10% drift.

2. Diagonal cracking initiated at 0.10% drift.


212

3. Vertical cracks initiated at 0.35% drift.

4. Compressive yield of extreme vertical reinforcing bars was indicated by strain gauges

at 0.20% drift.

5. Tensile yield of extreme vertical reinforcing bars was indicated by strain gauges at

0.42% drift.

6. Cover spalling initiated above the splice during the first cycle to 0.75%.

7. Concrete damage extensive enough to expose the longitudinal reinforcement occurred

above the splice during the third cycle to 0.75% drift.

8. Bar buckling was observed above the splice during the third cycle to 0.75% drift.

9. Damage (crushing) of the confined core of the boundary element was observed above

the splice during the first half cycle (ER+) to 1.5% drift.

10. Failure occurred due to extensive bar buckling and core crushing in the west (left)

boundary element above the splice at 1.05% drift in the during the first WL- cycle to

1.5% drift.

Figure 4.13 shows the crack pattern in Specimen PW2 following three cycles to 0.21%

drift. Damage other than horizontal and diagonal cracking primarily occurred at the top of

the splice (24 − 36 inches (61.0 − 91.4 cm)), although some minor spalling of the concrete

was observed at the toes of the wall. Damage above the splice was greater in the west (left)

boundary element than in the east (right) boundary element; the damage to both boundary

elements following cycles to −0.75% and 1.0% drift are shown in Figure 4.14.

Specimen PW2 successfully sustained one half cycle to 1.5% drift in the ER+ direction.

When the specimen was loaded in the WL- direction, bar buckling and core crushing pro-

gressively increased above the west (left) boundary element splice. Damage at this location

is shown in Figure 4.15a at −1.06% drift (one step beyond the previous maximum drift in

the WL- loading direction). While loading the wall to the next step (Step 1305, 1.10% drift,

the axial load dropped from 560 to 220 kips (2491 to 979 kN), all of the longitudinal bars in
213

Figure 4.13: Specimen PW2: Crack pattern in lower two stories following three cycles to
0.21% drift [44].

(a) Bar buckling in west boundary element (b) Bar buckling above east boundary ele-
following third WL- peak at −0.75% drift ment splice at third ER+ peak at 1.0% drift
(Step 1070) (Step 1190).

Figure 4.14: Specimen PW2: Comparison of damage above splice in west (left) and east
(right) boundary elements.
214

(a) −1.06% drift (Step 1304); one step prior to loss of load carrying capacity.

(b) −1.10% drift (Step 1305) following loss of lateral load carrying capacity.

Figure 4.15: Specimen PW2: Damage to in the first floor before and after loss of lateral
load carrying capacity.
215

(a) (b)

Figure 4.16: Specimen PW2: Final damage state and crack pattern.

the west (left) boundary element buckled2 , the region of crushed concrete extended into the

web, and longitudinal and horizontal bars in the web buckled. The axial load was reduced

to complete the step. Figure 4.15b shows the damage to the wall after loss of lateral load

carrying capacity. Figure 4.16 shows a photograph and the final crack pattern of the wall

at the end of the test.

4.3 Specimen PW3

Specimen PW3 had uniformly distributed reinforcement and a lap splice at the base of the

wall. Figure 4.17 and 4.18 show the cross section and elevation drawings of the as-built

specimen.

An average axial load of N = 360 kips (1601 kN) or λN = 0.10, was applied to the

2
Previously only the outer bars were buckled.
216

Figure 4.17: Specimen PW3 cross-section. Drawing provided by Chris Hart [44].

Figure 4.18: Specimen PW3 elevation. Drawing provided by Chris Hart [44].
217

wall. A lateral force (Vapp ) and overturning moment (Mapp ) were applied to the top of the

specimen. Throughout the test, the ratio of the lateral force to the overturning moment was

held constant such that the base reactions (base shear Vb and base moment Mb ; measured

at the wall-foundation interface) were equivalent to those of a 10-story wall with a uniform

lateral load distribution. The applied load ratio (Mapp /Vapp ) was 11.8 feet (0.30 m), ancillary

actuators provide first and second floor lateral loads of 12.4% and 12.5% of Vapp , respectively,

and the ratio of the base reactions was 20.2 feet (0.51 m). The effective height of the load

was 0.51h10 , or 1.70hw . Table 4.3 summarizes the applied displacement history.

The following sections summarize general response and damage progression of the test.

Section 4.3.1 provides a general overview of the test, including load-drift response, failure

mode, and comparison of the maximum strength to the expected strength. Section 4.3.2

provides an overview of the wall damage.

Table 4.3: Specimen PW3 applied displacement history.

Displacement, in (mm) Drift Cycles SimCor Steps


0.02 (0.5) 0.014% 1-3 10-70
0.04 (1.0) 0.028% 4-6 70-130
0.2 (5.1) 0.139% 7-9 130-250
0.3 (7.6) 0.208% 10-12 250-370
0.5 (12.7) 0.347% 13-15 370-610
0.72 (18.3) 0.50% 16-18 610-850
1.08 (27.4) 0.75% 19-21 850-1090
1.44 (36.6) 1.0% 22-23 1090-1250
2.16 (54.9) 1.5% 24-25 1250-1287

4.3.1 General Response

The load-displacement hystereses for Specimen PW3 are shown in Figure 4.19. Figure 4.19a

shows the base shear versus drift. Figure 4.19b shows the base moment versus drift. The

wall lost lateral load carrying capacity in the first cycle to 1.5% drift (at approximately

1.25% drift) due to extensive damage to the concrete and severe buckling of the steel in the
218

east (right) boundary element above the top of the splice.

300 6000 8000


1200
6000
200 900 4000

600 4000
100 2000
300 2000

M b , kN-m
M b , k-ft
Vb , kip

Vb , kN
0 0 0 0

−300 −2000
−100 −2000
−600 −4000
−200 Test −900 −4000 Test
−6000
Vn Mn
−1200
−300 −6000 −8000
−1.5 −1 −0.5 0 0.5 1 1.5 −1.5 −1 −0.5 0 0.5 1 1.5
% Drift % Drift
(a) Base shear (b) Base moment

Figure 4.19: Specimen PW3 load-drift response.

At the base of the wall, the maximum shear demand (Vb ) was approximately 0.90Vn ,

where Vn is the expected shear strength of the cross section (see Section 3.3.1). The max-

imum base moment demand (Mb ) was approximately 1.16Mn , where Mn is the maximum

expected moment strength of the cross section (see Section 3.3.2). The expected nominal

moment strength at the base of the wall was first reached at approximately ±0.43 drift; for

the purpose of data analysis, the nominal strength was determined to occur at Step 630 in

the ER+ direction and Step 670 in the WL- direction.

At the center of the splice (elevation of 12 inches (30.5 cm)), the maximum moment de-

mand Msp,mid was approximately 0.73Mn,sp , where Mn,sp is the expected moment strength

of the spliced cross section (see Section 3.3.2). At the top of the splice (elevation of 24

inches (61.0 cm)), the maximum moment demand (Msp,top ) was approximately 1.04Mn .

The expected nominal moment strength at the top of the splice was first reached at approx-

imately 0.63% and −0.68% drift; for the purpose of data analysis, the nominal strength was
219

determined to occur at Step 870 in the ER+ direction and Step 910 in the WL- direction.

4.3.2 Damage

A summary of the primary damage states and their progression is presented here. Ap-

pendix A.3 provides a detailed description of the damage progression in Specimen PW3,

including a large set of images depicting the damage. A time-lapse movie of damage in

the first floor (lower one-third of the specimen) can be viewed at http://www.youtube.

com/user/NEESRWallProject#p/u/3/WVtKmeC0FPU. The key damage states and the corre-

sponding drift at which it occurred follows.

1. Horizontal cracking initiated at 0.06% drift.

2. Diagonal cracking initiated at 0.06% drift.

3. Vertical cracking initiated at 0.21% drift.

4. Tensile yield of extreme vertical reinforcing bars was indicated by strain gauges at

0.17% drift.

5. Compressive yield of extreme vertical reinforcing bars was indicated by strain gauges

at 0.32% drift.

6. Cover spalling initiated above the splice at 0.52% drift.

7. Web crushing was observed at 0.75% drift.

8. Bar buckling was observed above the splice during the first cycle to 1.0% drift.

9. Damage (crushing) of the confined core of the boundary element was observed above

the splice during the second (and final) cycle to 1.0% drift.

10. Loss of lateral load capacity occurred due to extensive bar buckling and core crushing

in the east (right) boundary element above the splice at 1.28% drift during the first

ER+ half cycle to 1.5% drift.

Figure 4.20 shows the crack pattern in Specimen PW3 following three cycles to 0.21%

drift. Damage other than horizontal and vertical cracking was concentrated above the top
220

Figure 4.20: Specimen PW3: Crack pattern in lower two stories following three cycles to
0.21% drift [44].

of the splice in the boundary elements and in the web3 . The initial spalling occurred along

the thickness of the wall above the top of the splice in the boundary elements. The most

apparent concrete damage was in the web of the wall. The cover concrete cover in the web

spalled but did not initially fall off the wall, forming a bulged region in the web at the top of

the splice; the bulged region was identified by tapping the wall and listening for a change in

the sound (denoting solid to hollow), and was marked on the wall with a dashed blue line.

Figure 4.21a shows the wall after the second ER+ peak at 1.0% drift; Figure 4.21b and 4.21c

show the damage in the web and boundary element.

Specimen PW3 failed during the first half cycle (ER+ direction) to 1.5%. At 1.28%

drift, the axial load dropped from 360 kips (1601 kN) to 190 kips (845 kN). This drop in

load was accompanied by extensive bar buckling and core crushing in the east boundary

element and increased damage to the concrete in the web. Figure 4.22 shows a photograph

and the final crack pattern of the wall at the end of the test.

3
Splice height was uniform along the length of the wall
221

(b) (c)

(a) Damage to first two floors

(b) Close up of damage in east (right) side of web above the splice. (c) Close up of damage in
east (right) boundary element
above the splice.

Figure 4.21: Specimen PW3 damage at second WL- peak at 1.0% drift (Step 1230).
222

(a) Damage at 1.5% drift in ER+ direc- (b) Crack map at failure (1.28% drift). Crack map
tion. not available for final damage state of the test.

Figure 4.22: Specimen PW3: Final damage state and crack pattern.

4.4 Specimen PW4

Specimen PW4 had heavily reinforced boundary elements; reinforcement was continuous

from the foundation to the wall cap. Figure 4.23 and 4.24 show the cross section and

elevation drawings of the as-built specimen.

An average axial load of N = 360 kips (1601 kN) or λN = 0.117, was applied to the

wall.A lateral force (Vapp ) and overturning moment (Mapp ) were applied to the top of the

specimen. Throughout the test, the ratio of the lateral force to the overturning moment was

held constant such that the base reactions (base shear Vb and base moment Mb ; measured

at the wall-foundation interface) were equivalent to those of a 10-story wall with a uniform

lateral load distribution. The applied load ratio (Mapp /Vapp ) was 11.2 feet (0.28 m), ancillary

actuators provided first and second floor lateral loads of 12.2% of Vapp , and the ratio of the
223

Figure 4.23: Specimen PW4 cross-section. Drawing provided by Chris Hart [44].

Figure 4.24: Specimen PW4 elevation. Drawing provided by Chris Hart [44].
224

Table 4.4: Specimen PW4 applied displacement history.

Displacement, in (mm) Drift Cycles SimCor Steps


0.02 (0.5) 0.014% 1-3 10-70
0.04 (1.0) 0.028% 4-6 70-130
0.2 (5.1) 0.139% 7-9 130-250
0.3 (7.6) 0.208% 10-12 250-370
0.5 (12.7) 0.347% 13-15 370-610
0.72 (18.3) 0.50% 16-18 610-850
1.08 (27.4) 0.75% 19-21 850-1090
1.44 (36.6) 1.0% 22-23 1090-1250

base reactions was 21.0 feet (0.53 m). The effective height of the load was 0.52h10 , or

1.73hw . Table 4.4 summarizes the applied displacement history.

The following sections summarize general response and damage progression of the test.

Section 4.4.1 provides a general overview of the test, including load-drift response, failure

mode, and comparison of the maximum strength to the expected strength. Section 4.4.2

provides an overview of the wall damage.

4.4.1 General Response

The load-displacement hystereses for Specimen PW4 are shown in Figure 4.25. Figure 4.25a

shows the base shear versus drift. Figure 4.25b shows the base moment versus drift. The

wall lost lateral load carrying capacity during the second cycle to 1.0% drift due to extensive

damage to the concrete and buckling of the steel the east (right) boundary element at the

base of the wall.

At the base of the wall, the maximum shear demand (Vb ) was approximately 0.84Vn ,

where Vn is the expected shear strength of the cross section (see Section 3.3.1). The maxi-

mum moment demand (Mb ) was approximately 1.02Mn (0.94Mn in the ER+ direction and

1.02Mn in the WL- direction), where Mn is the expected moment strength of the cross sec-

tion (see Section 3.3.2). The expected nominal moment strength of the wall was first reached

at approximately −0.89% drift; for the purpose of data analysis, the nominal strength is
225

considered to occur at Step 1150 in the WL- direction.

4.4.2 Damage

A summary of the primary damage states and their progression is presented here. Ap-

pendix A.4 provides a detailed description of the damage progression in Specimen PW4,

including a large set of images depicting the damage. A time-lapse movie of damage in

the first floor (lower one-third of the specimen) can be viewed at http://www.youtube.

com/user/NEESRWallProject#p/u/4/auU03q70EVg. The key damage states and the corre-

sponding drift at which it occurred follows.

1. Horizontal cracking initiated at 0.06% drift.

2. Diagonal cracking initiated at 0.07% drift.

3. Compressive yield of extreme vertical reinforcing bars was indicated by strain gauges

at 0.19% drift.

300 6000 8000


1200
6000
200 900 4000

600 4000
100 2000
300 2000 M b , kN-m
M b , k-ft
Vb , kip

Vb , kN

0 0 0 0

−300 −2000
−100 −2000
−600 −4000
−200 Test −900 −4000 Test
−6000
Vn Mn
−1200
−300 −6000 −8000
−1.5 −1 −0.5 0 0.5 1 1.5 −1.5 −1 −0.5 0 0.5 1 1.5
% Drift % Drift
(a) Base shear (b) Base moment

Figure 4.25: Specimen PW4 load-drift response.


226

4. Tensile yield of extreme vertical reinforcing bars was indicate by strain gauges at

0.30% drift.

5. Vertical cracking initiated during the first cycle to 0.50% drift.

6. Cover spalling initiated at the toe of the wall during the first cycle to 0.50% drift.

7. Concrete damage extensive enough to expose the longitudinal reinforcement occurred

at the toe of the wall at 0.75% drift.

8. Bar buckling was observed at the toe of the wall during the third cycle to 0.75% drift.

9. Damage (crushing) of the confined core of the boundary element was observed during

the third cycle to 0.75% drift.

10. Failure occurred due to extensive bar buckling and core crushing in the east (right)

boundary element at the toe of the wall during the second ER+ half-cycle to 1.0%

drift.

Figure 4.26 shows the crack pattern in Specimen PW4 following three cycles to 0.21%

drift.

Figure 4.26: Specimen PW4: Crack pattern in lower two stories following three cycles to
0.21% drift [44].
227

Damage to the concrete and reinforcement primarily occurred in the boundary elements

at the wall-foundation interface. In the east (right) boundary element, at a height of ap-

proximately 4 inches (10.2 cm), cracking and spalling occurred primarily where the cover

concrete was patched. Throughout the test, damage in the east (right) toe of the wall was

greater than that in the west (left) toe of the wall. Figure 4.27 shows the damage to the

toes of the wall during the first cycle to 0.75% drift.

Specimen PW4 lost lateral load carrying capacity during the second cycle to 1.0% drift,

at 0.85% drift (Step 1187). The loss of lateral load carrying capacity was associated with

a significant increase in bar buckling and core crushing in the east (right) boundary ele-

ment; the damage extended the full length of the boundary element, where on the previous

cycle, the damage extended only a few inches in from the edge of the wall. On the back

(north) of the wall, buckled horizontal and longitudinal steel was observed in the web. Fig-

ure 4.28 shows the base of the east (right) and west (left) boundary elements at the second

ER+ peak. The test was terminated following a return to zero displacement (Step 1250).

Figure 4.29 shows a photograph and the final crack pattern of the wall at the end of the

test.

4.5 Comparison of Walls

Sections 4.1 thru 4.4 provide an overview of the observed response of the four planar wall

specimens. This section provides a brief comparison of those observations. Section 4.5.1

compares the observed and expected strengths for the walls. Section 4.5.2 compares the

observed damage.

4.5.1 General Response

Load-drift response provides insight into stiffness, strength and nonlinear behavior. Enve-

lope curves of the load-drift response were used to compare these characteristics for each

specimen. The envelopes were created using the first peaks (ER+ and WL- loading direc-

tions) for each drift level: where prudent, additional points were included to capture the
228

(a) Bar buckling and core damage in east (right) toe of wall.

(b) Exposed longitudinal reinforcement in west (left) toe of wall.

Figure 4.27: Specimen PW4: Damage to wall following three cycles to 0.75% drift (Step
1070).
229

(a) West (left) boundary element.

(b) East (right) boundary element

Figure 4.28: Specimem PW4 damage at base of wall at the second ER+ peak at 1.0% drift
(Step 1230).
230

(a) (b)

Figure 4.29: Specimen PW4: Final damage state and crack pattern.

exact shape of the load-drift hystereses. The envelopes are normalized by the expected

strengths (see Section 3.3) to account for the differences in material properties and ap-

plied loads. In addition to the envelopes, the maximum values of each response value are

tabulated and compared to the expected strengths. The sections that follow present this

information for base shear demand, moment demand at the base and within the splice, and

finally, the demand at the critical section.

4.5.1.1 Moment Demand and Capacity

The intended response mode of the specimens was flexure. Comparison of the flexural

response is made by comparing the measured moment-drift envelopes. For each wall, the

expected moment capacity at the base of the wall and above the splice is the nominal moment

strength of the cross section, Mn , as calculated from fiber analysis of a cross section (see
231

Section 3.3.2). Within the spliced region of the specimens, the expected moment capacity

of the cross section was estimated as the nominal moment strength of the cross section

with twice the longitudinal steel, Mn,sp (see Section 3.3.2). The base moment demand

(Mb ), the moment demand at the top of the splice (Msp,top ), and the moment demand at

the center of the splice (Msp,mid ) are calculated from the applied lateral forces and the

specimen geometry.

Figure 4.30 shows the envelopes of the base moment normalized by the expected nominal

strength of the walls. Figures 4.31 shows the normalized envelopes at the top of the splice4 .

Table 4.5 provides the maximum (ER+, WL-, and average of both) moment values, as well

as the normalized values, at the base of the wall for all four specimens and at the center

and top of the splice for specimens with spliced longitudinal reinforcement.

The impact of lateral load distribution is observed by comparing Specimens PW1 and

PW2. At the base of the wall, the normalized strength ratio is the same (Mb /Mn = 1.07) for

both walls in the WL- direction (failure direction), but was slightly larger (Mb /Mn = 1.08

versus 1.03) for PW1 in the ER+ direction. In the WL- direction, the envelopes of the

two walls are similar until Specimen PW2 reaches its lateral drift capacity. In the ER+

direction, the envelopes have similar initial stiffness, but as each specimen approaches its

strength, the envelope of Specimen PW2 falls below that of Specimen PW1. Comparing

the responses at the top of the splice, it is noted that Specimen PW2 does not reach the

nominal moment strength of the wall at that elevation. The comparison of the specimens

at the top of the splice are similar to the comparison at the base of the wall, however,

Specimen PW2 does not reach, in either direction, the expected nominal moment strength

of the wall at that elevation. Specimen PW1 reached 1.5% drift in both the ER+ and WL-

directions, while Specimen PW2 reached this drift only in the ER+ direction; in the WL-

directions, Specimen PW2 reached just more than 2/3 the drift of Specimen PW1.

4
Specimen PW4 is included on this plot even though there is no splice in this wall. The envelopes are
calculated at the same elevation for all specimens, thus the PW4 envelopes serve as a reference for the
demand-capacity ratio in a wall with continuous reinforcement.
232

1.2
1
0.8
0.6
0.4
0.2
Mb /Mn

0
−0.2
−0.4
−0.6
−0.8 PW1
PW2
−1 PW3
PW4
−1.2
−1.5 −1 −0.5 0 0.5 1 1.5
Drift, %

Figure 4.30: Base moment (Mb ) envelopes normalized by nominal moment strength (Mn ).

Table 4.5: Summary of maximum moment demands at the base of the wall (Mb ), center of
top
splice (Msp,mid ), and top of splice (Msp ).

Specimen Dir. Mb , k-ft Mb /Mn Msp,mid , k-ft Msp,mid /Mn,sp Msp,top , k-ft Msp,top /Mn
PW1 ER+ 5250 1.08 4998 0.63 4868 1.00
WL- 5175 1.07 4954 0.62 4818 0.99
Avg. 5212 1.08 4976 0.62 4843 1.00
PW2 ER+ 5726 1.03 5392 0.62 5185 0.94
WL- 5913 1.07 5571 0.64 5355 0.97
Avg. 5819 1.05 5482 0.63 5270 0.95
PW3 ER+ 4529 1.16 4265 0.73 4087 1.05
WL- 4496 1.16 4226 0.72 4042 1.04
Avg. 4512 1.16 4246 0.73 4064 1.04
PW4 ER+ 4093 0.94 - - - -
WL- 4404 1.02 - - - -
Avg. 4248 0.98 - - - -
233

1.2
1
0.8
0.6
0.4
0.2
/Mn

0
top
Msp

−0.2
−0.4
−0.6
−0.8 PW1
PW2
−1 PW3
PW4
−1.2
−1.5 −1 −0.5 0 0.5 1 1.5
Drift, %

Figure 4.31: Normalized moment envelopes at the top of the spliced region of the wall
specimens. Specimen PW4 did not have a splice; the envelopes at the same height are
shown as a reference.

The impact of reinforcement distribution is observed by comparing of Specimens PW2

and PW3. At the base of the wall, the normalized strength ratio is larger for Specimen

PW3 (Mb /Mn = 1.16 versus 1.05). The envelopes of the two specimens have the same

initial stiffness, but as they approach the peak strength, Specimen PW2 is more flexible.

Specimen PW2 reached 1.5% drift in the ER+ direction and 1.05% drift in the WL- direc-

tion, while Specimen PW3 reached 1.28% drift in the ER+ direction and 1.0% drift in the

WL- direction.

The impact of spliced longitudinal reinforcement is observed by comparing of Specimens

PW2 and PW4. The values at the base of the wall are compared. Specimen PW4 had

a lower normalized flexural strength ratio (Mb /Mn = 0.98 versus 1.05). The normalized

base moment envelopes are nearly identical for the initial stiffness. As the response curves
234

approach the peak strength, the responses are also similar. Specimen PW2 reached a larger

drift (1.5% versus 1.0%) in the ER+ direction, but the drift capacities were approximately

the same (1.05% and 1.0% drift) in the WL- direction.

4.5.1.2 Critical Section Demand and Capacity

Evaluation of the maximum normalized flexural strength ratios for the walls shows that, for

specimens with spliced longitudinal reinforcement, the base moment demands exceed the

nominal moment strength of the cross section and the moment demand at the top of the

spliced region is approximately equal to the nominal moment strength. This observation,

along with the observed concentration of damage at the top of the spliced regions, suggests

that the splice has the affect of “shifting” the critical cross section of the wall from the base

of the wall (as is the case for the unspliced Specimen PW4) to the top of the spliced region.

To account for the different critical sections (top of the splice for Specimens PW1,

PW2, and PW3; base of the wall for Specimen PW4), one more comparison of the force-

drift envelopes is made. Figure 4.32 shows the envelopes of the moment demand at the

critical section (Mcrit ), normalized by the nominal moment strength at that section (Mn,crit ).

Table 4.6 specifies the location of the critical section, the maximum normalized values,

Mcrit /Mn,crit , and the drift capacity. When the wall demand-capacity ratios are compared

in this manner, the envelopes and normalized strengths are very similar; however, the drift

capacities of the walls are not.

4.5.1.3 Shear Demand and Capacity

For each specimen, the expected shear capacity is the nominal shear strength of the wall,

Vn , as calculated per ACI 318-08 [1] (see Section 3.3.1). The base shear demand, Vb , is

equal to the total of all applied lateral forces. Because one of the planar wall test program

study parameters is the shear demand (achieved by different lateral load disributions), it

is expected that Specimen PW1, loaded to create base reactions for an ASCE 7 lateral
235

1.2
1
0.8
0.6
0.4
Mcrit/Mncrit

0.2
0
−0.2
−0.4
−0.6
−0.8 PW1
PW2
−1 PW3
PW4
−1.2
−1.5 −1 −0.5 0 0.5 1 1.5
Drift, %

Figure 4.32: Critical moment (Mcrit ) envelopes normalized by the computed moment
strength at the critical section (Mcompcrit ).

Table 4.6: Summary of maximum moment demands at the critical section of the wall (Mcrit ).

Specimen Direction Section Mcrit /Mn,crit Max. Drift


PW1 ER+ Top Splice 1.00 1.5%
WL- Top Splice 0.99 1.5%
Avg. Top Splice 1.00 -
PW2 ER+ Top Splice 1.04 1.5%
WL- Top Splice 1.07 1.05%
Avg. Top Splice 1.05 -
PW3 ER+ Top Splice 1.05 1.25%
WL- Top Splice 1.04 1.0%
Avg. Top Splice 1.04 -
PW4 ER+ Base 0.94 1.0%
WL- Base 1.02 1.0%
Avg. Base 0.98 -
236

load distribution (inverted triangle), would have a lower shear demand relative to the shear

capacity than would the other three specimens which were loaded with a uniform lateral

load distribution.

Figure 4.33 shows the envelopes of the base shear normalized by the expected nominal

strength of the walls. Table 4.7 provides the maximum (ER+, WL-, and average of both)

base shear values, as well as the normalized value (Vb /Vn ) and the shear stress demand (νb ),
p
expressed in terms of fc0 .

The impact of lateral load distribution is observed by comparing Specimens PW1 and
p p
PW2. As expected, PW1 has a lower shear demand (3.5 fc0 versus 4.9 fc0 ) and shear-

demand capacity ratio (Vb /Vn = 0.73 versus 1.06). The slope of the envelopes is initially

very similar, but Specimen PW1 has a change in stiffness (becomes more flexible) sooner

than does Specimen PW2 and reaches an approximate strength plateau sooner.

The impact of reinforcement distribution is observed by comparing Specimens PW2 and


p p
PW3. Specimen PW3 has a lower shear demand (4.4 fc0 versus 4.9 fc0 ) and shear-demand

capacity ratio (Vb /Vn = 0.90 versus 1.06). The initial slopes of the envelopes, the initial

change in stiffness, and the drifts at which the maximum shear demands reached are very

similar for the two specimens.

The impact of spliced longitudinal reinforcement is observed by comparing Specimens


p p
PW2 and PW4. Specimen PW4 has a lower shear demand (4.5 fc0 versus 4.9 fc0 ) and

shear-demand capacity ratio (Vb /Vn = 0.90 versus 1.06). The initial slopes of the envelopes

and the initial change in stiffness are very similar for the two specimens, however, Specimen

PW4 reaches the maximum shear at a lower drift than does Specimen PW2.

In comparing the specimens, it is important to note a key characteristic that was not

intended by the test program. Specimen PW2 was loaded with a larger axial force than

intended, resulting in an axial load ratio of λN = 0.13, as opposed to the intended ratio of

λN = 0.10. While Specimen PW4 had a similar axial load ratio (due to concrete compressive

strength below the design strength), Specimen PW2 is unique in that the increased axial
237

1.2
1
0.8
0.6
0.4
0.2
Vb /Vn

0
−0.2
−0.4
−0.6
−0.8 PW1
PW2
−1 PW3
PW4
−1.2
−1.5 −1 −0.5 0 0.5 1 1.5
Drift, %

Figure 4.33: Base shear (Vb ) envelopes normalized by nominal shear strength (Vn ).

Table 4.7: Summary of maximum shear demands at the base of the wall (Vb ).
p
Specimen Direction Vb , kips Vb /Vn νb , fc0 Max. Drift
PW1 ER+ 188 0.75 3.60 1.50%
WL- 180 0.71 3.45 1.50%
Avg. 184 0.73 3.53 -
PW2 ER+ 267 1.04 4.85 1.50%
WL- 276 1.07 5.02 1.05%
Avg. 272 1.06 4.94 -
PW3 ER+ 224 0.90 4.41 1.25%
WL- 226 0.91 4.45 1.05%
Avg. 225 0.90 4.43 -
PW4 ER+ 203 0.84 4.32 1.0%
WL- 218 0.90 4.63 1.0%
Avg. 210 0.87 4.47 -
238

force results in a large enough increase in the expected nominal moment strength of the wall

such that the nominal shear strength of the wall can be reached prior to or approximately at

the nominal moment strength being reached; in the other specimens, the maximum moment

strength is reached prior to the full shear capacity being reached. It is reasonable to assume

that, had Specimen PW2 been loaded with a smaller axial force, the shear demand values

may have more closely resembled those of Specimens PW3 and PW4.

4.5.2 Damage

This section, a comparison of the damage to the walls is presented. First, a summary of

key damage states (cracking, spalling, etc.) is made and response quantities are compared

between the four specimens. Next, the initial yield of the longitudinal reinforcement is

compared between the walls. This is followed by a discussion of the crack patterns in the

walls. Finally, a discussion on the locations of the primary damage is provided.

4.5.2.1 Damage State Comparison

This section presents a comparison of basic response quantities (normalized reactions and

drift) for several key damage states and specify the steps at which the damage occur. These

values are presented in Tables 4.8 through 4.10. Generally, response quantities are provided

for the step at which the damage occurs, however, in some instances, the response quantities

are reported for the step associated with the maximum historic drift in the direction of

loading5 ; in these case, the step numbers associated with the damage and the maximum

historic drift are both reported. For each damage state, the step number(s), specimen drift,

base shear demand-capacity ratio (Vb /Vn ), base moment demand-capacity ratio (Mb /Mn ),
crit )
and the moment demand-expected capacity ratio at the critical cross section (Mcrit /Mcomp

are presented.

5
It is necessary to report response quantities for the maximum historic drift for damage such as bar
buckling if the damage occurs at a low drift after a higher drift demand has been applied to the wall in
the same direction.
239

Table 4.8 presents the response quantities for crack damage states (horizontal cracks,

diagonal cracks and vertical cracks). The following observations can be made:

Table 4.8: Summary of response at crack damage states. Where two step numbers are
presented, the first is the step at which response values are reported and the number in ( )
is the step the damage occurred.

Damage Value PW1 PW2 PW3 PW4


Horizontal cracks Drift, % 0.06 0.10 0.05 0.06
Vb /Vn 0.23 0.41 0.30 0.31
Mb /Mn 0.33 0.41 0.39 0.36
Mcrit /Mcomp 0.30 0.41 0.35 0.36
Step 73 137 134 134
Diagonal cracks Drift, % 0.10 0.10 0.07 0.07
Vb /Vn 0.29 0.41 0.36 0.33
Mb /Mn 0.43 0.41 0.44 0.39
Mcrit /Mcomp 0.40 0.41 0.39 0.39
Step 75 137 135 135
Vertical cracks Drift, % −0.54 0.35 −0.30 0.51
Vb /Vn −0.63 0.76 −0.69 0.83
Mb /Mn −0.94 0.75 −0.87 0.92
Mcrit /Mcomp −0.87 0.75 −0.77 0.92
Step 420 (456) 470 427 630

1. Horizontal cracks initiated later in Specimen PW2 than in the other three specimens.

This is indicated by the drift at which horizontal cracks were first observed (0.10% for
crit = 0.41
PW2, 0.05 − 0.06% for others) and the demand-capacity values (Mcrit /Mcomp

for PW2, 0.30 − 0.36 for others). The observed delay in horizontal cracks may be

explained by the larger axial force applied to Specimen PW2.

2. Horizontal cracks were observed in Specimen PW1 at a lower normalized moment


crit = 0.30 for PW1, 0.35 − 0.41 for others).
ratio than for other specimens (Mcrit /Mcomp

3. Diagonal (or inclined) cracks initiated at approximately the same demands for all

specimens. Diagonal cracks were first observed at 0.10% drift for Specimens PW1 and

PW2 and at 0.07% for Specimens PW3 and PW4. The moment demand-expected
crit ) was approximately 0.40 for all
capacity ratio at the critical section (Mcrit /Mcomp
240

four specimens.

4. In Specimens PW1, PW3, and PW4, horizontal cracks were observed at slightly lower

drifts than the drift at which diagonal cracks were observed. In Specimen PW2,

horizontal and diagonal cracks were observed at the same drift.

5. The demand-capacity ratios associated with the first observed horizontal and diagonal

cracks are approximately equal to the demand-capacity ratios associated with change

in stiffness of the envelopes presented in Figures 4.30 thru 4.33.

Table 4.9 presents the response quantities for concrete damage (spalling and core crush-

ing). Two types of concrete spalling are identified; initial spalling of the concrete and

spalling of the concrete to the extent that longitudinal reinforcement is exposed. The fol-

lowing observations can be made:

Table 4.9: Summary of response at concrete damage states. Where two step numbers are
presented, the first is the step at which response values are reported and the number in ( )
is the step the damage occurred.

Damage Value PW1 PW2 PW3 PW4


Spalling Drift, % 0.56 0.76 −0.41 0.51
Step 478 870 666 630
Exposed long. reinf. Drift, % −0.79 −0.75 0.75 −0.76
Step 500 (535) 990 (1055) 1030 910
B.E. core crushing Drift, % 1.47 1.50 −1.00 −0.75
Step 720 1270 1230 1070

1. Initial spalling occurs at drifts ranging from 0.41 − 0.76%. Spalling occurs at the

lowest drift in Specimen PW3 and at the highest drift in Specimen PW2.

2. Spalling revealing longitudinal reinforcement occurs during the 0.75% drift cycles for

all four specimens.

3. The drift demands at which core crushing is first observed ranges from a low of −0.75%

drift for Specimen PW4 to a high of 1.5% drift for Specimens PW1 and PW2.
241

Table 4.10 presents the response quantities for rebar damage (bar buckling and bar

fracture). Bar buckling is first observed at ±0.75% drift for Specimens PW2 and PW4 and

at ±1.0% drift for Specimens PW1 and PW4. Bar fracture occurs only in Specimen PW1.

Table 4.10: Summary of response at rebar damage states. Where two step numbers are
presented, the first is the step at which response values are reported and the number in ( )
is the step the damage occurred.

Damage Value PW1 PW2 PW3 PW4


B.E. bar buckling Drift, % −1.04 −0.76 1.00 0.75
Step 620 1070 1110 950
B.E. bar fracture Drift, % −1.39 - - -
Step 739 - - -

4.5.2.2 Initial Yield

This section presents data related to initial yield of the longitudinal reinforcement. Addi-

tional analysis and discussion of the yield behavior of the walls is presented in Chapter 6.

Table 4.11 presents basic information on when initial yield was measured by the strain

gauges on the extreme longitudinal reinforcing bars. Initial yield corresponds to a gauge

measurement of the yield strain (±εy ). Data is presented for yield in tension (+εy ) and

compression (−εy ) for each direction of loading (ER+ and WL-). Initial yield in the rein-

forcing bars was measured at different heights in each specimen, thus, the elevation at which

yield first occurred is provided, as well as the moment ratio at that elevation (Melev /Mn ).

The following observations are made:

1. In walls with splices (Specimens PW1, PW2, and PW3), tensile yield was first mea-

sured above the splice. The drift values at initial tensile yield cannot be compared

among all three specimens as the yielding was measured at different heights (in Speci-

mens PW1 and PW2, no gauges were provided at an elevation of 24 inches (61.0 cm),

the elevation at which at which yield was measured in Specimen PW3). Specimen
242

PW1 and PW2 yielded at the same height, with yield in PW1 having occurred at

lower drifts than PW2 (0.36% versus 0.43%).

2. In Specimen PW4 (no splice), tensile yield was first measured at the base of the wall.

Yield occurred at a lower drift demand in the ER+ loading direction (0.30%) than in

the WL- loading direction (0.35%).

3. In walls with splices (Specimens PW1, PW2, and PW3), compression yield was first

measured within the spliced region. In Specimens PW1 and PW3, compression yield

occurred after tension yield. In Specimen PW2 (higher axial demand), compression

yield occurred after tension yield in the WL- loading direction and after tension yield

in the ER+ loading direction.

4. In the wall without splices (Specimen PW4), compression yield was first measured at

the base of the wall. The drift demand at compression yield (≈ 0.20%) was less than

the drift demand at first tension yield (0.30 − 0.35%).

Table 4.11: Initial Yield of Longitudinal Reinforcement

Melev
Wall Description Step (ER+/WL-) Drift (%) Mn
Elev. (in)
PW1 +εy (B.E.) 358 / 259 0.37 / -0.35 110.27 / -95.82 35.5 / 35.5
−εy (B.E.) 479 / 498 0.64 / -0.64 142.12 / -142.64 18.5 / 7.5
PW2 +εy (B.E.) 627 / 667 0.43 / -0.43 102.98 / -105.02 35.0 / 35.0
−εy (B.E.) 260 / 910 0.21 / -0.75 82.22 / -139.55 7.5 / 18.5
PW3 +εy (B.E.) 258 / 279 0.17 / -0.19 82.29 / -89.20 24.0 / 24.0
−εy (B.E.) 390 / 428 0.35 / -0.32 128.56 / -115.04 7.5 / 24.0
PW4 +εy (B.E.) 387 / 430 0.30 / -0.35 105.78 / -113.25 2.0 / 2.0
−εy (B.E.) 300 / 279 0.21 / -0.19 87.85 / -83.33 2.0 / 2.0

Table 4.12 presents basic information on when initial yield was measured by the strain

gauges on the extreme longitudinal reinforcing bars within the foundation (gauges were

located 2 inches (5.08 cm) below the wall-foundation interface). This location was selected to

provide a consistent comparison between all four wall specimens. The following observations

can be made:
243

1. Specimen PW1 yielded in tension but not compression in the foundation.

2. Specimen PW2 yielded in both tension and compression on the east (right) side of the

foundation but did not yield on the west (left) side of the foundation.

3. Specimen PW3 did not yield in tension and yielded in compression only on the east

(right) side of the foundation; this gauge yielded very early in the test at 0.01% drift,

suggesting an error in the measurement.

4. Specimen PW4 yielded in both tension and compression on both sides of the founda-

tion.

5. Yield of the reinforcement in the foundation typically occurs at lower drift demands

those associated with initial yield in the wall (Table 4.11). Exceptions to this are

compressive yield in Specimens PW2 and PW3.

Table 4.12: Initial Yield of Longitudinal Reinforcement in Foundation

Mb
Wall Description Step (ER+/WL-) Drift (%) Mn
Elev. (in)
PW1 +εy (Fnd) 359 / 620 0.42 / -1.04 132.51 / -153.70 -2.0 / -2.0
−εy (Fnd) -/- -/- -/- -/-
PW2 +εy (Fnd) - / 1144 - / -0.70 - / -137.81 - / -2.0
−εy (Fnd) 258 / - 0.17 / - 78.33 / - -2.0 / -
PW3 +εy (Fnd) -/- -/- -/- -/-
−εy (Fnd) 14 / - 0.01 / - 18.84 / - -2.0 / -
PW4 +εy (Fnd) 626 / 669 0.41 / -0.48 122.09 / -133.35 -2.0 / -2.0
−εy (Fnd) 709 / 669 0.48 / -0.48 123.94 / -133.35 -2.0 / -2.0

4.5.2.3 Crack Patterns

Two sets of crack maps are presented here to compare the four wall specimens. Figure 4.34

shows the crack patterns for the walls after completion of three cycles to 0.35% drift; this

drift was selected because it was the lowest peak drift that all specimens had in common

and the crack patterns were well established. Figure 4.35 shows the crack patterns for the

walls after completion of the tests. Comparison of the crack maps here is done visually;
244

analysis of crack spacing, orientation, length and width is presented by Hart [44].

(a) PW1 (b) PW2

(c) PW3 (d) PW4

Figure 4.34: Crack patterns in lower two floors of the walls following three cycles to 0.35%
drift. Crack maps provided by Chris Hart [44].

Specimens PW1 and PW2 were subjected to different load distributions. The crack

pattern in the boundary elements are similar. The horizontal cracks extended the length

of the boundary element and were spaced at approximately the spacing of the confining
245

(a) PW1 (spalling not indicated) (b) PW2

(c) PW3 (prior to end of test) (d) PW4

Figure 4.35: Crack and spall patterns in lower two floors of the walls at end of test. Crack
maps provided by Chris Hart [44].

reinforcement. The crack patterns are different in the web of the walls. For Specimen PW1,

which had a lower shear demand, the horizontal cracks extend into the web. Diagonal cracks

initiate within the web of the wall. In Specimen PW2, the diagonal cracks initiate at the

interface of the boundary element and web.


246

Specimens PW2 and PW3 have different longitudinal reinforcement layouts. At 0.35%

drift, the spacing of horizontal cracks in the confined boundary elements was greater for

Specimen PW3 than for Specimen PW2. The length of the horizontal cracks is larger

for Specimen PW3. Diagonal cracks initiate at the boundary element-web interface for

Specimen PW2 and near the center of the web for Specimen PW3.

Specimens PW2 and PW4 provided a comparison the use of a lap splice in the critical

region. After cycles to 0.35%, the cracks in the bottom half of the first floor (spliced region)

boundary elements of Specimen PW2 are less dense than the cracks above the splice. Such

an observation is logical due to the increased stiffness of the spliced region. However, in the

unspliced Specimen PW4, the density of cracks in the bottom half of the first floor is also

less than higher up in the wall. Looking at the end of the tests instead of at 0.35% drift,

the number, spacing, and orientation of cracks in the boundary elements of PW4 are the

same the entire height of the first two floors; this is not the case for Specimen PW2. The

difference in the final crack pattern between the two walls suggests that a spliced boundary

element does not have the purely flexure response that an unspliced boundary element has.

4.5.2.4 Locations of Primary Damage

The following paragraphs discuss the impact of the test program study parameters on the

location of the primary damage in the walls.

Specimens PW1 and PW2 were subjected to different load distributions. In both speci-

mens, there was a concentration of damage (spalling and bar buckling) above the top of the

splice in the west (left) boundary element and to a lesser extent above the top of the splice

in the east (right) boundary element. In Specimen PW1, the spalling also occurred at the

base of the wall and extending the height of the splice along the edge. The failure mode and

location of the failure were different for the two specimens, but both failed while loading in

the WL- direction; PW1 failed due to bar fracture at the base of the east (right) boundary

element and PW2 failed due to shear compression failure of the west (left) boundary element
247

above the top of the splice.

Specimens PW2 and PW3 have different longitudinal reinforcement layouts. In both

specimens, the damage was concentrated above the top of the splice, and spalling first

occured at the edges of the wall. However, Specimen PW3 (uniform distribution of steel)

had the most damage occur at the top of the splice in the unconfined web of the wall, while

the damage prior to failure in Specimen PW2 was restricted to the boundary element. Both

specimens failed due to shear-compression damage (bar buckling and core crushing) to the

boundary elements above the splice.

Specimens PW2 and PW4 provided a comparison the use of a lap splice in the critical

region. In both specimens, the damage was concentrated in the boundary elements. In

Specimen PW2, this is above the top of the splice, while in Specimen PW4 this was at the

base of the wall. Failure of both walls was due to shear-compression failure (bar buckling

and core crushing) where the damage was concentrated.

4.6 Conclusions

This chapter presents a summary of experimental results for the “Design, Behavior, and

Analysis of Complex Walls” project planar wall test program. The parameters studied by

the test program were lateral load distribution (shear demand), distribution of longitudinal

reinforcement ratio, and the use of lap splices at the base of the walls. For each specimen,

load-drift hysteresis and a brief summary, including photos, of damage progression and

failure mode are presented. Normalized load-drift envelopes are compared from the four

specimens and the load and drift values for key damage states were tabulated. The following

observations are made:

1. Walls with lap splices located in the expected region of highest moment demand had

damage concentrated above the top of the spliced region. For these walls, the critical

section (that controlling the response of the specimen) was shifted from the base of

the wall to the top of the splice.


248

2. Walls with higher shear demands (achieved through a lower effective height of the

applied lateral load) had a loss of load carrying capacity due to shear-compressive

failure of the confined boundary elements, while the wall with lower shear demand

failed due to bar fracture at the base of the wall.

3. When the moment demand envelopes at the critical section of the wall (base or top

of splice) are normalized by the computed moment strength (lesser of the nominal

moment and thmoment corresponding to nominal shear), the response of the four

different specimens are nearly identical. While this observation is not necessarily of

importance to the overall conclusions of the test program, it is important because

it validates that the fundamental response of the four specimens are indeed similar

despite the different design, complex loading conditions and non-ideal boundary con-

ditions at the base of the wall. Thus, analysis of the collected data can be confidently

evaluated to seek an improved understanding of the impact of the test program study

parameters.

4. In walls with lap splices, crack patterns suggest that boundary element does not have

the same flexural response of a boundary element in an unspliced wall.

These observations are made only on the observed response of the tests. Analysis of the

data collected from instruments provided on the test are used to understand and build on

these observations. The methods used to analyze the experimental data are presented in

Chapter 5 and the analysis results are presented in Chapter 6.


249

Chapter 5

METHODS FOR ANALYSIS OF EXPERIMENTAL DATA

This chapter discusses the methods used to analyze data collected from the planar wall

test program of the “Design, Behavior, and Analysis of Complex Walls” project. The test

program and the experimental results are presented in Chapters 3 and 4. Detailed results

from the data analysis are not presented in this chapter. Rather, the results are presented

in Chapter 6. Presentation of the analysis methods here allows for a clean, free-flowing

presentation of the results in Chapter 6.

Section 5.1 discusses the calculation and presentation of displacements and rotations

along the height of the wall. Section 5.2 discusses the calculation of effective stiffness values.

Section 5.3 discusses the estimation of drift at the effective loading height (hef f ) and the

theoretical 10th story (h10 ) of the wall specimens. Section 5.4 discusses the quantification

of yielding behavior of the wall specimens using the strain gauge data. Section 5.5 discusses

the calculation of strain fields using the Krypton/Metris displacement data. Section 5.6

discusses the calculation of fixed-end, flexure and shear deformations and the respective

contributions to the total deformation of the wall.

5.1 Displacement and Rotation Calculations

Lateral displacements and in-plane rotations are useful in that they can be used to i) show

the displaced shape of the wall, ii) calculate effective stiffness values, and iii) determine shear,

flexure, and fixed-end deformation components. Due to the large number of instruments

provided on the specimens, redundant displacements and rotations are available or can be

calculated. The following sections provide a brief analysis of the data available and explains

how they are ultimately used. Section 5.1.1 addresses lateral displacements and Section 5.1.2
250

addresses in-plane rotations.

5.1.1 Displacements

Lateral displacements were measured using string pots and Krypton/Metris LED targets

(see Section 3.6). Most string pots were attached to threaded rods embedded approximately

2 inches (5.08 cm) from the west (left) edge of the wall, with the exception of the control

string pot, which was attached to the center bottom of the wall cap. The Krypton targets

were only available in the first two floors (bottom two-thirds) of the specimens. Within each

floor, 5 rows of up to 15 LEDs each were provided. Some rows of Krytpon targets were at

approximately the same elevation as the string pots and individual targets were attached to

the same rods as the string pots, allowing for comparison of the two measurement systems.

Review of the lateral displacement data available led to the following observations re-

garding the use of lateral displacements:

1. The displacement at the top of the specimen is only available from a string pot at-

tached to the bottom center of the wall cap. This is the control displacement and is

used to define the third floor drift frequently used in the presentation of the experi-

mental results and data analysis results.

2. The displacements measured by string pots attached to the west (left) side of the wall,

when used to plot displacement profiles that include the control string pot, result in

a “kink” in the displacement profile near the top of the specimen.

3. The displacements measured by the center column of Krypton targets, when used to

plot displacement profiles that include the control string pot, present a more reason-

able displacement profile (the “kink” mentioned in the item above is not present).

4. The displacements measured by the center target in a row of Krypton targets is

approximately the same as the average displacements measured by all Krypton targets

in the row.

5. The displacements measured by string pots attached to the west (left) side of the
251

0.7
String Pot
Avg. LED
0.6 West LED
Center LED
Horizontal displacement, in.

East LED
0.5

0.4

0.3

0.2

0.1

0
0 200 400 600 800 1000 1200 1400
Step #

Figure 5.1: Comparison of string pot and Krypton measured lateral displacements in Spec-
imen PW2 at the top of the first floor (ER+ peaks).

wall are approximately the same as the displacements measured by the western most

column of Krypton targets. These displacements on the west (left) side of the wall

are lower in magnitude than the center displacements when the wall is pushed in the

ER+ direction and greater in magnitude when the wall is pushed in the WL- direction.

Figure 5.1 illustrates the difference in measured displacement along the length of the

wall.

6. Lateral displacements of the foundation block (measured by a string pot attached to

the west (left) side of the block, indicate significant difference in the values measured

in the ER+ and WL- loading directions. The amount of measured slip was no more

that 2.0% of the total wall displacement.

Based on these observations, the following decisions were made regarding the use of

lateral displacement:
252

1. String pots attached to the west (left) side of the wall would not be used unless

alternative data was not available.

2. Krypton lateral displacements would be averaged across each row to provide a dis-

placement for the wall at that elevation. This should result in a displacement approx-

imately equal to that at the center of the wall, providing consistency with the control

string pot. Averaging the displacements provides an added benefit of reducing the

effect of any missing targets or bad data points.

3. When the displacements for the first and second floors were used, these were deter-

mined using the row of Krypton targets at approximately 2 inches below the actual

floor elevation. This is because no targets are available at the actual floor levels. For

the first floor level, targets were provided above and below this level, so the floor dis-

placement could reasonably be interpreted, but this could not be done for the second

floor as no targets were provided in the third floor. Thus, use of the row of targets two

inches below the floor level provided consistent results. In calculating drifts or using

the floor level displacements to calculate effective stiffnesses and drift contributions,

the height of the targets, not the floor will be used.

4. Measured displacements would not be modified to account for the lateral displacement

of the foundation block due to inconsistent readings and the negligible impact on the

total deformations.

5.1.2 Rotations

In-plane rotations were not measured directly but can be computed using the vertical dis-

placements measured by the Krypton LED targets and the LVDTs and linear potentiometers

measuring relative vertical displacements of the wall.

The general procedure in calculating the rotations was to fit a linear best-fit line to the

measured vertical displacements at a section of the wall. The angle of the best-fit line to the

horizontal was considered the rotation for that section. When calculating the rotation at
253

a section using the linear potentiometers that measured relative displacement, the vertical

displacements were determined from the sum of the measured relative displacements at all

instruments that terminate at or below that section in question. For calculations of the

rotation at a section using the Krypton LED targets,

Rotations calculated using Krypton and traditional instrumentation produced similar,

but not the same, values. No consistent observations between specimens were made with

respect to which values were larger. It was decided to use, whenever possible, values calcu-

lated from the Krypton measurement system as the rotations were calculated using more

data points and were available at more locations in the wall. In the third floor, traditional

instrumentation was used as no Krypton data was available.

Rotation of the foundation block was calculated using instruments measuring vertical

displacement of the foundation relative to the strong floor. The rotation of the foundation

block was determined to produce less than 1.0% of the total wall displacement. As with the

lateral displacement of the foundation block, the effects of the rotation of the foundation

block were not removed from the measured and calculated wall deformations.

5.2 Calculation of Effective Stiffness Values

Effective stiffness values are a fundamental tool necessary for the elastic analysis of struc-

tures. The large amount of data collected from the planar walls tests provides a unique

opportunity to calculate axial, flexural, and shear stiffness not only for the full wall speci-

men, but also for individual portions of the wall. Redundant measurements of axial defor-

mation, lateral deformation, and in-plane rotations make it necessary to consider multiple

approaches to calculating effective stiffness values. This section discusses the methods used

to calculate the effective stiffness values and the decisions that were made as to what mea-

surements were necessary for such calculations. Section 5.2.1 discusses axial stiffness. Sec-

tion 5.2.2 discusses calculation of shear and flexural stiffness using the Timoshenko beam

theory. Section 5.2.3 discusses calculation of flexural stiffness using the Euler-Bernoulli

beam theory for comparison to data available from other experimental test programs for
254

which the methods described in Section 5.2.2 cannot be used.

5.2.1 Axial Stiffness

measure ) of the walls was calculated as the axial force (N ) divided by


The axial stiffness (kaxial

the vertical displacement of the wall (∆z ). The calculated stiffness was normalized by the
gross
gross axial stiffness (kaxial ) to determine the axial stiffness factor, αaxial .

measure N
kaxial = (5.1)
∆z
gross Ag Ec
kaxial = (5.2)
hw
measure
kaxial
αaxial = gross (5.3)
kaxial

The axial force was constant throughout the tests. The axial deformation of the wall

was calculated in two ways1 . The first method was to average the total displacement of

each column of linear pots measuring relative displacements on the north (back) face of the

walls. The second method was to average the total vertical displacements of the string pots

measuring the vertical displacement of the wall cap relative to the foundation block. The

first method produced lower axial deformation and higher axial stiffness values. Ultimately

the vertical string pots were used to calculate the final axial stiffness values.

Calculation of the axial stiffness at the cycle peaks produced reasonable values (αaxial ≤

1.0) only at very low drift demands.

At steps where lateral load and overturning moment were applied to the walls, the

vertical deformation resulting from the compressive axial load could not be separated from

the deformation resulting from the flexural demand, thus the axial stiffness was calculated

only at the steps closest to zero lateral force. At low drift levels, the axial deformation at

zero lateral force grew in magnitude (negative deformation due to the compressive load)

from that when the axial load was first applied. In all specimens, a drift level was reached

1
The krypton system was not considered for the calculation of the axial stiffness of the wall because no
data was available in the third floor of the specimen.
255

in which the magnitude of the axial deformation decreased, resulting in an increase in the

calculated axial stiffness factor (αaxial ). This is likely the result residual crack openings and

does not represent a realistic behavior of the response of the wall under axial load, thus the

axial stiffness values beyond these drifts were disregarded. For Specimens PW1 and PW3,

this wall was drift of 0.50% and beyond. The other two specimens had reasonable values

to larger drifts, which is presumably a function of the slightly higher axial load ratios; data

for Specimen PW2 and PW4 are disregarded at 1.0% and 0.75% drift, respectively.

5.2.2 Shear and Flexural Stiffness: Timoshenko Beam Theory

Timoshenko beam theory for static beams accounts for shear and flexural deformations in

a beam; it is represented by the following set of governing equations:

 
∂2
∂θ
−w = EI (5.4)
∂x2
∂x
 
∂u 1 ∂ ∂θ
= θ− EI (5.5)
∂x κAG ∂x ∂x

where w is the distributed load on the beam, x is the length along the beam, θ is the

rotation, u is the deflection of the beam, E is the modulus of elasticity, I is the moment of

inertia, A is the area, G is the shear modulus, and κ is a shear coefficient dependent of the

cross-section shape of the beam. If the shear term of the governing equations is ignored,

the Timoshenko beam theory reduces to the classic, or Euler-Bernoulli beam theory (see

Section 5.2.3), which only requires deflection boundary conditions. The Timoshenko beam

theory requires deflections and rotations as boundary conditions, and thus, cannot be used to

calculate effective flexural stiffness values for many experimental test programs. Section 5.1

discusses the calculation of lateral deformation and in-plane rotations that allow calculation

of the effective shear and flexural stiffness for the planar wall test program.

Timoshenko beam theory is used to determine the flexural (EI) and shear stiffness (AG)

for each floor of the wall specimens. Within each floor, the moment distribution M (x) at a
256

distance x from the bottom of the floor is:

M (x) = Mbot − Vbot x (5.6)

where Mbot and Vbot are the moment and shear at the bottom of the floor, respectively.

Integrating Eqn 5.4 for this moment distribution and applying the boundary condition

θ(x = 0) = θbot results in Eqn 5.7; using the known rotation at the top of the story (x = h),

θ(x = h) = θtop , results in Eqn 5.8. The effective stiffness factor, α is calculated as the

effective flexural stiffness, EIef f , over the gross flexural stiffness, EIg (Eqn 5.9).

1 Vbot x2
θ(x) = [Mbot x − ] + θbot (5.7)
EI 2
1 Vbot h2
EIef f = [Mbot h − ] (5.8)
θt − θbot 2
EIef f
αf lex = (5.9)
EIg

Integration of Eqn. 5.5 with the boundary condition u(x = 0) = ∆bot results in the

expression for beam deflection given by Eqn. 5.10. The effective shear stiffness (GAef f )

of the beam can be found using Eqn 5.11 and the displacement at the top of the story,

u(x = h) = ∆top . The effective shear stiffness factor, αs is calculated as the effective shear

stiffness over the gross shear stiffness, GAg (Eqn. 5.12).

1 Mbot x2 Vbot x3 Vbot x


u(x) = [ − ]+ + θbot x + ∆bot (5.10)
EIef f 2 6 κAG
1 Vbot h
GAef f = (5.11)
κ ∆bot − ∆bot − θbot h − 1 [ Mbot h2 − Vbot h3 ]
EIef f 2 6
GAef f
αshear = (5.12)
GAg
257

5.2.2.1 Calculated Values

Shear and flexural stiffnesses were calculated for each floor of the wall (krypton was not

used to calculate third floor stiffness). The displacements used at the top and bottom of

the first and second floors were calculated using the krypton targets (see Section 5.1.1 for

explanation). Rotations calculated from both the krypton and traditional systems were

used to calculate the stiffness values. The effective stiffness for the first floor was calculated

twice, once assuming the base of the wall was fully fixed (lateral displacement and rotation

equal to zero) and once using the displacement and rotation calculated at two inches above

the wall-foundation interface.

In general, the stiffness values calculated using the krypton rotation values varied less

between the ER+ and WL- peaks than did the values calculated using rotation values from

traditional instrumentation. There was no consistent difference between values calculated

using the two instrumentation system; for some walls/floors, the krypton produced larger

values, for others the krypton produced similar or smaller values. It was decided to use the

values calculated using krypton due to the small variation between loading directions.

The change in value between peaks for each loading direction were also considered. For

loading in the ER+ direction (initial direction of loading), the effective flexural stiffness

typically decreased for each peak at a specified drift demand. For loading in the WL-

direction (second direction of loading), the effective flexural stiffness typically remained

unchanged for each peak at a specified drift demand. A similar observation holds true for

the effective shear stiffness. When the peaks for each cycle are averaged, there was little

change in the values for each drift level, and when there was, the variation was less than

0.1 at low drift demands when the stiffness is changing rapidly. As a result of this analysis,

it was concluded that averaging all values at peaks for a drift demand was a reasonable

approach when presenting effective stiffness values in Chapter 6.


258

5.2.3 Flexural Stiffness: Euler-Bernoulli Beam Theory

Although the data collected allowed calculation of shear and flexural stiffness of the wall

specimens using the Timoshenko beam theory (see Section 5.2.2), the ability to do so is

typically not found in the literature for other experimental test programs on structural walls,

thus, it was desired to provide a calculation of wall stiffness that would allow comparison to

values from other test programs. To provide an equal comparison of the stiffness to other

test programs, the displacement (∆ef f ) at the effective height (hef f ) of loading was used

(see Section 5.3). The effective flexural stiffness, αf lex , was calculated as:

Vb
ksecant = (5.13)
∆ef f
3Ec Ig
kgross = (5.14)
h2ef f
ksecant
αf lex = (5.15)
kgross

Calculation of the flexural stiffness in this manner based on classical beam theory and

thus does not consider shear deformations. As expected, the effective flexural stiffness values

calculated using the secant method are lower than those calculated using the Timoshenko

beam theory as they must also account for the shear deformation of the wall.

5.3 Drift Estimation

The four planar wall specimens represented the bottom three stories of a ten-story prototype

wall, therefore the specimen drifts are representative of the drifts at the third floor. To

compare the drift capacities to the drift capacities of other experimental test programs, it

was necessary to estimate the displacement at the effective height of loading (∆ef f ). To

evaluate the drift capacity with respect to typical design drifts used by engineers, it was

necessary to estimate the equivalent ten-story (or roof) displacement (∆10 ).

Estimation of the displacements in the theoretical upper seven stories of the wall spec-
259

imens was done using the Timoshenko beam theory. Stiffness values used were selected

based on the preliminary calculations of the flexural and shear stiffnesses as described in

Section 5.2.2; multiple stiffness values were considered to evaluate the impact on the drift es-

timations. Flexural stiffness of 0.3Ec Ig and 0.5Ec Ig were used. Shear stiffnesses of stiffnesses

of 0.1GAg and 0.2GAg were used. The displacement at each floor level where estimated

using Equation 5.10, where the forces at the top and bottom of each floor were determined

from the appropriate lateral load distribution. To simplify the calculation of the displace-

ments, the effective displacement for the specimen with the ASCE 7 lateral load distribution

(PW1) was taken at the top of the 7th floor and the effective displacement for the specimens

with the uniform lateral load distribution (PW2, PW3, and PW4) was taken at the top of

the 5th floor.

0.30Ec Ig , 0.10GAg
0.50Ec Ig , 0.20GAg
1.25
1
0.75
0.5
0.25
Mb , k-ft

0
−0.25
−0.5
−0.75
−1
−1.25
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2
10th Floor Drift, %

Figure 5.2: Specimen PW2 comparision of 10th floor drifts calculated using combinations
of shear and flexure stiffness estimated from experimental data.

Figure 5.2 shows the Specimen PW2 envelope of base moment (Mb ) versus the estimated

10th floor drift using two different combinations of flexural and shear stiffness. The difference
260

between the two estimations is minimal, with the largest difference at the maximum drift

being approximately 0.10% drift. In presenting estimations of upper story displacements

or drifts, flexural stiffness equal to 0.50Ec Ig and shear stiffness equal to 0.20GAg as it was

assumed that the stiffness of the upper stories is likely the stiffer of the values estimated

from the lower stories.

Figure 5.3 shows, for each specimen in the ER+ loading direction, the envelopes for the

specimen, effective, and roof drifts. For all specimens, the roof drift is the largest of the

three and the effective drifts are greater than the specimen drifts. The differences between

the specimen drift and the drifts in the theoretical upper stories are not considered in the

presentation of data analysis results (Chapter 6 but rather are used for comparison for to

data from other experimental test programs.

5.4 Yield

Strain gauges were provided on longitudinal and horizontal reinforcement in the wall speci-

mens (see Section 3.6). The measured strains can be used to indicate when certain thresh-

olds, such as yield strain (εy ), were reached. Additionally, the strains can be given as input

to a material model to estimate steel stresses. For analysis of the planar experimental data,

the measured strain data and the calculated cyclic stress-strain response are used to quan-

tify the extent of yield. Section 5.4.1 discusses the estimation of steel stresses. Section 5.4.2

discusses the types of yield defined for analysis of the experimental data. Section 5.4.3

provides examples of “yield maps” that are used to visualize the extent of yielding in the

walls.

5.4.1 Stress Calculation

Hysteretic stress-strain response of the steel was determined using the Hoehler-Stanton

cyclic material model for reinforcing steel [45]. The model was developed to provide an easy

to implement, nonlinear, cyclic material model for reinforcing steel that is comparable in

accuracy to more complex material models.


261

1.25 1.25

1 1

0.75 0.75
Mb /Mn

Mb /Mn
0.5 0.5

0.25 0.25
Specimen Specimen
Effective Effective
Roof Roof
0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Drift, % Drift, %
(a) PW1 (b) PW2

1.25 1.25

1 1

0.75 0.75
Mb /Mn

Mb /Mn

0.5 0.5

0.25 0.25
Specimen Specimen
Effective Effective
Roof Roof
0 0
0 0.5 1 1.5 2 0 0.5 1 1.5 2
Drift, % Drift, %
(c) PW3 (d) PW4

Figure 5.3: ER+ loading direction envelopes for the specimen, effective, and roof drifts.
262

The Hoehler-Stanton model uses the Raynor [90] equations to define a monotonic enve-

lope for reinforcing steel; this allows simple definition of a nonlinear envelope with a yield

plateau and strain hardening branch. Figure 5.4a shows the envelope, where E is the elastic

modulus, σy and εy are the stress and strain at the yield, σsh and εsh are the stress and

strain at the onset of strain hardening, σu and εu are the stress and strain at the peak

stress, and C1 is a parameter defining the strain hardening curve curvature. The cyclic

stress response of the steel is determined using branch curves that define the behavior of

steel between load reversals. A typical branch curve is shown in Figure 5.4b. The curvature

of the branch curve, R, is defined as a function of the variables R0 , a1 , and a2 . Hoehler

and Stanton verified the accuracy of the model using data from cyclic stress-strain found in

the literature; the model accurately captured the response of a variety of steel specifications

and loading conditions.

A Matlab implementation of the Hoehler-Stanton model [45] was used to determine the

cyclic stress response from the steel strain gauge readings. The model stress and strain

parameters used are provided in Table 3.3. The parameter C1 = 3 was selected as it

approximately reproduced the curvature of the strain hardening branch of the measured

stress-strain response of the reinforcing steel. The other parameters defined for the model

were based on the suggested model input parameters provided by Hoehler and Stanton [45]:

R0 = 25, a1 = 23, and a2 = 0.46. The stress values computed for each strain gauge are

available as a part of the complete set of experimental data found on the NEES Project

Warehouse (http://nees.org/warehouse/project/104).

5.4.2 Definitions of Yield

In considering the yield behavior of the wall, it is desirable to know i) when and where yield

was measured, ii) the extent of measured yield, and iii) the extent of plastic behavior of

the wall. Distinguishing between yield in tension and compression provides a more refined

analysis. To provide a detailed overview of the yield behavior in the wall specimens, the
263

(a) Raynor [90] envelope. (b) Typical branch curve.

Figure 5.4: Hoehler-Stanton cyclic material model [45]

measured strains and the calculated hysteretic stress-strain response were both considered.

Examination of the data indicated that yield was best defined using the following types of

yield:

1. Yield (or initial yield): Strain gauge measurement reaches or exceeds the yield

strain, εy , in either tension or compression. Figure 5.5a shows an example stress-

strain response of a gauge that indicates initial yield.

2. Strain hardening: Strain gauge measurement reaches or exceeds the strain harden-

ing strain, εsh , in either tension or compression. Evaluation of the cyclic stress-strain

response indicated that gauges typically did not have inelastic, or plastic, behavior

until the strains reached the end of the strain hardening plateau. Figure 5.5b shows

an example stress-strain response of a gauge that indicates strain hardening.

3. Stress yield: The cyclic stress-strain response indicated that a gauge reached the

tension or compression yield stress, fy , without reaching the yield strain, εy , of the

same magnitude. Theoretically, this could occur in either tension or compression,

however, it was only observed in compression. This behavior is a result of significant


264

1 1.5

1
0.5
0.5
f/fy

f/fy
0
0
−0.5

−0.5 −1
−0.05 0 0.05 0.1 0.15 0.2 0.25 0.3 −0.5 0 0.5 1 1.5 2 2.5 3 3.5 4
ε/εsh ε/εy
(a) Initial Yield (b) Strain Hardening

1
f/fy

−1

−1 0 1 2 3 4 5 6
ε/εy
(c) Stress Yield

Figure 5.5: Example cyclic stress-strain response for the three definitions of yield. Measured
strains, ε, are normalized by the yield strain of the material, εy . Calculated stresses, f , are
normalized by the yield stress of the material, fy .

cyclic response. Figure 5.5c shows an example stress-strain response of a gauge that

indicates stress yield.

5.4.3 Yield Maps

When consideration is made for the difference between tension and compression yield, the

three definitions of yield presented in Section 5.4.2 result in five different yield indicators. If

an additional distinction is made to indicate if a gauge yields first in tension or compression,

the total number of yield indicators is nine. With each specimen having approximately 100
265

or more strain gauges, the amount of information available is vast. To understand the

extent and progression of yield throughout the tests, a set of “yield maps” were created to

qualitatively evaluate this data.

For each specimen, a pair of yield maps were made for each of a number of specified

steps throughout the test (e.g. peak drifts and/or key damage states) to show the extent of

yielding in the longitudinal reinforcement. The outline of the wall and the first and second

floor levels are shown by solid black lines. The top of the splice in the boundary element and

web are indicated by dashed lines. The wall cap and foundation block are not shown; the

gauges appearing below the wall outline are located in the foundation block. The locations

of gauges are shown by dots. One map indicates the gauges that yield in tension (+εy and

+εsh ) and another indicates the gauges that yield in compression (−εy , −εsh , and −fy ).

On each map, the type of yield is indicated by the marker type; an open circle indicates

initial yield (±εy ), a filled circle indicates strain hardening (±εsh ), and a cross indicates

stress yield (−fy ). The color of the marker indicates if the yield strain εy was reached first

in tension or compression; red (appears light gray when printed in grayscale) indicates that

tensile yield occurred first and blue (appears dark gray when printed in grayscale) indicates

that compressive yield occurred first. Figure 5.6 provides a set of example maps from the

reference wall (Specimen PW2) at the end of the test. In the specimens, with splices,

there are instances where two gauges are provided at the same location. When necessary,

additional sets of yield maps showing the gauges on only the foundation or wall bars were

created. Yield maps were also created to indicate yield measured by the strain gauges on

horizontal reinforcing bars, however, these show only tension yield.

5.5 Strain Fields

Strain fields are an efficient way to visualize various types of demands in the wall specimens

and to compare these demands to numerical finite element models. The dense grid of Kryp-

ton/Metris targets in the first two floors of the wall specimens were provided primarily for

this purpose. Two approaches were used to calculate strains from the displacments measured
266

(a) Tension yield. (b) Compression yield.

Gauge location Gauge location


Initial yield (compression first)
Initial yield (tension first)
Strain hardening (compression first)
Strain hardening (tension first) Initial yield (tension first)
Initial yield (compression first) Strain hardening (tension first)
Strain hardening (compression first) Yield Stress (no yield strain)
(c) Tension yield legend. (d) Compression yield legend.

Figure 5.6: Sample yield maps indicating yield of longitudinal reinforcement for Specimen
PW2 at the end of the test. Red and blue markers indicate if tension or compression yield
occurred first, respectively. Open circles indicate yield strain was reached. Solid circles
indicate strain hardening was reached. ‘Plus’ signs indicate yield stress, but not yield
strain, was reached in compression.
267

by the krypton system. The first, discussed in Section 5.5.1, treats the displacement data

as nodal displacements for a finite-element mesh. The second, discussed in Section 5.5.2,

calculates strains from the change in geometry of a quadrilateral area; this method is neces-

sary i) when F.E.M. calculations cannot be performed due to missing displacement values,

and ii) for calculations of strains using the linear potentiometers that measured relative

displacements (these are used in calculation of deformation contributions, see Section 5.6).

5.5.1 Calculation Using Finite Element Method

The krypton system measured three-dimensional displacements of a grid of points in the

lower two floors of the wall specimens. The grid formed by the targets creates regions (or

elements) that are approximately square with an LED target in each corner. The in-plane

displacements of the corners (or nodes) were used to calculate strains for the element using

a four-node isoparametric quadrilateral finite element formulation. Details of the element

formulation can be found in any finite element textbook (e.g. [22]) and are not presented

here. Strains in an element were determined at the four corners and at the center of the

elements. The corner strains were used in plotting strain fields and the strains at the center

of the elements were used in calculating deformation components (see Section 5.6). At each

evaluation point, the 4-node quadrilateral element provide three strain values, i) horizontal

strain, εx , ii) vertical strain, εz , and iii) shear strain, γxz . These values can be used to

calculate the principal strains, ε1 and ε2 , and the maximum shear strain, γmax :
s 
εx + εz εx − εz 2
ε1,2 = ± + (γxz )2 (5.16)
2 2
s 
εx − εz 2
γmax = + (γxz )2 (5.17)
2

5.5.1.1 Smoothed Strain Fields

Figure 5.7a shows a sample vertical strain field for Specimen PW2. The distribution of

the strain is very coarse and is discontinuous between adjacent elements. At each node, as
268

many as four different values are available for each strain quantity. To improve the quality

of information provided, the strain fields are smoothed at the nodes (Krypton targets). To

smooth the strains, a weighted average of all values at a node was calculated based on the

tributary area of each element. The smoothed version of the strain field in Figure 5.7a is

shown in Figure 5.7b. The smoothed strain data is archived in the project warehouse (see the

project summary document [63] for details). Smoothed strain fields at select events in the

specimen tests are presented in Chapter 6. Animations showing the strain field variations

during the tests are available on the project website at https://nees.org/groups/nees_

2005_0104/wiki/StrainFields.

(a) Unsmoothed (b) Smoothed

Figure 5.7: Specimen PW2 vertical strain fields at the first ER+ peak at 0.75% drift.

5.5.2 Calculation Using Geometry

Calculation of strains using the 4-node quadrilateral elements were not always possible.

One such instance was when only three nodal displacements were available for a rectangu-

lar region. Another was when using the relative displacement measurements from linear

potentiometers attached to the backsides of the specimens. In these cases, strains were

calculated using the change in geometry of a quadrilateral region.


269

Figure 5.8 shows the notation used for determining the strain from region geometry.

When using the linear potentiometer data, the deformed lengths of the sides and diagonals

were the measured values. When using the Krypton data, the displacements were used to

calculate the change in length of the i) horizontal sides (H1 and H2), ii) vertical sides (V 1

and V 2), and iii) diagonals (D1 and D2).

D H1 C

D2

V1 V2

D1

A H2 B

Figure 5.8: Notation usedH1for Krypton elements. H1


α4 α3
D2
Vertical strains (εz ) were calculated D1 of the
V1 on both the left and right sides V2element
H 2  V 2  D2
 o )by
by dividing the change in length of the side (∆V
1
costhe original length of the side (V ).
2 HV
( H way
Horizontal strains (εz ) were calculated in the1same H for
) 2 the
(V top
Vand
) 2 bottom
( D  Dsides
) 2 of the
  cos
elements using the variables ∆H and H. 2( H  H )(V  V )
   
i were calculated for each corner (i = oA, B, C, D) of the element by
Shear strains γxy
V1 D2
determining the change in angle of the corner: D1 V2

α
i 1
γxy = αio − αi α2 (5.18)
H2 H2

where αio is the original angle and αi is the angle at any given step. The original and

deformed angles were calculated (see Equations 5.19) using the law of cosines and the

known lengths of the sides of the triangles shown in Figure 5.9, where H, V , D and ∆H,

∆V , ∆D are the original and change in length of the horizontal, vertical and diagonal sides

of the element. The shear strains at the corners were averaged to establish a mean shear
270

mean , to be used in calculating the shear deformations of the walls


strain for the element, γxy
D
(see Section 5.6.2). H1 C

D2

−1 H 2 + V 2 − D2
V1αio = cos 2HV V2
(5.19a)
+ ∆H)2 + (V + ∆V )2 − (D + ∆D)2
(H D1
αi = cos−1 (5.19b)
2(H + ∆H)(V + ∆V )

A H2 B

H1 H1

α3 α4
V1
D1 D1 D2 D2 V2
V2 V1
α1 α2
H2 H2

Figure 5.9: Notation used for calculating element shear strains. A shear strain value is
calculated for each corner of the element.

5.6 Deformation Calculations

Data collected from the dense instrumentation available on the wall enabled the calculation

of deformation components along the height of the wall. The deformation components

consisted were i) flexural deformation, ii) shear deformation, and iii) fixed-end (or base)

deformations. Calculation of these components provides an understanding of how much the

wall is deforming due to flexure and shear, as well as the distribution of these components

along the length of the wall. In calculating the deformation components, the work done by

previous researchers was considered.

Sections 5.6.1 and 5.6.2 discuss the calculation of flexural deformations and shear defor-

mations, respectively. Section 5.6.3 discusses the calculation of the deformation components
271

at the fixed-end of the wall. Section 5.6.4 discusses how the deformation components were

combined for comparison to the total measured deformation of the wall specimens; this

includes a discussion of which instrumentation systems were used and explanations of why.

The deformation contributions for the wall specimens are presented and discussed in Sec-

tion 6.4; additional figures are provided in Appendix B.

5.6.1 Flexural Deformations

The flexural deformations of the walls was determined from the calculated rotation val-

ues. The wall was broken up into layers between each calculated rotation from either the

traditional or krypton measurement systems. The change in rotation between the bottom

and top of each layer was used to determine the amount of flexural deformation resulting

from the layer. The center of rotation was assumed to act at the center of the layer; this is

equivalent to the factor α = 0.50 commonly used in determining flexural deformations. The

deformation for the layer was calculated as the change in rotation of the layer multiplied by

the height of the wall above the center of the layer.

5.6.2 Shear Deformations

Shear deformations were calculated using the shear strains calculated using the Krypton

LED targets and the linear potentiometers (see Section 5.5). For each region for which

shear strains were available, the shear deformation of the region was calculated as:

dshear = tan(γregion )hregion (5.20)

where hregion and γregion are the average height and shear strain for the region. When the

shear strain for a region was calculated using the finite element method (see Section 5.5.1),

the strain at the center of the region was used for γregion . When the shear strain for a region

was calculated using the region geometry (see Section 5.5.2), the average of the strains in

the four corners of the region was used for γregion .


272

The dense grid of Krypton LED targets provided the opportunity to explore multiple

methods of determining the total shear deformation of the wall. These included i) calculation

of the shear strain for the region formed by the targets at the bounds of each floor, ii)

calculation of the shear strain for the regions formed by the left and right most targets in

two adjacent rows, adding the total deformations of each layer of element to get total floor

shear deformation and iii) calculation of of the shear strain for the regions used in calculating

the strain field, averaging all regions in a row to get the shear deformation for that layer,

and then adding up the total deformation of each layer. Ultimately it was determined that

the most appropriate method to use was the third as it best accounted for the horizontal

and vertical distribution of the shear demands. In considering a row of regions, there is a

significant variation in shear strains, and thus shear deformations along the length of the

wall. This can be seen in the strain field plots for the walls (see Section 6.6.2). The shear

strain was typically much larger in the tension region of the wall. To calculate the shear

deformation for a layer of the wall, the deformations for the individual elements in the layer

were averaged. This is consistent with averaging the lateral displacement of each target,

as discussed in Section 5.1.1, and was found to produce reasonable estimates for the shear

deformation.

The linear potentiometers mounted on the back of the wall did not allow for the same

flexibility in calculating the shear deformations as did the Krypton LED targets. Thus, the

shear deformations using this set of instruments was available for each floor but could not be

broken down into layers within each floor. The shear deformations for each each floor was

taken as the average of the shear deformations for each region within the floor. For the 1st

floor, two regions were stacked vertically in the boundary elements; the shear deformations

from each pair of regions in each boundary element were added prior to averaging the

shear deformations along the length of the wall. For the 2nd floor, no shear deformations

were available in the boundary element and the shear deformations were averaged from the

deformations in the web.


273

5.6.3 Fixed-end Deformations

The fixed-end, or base, deformations were calculated at a height of 2.0 inches (5.08 cm)

above the wall-foundation interface, as this was the lowest elevation where instrumentation

was available. Two fixed-end deformations were calculated: base slip and base rotation.

The base slip was calculated as the lateral displacement of the bottom row of targets (see

Section 5.1.1). The base rotation was calculated as the rotation from either the bottom row

of Krypton targets or the two instruments measuring vertical displacement of the wall with

respect to the foundation block. To determine the total wall deformation resulting from the

base shear, the base rotation was assummed to act at the center of the region below the

measurement (1.0 inch or 2.54 cm).

5.6.4 Total Drift Contributions

The deformation components discussed in Sections 5.6.1 through 5.6.3 were calculated not

only to be presented independently, but more importantly to be presented as a percentage of

the total wall deformation so that the contribution of each mechanism could be understood.

In adding the individual contributions and comparing to the total measured deformation

(the control displacement of the wall), it was desired to have the most discretized data

possible, thus, the deformations calculated from the Krypton LEDs was used wherever

possible. However, this data was not available in the third floor of the specimen and did

not always provide reasonable estimates of deformations. The following list explains what

instrumentation was used for which calculation.

1. Fixed-end deformations

(a) Lateral deformation (or base slip): Krypton (traditional was not available)

(b) Base rotation: Krypton provided more measurements from which to determine

the rotation

2. Flexural deformations
274

(a) 1st floor: Krypton provided more measurements from which to determine the

rotation, as well as more locations at which to calculate the rotation.

(b) 2nd floor: Krypton provided more measurements from which to determine the

rotation, as well as more locations at which to calculate the rotation.

(c) 3rd floor: Traditional (Krypton was not available)

3. Shear deformations

(a) 1st floor: Krypton provided more shear deformations laterally to use in calculat-

ing the deformation of a layer and provided more layers at which to calculate the

shear deformation

(b) 2nd floor: Traditional instrument was used. For Specimen PW3, many targets in

the upper west (left) portion of the 2nd floor produced bad readings throughout

the test. This resulted in missing shear strain calculations for this region of the

wall, consequently resulting in over-estimation of the shear deformation when

this region was in compression and under-estimation of the shear demand when

this region was in tension. Thus, the traditional instrumentation was used for

all specimens to provided equal comparison.

i. 3rd floor: Traditional (Krypton was not available)

Typically the only the deformation components at the cycle peaks were considered.

Initially, the data for all cycles in the ER+ and WL- loading directions were considered and

evaluated using the methods described in Section 5.6. The plots evaluated are provided

in Appendix B. The following general observations were made regarding the results of the

deformation component calculations:

1. At drifts lower than 0.14% there was significant error (more than 50% is some cases)

and these contribution values were not considered valid. These errors are a result
275

of inaccuracy of the Krypton measurements at low displacements and division of the

deformation components by small total deformation values to calculate percentanges.

2. In general, the total specimen deformations were over-predicted by at much as 10%

when the deformation components were evaluated at the east peaks.

3. In general, the total specimen deformations were under-predicted by no more than

10% when the deformation components were evaluated at the west peaks.

4. The total first floor deformations were typically under-predicted for cycle peaks in

both the ER+ and WL- direction.

5. Typically, there was little variation between the deformation contributions at the first

and last peaks of a cycle (ER+ and WL- loading directions considered separately).

Based on the above observations, it was determined that the cleanest and most efficient

method for qualitatively presenting (see Section 6.4) the contribution of the various mecha-

nisms (shear, flexure, and fix-end deformation) to the total wall deformation was to average

all peaks within a cycle and present one data point per drift level for each specimen.

5.7 Conclusions

This chapter presented details on the methods used to analyze experimental data collected

from the four planar wall specimens discussed in Chapters 3 and 4. Section 5.1 discussed

calculation of displacement and rotation values that are used to calculate effective stiff-

ness values (Section 5.2), drift at effective height of loading (Section 5.3), and deformation

components (Section 5.6). Section 5.5 discusses the calculation of strain fields from the

Krypton non-contact measurement system. Results obtained from the data analysis meth-

ods discussed in this chapter are presented in Chapter 6.


276

Chapter 6

ANALYSIS OF EXPERIMENTAL DATA

This chapter presents the results of the analysis of experimental data collected for the

four planar wall specimens. Analysis of the data was done to i) provide an understanding

of the observed experimental behavior, ii) develop recommendations for use in wall design,

and iii) provide tools to validate numerical models of walls.

Section 6.1 provides a summary of the experimental results presented in Chapter 4, in-

cluding a summary of when key damage states occurred and a list of observations/hypotheses

to explore using the data analysis methods outline in Chapter 5.

The results of the data analysis are presented in Sections 6.2 - 6.7. In each section,

the results for the four walls are presented together followed by a discussion of i) general

trends, ii) impact of damage states, and iii) impact of study parameters. The data analysis

results are presented in order of investigation of global to local behavior. Section 6.2 presents

displacement and rotation profiles. Section 6.3 presents effective stiffness values. Section 6.4

presents deformation components. Section 6.5 presents yield patterns in the reinforcement.

Section 6.6 presents strain fields. Sections 6.7 and 6.8 summarize the strain and stress

demands, respectively. Section 6.9 provides an overall summary of the results of the data

analysis.

6.1 Summary of Experimental Observations

The planar wall test program sought investigate the impact of i) lateral load distribution

(and its impact on shear demand), ii) distribution of longitudinal reinforcement, and iii)

lap splices in the expected region of yield. Table 6.1 summarizes the damage seen in each

specimen at 0.50%, 0.75%, 1.0%, and 1.5% drift. Data analysis results are presented at
277

these drift levels in Sections 6.2 - 6.8 to consider how damage relates to the data analysis

quantities provided.

Table 6.1: Summary of Damage States

PW1 PW2 PW3 PW4

0.50% Drift Cracking Cracking Cracking Cracking


Spalling Spalling Spalling
0.75% Drift Cracking Cracking Cracking Cracking
Spalling Spalling Spalling Spalling
Exposed Rebar Exposed Rebar Exposed Rebar Exposed Rebar
Bar Buckling Bar Buckling
Core Crushing
0.75% Drift Cracking Cracking Cracking Cracking
Spalling Spalling Spalling Spalling
Exposed Rebar Exposed Rebar Exposed Rebar Exposed Rebar
Bar Buckling Bar Buckling Bar Buckling Bar Buckling
Web Crushing
Core Crushing Core Crushing
1.50% Drift Cracking Cracking Cracking -
Spalling Spalling Spalling -
Exposed Rebar Exposed Rebar Exposed Rebar -
Bar Buckling Bar Buckling Bar Buckling -
Web Crushing Web Crushing
Core Crushing Core Crushing Core Crushing -
Bar Fracture

Chapter 4 provides a summary of each wall test, including measured response, damage

behavior, and failure mode. Section 4.6 presents a list of the observed impact of the test

variables. From these observations, a set of hypotheses regarding behavior of the walls and

the impact of design parameters was made. These hypotheses are as follows:

• Lateral load distribution/shear demand:

1. Walls with higher shear demand (Specimens PW2, PW3 and PW4) had loss of
278

load carrying capacity due to shear-compressive failure of the boundary elements

at the critical section, while the wall with lower shear demand (Specimen PW1)

had loss of load carrying capacity due to bar fracture at the base of the wall.

2. The wall with lower shear demand had greater tensile strain demands at the

wall-foundation.

3. Increased shear demand results in a decrease in wall drift capacity.

• Longitudinal reinforcement layout:

1. Uniform distribution of longitudinal reinforcement resulted in web crushing.

Crushing of concrete in the web of the wall was not observed in walls with

concentrated boundary element steel. not seen in walls with boundary element

distribution of steel, suggesting high shear demands in the web of the wall.

• Lap splice:

1. The critical section of the wall (that which controls the response of the wall)

is located at the base of the wall for continuous reinforcement (Specimen PW4)

and at the top of the splice for walls with spliced reinforcement (Specimens PW1,

PW2, and PW3).

2. Spliced regions of walls do not have the same flexural response as unspliced

regions, as indicated by the crack patterns of the walls.

3. A wall without a splice has a lower drift capacity than a comparable wall with a

splice in the expected yield region.

6.2 Displacement and Rotation Profiles

Displacement and rotation profiles were determined primarily using the Krypton measure-

ment system; details are provided in Section 5.1. The displacement and rotation profiles at

cycle peaks are provided for each wall in Figures B.1 through B.8. The profiles presented

in this section provide an overview of the displaced shape of the wall specimens.

Figure 6.1 shows the wall displacement and rotation profiles at the first peaks to 0.50%
279

12
PW1
PW2
PW3
10 PW4

8
Height, feet

0
−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8
Displacement, in.
(a) Displacement

12
PW1
PW2
PW3
10 PW4

8
Height, feet

0
−8 −6 −4 −2 0 2 4 6 8
Rotation, rad. −3
x 10
(b) Rotation

Figure 6.1: Profiles at first ER+ and WL- peaks at 0.50% drift.
280

drift. At this drift level, walls had not yet reached the peak capacity but initial tensile

yielding had occurred. The displacement profiles shape and magnitude are essentially the

same.

The rotation profiles indicate a distinct difference in the first floor for the spliced spec-

imens (PW1, PW2, and PW3) versus the unspliced specimen (PW4). In Specimen PW4,

the rotation profile is linear. In the other specimens, the rotation at the base of the wall

was less relative to the wall without a splice

In the other specimens, the rotation at the base of the wall but there was a large increase

in the rotation demand at the top of the splice. The effect of the splice was more pronounced

in the ER+ loading direction.

Figure 6.2 shows the wall displacement and rotation profiles at the first peaks at 1.0%

drift. At this drift level, none of the walls had reached loss of lateral load carrying capacity.

The displacements are similar, however, the displaements in the spliced region for Specimen

PW3 were lower than in the other walls. The rotation profile differed primarily in the

first floor, with the rotation at the base of the wall greater for the wall without a splice

(PW4) than in the walls with a splice (PW1, PW2, and PW3). Amongst the spliced

specimens, the rotation profiles vary: Specimen PW3 (uniform distribution of longitudinal

reinforcement) had less rotation in the spliced region than did Specimens PW1 and PW2

(boundary element distribution of longitudinal reinforcement). The wall rotations were

largest in Specimen PW3 above the splice.

Residual displacement and rotation profiles at 0.50% and 1.0% drift are shown in Fig-

ures 6.1 and 6.2, respectively. Figure 6.3 shows the residual profiles following the first ER+

peaks to 1.0% drift; again, the impact of the splice is seen. The splice in the bottom half of

the first floor resulted in slightly less lateral displacement and significantly smaller rotations.

The observations that the rotations are larger above the splice region in Specimens PW1,

PW2, and PW3 but above the base in Specimens PW4 indicates that the nonlinearity

concentrates above the splice. This shifts the critical section for the spliced walls from the
281

12
PW1
PW2
PW3
10 PW4

8
Height, feet

0
−1.5 −1 −0.5 0 0.5 1 1.5
Displacement, in.
(a) Displacement profiles.

12
PW1
PW2
PW3
10 PW4

8
Height, feet

0
−0.015 −0.01 −0.005 0 0.005 0.01 0.015
Rotation, rad.
(b) Rotation profiles.

Figure 6.2: Profiles at first ER+ and WL- peaks at 1.0%.


282

12 12

10 10

8 8

Height, feet
Height, feet

6 6

4 4

PW1 2 PW1
2 PW2
PW2
PW3 PW3
PW4 PW4
0 0
0 0.2 0.4 0.6 0 2 4 6
Displacement, in. Rotation, rad. −3
x 10
(a) Displacement profiles. (b) Rotation profiles.

Figure 6.3: Profiles of residual displacements and rotations (approximately zero lateral load)
following the first ER+ peak at 1.0% drift.

base of the wall to above the splice.

6.3 Stiffness

Stiffness values are needed to conduct elastic analysis of walls to determine load and dis-

placement demands. Effective stiffness values for the four wall specimens were calculated

using the applied loads and calculated deformations. Section 6.3.1 presents axial stiffness

values. Section 6.3.2 presents flexural and shear stiffness values.

6.3.1 Axial Stiffness

Axial stiffness values of each wall specimen were calculated at the steps for which the lateral
measure , were calculated
force was approximately zero. The axial effective stiffness values, kaxial

using the applied axial load (measured) and the average vertical displacement (measured) of

the specimen; details are provided in Section 5.2.1. The measured effective stiffness values
283

gross
were normalized by the gross axial stiffness, kaxial , to determine an axial stiffness factor

αaxial . Figure 6.4 shows the axial stiffness parameter for each specimen versus the maximum

historic drift (MHD). Only data for cycles at 0.50% drift or less are shown; at larger drifts,

the calculated stiffness values were deemed unreliable (see Section 5.2.1).

The initial axial stiffness (calculated after application of the axial load and prior to the
gross
application of the cyclic displacement history) of the specimens varied from 0.60−0.75kaxial .
gross
At drifts less than 0.10%, the stiffness ranged from 0.45 − 0.65kaxial .

For Specimen PW1, the stiffness factor was similar for drifts exceeding 0.05%. The

last reliable factor was αaxial = 0.46 during the 0.50% drift cycles. In Specimens PW2

and PW4 (both had slightly higher axial load ratios), the stiffness factor decreased as the

maximum historic drift decreased; the axial stiffness factors at 0.50% drift were 0.36 and

0.41, respectively. The calculated values for Specimen PW3 are a bit more erratic, with the

lowest calculated value of αaxial = 0.52.

1.0

0.8

0.6
αaxial

0.4

0.2 PW1
PW2
PW3
PW4
0
0 0.1 0.2 0.3 0.4 0.5 0.6
Drift, %

Figure 6.4: Calculated axial stiffness at zero lateral force for cycles to 0.50% drift or less
(minimum stiffness for each displacement level is shown).
284

6.3.2 Flexural and Shear Stiffness

The measured lateral displacements and in-plane rotations at the floor levels were used to

calculate the effective flexural stiffness and effective shear stiffness values using the Timo-

shenko beam theory. Details of the calculations and the data used are found in Section 5.2.2.

The average stiffness value at each drift level are presented here for i) the first floor of

the wall assumming the base of the wall is fixed, ii) the first floor of the wall using the

measured rotation and displacement values at the base of the wall, and iii) the second floor

of the wall1 . The flexural stiffness factor (αf lex ) is the ratio of the measured effective flexural

stiffness (EIef f ) to the gross flexural stiffness (EIg ). The shear stiffness factor (αshear ) is

the ratio of measured effective shear stiffness (κGAef f ) to the gross shear stiffness (κGAg ),

where κ is the shear coefficient and is equal to 5/6 for a rectangular cross-section.

EIef f
αf lex = (6.1)
EIg
κGAef f
αshear = (6.2)
κGAg

Figure 6.5 presents the calculated flexural and shear stiffness factors for the first floor

assuming the base of the wall is fixed (zero displacement and zero rotation). This calculation

includes all base and first story deformation. The trend of the flexural stiffness related to

drift is similar for the four specimens. The shear stiffness factors vary slightly more, but are

also similar. This is consistent with the observations in Section 6.2 that indicate the defor-

mations at the top of the first floor are consistent between specimens but the deformations

within the first floor vary more. The variation within the floor is apparent in the stiffness

calculations for the first floor excluding the base deformations, shown in Figure 6.6. The

flexural stiffness was approximately the same as a function of drift for the specimens with

boundary element distribution of the longitudinal reinforcement (PW1, PW2, and PW4)

1
Additional figures showing the stiffness values for each specimen are presented in Appendix B
285

1.0
PW1
PW2
PW3
PW4
0.8

0.6
EIeff /EIg

0.4

0.2

0
0 0.5 1.0 1.5
Drift, %
(a) Flexural stiffness

1.0
PW1
PW2
PW3
PW4
0.8

0.6
AGeff /AGg

0.4

0.2

0
0 0.5 1.0 1.5
Drift, %
(b) Shear stiffness

Figure 6.5: Calculated effective stiffness values for the first floor assuming the base of the
wall is fixed.
286

1.0
PW1
PW2
PW3
PW4
0.8

0.6
EIeff /EIg

0.4

0.2

0
0 0.5 1.0 1.5
Drift, %
(a) Flexural stiffness

1.0
PW1
PW2
PW3
PW4
0.8

0.6
AGeff /AGg

0.4

0.2

0
0 0.5 1.0 1.5
Drift, %
(b) Shear stiffness

Figure 6.6: Calculated effective stiffness values for the first floor.
287

but was slightly lower for the wall with uniform distribution of reinforcement (PW3). The

variation in the shear stiffness factor was much greater. The largest values were for Spec-

imen PW3, apparently accounting for the lower flexural stiffness factor. The lowest shear

stiffness values were calculated for Specimen PW1, which had the lowest shear demand of

the four walls.

Figure 6.7 shows the flexural and shear stiffness factors calculated for the second floor

of the specimens. The flexural and shear stiffness are essentially the same between the four

specimens after approximately 0.25% drift. One noticeable exception to this is the flexural

stiffness for Specimen PW3 at drifts in excess of 0.50%, however, this is likely the result of

missing data points used in determining rotation values rather than the result of a different

physical mechanism.

6.3.3 Secant Stiffness at Effective Height

The flexural and shear stiffness factors presented in Section 6.3.2 were used to guide the

selection of the stiffness values used to estimate the drift at the effective height of loading and

roof (10th floor); details on the selection of values are provided in Section 5.2.3 and details

on the calculation of the displacement at the effective height are provided in Section 5.3.

Figure 6.8 presents the secant stiffness values as a function of the estimated drift.

6.4 Deformation Components

The displacement profiles presented in Section 6.2 are similar, however, the differed. To in-

vestigate this further, the contributions of shear, flexural, and fixed-end (or base) deforma-

tions to the total specimen drift were calculated using the methods described in Section 5.6.

These deformation components are discussed in this section; additional figures are provided

in Appendix B.

Figure 6.9 shows the contribution of the deformation mechanisms (flexure, shear, and

fixed-end deformations) to the total deformation of the wall. Each mechanism is represented

by a different color; a different shade of each color indicates the different floors. In the first
288

1.0
PW1
PW2
PW3
PW4
0.8

0.6
EIeff /EIg

0.4

0.2

0
0 0.5 1.0 1.5
Drift, %
(a) Flexural stiffness

1.0
PW1
PW2
PW3
PW4
0.8

0.6
AGeff /AGg

0.4

0.2

0
0 0.5 1.0 1.5
Drift, %
(b) Shear stiffness

Figure 6.7: Calculated effective stiffness values for the second floor.
289

1.0 1.0
PW1 PW1
PW2 PW2
PW3 PW3
PW4 PW4
0.8 0.8

0.6 0.6
αflex

αflex
0.4 0.4

0.2 0.2

0 0
0 0.5 1.0 1.5 0 0.5 1.0 1.5
Effective Drift, % Roof Drift, %
(a) (b)

Figure 6.8: Secant stiffness at the effective height of loading and roof. Stiffnesses were
calculated using estimated displacement values.

floor, a line separates the lower half (spliced region in spliced specimens) and upper half of

the floor.

For each drift level, the average values of all peaks is shown (Appendix B provides

additional figures that show the deformation components at each individual peak in the

ER+ and WL- directions). Details on the calculation of drift contributions can be found

in Section 5.6. Due to error in the calculation of deformation contributions at low drift

levels, only the contributions post-yield (yield is indicated by a solid vertical line in each

plot) are discussed here. For all specimens, Figure 6.9 indicates that following yield, there

is generally a small and gradual decrease in the deformation components in the second and

third floors, and a change in the distribution of the deformation components in the first

floor. Figure 6.10 shows the contributions of deformation mechanisms in the first floor,
290

relative to the total deformation in the first floor. In the following list, the contributions are

evaluated for each test variable by comparison of the reference specimen (Specimen PW2)

to the other specimens.

• Impact of shear demand (PW1 vs PW2)

– Specimen PW1 had a lower shear demand yet the deformation contributions

were similar to Specimen PW2. The amount of deformation from flexure was

only slightly higher for Specimen PW1.

– The primary difference between Specimens PW1 and PW2 is in the amount of

total wall deformation resulting from base rotation. In Specimen PW1, the base

rotation contribution was approximately the same from 0.50% to 1.0%, while it

decreased gradually in Specimen PW2. The largest difference, however, came in

the 1.5% drift cycles. Figure 6.10 shows i) for PW1 the base rotation contribution

increased while the flexural contribution in the splice and the shear deformations

in the first floor decreased, and ii) for PW2 the base rotation (and flexure in the

splice region) decreased while the flexure above the splice increased. For both

specimens the change in base rotation contributions occurred in cycles in the WL-

direction (loading direction associated with failure); drift contributions remained

largely unchanged in the ER+ loading direction (see Figures B.11 and B.12).

These observations are consistent with the observed test results that indicated

similar damage to the two specimens for a large portion of the test but failure

of the specimens occurred in different locations (at the base for PW1 and above

the splice in PW2).

• Impact of reinforcement distribution (PW2 vs PW3)

– At yield, the distribution of the deformations was essentially the same, but at

drifts beyond yield the distributions were different; the shift in mechanisms was

the same but rate at which the shift occurred differed. In Specimen PW2, the

distribution is fairly constant up to 0.75% drift, at which point there was a


291

120 120

100 100
% Total Deformation

% Total Deformation
80 80

60 60

40 40

20 20

0 0
0 0.5 1 1.5 0 0.5 1 1.5 2
% Drift % Drift
(a) PW1 (b) PW2

3rd story flexure 3rd story shear


2nd story flexure 2nd story shear
1st story flexure 1st story shear
Base slip Base rotation
(c)

120 120

100 100
% Total Deformation

% Total Deformation

80 80

60 60

40 40

20 20

0 0
0 0.5 1 1.5 0 0.5 1 1.5
% Drift % Drift
(d) PW3 (e) PW4

Figure 6.9: Average deformation components for the full wall specimens.
292

120 120

100 100
% Total Deformation

% Total Deformation
80 80

60 60

40 40

20 20

0 0
0 0.5 1 1.5 0 0.5 1 1.5 2
% Drift % Drift
(a) PW1 (b) PW2

3rd story flexure 3rd story shear


2nd story flexure 2nd story shear
1st story flexure 1st story shear
Base slip Base rotation
(c)

120 120

100 100
% Total Deformation

% Total Deformation

80 80

60 60

40 40

20 20

0 0
0 0.5 1 1.5 0 0.5 1 1.5
% Drift % Drift
(d) PW3 (e) PW4

Figure 6.10: Average deformation components in the first floor.


293

decrease in base rotation and an increase in flexure above the splice. In Specimen

PW3, the base rotation gradually decreases beyond yield while the flexure and

shear above the splice increased.

– The shear deformations increased in the splice in PW2 and above the splice in

PW3.

• Impact of lap splices (PW2 vs PW4)

– Specimen PW4 had a greater amount of base rotation2 than did Specimen PW2.

The relative contribution of this demand was fairly constant throughout the test,

whereas in Specimen PW2 the relative contribution of the base rotation decreased

as the specimen drift increased.

6.5 Yield

Yield of longitudinal and horizontal reinforcing bars was determined from strain gauges (see

Section 3.6.3). Three types of response were highlighted i) gauges indicated initial yield was

reached in tension or compression (±εy ), ii) gauges indicated strain hardening was reached

in tension or compression (±εsh )3 , and iii) the compressive stress of the steel (−fy ) was

reached due to the cyclic response of the steel (see Section 5.4.2). Section 6.5.1 discusses

yield of the longitudinal reinforcing bars. Section 6.5.2 discusses yield of the horizontal

reinforcing bars.

6.5.1 Yield of Longitudinal Reinforcement

Figure 6.11 and 6.12 show the pattern of tension and compression yield in longitudinal

reinforcement for the final four cycles of each test. Figure 6.13 and 6.14 show the pattern of

tension and compression yield in longitudinal reinforcement after completion of each test.

2
The calculation of the base rotation values accounts for both the rotation at the wall-foundation interface
and the flexural deformation in the lower 2.0 inches (5.08 cm) of the wall (see Section 5.6.1).
3
Strains that reached the strain hardening strain typically exhibited significantly greater plastic behavior
that those gauges that did not. See Section 5.4.2 for additional discussion.
294

For the spliced walls (Specimens PW1, PW2, and PW3), the gauges on both the foundation

and wall bars in the splice region are shown at the same locations and the gauges at two

inches above the interface are not shown4 . Figure 6.15 shows the tension and compression

yield of only the foundation bars.

Tensile yielding of the longitudinal reinforcement extends into the third floor for all

walls, while plastic yielding was concentrated to a region no greater than one-half the wall

length above the critical section of the wall (above the splice PW1-PW3, above the interface

in PW4). Compression yield was generally limited to the first floor except for PW3, where

compression yield extended into the second floor of the wall. The following sections discuss

the impact of the test program variables (shear demand, reinforcement distribution, and

spliced longitudinal reinforcement) on the pattern of reinforcement yield.

6.5.1.1 Impact of Shear Demand (PW1 vs PW2)

The impact of shear demand can be seen by comparing Specimens PW1 and PW2; both

walls were constructed using the same batch of reinforcing steel, thus, the only differences

in the yield patterns arise from the higher shear demand and higher axial load in Specimen

PW2.

The pattern of tensile behavior at the end of the test was very similar in the two walls.

Strain hardening behavior was concentrated above the spliced region; in Specimen PW1,

strain hardening extended into the second floor in the west (left) boundary element. As

indicated in Table 4.11, tensile yield was first measured at approximately 0.35% drift in

PW1 and approximately 0.43% drift in PW2; however, in PW1, few gauges indicated yield

following three cycles to 0.35% drift and after the first cycle to 0.50%, the amount of yield

in both walls was similar and extend to the top of the second floor. Strain hardening was

first measured in the boundary elements above the splice in the first cycle to 1.0% drift.

The extend of compression yield in the two walls was similar, but the type of yield was

4
These gauges were located near the end of the bars that terminated at the wall-foundation interface.
These gauges indicated that the ends of the bars carried compressive strains but very little tensile strains.
295

(a) PW1: (b) PW1: (c) PW1: (d) PW1:


0.5% 0.75% 1.0% 1.5%

(e) PW2: (f) PW2: (g) PW2: (h) PW2:


0.5% 0.75% 1.0% 1.5%

(i) PW3: (j) PW3: (k) PW3: (l) PW3:


0.5% 0.75% 1.0% 1.5%

(m) PW4: (n) PW4: (o) PW4: (p) PW4:


0.35% 0.50% 0.75% 1.0%

Gauge location
Initial yield (tension first)
Strain hardening (tension first)
Initial yield (compression first)
Strain hardening (compression first)
(q) Legend

Figure 6.11: Tensile yield in vertical reinforcement during final four drift levels.
296

(a) PW1: (b) PW1: (c) PW1: (d) PW1:


0.5% 0.75% 1.0% 1.5%

(e) PW2: (f) PW2: (g) PW2: (h) PW2:


0.5% 0.75% 1.0% 1.5%

(i) PW3: (j) PW3: (k) PW3: (l) PW3:


0.5% 0.75% 1.0% 1.5%

(m) PW4: (n) PW4: (o) PW4: (p) PW4:


0.35% 0.50% 0.75% 1.0%

Gauge location
Initial yield (compression first)
Strain hardening (compression first)
Initial yield (tension first)
Strain hardening (tension first)
Yield Stress (no yield strain)
(q) Legend

Figure 6.12: Compression yield in vertical reinforcement during final four drift levels.
297

(a) PW1 (b) PW2

(c) PW3 (d) PW4

Gauge location
Initial yield (tension first)
Strain hardening (tension first)
Initial yield (compression first)
Strain hardening (compression first)
(e) Legend

Figure 6.13: Gauges indicating tension yield and strain hardening in longitudinal bars at
end of test.
298

(a) PW1 (b) PW2

(c) PW3 (d) PW4

Gauge location
Initial yield (compression first)
Strain hardening (compression first)
Initial yield (tension first)
Strain hardening (tension first)
Yield Stress (no yield strain)
(e) Legend

Figure 6.14: Gauges indicating compression yield and strain hardening in longitudinal bars
at end of test.
299

Gauge location
Initial yield (tension first)
Strain hardening (tension first)
Initial yield (compression first)
Strain hardening (compression first)
(a) Legend Tension

(b) PW1 Tension (c) PW2 Tension (d) PW3 Tension

(e) PW1 Compression (f) PW2 Compression (g) PW3 Compression

Gauge location
Initial yield (compression first)
Strain hardening (compression first)
Initial yield (tension first)
Strain hardening (tension first)
Yield Stress (no yield strain)
(h) Legend Compression

Figure 6.15: Final yield and strain hardening pattern in foundation bars for spliced walls.
300

not. Above the top of the splice, compression yield occurred as a result of the yield stress,

rather than strain, having been reached. In Specimen PW2, compression yield occurred

prior to tension yield in the spliced regions of the wall. This is likely a result of the higher

axial load applied to the wall. As indicated in Table 4.11, compression yield occurred in

east (right) toe of PW2 at approximately 0.21% drift; compression yield in the west (left)

boundary element and in Specimen PW1 did not occur until the 0.75% drift cycles.

6.5.1.2 Impact of Reinforcement Distribution (PW2 vs PW3)

The impact of longitudinal reinforcement distribution can be seen by comparing Specimens

PW2 and PW3.

The pattern of tensile yield behavior of the walls was very similar in the spliced region

but significantly different in the unspliced portion of the walls. In PW2, strains beyond the

yield plateau were restricted to the first floor boundary elements. In PW3, strain hardening

effects extend from the top of the splice through the third floor. Initial tensile yield occurred

in PW3 at approximately 0.18% drift, and following the cycles to 0.35% drift yield was

measured as high as the bottom of the third floor; no tensile yield was yet measured in

Specimen PW2. At drifts when tensile yield was first measured in PW2, the first gauges

indicating strain hardening behavior were observed in PW3. The observed difference are a

combination of the higher axial load in Specimen PW2 and the lower yield (εy ) and strain

hardening (εsh ) strains for the steel used in Specimen PW3.

The final pattern of compression yield is similar for Specimens PW2 and PW3. In the

spliced regions, the compression yield typically occurred prior to tension yield. Above the

spliced regions, the tension yield typically occurred prior to compression yield. As with

tension yield, the compression yield extended into the second floor for PW3 but not for

PW2 and yield occurred at significantly lower drifts in PW3.


301

6.5.1.3 Impact of Spliced Reinforcement (PW2 vs PW3)

The impact of spliced longitudinal reinforcement can be seen by comparing Specimens PW2

and PW4.

The pattern of tensile yield behavior was significantly different as a result of the splice.

In PW4 (no splice) the region where strain hardening effects were measured was located

at the base of the wall. In PW2 (splice) the region where strain hardening effects were

measured was located at the top of the splice. The height of the strain hardening effects

was greater in Specimen PW4. The steel used in Specimen PW4 had a shorter yield plateau

for the #4 reinforcing bars than that for Specimen PW2.

The final pattern of compression yield is also different for PW2 and PW4. As with

the tension yield, the yield behavior is concentrated at the base in PW4 and at the top

of the splice in PW2. In PW4, the compression yield occurred before tension yield in the

east (right) boundary element, but tension yield occurred first in the west (left) boundary

element, where significant stress yield was measured. The pattern of compression yield

at the base of the west (left) boundary element in PW4 closely resembles the pattern of

compression yield above the splice in PW2.

6.5.2 Yield of Horizontal Reinforcement

Figure 6.16 indicates the pattern of tension yield in horizontal gauges following the final four

sets of cycles (drift levels) in each wall. It is noted that plastic yield of the reinforcement

occurred in few gauges and compression yield did not occur in any of the gauges.

The impact of shear demand can be seen by comparison of Specimens PW1 and PW2.

The amount of yielding is similar following failure of the walls, however, Figures 6.16a-c

and e-g indicate that yielding occurred at lower drift demands in PW2 than in PW1. The

difference is reasonable considering the increased shear demand in PW2.

The impact of longitudinal reinforcement distribution can be seen by comparison of

Specimens PW2 and PW3. In both specimens, yielding in the web occurs above the top of
302

the splice in the web; this is lower in Specimen PW2 due to the lower height of the splice.

In PW3, there was yield of just one gauge in the horizontal bars below the top of the splice,

giving the appearance that less overall yield of the horizontal reinforcement occurred.

The impact of spliced longitudinal reinforcement can be seen by comparison of Specimens

PW2 and PW4. The amount of horizontal yield was significantly less for Specimen PW4.

The few gauges that did indicate yield in PW4 occurred near the end of the test at 1.0%

(Figure 6.16p). Comparison at the same drift level in PW2 (Figure 6.16g) indicated less

yield of the horizontal steel at 1.0% drift.

6.6 Strain Fields

The collection of high-resolution data from the planar wall test program enabled calculation

of high-resolution strain fields that can be used to investigate response and to validate

numerical models. Strain fields were calculated for the lower two floors of the wall specimens

using the displacements measured by the Krypton system. Details for the calculation and

smoothing of the strain fields is presented in Section 5.5. Six types of strains were calculated:

i) vertical strain (εz ), ii) horizontal strain (εx ), iii) in-plane shear strain (γxz )5 , iv) first

principal strain (ε1 ), v) second principal strain (ε2 ), and vi) maximum in-plane shear strain

(γxz ). If the cyclic behavior of the walls are considered, the amount of strain data is vast

and cannot be presented in its entirety here, thus, animations showing the cyclic strain

fields are available at https://nees.org/groups/nees_2005_0104/wiki/StrainFields.

The jpeg images of the strain fields and text files containing the strain data are available in

the project warehouse [63].

Here a reduced set of the strain field images (at cycle peaks) presented. The selected

strain fields are used to illustrate the general strain behavior of the specimens and demon-

strate the impact of the test variables (shear demand, reinforcement distribution, and spliced

longitudinal reinforcement). Section 6.6.1 presents vertical strain. Section 6.6.2 presents

5
The absolute values of these strains are used in displaying the strain fields.
303

(a) PW1: (b) PW1: (c) PW1: (d) PW1:


0.5% 0.75% 1.0% 1.5%

(e) PW2: (f) PW2: (g) PW2: (h) PW2:


0.5% 0.75% 1.0% 1.5%

(i) PW3: (j) PW3: (k) PW3: (l) PW3:


0.5% 0.75% 1.0% 1.5%

(m) PW4: (n) PW4: (o) PW4: (p) PW4:


0.35% 0.50% 0.75% 1.0%

Figure 6.16: Tensile yield in horizontal reinforcement during final four drift levels.

shear strain. Section 6.6.3 presents 2nd principal strains. In each section, the strain fields for

each specimen at the first ER+ peak at 0.75% drift are shown; this cycle peak was selected

as the strain patterns were well established (all walls had reached maximum strength) but

the strains associated with severe damage were not yet established. In addition to these,

additional strain fields are provided to illustrate the impact of the indivudal test variables.
304

6.6.1 Vertical Strain

Figure 6.17 shows the vertical strain, εz , fields for the four wall specimens at the first ER+

peaks at 0.75% drift. The impact of the test variables affected the tensile and compressive

vertical strains in different ways. The most noticable difference was in the effect of the splice

on the location of the largest magnitude tensile demands. In the sections that follow, the

impact of the test variables on the shear strain distribution are discussed.

(a) PW1 (b) PW2

(c) PW3 (d) PW4

Figure 6.17: Vertical strain fields at first ER+ peak at 0.75% drift.
305

6.6.1.1 Impact of Shear Demand (PW1 vs PW2)

The impact of shear demand can be evaluated by comparing Specimens PW1 and PW2. In

both walls, the tensile strain is largest in the first floor above the splice; at the base of the

wall, the tensile strain is much lower. In the compressive region, strains extended from the

base of the wall to the first ‘element’ above the splice.

(a) PW1 (b) PW2

Figure 6.18: Vertical strain for Specimens PW1 and PW2 at last WL- peak at 1.0% drift.

Figure 6.18 shows the strain fields of the two specimens at the last peak at 1.0% drift.

These figures illustrate that i) the compressive strains were larger in Specimen PW2 and

i) the tensile strains were larger in the Specimen PW1. These observations hold for the

majority of the test. The first supports the hypothesis that higher shear demand (PW2)

results in compression control behavior of the wall. The second seems to support the

hypothesis that lower shear demand (PW1) results in a tension control behavior of the

wall. However, as Figure 6.19 indicates, the tensile demand above the splice decreases

dramatically from the first to last cycle at 1.5% drift, and supports the hypothesis that the

failure section of the wall with the lower shear demand is at the base of the wall where there

is significant uplift.

Comparison of Figure 6.17 (ER+ at 0.75% drift) and Figure 6.18 (WL- at 1.0% drift)
306

(a) First cycle (b) Last cycle

Figure 6.19: Vertical strain for Specimen PW1 at first and last WL- peak at 1.5% drift.

illustrate another behavior that is common to both wall specimens. At drifts up to 0.75%

drift, the magnitude and distribution of compressive strains in the boundary elements were

essentially the same regardles of loading direction. However, beginning in the 1.0% drift

cycles, the location of the largest magnitude compressive strains was located at the base of

the wall in the east boundary element under ER+ loading and at the top of the splice in

the west boundary element under WL- loading. This suggests that the tensile response of

the dog-leg detail at the top of the splice (occurs first in the west boundary element) may

be affecting the compressive response of the boundary elements on the reverse cycle.

6.6.1.2 Impact of Reinforcement Distribution (PW2 vs PW3)

The impact of reinforcement distribution can be evaluated by comparing Specimens PW2

and PW3. In both walls, the tensile demand is largest in the first floor above the splice and

the largest compressive strains were present in and above the spliced region.

Figure 6.17 shows the concentration of the tensile strains above the splice in Specimens

PW2 and PW3. Close examination of the strains for PW2 indicate that in the web the

location of this strain concentration is lower than in the boundary elements as a result of the

different splice heights. In PW3, the splice height was uniform across the length of the wall
307

and the concentration of vertical strains is at the same height in both the boundary element

and web. The strain concentration at different heights is seen more clearly in Figure 6.20,

which shows the vertical strains at the two peaks prior to failure of each specimen6 . These

also indicate that the compressive demands were larger in the boundary elements in PW2

and that the tensile strain demands are larger and extend further into the web in PW3.

(a) PW2: Last WL- peak at 1.0% (b) PW2: First ER+ peak at
drift 1.5% drift

(c) PW3: Last ER+ peak 1.0%(d) PW3: First WL- peak 1.0%
drift drift (no data at last peak)

Figure 6.20: Vertical strain for Specimens PW2 and PW3 at the last two peaks prior to
failure.

6
The strain fields at failure are not shown due to the loss of many targets in the regions of the walls
where significant compressive damage occurred.
308

6.6.1.3 Impact of Spliced Reinforcement (PW2 vs PW4)

The impact of spliced longitudinal reinforcement can be evaluated by comparing Specimens

PW2 and PW4.

In Specimen PW2, the largest tensile strains were located above the top of the splice.

In Specimen PW4, the largest tensile strains were located at the base of the wall. The

compressive strains were larger in the toe of the wall in PW4 than in PW2. Figure 6.17

shows the strain demands at 0.75% drift in the ER+ loading direction. Figure 6.21 shows

the strain demands at the last WL- peaks at 1.0% drift and illustrate that the significant

compressive strains are located at the top of the splice in Specimen PW2 and at the base

of the wall in the unspliced Specimen PW4.

(a) PW2 (b) PW4

Figure 6.21: Vertical strain for Specimens PW1 and PW2 at last WL- peak at 1.0% drift.

6.6.2 Shear Strain

Figure 6.22 shows the shear strain, γxz , fields for the four wall specimens at the first ER+

peaks at 0.75% drift. The pattern of the shear distribution is fundamentally the same for

all specimens, with the shear demand having occurred primarily in the portion of the web

in tension. A region of highest shear strain was located in the web adjacent to the tensile
309

boundary element in the second floor before being changing to a diagonal region in the web

of the first floor and finally terminating near the compression boundary element at the base

of the wall. An additional concentration of shear strains was located in the compression

boundary element at the base of the wall. The most noticable difference was in the effect

of reinforcement distribtuion on the distribution of shear strains in the web. In the sections

that follow, the impact of the test variables on the shear strain distribution are discussed.

(a) PW1 (b) PW2

(c) PW3 (d) PW4

Figure 6.22: Shear strain fields at first ER+ peak at 0.75% drift.
310

6.6.2.1 Impact of Shear Demand (PW1 vs PW2)

The impact of shear demand can be evaluated by comparing Specimens PW1 and PW2.

Prior to drifts of 0.75% drift, the shear strains at the base of the compression boundary

element are greater magnitude and extend higher in PW1 than in PW2. At larger drifts,

the shear strains in the compression toe is greater in Specimen PW2. Figure 6.23 shows the

shear strain fields at 0.50% and 1.0% drift. At all drift levels, shear strains in the tension

boundary elements are greater in Specimen PW2.

6.6.2.2 Impact of Reinforcement Distribution (PW2 vs PW3)

The impact of reinforcement distribution can be evaluated by comparing Specimens PW2

and PW3. In Specimen PW2 the shear strains are concentrated near the boundary element

in the second floor; in Specimen PW3, the shear strains in the second floor are spread

out more uniformly in the web7 . The same holds for the first floor. For Specimen PW2

the region of highest shear strain is diagonal, spanning the length of the web from tension

to compression boundary elements and top to bottom. For Specimen PW1, the region of

highest shear strain is horizontal along the top of the splice. Because the distribution of

strain in both the first and second floors is affected in PW3, this is likely a result of the

uniform distribution of reinforcement rather than the uniform height of the splice.

The strain fields in Figure 6.22 show the observations for loading in the ER+ direction;

the same holds true for loading in the WL- direction, as shown in Figure 6.24 for which

more of the tension region had valid data from which to calculate the strain fields. The

distribution of shear strain and magnitude above the splice in Specimen PW3 is consistent

with the experimentally observed web crushing, a damage state associated with shear.

7
This observation might suggest that there is greater shear deformation in the second floor of Specimen
PW2, but as the deformation components presented in Section 6.4 shows, the shear deformation in the
two specimens in essentially the same. Thus the observed strain fields indicate only a different distribution
of strain demands
311

(a) PW1 WL- at 0.50% drift (b) PW2 WL- at 0.50% drift

(c) PW1 WL- at 1.0% drift (d) PW2 WL- at 1.0% drift

Figure 6.23: Specimens PW1 and PW2 shear strain fields at last WL- peaks at 0.50% and
1.0% drift.

6.6.2.3 Impact of Spliced Reinforcement (PW2 vs PW4)

The impact of spliced longitudinal reinforcement can be evaluated by comparing Specimens

PW2 and PW4. The primary difference in the shear strain fields is in the compression toes

of the walls. In Specimen PW4, the shear strains in the east (right) toe of the wall were

consistently greater than those in the west (left) toe of the wall (see Figure 6.24). When

the strains in the compression toe of PW4 are compared to PW2, the shear strain extends
312

(a) PW2 (b) PW3

Figure 6.24: Shear strain fields at first WL- peak at 0.75% drift.

higher in the compression boundary elements of PW2. This is shown in the ER+ direction

in Figure 6.22 and in the WL- loading direction in Figure 6.26. It is also observed that the

shear strains in the tension region of the wall is greater in Specimen PW2 than in Specimen

PW4.

(a) ER+ (b) WLi

Figure 6.25: Specimen PW4 shear strain fields at last peaks at 0.50% drift.
313

(a) PW2 (b) PW4

Figure 6.26: Shear strain fields at first WL- peak at 0.75% drift.

6.6.3 Second Principal Strain

Figure 6.27 shows the 2nd principal strain, ε2 , fields for the four wall specimens at the first

ER+ peaks at 0.75% drift. In the sections that follow, the impact of the test variables on

the 2nd principal strain distribution are discussed.

6.6.3.1 Impact of Shear Demand (PW1 vs PW2)

The impact of shear demand can be evaluated by comparing Specimens PW1 and PW2.

There is little difference between the distribution and magnitude of strains throughout the

tests in the ER+ loading direction. In the WL- loading direction (the direction of failure

for both specimens), the principal compressive strains are more distributed for PW1. At

drifts of 0.75% and higher, there is a larger concentration of strains above the top of the

splice in PW2, as shown in Figure 6.28.


314

(a) PW1 (b) PW2

(c) PW3 (d) PW4

Figure 6.27: 2nd principal strain fields at first ER+ peak at 0.75% drift.

6.6.3.2 Impact of Reinforcement Distribution (PW2 vs PW3)

The impact of reinforcement distribution can be evaluated by comparing Specimens PW2

and PW3.

At cycles to 0.75% drift or less, the principal compressive strains are slightly greater

and more distributed in Specimen PW2 than in Specimen PW3, but these difference are

relatively minor. During the cycles to 1.0% drift, the difference between the two specimens
315

(a) PW1 (b) PW2

Figure 6.28: Specimen PW1 and PW2 2nd principal strain fields at first WL- peak at 1.0%
drift.

become significant. Figure 6.29 shows the strain fields at the last WL- peak at 1.0% drift. In

PW3, the region of largest principal compressive strain is located above the splice in the web

and in the compression boundary region, while the pattern of strains remains unchanged in

Specimen PW2. This is consistent with the observed web crushing and high shear strains

in the web of wall for PW3.

6.6.3.3 Impact of Spliced Reinforcement (PW2 vs PW4)

The impact of spliced longitudinal reinforcement can be evaluated by comparing Specimens

PW2 and PW4. The largest 2nd principal compressive strains are located near the top of

the splice in Specimen PW2 and at the base of the wall in Specimen PW4. This reflects the

observation from Section 6.6.1 that compressive strains were concentrated in these regions.

Despite the lower shear strain demands in the compression toes of PW4 (see Section 6.6.2),

the principal compressive strains were larger and more concentrated in PW4 at any given

drift, as shown in Figures 6.27 and 6.30. This is a result of the maximum compressive

and shear strains occurring in the same location in the unspliced specimen (PW4) and at
316

(a) PW2 (b) PW3

Figure 6.29: Specimen PW2 and PW3 2nd principal strain fields at last ER+ peak at 1.0%
drift.

different heights in the spliced specimen (PW2).

(a) PW2 (b) PW4

Figure 6.30: Specimen PW2 and PW4 2nd principal strain fields at first WL- peak at 1.0%
drift.
317

6.7 Strain Demands

The presentation of strain fields in Section 6.6 provides a qualitative evaluation of the strain

distributions in the four wall specimens which are useful in determining the basic response

of the walls and how the responses were affected by the test program variables. In using the

strain fields for validation of numerical models, the usefulness of the strain fields is limited

to visual comparison to the strain fields from continuum models. For simpler numerical

models, it is necessary to have discrete measures of strains for comparison to strains in

those models. This section seeks to do so.

The type and location of strains evaluated were selected for the purpose of supporting

numerical models and were guided by the regions of highest damage and strain demands

as indicated by the observed experimental results and the strain fields. Thus, vertical and

2nd principal strains at the extreme fibers are presented, as well as shear strains in the web

and boundary elements. In the specimens with splices (PW1, PW2, and PW3), strains are

reported at the top of the splice region and at the base of the wall. In the unspliced specimen

(PW4), strains are reported on at the base of the wall. The sections that follow present the

strains at points in the tests (nominal and failure) for each strain measure considered.

6.7.1 Nominal Strains

ACI 318-08 defines sections as being one of three types of control based on steel strain mea-

sures (εt ) at nominal strength (εc = 0.003): i) tension-control (εt ≥ 0.005), ii) compression-

control (εt ≤ 0.002), or iii) transition-control (0.002 < εt < 0.005). To evaluate these limits

for the four specimens, the compression and tension strains when the nominal strain of

−0.003 is reached at the critical sections is presented in Table 6.2.

Specimen PW4 provides the most straightforward evaluation of tension- or compression-

controlled behavior as there are no complications lap splices. Table 6.2 indicates that at

nominal compressive strain, the steel strain gauges provide readings in excess of 0.005 in/in

and the walls can be considered tension-controlled. The same observation can be made for
318

Table 6.2: Critical section extreme fiber strains when nominal compressive strain first
reached.

comp
Specimen Direction −εz , in/in +εz , in/in εt , in/in MHD, % Mcrit /Mncrit
PW1 ER+ -0.0030 0.0108 0.0028 0.64 0.98
WL- -0.0033 0.0184 0.0220 -0.93 -0.97
PW2 ER+ -0.0031 0.0260 0.0270 0.91 0.93
WL- -0.0030 0.0198 0.0215 -0.97 -0.97
PW3 ER+ -0.0030 0.0133 0.0250 0.72 1.03
WL- -0.0030 0.0178 0.0302 0.81 0.97
PW4 ER+ -0.0031 0.0035 0.0150 0.51 0.88
WL- -0.0031 0.0027 0.0066 -0.45 -0.90

Specimen PW3. If the steel strains for these two specimens are compared to the respective

tensile strains from the krypton data, the steel strains are greater than those from the

Krypton data. This observation is useful in examining the strains for Specimens PW1 and

PW2 as no steel strain gauges were provided at the critical section for these specimens. If

it is assummed that the steel strains in these two specimens are greater than the Krypton

tensile strains, both specimens can be considered tension-controlled. The classification of all

four specimens as tension-controlled based on the experimental strains are consistent with

i) the intent of the ACI 318-08 code definition that the cross-sections yield prior to reaching

the maximum usable strain, and ii) the results of cross-sectional analysis of the walls using

actual material properties. However, the evaluation of the tension-/compression-control

limits does not account for i) the shear-compression failures of PW2, PW3, and PW4 and

ii) compression yield prior to tension yield in Specimen PW2 loaded in the ER+ direction

and Specimen PW4 in the ER+ and WL- directions. Section 6.7.2 present additional strain

data to determine the limits controlling the behavior of the walls.


319

6.7.2 Failure Strains

The strain demands associated with failure are of importance in developing numerical mod-

els as they are used in indicating maximum drift capacities. Section 6.7.2.1 presents tensile

strain demands at the wall interface where bar fracture occurred in Specimen PW1. Sec-

tion 6.7.2.2 presents compressive strain demands at the critical sections of the walls, where

shear-compressive failure occurred in Specimens PW2, PW3, and PW4.

6.7.2.1 Tensile Interface Strains

Strains at the wall-foundation interface are of interest due to the flexural failure of Specimen

PW1 when longitudinal reinforcement fractured at this location in the east boundary ele-

ment. The interface strains were calculated using the bottom corner krypton LED targets

and by assumming the vertical displacement at the actual interface were zero; the interface

strain presented is thus the average strain in the bottom two inches of the wall. Table 6.3

provides the maximum tensile interface strains for each specimen.

Table 6.3: Maximum tensile strains at wall-foundation interface.

Specimen Location +εmax


z , in/in MHD, %
PW1 EI 0.291 -1.5
WI 0.14 1.5
PW2 EI 0.085 -1.0
WI 0.096 1.5
PW3 EI 0.011 -0.35
WI 0.016 0.75
PW4 EI 0.14 -1.0
WI 0.090 1.0
1 Maximum strain measured prior to bar fracture.

For Specimen PW1, the maximum tensile strain at the interface in the east boundary

element (where bar fracture occurred) was well in excess of the ultimate strain εu = 0.12

of the steel. The strain limit was first exceed for both sides of the wall in the first cycle
320

to 1.5% drift; bar fracture occurred in the second cycle. By contrast, the maximum strain

on either side of the wall for Specimen PW2, which was built using the same steel, was

approximately 0.10 in/in and below the ultimate strain of the steel. In Specimen PW3, the

maximum interface strains occurred early in the test (−0.35% and 0.75% drift) and were

not a factor in limiting the drift capacity of the wall. In Specimen PW4, the maximum

interface strain exceeded the ultimate strain εu = 0.13 at the east interface but did do so

until after the boundary element had failed.

The interfaces strains presented in Table 6.3 are average surface strains for the lower

two inches of the wall, thus, it is desirable to compare these strains to strains measured by

the steel strain gauges. Unfortunately, in the east (right) boundary element of Specimen

PW1 where bar fracture occurred, the only gauges on bars that fractured were located at

least 7.5 inches from the interface and cannot provide a helpful comparison.

6.7.2.2 Compressive Strains

For Specimen PW1, the strain demands associated with failure were analyzed by examining

the tensile interface strains. To examine strain demands at failure in the other specimens,

it is necessary to consider compressive strains. The location of damage and teh regions

of largest magnitude strains in the strain fields (see Section 6.6) indicate that the strains

should be evaluated at the top of the splice in Specimens PW2 and PW38 and at the base

of the wall for Specimen PW4. Two strain measures are reported, compressive vertical

strain (−εz ) and 2nd (compressive) principal compressive strain (ε2 ). The first is a value

commonly reported as a demand for performance based design. The second is considered

based on the observations of Section 6.6.3 that shear demands contribute to the ultimate

demands on the walls. The strain demands are normalized to account for the difference in

concrete properties of the wall. The strain and normalized values are presented in Table 6.4;

8
Strains are reported at the top of the splice in Specimen PW1 as well because damage and strain fields
suggest that failure under compressive demands would have occurred here if bar fracture at the base of
the wall had not first occurred.
321

Table 6.4: Critical compressive strains at failure. Gray columns indicate the strains at
failure. White columns indicate the maximum historic strains on the opposite side of the
wall.

Value PW1-Fail PW1-Opp PW2-Fail PW2-Opp PW3-Fail PW3-Opp PW4-Fail PW4-Opp


Location WS ES WS ES ES WS EB WB
εZ , mµ -16/-16 -4 -53/-112 -29 -21/-47 -13 -12/-19 -10
εZ /εc0 -7.0/-7.1 -1.9 -21.4/-44.9 -11.4 -7.6/-16.9 -4.5 -5.6/-9.3 -4.7
ε2 , mµ -21/-22 -5 -55/NaN -29 -22/-49 -13 -14/-28 -10
ε2 /εc0 -9.3/-9.4 -2.0 -22.0/NaN -11.4 -7.7/-17.4 -4.7 -6.7/-13.4 -5.0

the gray columns indicate values associated with failure of the specimens (two values are

provided: immediately before and immediately after the loss of load carrying capacity) and

the white columns indicate values associated with the opposite direction of loading.

Ideally, the strain values listed in Table 6.4 would indicate a level of demand that is i)

reached in Specimens PW2, PW3, and PW4 for the direction of loading associated with

failure but not in the opposite direction of loading, and ii) not reached in either loading

direction for Specimen PW1.

If the compressive strains are considered, this is not fully observed. Strains for PW1 and

PW4 suggest an appropriate limit for −εz would be in the range of 16 − 19 mµ; however

these strains were exceeded in Specimens PW2 and PW3 prior to failure. If the compressive

strains are normalized, the values for Specimens PW1, PW3, and PW4 suggest that values

of −εz /εc0 in the range of 7.5 − 9.5 are indicative of failure. Again, the data for Specimen

PW2 does not support this, as the normalized strain in the non-failure direction exceeded

this value. However, strains in PW2 do not exceed the limits suggested by other tests until

the final cycle leading up to failure, and the strain limit was reached prior to 0.0% drift when

the wall was being loaded from the final ER+ to the next WL- peak; the wall eventually

failed when teh wall was loaded to a drift demand in excess of the previously obtained in

the WL- direction. Thus, if the limit of 7.5 − 9.5 had been applied to a numerical model,

the simulated drift capacity of the wall would have essentially been that which was observed
322

experimentally.

If the principal compressive strains are considered, the same characteristics hold as the

for the normalized compressive strains. Values from Specimens PW1, PW3, and PW4,

plus PW2 in the non-failure loading direction, suggest that limits of 22mµ ≤ ε2 ≤ 29mµ

and/or 11.5 ≤ ε2 /εc0 ≤ 13.5 are appropriate values to consider for failure of walls. Although

the values for Specimen PW2 under loading the the failure direction appear to refute this

conclusion, it must be considered that the suggested strain limits were first reached when

the load in the region of failure was much lower due to the location is the displacement

history. Once the wall had significant loading in this region of strain, failure did not occur.

6.8 Stress Demands

Table 6.5 provides an estimation of the shear stress and axial load demands assuming the

load is carried only by the region of the wall in compression. Values are reported at the

end of the test. The area of compression was determined by using the neutral axis depth

(calculated using strains determined from the Krypton displacement) at the critical sections

(at the top of the splice for Specimens PW1, PW2 and PW3 and at the base of the wall

for Specimen PW4). Three areas were considered i) all concrete in the compression region

contributes to the strength (Ac ), ii) all concrete in the compression region minus the cover

concrete contributes to the strength (Ac,core ), and iii) the confined portion of the compression

boundary element contributes to the strength (Abe,core ).

6.9 Conclusions

Section 6.6 presents strain fields that show i) vertical strain, ii) shear strain, and iii) 2nd

principal strains. These strains can be used to validate numerical models of walls9 . Based

on evaluation of these strain fields, the following observations were made:

• Regarding the influence of shear demand:

9
An expanded set of these images and the raw data will be provided on the NEES Project Warehouse.
323

Table 6.5: Shear and axial demands assuming load carried only in the compressive region.

PW1 PW2 PW3 PW4


Value ER+ WL- ER+ WL- ER+ WL- ER+ WL-
1
MHD 1.47% -1.54% 1.50% -1.06% 1.21% -0.76% 1.00% -1.01%
c, in 23.3 26.0 43.4 55.2 35.0 18.5 53.7 33.3
Ac 0.19Ag 0.22Ag 0.36Ag 0.46Ag 0.29Ag 0.15Ag 0.45Ag 0.28Ag
Ac,core 0.14Ag 0.16Ag 0.27Ag 0.34Ag 0.21Ag 0.11Ag 0.33Ag 0.20Ag
Abe,core 0.12Ag 0.12Ag 0.12Ag 0.14Ag 0.12Ag 0.12Ag 0.12Ag 0.12Ag
N , kips 362 kips 361 kips 556 kips 549 kips 358 kips 364 kips 356 kips 364 kips
λN 0.10Ag fc0 0.10Ag fc0 0.13Ag fc0 0.13Ag fc0 0.10Ag fc0 0.10Ag fc0 0.12Ag fc0 0.12Ag fc0
λcN 0.50Afc0 0.44Afc0 0.37Afc0 0.28Afc0 0.34Afc0 0.66Afc0 0.26Afc0 0.43Afc0
λc,core
N 0.68Afc0 0.61Afc0 0.50Afc0 0.38Afc0 0.47Afc0 0.92Afc0 0.35Afc0 0.58Afc0
λbe,core
N 0.83Afc0 0.83Afc0 1.14Afc0 1.13Afc0 0.73Afc0 0.88Afc0 1.0Afc0 1.02Afc0
Vb , kips 172 kips
√ 143 kips
√ 254 kips
√ 245 kips
√ 204 kips
√ 216 kips
√ 159 kips
√ 215 kips

νb 3.3A √fc0 2.7A √fc0 4.6A √fc0 4.5A√fc0 4.0A √fc0 4.3A √fc0 3.4A√fc0 4.6A √fc0
νbc 17.0A√fc0 -12.7A√ fc0 12.8A√fc0 9.7A √fc0 13.8A√fc0 27.6A√fc0 7.5A √fc0 16.4A√fc0
νbc,core 23.5A fc0 17.4A fc0 17.3A fc0 13.1A fc0 18.8A fc0 38.4A fc0 10.2A fc0 22.4A fc0
√ √ √ √ √ √ √ √
νbbe,core 28.6A fc0 23.8A fc0 39.9A fc0 38.5A fc0 29.6A fc0 36.8A fc0 29.1A fc0 39.4A fc0

– In a wall with lower shear demand, the tensile vertical strains are larger and more

concentrated than in a wall with higher shear demand.

– In a wall with higher shear demand, the compressive vertical strains extend

further into the boundary region than in a wall with lower shear demand.

– In a wall with higher shear demand, the shear strains are more distributed hor-

izontally within the tensile region of the wall than in a wall with lower shear

demand.

• Regarding the influence of reinforcement distribution:

– In a wall with uniform longitudinal reinforcement distribution, the region of max-

imum tensile strain extends further into the web than in a wall with longitudinal

reinforcement concentrated in boundary elements.

– In a wall with uniform longitudinal reinforcement distribution, shear strains are

distributed evenly in the tension region of the wall. In a wall with longitudinal

reinforcement concentrated in boundary elements, shear strains are concentrated


324

near the boundary element-web interface.

• Regarding the influence of spliced longitudinal reinforcement at the critical section:

– The use of a splice results in a concentration of tensile vertical strains at the

top of the splice. In wall with no splice, the tensile vertical strains are more

distributed.

– The use of a splice increases the region over which the compressive vertical strains

and 2nd principal strains are distributed. In a wall with no splice, these strains

are concentrated at the toe of the wall.

– In a spliced wall, the region where compressive strains are concentrated appears

to be affected by the cyclic strain response. This does not appear to be a factor

for a wall with no splice.

– The distribution of shear strains in a wall with no splice are less distinct than in

the wall

The strain fields provided not only a tool for validating numerical models, but also

indicate the strains and locations that can best quantify the demands on the walls at key

damage states. The following observations were made:

• The tensile strains in the extreme fibers when the concrete reaches the strain typically

associated with nominal strength (−0.003 in/in) indicate that the walls are tension-

controlled per the ACI 318-08 [1] definition. Although this is the desired design of the

walls, the failure mode of the walls with lower shear demands indicate that the walls

should be classified as either compression-controlled or as falling within the transition

region between tension- and compression-controlled.

• The average tensile strain measured at the wall-foundation interface exceeded the

ultimate strain limit of the reinforcing steel prior to failure at the location where bars

fractured in the wall with lower shear demand (Specimen PW1). In walls that did

not have bar fracture, the average interface strains were less than the ultimate strain
325

of the reinforcing steel.


326

Chapter 7

FRAGILITY FUNCTIONS

This chapter discusses the development of and presents compiled fragility functions for

slender reinforced concrete walls. The presented fragility functions are part of the ATC-

58 project and were developed following the procedures outlined in Seismic Performance

Assessment of Buildings Volume 1 - Methodology, ATC-58 75% draft [5].

Fragility functions for performance based earthquake engineering (PBEE) are functional

relationships defining the likelihood that a specific method of repair (MOR) will be required

given a specific magnitude of engineering demand parameter (EDP) experienced by a struc-

tural component or system. The methods of repair are associated with specific damage

states (DS). The MORs are intended to provided a basis for quantifying the economic

impact of structural damage resulting from a seismic event. The fragility functions were

developed using experimental data and expert opinion.

The layout of this chapter is as follows: Section 7.1 describes the database of experimen-

tal tests used to develop the fragility functions; Section 7.2 evaluates suitability of a suite

of engineering demand parameters (EDPs); Section 7.3 describes the damage states (DSs)

and associated methods of repair (MORs); Section 7.4.6 presents the development of the

fragility functions recommended for use in the ATC-58 project. A summary of the work

and recommendations is provided in Section 7.5.

7.1 Experimental Data Used to Develop Fragilities

The objective of this study was to develop fragility functions for slender walls. Fragilities

were developed using data from experimental tests of slender wall subassemblages. Wall

slenderness was defined on the basis of the shear span ratio.


327

Typically, wall subassemblages are tested to assess earthquake response by applying a

lateral (shear) load at the top of the laboratory test specimen. However, in some cases, to

provide improved simulation of the earthquake load distribution in a wall, lateral load and

moment are applied at the top of the test specimen. To capture the effect of these different

loading schemes, “slenderness” was assessed on the basis of shear span ratio, defined as

M/(V lw ) where lw is the length of the wall, M is the base moment, and V is the base shear.

The ratio M/V is the effective height of loading. Wall subassemblages were considered

“slender” if the shear span ratio was greater than or equal to 2.0.

Any wall determined slender, that did not contain openings, and for which damage and

load-displacement data were provided was included in the dataset. The database com-

promised 66 walls from 18 test programs, with 42 rectangular, 13 barbell, 5 C-shaped, 2

H-shape, and 4 T-shape. Fifty walls were tested with cyclic uni-directional loading, 11 with

monotonic uni-directional loading, and 5 with bi-directional cyclic loading.

The experimental data set is comprised of the test programs summarized in detail in

Chapter 2 and the four planar specimens documented in Chapters 3 through 6. The reader

is referred to these sections for details on the specimens. Section 7.1.1 provides a description

of the data collected from each test. Section 7.1.2 provides a discussion of the design and

response parameters included in the database.

7.1.1 Assembled Test Data

For each specimen, the following data were included in the database: i) measured material

properties (such as concrete compressive strength and steel yield strength), ii) specimen

geometry and reinforcement layout, iii) positive and negative load-displacement envelopes

and iv) damage information, such as cracking, cover spalling, and yielding, buckling and

fracture of reinforcing bars. The following sections provide a more detailed discussion on

each type of data included in the database. The data are summarized in Table 7.1 and

provided in detail in Appendix C.1.


328

Table 7.1: Summary of Database Properties

Parameter Mean Min. Max. Std. Dev. Coeff. of Var.


Scale 0.37 0.16 0.98 0.18 0.48
fc0 , ksi 5.5 3.0 11.3 1.6 0.3
fy , ksi 67 40 87 10 0
ρbe , % 3.5 0.8 11.4 2.3 0.7
ρweb , % 0.71 0.17 2.49 0.70 0.99
ρh , % 0.46 0.17 1.38 0.22 0.49
ρcon , % 1.62 0.00 5.93 1.64 1.01
λN 0.05 0.00 0.21 0.05 1.04
M/(V `w ) 2.46 1.99 5.00 0.53 0.21
vn 6.4 3.5 14.3 2.3 0.4
vu 4.8 1.1 11.0 2.3 0.5
Vu /Vn 0.7 0.2 1.4 0.3 0.4

7.1.1.1 Material Properties

Measured concrete compressive strength, fc0 , and measured steel tensile yield strength,

fy , were reported by researchers. For walls constructed in multiple lifts, the compres-

sive strength of the concrete in the lowest lift of the wall was used. The distribution of

compressive concrete strengths are shown in Figure 7.1a and the distribution of tensile steel

strengths are shown in Figures 7.1b.

In addition to tensile yield strength of the steel, values useful in reproducing the stress-

strain response of the steel were also recorded, including i) modulus of elasticity (Es ), ii)

yield strain (εy ), iii) strain at onset of strain hardening (εsh ), iv) tangent stiffness at on-set

of strain hardening (Esh ), v) ultimate stress (fu ) and vi) ultimate strain (εu ). These values

are provided in Tables C.3 and C.4 in Appendix C.1. When available, tabulated values

were recorded. Occasionally, only figures showing the stress-strain response of the steel

were provided. In these cases, the values were obtained by digitizing the response curve and

selecting the appropriate values.


329

20 45
40
35

# Specimen
# Specimen
15
30
10 25
20
15
5 10
5
0 0
2 4 6 8 10 12 40 50 60 70 80 90
fc′ fy
(a) Concrete compressive strength (b) Steel yield strength

Figure 7.1: Histograms of material strengths

7.1.1.2 Geometry and Reinforcement Layout

In recording the geometry of the walls, the height of the wall (hw ) was defined as the distance

from the bottom of the wall (top of foundation block) to the height of the instrument

recording the lateral displacement of the wall. Wall length, lw , and wall thickness, tw ,

were recorded for all walls. For barbell walls, the width, bf , and the thickness, hf , of the

boundary elements were recorded. For flanged walls, hf , is the thickness of the flange and

bf is the width of the flange.

Cross sections showing the reinforcement layouts of the walls were provided for all test

specimens. However, for some walls, information such as reinforcement spacing or cover

dimensions was not specified. In this case, dimensions were estimated from the provided

drawings. The reinforcing bar properties and layout/spacing were recorded for longitudinal

boundary element and web reinforcement, horizontal reinforcement, and boundary element

confining steel. The length of the boundary elements, `be , was defined as the distance be-

tween the centers of the extreme boundary element longitudinal bars plus twice the distance

from the center of the outermost bar to the surface of the wall.
330

7.1.1.3 Load-Displacement Response

Load and displacement points defining the positive and negative envelopes to the load-

displacement histories were recorded. For some walls, the forces and displacements/drifts

were tabulated by the researcher; for these walls, the values at the peaks for the first cycle

for each displacement demand level in the load history were recorded in the database. If

the values were not tabulated, the load-displacement plots were used to determine load-

displacement points defining the response envelope with the aide of digitizing software.

The factor αload was defined as the factor by which the height of the wall, hw , was

multiplied to get the effective height, `ef f , of loading for the walls, and thus the shear span

ratio was defined by Eq. 7.1. For most walls, this factor was equal to unity, as the load was

applied at the same height where the lateral displacement was measured. For walls loaded

with a moment and shear at the top of the wall, αload was greater than unity (Section 7.2

discusses how the drift at the effective loading height is determined for these walls).

M αload hw hef f
= = (7.1)
V `w `w `w

7.1.1.4 Damage Information

For each test, damage information was recorded if any of the following were reported by

the researchers: i) initiation of cracking (distinctions were made between horizontal and

inclined, or diagonal, cracking), ii) yielding of reinforcement (tensile or compressive), iii)

spalling of concrete, with distinctions made between cover spalling and spalling exposing

longitudinal reinforcement, iv) vertical cracks or splitting of concrete, v) crushing of core

concrete in the boundary region, vi) web crushing, vii) bar buckling, viii) bar fracture, and

ix) shear failure. The organization of the reported damage into damage states and the

association with methods of repair (MORs) is discussed in further detail in Section 7.3.

For each recorded instance of damage, the load-displacement point associated with the

damage occurrence was recorded. Whenever available, the load-displacement at the ex-
331

act occurrence of the damage was recorded, however, in many instances, only the load-

displacement point at the peak of the cycle in which the damage occurred was reported. If

the damage occurred during the unload phase of a cycle, the load-displacement information

corresponding to the maximum historic displacement was recorded.

7.1.1.5 Calculation of Flexural Yielding

One of the damage states used to characterize performance of the walls was the onset of

tensile yielding of the extreme longitudinal reinforcement in the wall. Several different

approaches may be used to determine the onset of yield, e.g. measurement of steel strains

in excess of the yield strain or the application of load in excess of a calculated yield load.

To maintain a consistent definition of yield for the development of the fragility functions,

a theoretical yield force was calculated for each wall and the yield drift was defined as the

drift corresponding to this measured force.

The theoretical yield forces were determined from moment-curvature analysis of the

wall cross section using a fiber model. These analyses were conducted using OpenSees [65].

The cross section models used the measured material strengths and reinforcement layouts

reported in the literature. The concrete fiber response was modeled using the modified

Kent-Park [83] response curve (Concrete01 material model in OpenSees). The measured

compressive strength was assumed to correspond to a compressive strain of −0.002. The

impact of confining reinforcement on the compressive response of the concrete was not

considered, and it was assumed that concrete in tension did not contribute to the strength.

The discretization of concrete fibers was 50-by-10 for wall webs or rectangular cross-sections,

10-by-30 for barbell boundary elements or C-wall flanges, and 10-by-40 for other flanges.

The reinforcing bars for the boundary and web longitudinal steel were modeled using ei-

ther the Steel01 (bilinear response) or ReinforcingSteel (reflects the shape of typical rebar

stress-strain response) uniaxial material models available in OpenSees, depending upon the

measured steel stress-strain response provided in the reference material. In some cases,
332

insufficient steel material information was provided in the literature and it was necessary

to make assumptions about various steel material properties. Tables C.3 and C.4 in Ap-

pendix C.1 contain the steel properties for each specimen, with assumed values denoted

with a *.

7.1.1.6 Crack Width Information

For each recorded damage state, crack widths were recorded when reported. Whenever

possible, a distinction was made between crack widths of horizontal and diagonal cracks.

Additionally, any reported cracks widths corresponding to both peak displacements and

residual displacement of the wall (at approximately zero force) were recorded. This was done

to allow for later consideration of damage states based on widths of cracks. When sufficient

information was available, crack spacing and the height over which cracking occurred was

recorded. The recorded crack widths are provided in Appendix C.3.

Unfortunately, there is not an extensive amount of crack width data for slender walls,

especially residual crack widths, which are the most applicable to development of fragility

functions as observation of damage following an earthquake will only provide data on resid-

ual crack widths. Only 59 instances of residual crack widths were available for the entire

data sets, with all but 6 coming from walls tests by Lowes et al. [61]1 . All residual crack

widths were less than 1/16 in (1.6 mm), the largest measuring 0.05 inches (1.3 mm) at

maximum historic drifts ranging from 0.2%-2.2%.

Figure 7.2a shows residual crack widths as a function of maximum historic drift. Fig-

ure 7.2b shows the crack widths at cycle peaks with markers denoted by damage state; two

cracks were measured as having widths greater than 0.5 inches (12.7 mm) at drifts of 3.2%

and 7.0% and are not shown.


1
The crack width data included for the walls tested by Lowes et al. were measured and recorded during
testing. Hart [44] provides additional information on crack widths and spacing obtained from digitized
crack maps and Metris displacement measurements.
333

0.05 0.5
0.045 0.45
0.04 0.4
0.035 0.35
Crack Width, in

Crack Width, in
0.03 0.3
0.025 0.25
0.02 0.2
0.015 0.15
0.01 0.1
0.005 0.05
0 0
0 0.5 1 1.5 2 2.5 0 1 2 3 4 5 6 7 8
Specimen Drift, % Specimen Drift, %
(a) Residual (b) Cycle max

Figure 7.2: Reported crack widths

7.1.2 Design Parameters

Various design parameters could be expected to affect damage progression in walls. The

parameters discussed in the following sections were computed for each specimen. Their

impact on damage progression is presented in Section 7.4.4. Tables C.1 and C.2 in Ap-

pendix C.1 provide values for each wall subassembly. Table 7.1 provides summary statistics.

Figures 7.3 - 7.6 are histograms showing the distributions of the design parameters.

Reinforcement ratios

Longitudinal reinforcement ratios for the boundary element (the boundary element length,

`be , was defined as the length between the centers of the extreme longitudinal bars in the

boundary element plus twice the distance from the center of the outermost bar to the

surface of the wall), ρbe , and web, ρweb , were calculated as the ratio of the total area of steel

for the respective area to the respective concrete area. The boundary element confining
334

reinforcement ratio, ρcon , was calculated as the ratio of the volume of steel to the volume

of concrete. The horizontal reinforcement ratio, ρh , was calculated as the area of steel in

one row of reinforcement to the area of concrete formed by the thickness of the wall and the

spacing of the horizontal reinforcement. Figure 7.3 shows the distribution of reinforcement

ratios for the database.

40 50
35
40
# Specimen

# Specimen
30
25 30
20
15 20
10 10
5
0 0
0 2 4 6 8 10 12 0 0.5 1 1.5 2 2.5
ρbe (%) ρweb (%)
(a) Boundary region (b) Web

25 45
40
20 35
# Specimen

# Specimen

30
15 25
10 20
15
5 10
5
0 0
0 2 4 6 0 0.5 1 1.5 2 2.5
ρcon (%) ρh (%)
(c) Confining (d) Horizontal

Figure 7.3: Histograms of reinforcement ratios

Scale

It was found that different researchers defined the scale of their test specimens in different

ways. To provide a consistent measure of specimen scale, a full-scale wall was assumed

to be 12 inches (305 mm) thick. Brown [12] reviewed existing structures and determined
335

that 12 inches represented a lower bound on typical wall thickness. Gulec et al. [41] used a

minimum thickness of 8 inches to define a scale for their squat wall database. For this study,

specimen scale was defined as web thickness, tw , divided by 12 inches; an upper bound of

1.0 was specified for the scale. Figure 7.4 shows the distribution of the wall scales used here.

35
# Specimen 30
25
20
15
10
5
0
0 0.2 0.4 0.6 0.8 1
Scale

Figure 7.4: Histogram of wall scale

Shear Span Ratio

The shear span ratio, M/(V `w ), is the ratio of the effective load height to the wall length

and is used to characterize the slenderness of the wall. For flanged walls loaded subjected to

weak-axis bending, the length of the wall `w was replaced by the flange width, bf , however,

the value corresponding to strong-axis bending is typically the value considered, and is

shown in Figure 7.5a.

M αload hw
( )strong = (7.2a)
V `w `w
M αload hw
( )weak = (7.2b)
V `w bf
336

Axial Load Ratio

The axial load ratio, λn , or ratio of applied axial load (P ), to gross section capacity (Ag fc0 )

varied from 0.0 to 0.21. Figure 7.5b shows the distribution of axial load ratio for the

database.

P
λn = (7.3)
Ag fc0

60 35
50 30
# Specimen

# Specimen

40 25
30 20
15
20
10
10 5
0 0
2 3 4 5 0 0.05 0.1 0.15 0.2 0.25
M/(Vℓw ) λN
(a) Shear span ratio (b) Axial load ratio

Figure 7.5: Histograms of load characteristics.

Shear Capacity

Nominal shear strength of the walls was defined using ACI 318-08 Chapter 21 (21.9.4.1) [1]:

 p 
Vn = Acv αc fc0 + ρh fyh (7.4)

where fc0 is the reported concrete compressive strength, fyh is the reported yield strength

of the horizontal reinforcement, ρh is the horizontal reinforcement ratio. The variable αc is

intended to account for the wall aspect ratio and is equal to 2.0 for the walls in the database.
337

Acv is the shear area of the wall, computed as the length of the wall, `w , multiplied by the

web thickness, tw ; for flanged walls under weak-axis loading, the wall length, `w , is replaced

by the flange width bf .

To enable comparison of the shear capacity of different walls, a normalized shear stress

capacity was also computed as the nominal shear strength normalized by the shear area,
p
Acv , and the square of the concrete compressive strength, fc0 .

Vn
vn = p (7.5)
Acv fc0

The distribution of shear stress capacity ratio is plotted in Figure 7.6a.

Shear Demand

The shear demand, Vu , was the maximum shear force reported by the researchers. To enable

comparison of walls, a normalized shear stress demand, vu , was calculated

Vu
vu = p (7.6)
Acv fc0

where all parameters are as defined previously. Figure 7.6b shows the distribution of shear

stress demand ratio.

Shear Demand-Capacity Ratio

The shear demand-capacity ratio was computed as Vu /Vn . The distribution of this ratio in

the database is shown in Figure 7.6c.


338

25 15
20

# Specimen
# Specimen

10
15
10
5
5
0 0
3 4 5 6 7 8 9 10 11 12 13 14 15 1 2 3 4 5 6 7 8 9 10 11
vn vu
(a) Shear capacity (b) Shear demand

20
# Specimen

15

10

0
0.2 0.4 0.6 0.8 1 1.2 1.4
Vu /Vn
(c) Shear demand-capacity ratio

Figure 7.6: Histograms of shear demand and capacity

7.2 Demand Parameters

Multiple earthquake demand parameters were investigated with the objective of determin-

ing the most efficient predictor of earthquake damage. All parameters considered can be

computed using analysis software employed commonly in practice, and thus could easily

be employed by practitioners considering a performance-based design or evaluation. The

parameters considered are discussed in the following sections.

7.2.1 Drift

The specimen drift was computed as the lateral displacement at the plane of loading divided

by the height above the fixed base of the wall. No consideration was made for the type of
339

load (i.e. pure shear or shear and moment) applied at the load plane.

Drift is the simplest demand measure and can be computed using either a linear or

nonlinear model.

7.2.2 Rotation Demand

The rotation demand was calculated as the rotation required to achieve the experimental

drift given the applied loading and using a lumped-plasticity model of the wall. The lumped-

plasticity model comprised a rotational hinge at the base of the wall with a hinge length of

one-half the length of the wall (`w ) with the remainder of the wall modeled as elastic. The

region of the wall above the plastic hinge was assumed to deform in flexure and shear. An

effective flexural stiffness of 0.5Ec Ig and an effective shear stiffness of 0.4Ec Ig were used for

computing rotations for all damage states except DS1 (initial cracking and initial yielding).

For DS1, these effective stiffnesses resulted in overly large elastic deformation and negative

rotation demands. Thus, for DS1, the gross flexural stiffness, Ec Ig , was used.

Rotation demand is the most complicated demand parameter considered in the study

and requires use of a nonlinear response model for application in practice.

7.2.3 Drift at Effective Height

The experimental test programs used to create the slender wall database loaded specimens

using one of three different configurations, shown in Figure 7.7: i) lateral force applied at the

top of the specimen (most common load configuration), ii) lateral forces applied at the top

of the specimen and at one or more additional locations along the height of the specimen,

and iii) lateral force and overturning moment applied at the top of the specimen. Because

these different load distributions imply specimens represent different portions of the real

wall, it could be expected that the drift at the top of the specimen would be different for

the different load distributions. The only consistent known height for the specimens is the

effective height of the applied load (hef f ), calculated as the ratio of the base moment to the
340

base shear. Thus, drift at the effective height (referred to as “effective drift”) was computed

for all specimens in an effort to provide a more consistent engineering demand parameter.

hheffh
effeff hheffh
effeff
0.5E
0.5E
0.5c
0.4E
0.4E
0.4c
hhwh
w=
w=h
=heffh
effeff hheffh
effeff
hhwh
ww hhwh
ww

lwlwlw lwlwlw lwlwlw


0.5l
0.5
0.

(a) Lateral force at top of specimen (b) Distribution of lateral forces (c) Lateral force and overturning
moment at top of specimen

Figure 7.7: Equal base reactions for a real wall (white box) can be achieved for different
specimen (gray box) heights (hw ) by changing the configuration of the applied loading.

For walls loaded with only a lateral force at the top of the specimen, the specimen

height is equal to the effective height, thus, the specimen drift is equal to the effective drift.

For walls with other loading conditions, the drift at the effective height of the loading was

calculated. To estimate the effective drift, the model used to calculate the rotation demand

(see Section 7.2.2) was extended to a height equal to the effective height of loading. Thus, the

model (shown in Figure 7.8) compromised a plastic hinge with a hinge length equal to half

the wall length and rotation equal to the “rotation demand” computed above, topped by an

elastic element with an effective shear stiffness of 0.4Ec Ag and an effective flexural stiffness

of Ec Ig for DS1 and 0.5Ec Ig for all other damage states. This method of estimating the roof

drift ensured that the model used for the estimation accurately simulated the experimental

drift and provided a reasonable estimate of rotation at the top of the experimental specimen,

as this information was not available for all tests.


341

heff heff
0.5EcIg
0.4EcAg
heff
hw hw

lw lw
0.5lw

Figure 7.8: Model used to estimate the drift at the effective height of the wall.

7.3 Damage States and Methods of Repair

In performance based design, there is an interest in being able to estimate the cost of re-

turning the structure to its pre-damaged condition. This is achieved by relating methods

of repair, for which costs can be estimated, to engineering demand parameters. This re-

lationship is typically established using experimental data. The primary focus during an

experimental test is on damage, therefore a critical component of performance based design

is linking the salient damage state(s) with the methods of repair. The methods of repair

and associated damage states established for the present study is presented in this chapter.

FEMA 308 Repair of Earthquake Damaged Concrete and Masonry Wall Buildings [4],

ACI 546-98 Concrete Repair Guide [2], and previous research to develop fragility functions

for RC components ([12, 14, 40, 62, 82]) were used to guide the selection of repair methods

and to link damage states to methods of repair. Four methods of repair, ranging from

cosmetic repair to wall replacement were identified. Table 7.2 provides a summary of the

methods of repair identified and the associated damage states. Figure 7.9 shows the damage

states versus the EDPs of drift, effective drift, and rotation demand2 .

2
Outliers are indicated with gray shading. Section 7.4.2 discusses how outliers are determined
342

In the sections that follow the four methods of repair are discussed. Each section in-

cludes i) a description of the method of repair, ii) the reason the repair is necessary to restore

structural integrity, iii) an overview of the repair process, iv) data necessary to estimate

the potential cost of repair, and v) a summary of each damage state associated with the

method of repair. Section 7.3.1 discusses MOR1, cosmetic repair, which is associated with

initial concrete cracking and initial yield of longitudinal and/or horizontal reinforcement.

Section 7.3.2 discusses MOR2, epoxy injection and patching, which is associated with verti-

cal cracks and spalling of cover concrete that does not reveal the longitudinal reinforcement.

Section 7.3.3 discusses MOR3, replacement of concrete, which is associated with spalling of

cover concrete that exposes the longitudinal reinforcement. Section 7.3.4 discusses MOR4,

replacement of steel and concrete, which is associated with web crushing, boundary element

core crushing, bar fracture, bar buckling, and bond slip.

7.3.1 MOR 1: Cosmetic Repair

The first method of repair, MOR1, cosmetic repair, is associated with damage states for

which structural integrity is not compromised. Cosmetic repair consists of the painting or

application of alternative surface coating/architectural finishes to conceal the presence of

cracks and to provide a barrier against water. FEMA 308 [4] identifies this repair as CR1,

cosmetic patching, and limits the effectiveness of painting to cracks equal to or less than

0.06 inches (1.5 mm) in width.

Cosmetic repair requires repair of the surface area of a wall only. The amount of surface

area cracked could be as large as the entire surface area of the wall. The cost of material

needed for cosmetic repair is largely dependent on the type of surface covering required for

the wall, e.g. paint, drywall, etc.


343

Table 7.2: Damage States and Methods of Repair

Method of Repair Damage State Description


DS1a Initial cracking
DS1b Initial flexural cracking
DS1c Initial shear cracking
Cosmetic repair
DS1d Tensile yield of extreme longitudinal steel
DS1e Compression yield of longitudinal steel
DS1f Tensile yield of horizontal reinforcement
DS2a Spalling of boundary region cover
Epoxy injection
concrete (not revealing longitudinal
and
reinforcement)
concrete patching
DS2b Spalling of patched concrete
DS2c Spalling of web concrete
DS2d Vertical cracks/splitting
Replace concrete DS3a Spalling revealing longitudinal
reinforcement
DS4a Core crushing (boundary element)
DS4b Bar buckling
DS4c Compressive failure of boundary element
DS4d Failure by core crushing
(boundary element)
DS4e Bar fracture
Replace wall
DS4f Failure due to bar buckling
DS4g Failure due to bar fracture
DS4i Shear failure
DS4k Web crushing
DS4m Failure due to web crushing
DS4o Failure by bond slip
DS4p Core crushing in confined boundary ele-
ment of flange tips
(bi-directional tests only)
DS4q Confining reinforcement open or fractured

7.3.1.1 Concrete Cracking

Concrete cracking is the first damage state observed in walls subjected to earthquake load-

ing. The presence of cracks significantly impacts the stiffness of a wall. In evaluating post-

earthquake damage of a building, the width of the residual cracks determines the repair
344

DS

DS
DS

DS
DS4
DS4 DS4
DS4

DS

DS
DS3 DS4 DS3
DS3 DS4
DS2 DS3 DS2
DS2 DS3
DS2 DS2
DS1 DS1
DS1
0 22 DS144 66 00 22 DS1 4 4 66
Specimen Drift,
Specimen Drift,0%
% 2 4 Effective
Effective
6 Drift,
0%%
Drift, 2 4
(a) Specimen Drift, % (b) Effective Drift, %

AllAllWalls
Walls
DS

DS4 All Walls


DS

DS3 DS4
DS3
DS2
DS2
DS1
0 0.02
0.02 DS1
0.04
0.04 0.06
0.06
Rotation 0
Demand
Rotation Demand 0.02 0.04 0.06
Rotation Demand
(c)

Figure 7.9: Damage state (DS) versus engineering demand parameter (EDP). Outliers are
indicated by filled marker.

method necessary. Experimental tests typically report crack widths at peak displacement

but rarely report residual crack widths. Thus, it can be difficult to pair damage states with

methods of repair based solely on the crack width.

For the data set used in this study, maximum crack widths associated with initial crack-

ing (DS1a, DS1b, and DS1c) do not exceed 0.012 inches (0.3 mm). As this is well below the

0.06 inches (1.5 mm) limit provided by FEMA 308 [4], these damage states are associated

with MOR1.
345

7.3.1.2 Yield of Reinforcement

Yield of reinforcement in an experimental test can be reported in a number of ways, such

as when steel strain gauges indicate strains in excess of the yield strain or when the applied

load exceeds a predetermined yield load. Because yield definitions can vary between test

programs, a theoretical yield point was identified for each specimen to provide consistency

among the data. To determine this theoretical yield point, the yield force corresponding to

initial yield was determined from a moment-curvature analysis of the cross-sections. Details

of this calculation are provided in Section 7.1.1.5.

Although there are no crack width measurements directly associated with the theoretical

yield, it was observed that for most walls, the drift at the theoretical yield is similar to

the experimentally observed drift, for which some crack width observations are available.

The maximum observed crack widths do not exceed 0.05 inches (1.3 mm), with only two

measurements exceeding 0.025 inches (0.6 mm). These widths are consistent with the widths

associated with cosmetic repair of walls, and thus association of the yield damage state with

MOR1 is appropriate.

7.3.2 MOR 2: Epoxy Injection and Patching

MOR2, epoxy injection of cracks and patching of spalled cover concrete, is necessary to

restore the structural integrity of the wall but does not require replacement of the steel and

concrete.

Epoxy injection of cracks is necessary to restore the stiffness of a wall, to prevent cor-

rosion of steel, and to restore fire resistance. The literate contains a range of cracks widths

at which it is believed that epoxy injection is required. FEMA 308 [4] states that epoxy

injection is required for cracks wider than 0.06 inches (1.5 mm), but can be achieved at

crack widths as low as 0.002 inches (0.05 mm). Gulec et al. [40] suggested epoxy injection

for squat walls is necessary for cracks equal to or exceeding 0.02 inches (0.5 mm). Dis-

cussions by Pagni and Lowes [82] for the repair of reinforced concrete beam-column joints
346

suggests that widths of 0.025 inches (0.6 mm), 1/32 inches (0.8 mm) or 1/16 inches (1.6

mm) are commonly used by structural engineers. In this study, experimentally observed

crack widths were not explicitly used in establishing methods of repair and it is suggested

that the FEMA 308 guidelines be used.

To estimate the cost of repairing a wall by epoxy injection of the cracks, it is necessary to

provide an estimate of the linear length of cracks for a given surface area of wall. This can

be determined from the descriptions of damage and crack maps provided by researchers3 .

The amount of linear crack spacing requiring epoxy injection was estimated as 150 − 220

feet of linear cracks per 100 square feet of surface area and was based on the following

experimental observations:

• Cracks extend the length of the wall, `w

• Cracks extend as high as `w -2`w from the point of maximum moment demand

• Crack spacing ranges from 3-10 inches (76-254 mm), with a mean and median of 5.5

inches (140 mm)

Patching of cover concrete is necessary for partial spalling of the cover concrete. This

repair is primarily necessary to restore corrosion protection of the steel and assumes the

bond between steel and concrete is still intact. Patching of the cover concrete should not

require the use of course aggregates, thus, the repair method is similar to SR3 for shallow

spalls (less than 0.75 inches (19 mm) deep) in FEMA 308 [4].

The information provided in the literature related to cover spalling for slender walls

suggests that the volume of concrete requiring patching is no more than 0.625 cubic feet

per 100 square feet of wall. This was determined assuming a depth of 0.75 inches (19 mm),

upper limit of depth defining shallow spalling in FEMA 308, and a surface area of no more

than `2w /10, where `w is the length of the wall.


3
The information reported for DS2 damage states was used to estimate the linear footage of cracks
requiring epoxy injection because experimental damage states corresponding to crack widths were not
used.
347

7.3.2.1 Cover Spalling

Three damage states (DS2a, DS2b, and DS2c) were considered to be spalling of the cover

concrete in which the longitudinal reinforcement is not exposed. Because the longitudinal

reinforcement is not exposed, it is assumed that the bond between the steel and concrete is

still intact. Thus, these damage states were associated with MOR2. Figure 7.10a shows an

example of cover spalling.

7.3.2.2 Vertical Cracks

The vertical crack damage state, DS2d, was grouped with cover spalling, and consequently,

MOR2, because the two damage states were often reported together by researchers and

the median drifts for vertical cracks and cover spalling were very similar. Vertical cracks

reported by researchers typically occurred while the wall was in compression.

7.3.3 MOR 3: Replace Concrete

Replacement of concrete was considered necessary for spalling revealing longitudinal rein-

forcement (see Figure 7.10b). Simply patching was considered inadequate because most of

the cover is gone, which necessitates restoration of the bond and compressive capacities.

Consequently, this repair state requires the use of coarse aggregate (FEMA 308 [4]).

To replace the concrete, the following steps are necessary:

1. Shore wall.

2. Remove all concrete in damaged regions.

3. Replace damaged concrete.

The volume of concrete requiring replacement was estimated to be 4-10 square feet per

100 square feet of walls. This was computed assuming a wall thickness of 12 inches (305

mm). The surface area requiring replacement of concrete was determined based on damage
348

information reported by researchers, was approximately the same as that for patching of

cover concrete, and ranged from `2w /25 to `2w /10.

(a) Cover spalling (b) Exposed longitudinal reinforcement

Figure 7.10: Examples of spalling damage states.

7.3.4 MOR 4: Replace Steel and Concrete

MOR4 comprises the replacement of steel and concrete. This repair state was associated

with web crushing, damage to confined core concrete, bar buckling or fracture, shear failure,

and bond-slip failure. This repair method is required when damage to the steel and/or

concrete limits the remaining strength and stiffness of the wall. The extent of the wall area

in need of repair depends on the extent of the damage.

To replace the steel and concrete, the following steps are necessary:

1. Shore floor and wall.

2. Remove all concrete and steel from damaged region.


349

3. Remove all concrete and steel approximately one development length away from the

damaged region (either direction).

4. Replace removed steel and concrete. Jacket dimensions must be adequate to ensure

anchorage of the steel.

The costs associated with replacing the steel and concrete is dependent on i) partial or full

wall replacement, ii) the height and length over which the damage occurs, iii) the size of the

reinforcement and the development length needed to anchor the steel, iv) use of lap splices

or mechanical splices, and v) reinforcement ratios.

7.3.4.1 Core Damage

Damage to the confined core of the boundary region is shown in Figure 7.11a. Like deep

spalling, this requires replacement of the concrete. However, core concrete damage was

associated with MOR4, replacement of steel and concrete. In walls where core damage

occurs prior to visible damage of longitudinal reinforcement, the stresses required to damage

the core likely have also compromised the integrity of the steel and replacement of the steel

is likely necessary.

7.3.4.2 Web Crushing

Because the concrete in the web is not confined and typically does not experience crushing

until the test has reached drift capacity or significant damage has occurred in the boundary

region, crushing of web concrete are associated with damage to core concrete. Figure 7.11b

shows an example of web crushing.

7.3.4.3 Bar Buckling/Fracture

Buckling or fracture of the longitudinal reinforcement (DS4b, DS4e, DS4f, and DS4g), re-

quires replacement of the steel to restore the strength of the wall and replacement of concrete
350

to ensure bond and shear capacity is available. Thus, any damage to the wall reinforcing

steel is associated with MOR4, replacement of steel and concrete. Figures 7.11c and 7.11d

show examples of fractured and buckled bars, respectively.

(a) Core damage (b) Web crushing

(c) Bar fracture (d) Bar buckling

Figure 7.11: Examples of damage states associated with MOR4. The damage states shown
are severe examples to provide clear illustration of the damage states.

7.3.4.4 Bond Failure

There only two instances of bond failure in the data set used for this study. These walls

were designed to represent older construction (they do not satisfy the current ACI design

provisions) and were tested as control walls for evaluating retrofitting methods. Because a

bond failure is likely to only occur in older walls, repair of the wall would likely require a

full replacement of the wall. Thus, bond failure is associated with MOR4, replacement of

steel and concrete. It should be noted that the bond failure damage states occur at EDPs

that are sufficiently low such that they were ultimately determined as outliers.
351

7.3.4.5 Shear Failure

Walls that fail in shear typically sustain damage the full width of the wall and have reduced

strength and limited drift capacity. Thus, shear failure damage states (DS4i) require MOR4,

replacement of steel and concrete, to restore strength and stiffness of the wall. Shear failure

can include i) diagonal shear failure, ii) fracture of shear reinforcement, and iii) sliding shear

failure. Figure 7.12 shows examples of diagonal shear and sliding shear failures.

7.4 Development of Fragility Functions

In this study, fragility functions were developed to quantify the probability of a specific dam-

age state occurring and method of repair being required. Fragility functions are essentially

CDF’s and can be empirical or functional. Here, lognormal cumulative probability distri-

butions are used to define functional CDFs. The engineering demand parameters described

earlier were used to define the fragility functions.

To develop the proposed fragility functions for slender walls subject to earthquake mo-

tion, the full data set described in Section 7.1 was analyzed to remove specific data points

that might inappropriately skew the results. First, outliers were identified using Pierce’s cri-

terion and removed (see Section 7.4.2). Then data set statistics (mean, median, coefficient

of variation) were computed for all walls and for walls grouped by load history (mono-

tonic or cyclic) and test specimen scale. Ultimately, data for walls subjected to monotonic

loading were eliminated. The reduced data set for the proposed fragilities is discussed in

Section 7.4.3. The reduced data set was then used to assess the impact of the design pa-

rameters (shape, aspect ratio, shear demand, axial demand, etc) on the damage progression

(Section 7.4.4).

Section 7.4.5 provides a brief overview of the theory used in developing fragility functions.

Section 7.4.6 discusses the calculate and evaluation of fragility functions for the slender walls

evaluated in this study.


352

(a) Diagonal shear failure [53].

(b) Sliding shear failure [100].

Figure 7.12: Examples of shear failures associated with MOR4.


353

7.4.1 Analysis of the Full Dataset

Table 7.3 shows statistics for the full database, excluding outliers. The median EDP (drift,

effective drift, or rotation demand) is shown for each data set, as well as the mean, standard

deviation, coefficient of variation, minimum and maximum.

For all EDPs, there is a clear increase in the median value as the severity of the damage

states increase, although the separation between DS2 and DS3 is less than elsewhere. For

DS1 (initial cracking and initial yielding), the coefficients of variation range from 1.0-1.2;

for other damage states, the coefficient of variation ranges from 0.3-0.45. Comparison of

the specimen drift and the effective drift indicates that there is not too much statistical

difference between the two EDPs for the full data set.

Table 7.3: Median EDP values of damage states using full database. Outliers not removed.

All
# Median Mean Std. Dev. Coeff. of Var. Min. Max.
Specimen DS1 66 0.0836 0.133 0.132 0.991 0.0223 0.658
Drift DS2 48 1.02 1 0.423 0.424 0.283 2.09
DS3 14 1.12 1.29 0.513 0.399 0.753 2.23
DS4 60 1.87 1.93 0.866 0.448 0.212 4.97
Effective DS1 66 0.0856 0.138 0.131 0.949 0.0223 0.658
Drift DS2 48 1.02 1.02 0.426 0.417 0.283 2.09
DS3 14 1.13 1.35 0.455 0.339 0.843 2.23
DS4 60 1.89 1.97 0.85 0.431 0.449 4.97
Hinge DS1 66 0.000658 0.00113 0.00133 1.17 0.000139 0.00674
Rotation DS2 48 0.00916 0.00924 0.00428 0.463 0.00157 0.0194
DS3 14 0.0111 0.0132 0.00517 0.393 0.00746 0.023
DS4 60 0.0189 0.0197 0.00947 0.481 0.00201 0.0545

7.4.2 Removal of Outliers

When developing fragility functions using experimental data, it is possible to have a value,

referred to as an outlier, that does not reflect the true behavior of real walls. This can be

the result of experimental error or a test specimen that does not accurately represent real

walls. The ATC-58 [5] employed Pierce’s criterion to identify outliers. A detailed outline of
354

the procedure can be found in the ATC-58 guidelines. Using Pierce’s criterion, the general

approach to checking for outliers is the following:

1. Determine the median, θ, and dispersion, β, of the data set

2. Assume one doubtful observation

3. Determine the value R (R = |ln(d) − ln(θ)|/β). This is a tabulated value based on

the number of doubtful observations and the number of walls in the data set being

evaluated.

4. Determine the maximum allowable deviation of a measurement d from the median:

|ln(d) − ln(θ)|

5. Determine the distance of each measurement, di , from

6. Eliminate the data point di if the distance from the median exceeds the allowable

7. If any outliers are detected, return to Step 2 and increase the number of doubtful

observations, D. Repeat until no more outliers are detected.

8. Determine new median and dispersion for the data set

Table 7.4 provides the same information as Table 7.3, but values were determined after

outliers were removed from the data set. When considering the full dataset, few outliers

were detected, and when removed, did not have much effect on the median EDP values but

did reduce the coefficient of variation some. Although minimal impact is seen in the outliers

here, it is import to remove them as the Method of Maximum Likelihood, which is sensitive

to outliers, was used to determine the fragility functions in this study.


355

Table 7.4: Median EDP values of damage states using the full database, with outliers
removed.

All
# Median Mean Std. Dev. Coeff. of Var. Min. Max.
Specimen DS1 66 0.0836 0.133 0.132 0.991 0.0223 0.658
Drift DS2 48 1.02 1 0.423 0.424 0.283 2.09
DS3 14 1.12 1.29 0.513 0.399 0.753 2.23
DS4 58 1.89 1.99 0.822 0.413 0.688 4.97
Effective DS1 66 0.0856 0.138 0.131 0.949 0.0223 0.658
Drift DS2 48 1.02 1.02 0.426 0.417 0.283 2.09
DS3 14 1.13 1.35 0.455 0.339 0.843 2.23
DS4 59 1.89 2 0.833 0.417 0.678 4.97
Hinge DS1 66 0.000658 0.00113 0.00133 1.17 0.000139 0.00674
Rotation DS2 47 0.00951 0.0094 0.00418 0.444 0.00302 0.0194
DS3 14 0.0111 0.0132 0.00517 0.393 0.00746 0.023
DS4 58 0.0191 0.0203 0.00907 0.447 0.00655 0.0545

7.4.3 Reduction of Dataset

The check for outliers discussed in Section 7.4.2 removes data based on statistical analysis.

It is also desired to remove any data that is not representative of full scale walls subjected

to seismic loading. Thus, the impact of load history and test specimen scale on damage

progression were evaluated. When evaluating subsets of the data set, outliers were removed,

following the procedure outlined in Section 7.4.2, prior to presenting the final statistics.

First, the impact of load history on damage progression was investigated. The data were

separated into three categories: i) walls subjected to monotonic loading, ii) walls subjected

to uni-directional cyclic loading, and iii) walls subjected to bi-directional cyclic loading.

Dataset statistics are listed in Table 7.5. Eleven wall specimens (the full dataset comprised

66 specimens) were subjected to monotonic loading; ten were from same test program

[54]. The drift corresponding to DS5 was expected to higher for these walls than for those

subjected to cyclic loading. However, the data in Table 7.5 show that this was not the case,

with the median EDPs for onset of DS5 for monotonic loading was approximately 80% of

the EDPs for cyclic loading. On the basis of this unexpected behavior, the monotonic tests
356

were considered to provide damage data unrepresentative of earthquake damage and it was

decided to remove the monotonic tests from the data set. The median EDP was higher for

bi-direction tests than for uni-direction tests, which is likely because there were relatively

few data points available for for bi-directional tests. Additionally, the bi-directional tests

in the data set tended to have fewer cycles at larger displacement increments than did

uni-directional tests, and the researchers tended to report damage only at the cycle peaks.

Thus, the damage was not reported at the drift at which damage actually initiated, but

rather at the much larger drifts, thus resulting in unrealistically high median EPPs. It

was concluded that there was insufficient information to create a seperate set of fragility

functions for bi-direction loading.

Table 7.5: Median EDP values of damage states (walls sorted by load history)

Monotonic Cyclic Bi-Dir.


# Median c.v. # Median c.v. # Median c.v.
Specimen DS1 11 0.0462 0.31 50 0.097 0.715 5 0.494 0.751
Drift DS2 4 0.978 0.0863 40 1.01 0.453 4 1.4 0.274
DS3 0 NaN NaN 13 1.08 0.411 1 1.8 0
DS4 9 1.61 0.196 43 2 0.399 4 2.11 0.0735
Effective DS1 11 0.0462 0.31 50 0.103 0.672 5 0.494 0.729
Drift DS2 4 0.978 0.0863 40 1.01 0.43 4 1.41 0.349
DS3 0 NaN NaN 13 1.11 0.346 1 1.8 0
DS4 9 1.61 0.196 44 1.95 0.404 5 2.26 0.154
Hinge DS1 11 0.000244 0.386 50 0.000711 0.798 5 0.00514 0.777
Rotation DS2 4 0.0086 0.17 39 0.00951 0.469 4 0.0131 0.303
DS3 0 NaN NaN 13 0.0106 0.403 1 0.0184 0
DS4 9 0.0153 0.26 43 0.0191 0.419 4 0.0214 0.0839

Next, the impact of test specimen scale was investigated, as it is often thought that

small scale specimens result in unrealistic damage progression. Four subdivisions, shown in

Table 7.6, were considered. The median values of walls with scale less than or equal to 0.25

(which were primarily the result of one test program), were not consistently less than or

greater than those walls with a scale less than 0.33, and the medians were often very similar

to those with a scale greater than 0.33. Due to this lack of consistency, it was determined
357

that there was insufficient reason to develop suites of fragility functions based on the scale

of the wall. No walls were removed from the data set on the basis of scale.

Table 7.6: Median EDP values of damage states for reduced data set (walls sorted by scale)

Scale ≤ 0.25 0.25 < Scale ≤ 0.33 Scale > 0.33


# Median c.v. # Median c.v. # Median c.v.
Specimen DS1 11 0.0833 0.348 19 0.135 0.937 24 0.105 0.832
Drift DS2 10 1.33 0.254 15 0.75 0.323 18 1.08 0.5
DS3 2 1.73 0.222 3 1.8 0.154 9 0.843 0.314
DS4 6 1.83 0.139 17 2.5 0.236 22 1.78 0.427
Effective DS1 11 0.0833 0.348 19 0.135 0.937 24 0.153 0.737
Drift DS2 10 1.33 0.254 15 0.75 0.323 18 1.08 0.472
DS3 2 1.73 0.222 3 1.8 0.154 8 0.986 0.108
DS4 6 1.83 0.139 17 2.5 0.236 23 1.85 0.443
Hinge DS1 11 0.000651 0.276 19 0.00097 1.17 24 0.00105 0.949
Rotation DS2 9 0.0128 0.221 15 0.00661 0.39 18 0.00975 0.502
DS3 2 0.0176 0.255 3 0.0184 0.175 9 0.00896 0.294
DS4 6 0.0184 0.177 18 0.0241 0.301 22 0.0174 0.418

7.4.4 Impact of Design Parameters on Damage Progression

The impact of various design parameters on damage progression was investigated. In part,

this was conducted to investigate the development of multiple suites of fragility functions.

The reduced data set (monotonic tests excluded) was used. The following parameters were

considered: a) shape, b) shear-span ration, c) axial load ratio, d) shear stress demand, and

e) shear demand/capacity ratio. The specific design parameter values used to sub-divide the

data set were determined via an iterative process. Only the final subdivisions are presented

here. For each sub-division, an outlier check was performed using the procedure discussed

in the previous section. In evaluating the sub-divisions, the following were considered: i)

increasing median value as the damage level increases, ii) difference between the median

values for different sub-divisions, iii) the number of outliers in each sub-division, and iv) are

there sufficient data points in each sub-division to ensure confidence that the observations

are a function of the design parameter considered and is not a misrepresentation of the
358

behavior due to a few data points significantly affecting the median.

Table 7.7 shows the median values and coefficients of variation for the database sorted by

shape (rectangular (32), barbell(12), and flange (11; C-walls, T-walls, and H-walls)). While

there are distinct differences in the median values for each shape, there are not a sufficient

number of data points for the barbell and flanged walls to consider them independently.

Table 7.7: Median EDP values of damage states (walls sorted by shape)

Rect. Barbell Flange


# Median c.v. # Median c.v. # Median c.v.
Specimen DS1 32 0.0848 0.781 12 0.105 0.64 11 0.245 0.787
Drift DS2 26 1.03 0.436 11 0.597 0.507 7 1.11 0.289
DS3 10 0.927 0.4 3 1.67 0.223 1 1.8 0
DS4 26 1.83 0.429 10 2.21 0.296 10 2.04 0.197
Effective DS1 32 0.0946 0.712 12 0.105 0.64 11 0.245 0.775
Drift DS2 26 1.03 0.4 11 0.597 0.507 7 1.11 0.359
DS3 10 1.05 0.323 3 1.67 0.223 1 1.8 0
DS4 28 1.83 0.465 10 2.21 0.296 10 2.04 0.207
Hinge DS1 31 0.000703 0.636 12 0.00052 0.795 11 0.00214 0.87
Rotation DS2 26 0.00975 0.447 10 0.00537 0.499 7 0.0104 0.333
DS3 10 0.00933 0.392 3 0.0164 0.249 1 0.0184 0
DS4 26 0.0184 0.432 10 0.0223 0.333 9 0.0216 0.151

Shear span ratio, Eq. 7.2, was considered because one could expect that walls with a

lower shear span ratio to have lower drift demands. The data in Table 7.8 shows that, for

walls with a shear span ratio less than or equal to three, the median EDPs for all damage

states (except DS1 for drift and rotation) are significantly smaller than those for walls with

shear span ratios greater than three. However, there are at present too few experimental

data on walls with larger aspect ratios to conclude that multiple suites of fragility functions

are necessary for slender walls based on shear span ratio.

Axial load ratio was considered because one could expect that higher compressive loads

will delay the onset of initial cracking and yield but could decrease the demand at which

bar buckling and core crushing occur. Axial load ratios were divided as less than or greater

than 10% of the gross capacity, with most walls falling into the former category. Table 7.9
359

Table 7.8: Median EDP values of damage states for reduced data set (walls sorted by shear
span)

M/(V `w ) ≤ 3 M/(V `w ) > 3


# Median c.v. # Median c.v.
Specimen DS1 47 0.101 0.915 7 0.0941 0.629
Drift DS2 37 0.905 0.416 5 1.6 0.0414
DS3 14 1.12 0.399 0 NaN NaN
DS4 42 1.88 0.393 6 2.78 0.148
Effective DS1 47 0.104 0.883 7 0.106 0.504
Drift DS2 37 0.905 0.39 7 1.6 0.172
DS3 14 1.13 0.339 0 NaN NaN
DS4 42 1.9 0.385 6 2.78 0.115
Hinge DS1 47 0.000719 1.08 7 0.000666 0.558
Rotation DS2 36 0.00857 0.456 7 0.0137 0.239
DS3 14 0.0111 0.393 0 NaN NaN
DS4 42 0.0189 0.42 6 0.0254 0.144

provides the median values for each damage state and EDP. Walls with lower axial loads

have significantly higher EDPs for DS5 and DS4, and lower drift and roof drift for DS1.

However, there are relatively few data for walls with axial loads in excess of 10% of the

gross capacity and more experimental data would be needed to support the developement

of multiple suites of fragility functions on the basis of axial load ratio.

No trends were observed for the impact of shear demand. Table 7.10 provides three
p
groups of shear demand, walls with a demand less than or equal to 2.5Acv fc0 , walls with
p
a demand greater than or equal to 7Acv fc0 , and those between. The walls with low shear

demands have a higher median drift for DS3 and DS5 than do the walls with a mid-range

shear demand. However, the median drift for DS5 has a very similar value for low and high

shear demands. The higher drift for high shear demands is a result of all but one of the

walls with this higher shear demand coming from the same test program (Oesterle et al.

[80, 78]), which had a median drift at DS5 higher than than for the full data set.

Table 7.11 contains the median values for the database sub-divided by shear demand-

capacity ratio of the walls. There are no trends observed that indicate a need for separate
360

Table 7.9: Median EDP values of damage states for reduced data set (walls sorted by axial
load ratio)

λN < 0.10 λN ≥ 0.10


# Median c.v. # Median c.v.
Specimen DS1 45 0.0941 0.913 9 0.104 0.526
Drift DS2 34 1.1 0.375 9 0.566 0.584
DS3 9 1.67 0.308 5 0.756 0.17
DS4 38 2.18 0.345 10 1.27 0.435
Effective DS1 45 0.102 0.871 9 0.108 0.451
Drift DS2 34 1.1 0.38 9 0.597 0.512
DS3 9 1.67 0.267 5 0.921 0.0942
DS4 39 2.17 0.362 10 1.32 0.394
Hinge DS1 45 0.000719 1.05 9 0.000719 0.662
Rotation DS2 34 0.0105 0.408 9 0.00453 0.453
DS3 9 0.0164 0.31 5 0.00853 0.13
DS4 38 0.0217 0.369 10 0.0139 0.36

Table 7.10: Median EDP values of damage states for reduced data set (walls sorted by shear
demand)

vu ≤ 2.5 2.5 < vu < 7 vu ≥ 7


# Median c.v. # Median c.v. # Median c.v.
Specimen DS1 13 0.0522 0.636 32 0.105 0.829 9 0.108 0.541
Drift DS2 12 1.12 0.407 24 1.01 0.429 8 0.615 0.462
DS3 1 1.15 0 12 1.05 0.43 1 1.67 0
DS4 9 2.64 0.442 30 1.87 0.339 8 2.41 0.221
Effective DS1 13 0.0522 0.663 32 0.139 0.782 9 0.108 0.541
Drift DS2 12 1.12 0.407 24 1.01 0.415 8 0.615 0.462
DS3 1 1.15 0 12 1.1 0.361 1 1.67 0
DS4 10 2.42 0.511 30 1.9 0.337 8 2.41 0.221
Hinge DS1 13 0.000512 0.648 32 0.00105 0.975 8 0.000818 0.586
Rotation DS2 12 0.0115 0.405 24 0.00911 0.444 8 0.00466 0.592
DS3 1 0.0117 0 12 0.0101 0.424 1 0.0164 0
DS4 9 0.0262 0.462 30 0.0188 0.342 8 0.0244 0.247

suites of fragility functions.


Table 7.11: Median EDP values of damage states for reduced data set (walls sorted by shear demand ratio)

Vu /Vn < 0.5 0.5 ≤ Vu /Vn ≤ 0.75 0.75 ≤ Vu /Vn ≤ 1.0 Vu /Vn > 1.0
# Median c.v. # Median c.v. # Median c.v. # Median c.v.
Specimen DS1 14 0.0736 0.648 17 0.23 0.737 12 0.0833 1.14 11 0.101 0.266
Drift DS2 11 1.09 0.37 13 1.01 0.435 12 0.828 0.447 8 0.871 0.553
DS3 1 1.46 0 5 1.08 0.293 7 1.67 0.442 1 0.756 0
DS4 11 2.2 0.432 13 2.03 0.339 11 1.89 0.28 12 2.1 0.39
Effective DS1 14 0.0851 0.63 17 0.23 0.696 12 0.0833 1.11 11 0.108 0.26
Drift DS2 11 1.09 0.37 13 1.01 0.436 12 0.828 0.412 8 0.871 0.528
DS3 1 1.46 0 5 1.11 0.229 7 1.67 0.391 1 0.96 0
DS4 12 2.03 0.487 13 2.03 0.33 12 1.88 0.306 12 2.1 0.396
Hinge DS1 13 0.000456 0.611 17 0.00174 0.82 12 0.000719 1.27 11 0.00085 0.309
Rotation DS2 11 0.00978 0.384 13 0.00973 0.481 12 0.00767 0.486 8 0.00772 0.619
DS3 1 0.0144 0 5 0.0106 0.286 7 0.0164 0.438 1 0.00841 0
DS4 11 0.0238 0.464 13 0.0213 0.325 12 0.0195 0.336 11 0.0212 0.383

361
362

7.4.5 Fragility Function Theory

The fragility functions presented here are defined by the lognormal cumulative probability

distribution (CDF). Experimental data provides discrete information that can be used to

create an empirical CDF; however, the empirical CDF is defined by the entire dataset and

use of the empirical CDF requires the entire dataset. The lognormal CDF is defined by two

parameters. These parameters may be estimated in two ways performance-demand data set

[42]. The first method, the Method of Moments, uses the mean and variance of the empirical

data set. The Method of Moments introduces error associated with estimating population

statistics (i.e. distribution parameters) from the sample that is the data set. The second

method, the Method of Maximum Likelihood, uses a likelihood function to estimate the

distribution parameters which are most likely to produce the empirical values. Unlike the

Method of Moments, this produces a higher probability that the CDF accurately represents

the empirical data. However, because the Method of Maximum Likelihood considers the

individual empirical values, it is more sensitive to the presence of outliers.

The lognormal distribution was used for the development of fragility functions in this

study. Researchers have investigated other distributions (beta, gamma, Weibull) for use in

earthquake damage predictions ([82], [41]), however, these have not been shown to provide

significant improvement, over the lognormal distribution for damage prediction in structural

elements. Further, the lognormal distribution is being used in the ATC-58 project for all

structural and non-structural element.

The appropriateness of a chosen distribution can be evaluated using goodness-of-fit tests.

Pagni and Lowes [82] discuss in detail three such tests: the chi-squared (χ2 ), Kolmogorov-

Smirnov (K-S), and the Lilliefors [58] tests. The χ2 test is not appropriate for small sample

sizes. The K-S test is appropriate for any sample size but is not exact and provides un-

conservative results if the distribution parameters are defined by the empirical data set.

The K-S test can be modified to provide exact results for the case of distribution parame-

ters defined by the empirical dataset and a small data set; the Lilliefors [58] test does this
363

for the case of the normal distribution. Here, the Lilliefors test was used to evaluate the

appropriateness of the lognormal distributions used to define the fragility functions.

To perform the Lilliefors test, the maximum difference between the theoretical CDF,

FX , and the emperical CDF, Sn , is determined:

Dn = max(|FX (xi ) − Sn |) (7.7)

where xi is the ith observation of each CDF. If value of Dn is less than or equal to a

tabulated value of Dnα , then the distribution tested is acceptable at the significance level α.

The difference between the K-S and Lilliefors tests are the tabulated values of Dnα .

7.4.6 Fragility Functions for Slender Walls

The proposed fragility functions (functional CDF’s) were developed using the reduced data

set (no monotonic tests). Outliers were removed using Pierce’s criterion. As discussed in

Section 7.4.5, a log-normal probability distributions were used to define the fragilty functions

and the distribution parameters were determined using the method of maximum likelihood.

This was achieved through the use of the Matlab statistics toolbox lognfit function, which

returns the mean (µ) and standard deviation (σ) of the associated normal distribution.

The median of the lognormal distribution (θ) is equal to eµ and the dispersion (βd ) can be

approximated as σ.

Dispersion is a measure of uncertainty in the data. ATC 58 [5] requires consideration

for two sources of dispersion. The first of these, βd , is the dispersion of the experimental

data determined using the method of maximum likelihood. The second, βu , accounts for

uncertainty that experimental tests represent the conditions of a real building component.

ATC 58 specifies that the fragility function dispersion, β, be calculated as the square root

of the sum of the squares (SRSS) of βd and βu :

q
β= βu2 + βd2 (7.8)
364

For the slender wall dataset, the only condition for which βu was not equal to 0.10 was

when five or fewer specimens were used for a particular damage state, in which case 0.25

was used per the ATC 58 guidelines. The lognormal distribution parameters are provided

in Table 7.12. Figure 7.13 shows empirical data and the fitted lognormal curve (including

consideration for the uncertainty parameter).

The Lilliefors test with a significance level of 5% was used to evaluate the appropriate-

ness of the lognormal distribution. The test was done using the Matlab function lillietest.

Because the Lilliefors test was developed for a normal distribution, the log of the emperical

data was used. The test indicated that the log-normal distribution was appropriate for all

EDPs and MOR combinations except MOR2 and MOR4 for experimental drift.

Table 7.12: Lognormal distribution parameters for cyclically loaded walls

All Walls
θ βd βu β Correct CDF p
Specimen DS1 0.111 0.779 0.1 0.786 T 0.0763
Drift DS2 0.897 0.495 0.1 0.505 F 0.00918
DS3 1.19 0.398 0.1 0.41 T 0.348
DS4 1.86 0.419 0.1 0.431 F 0.0339
Effective DS1 0.118 0.755 0.1 0.762 T 0.5
Drift DS2 0.927 0.465 0.1 0.476 F 0.0469
DS3 1.28 0.326 0.1 0.341 T 0.128
DS4 1.86 0.43 0.1 0.441 F 0.0196
Hinge DS1 0.000869 0.877 0.1 0.882 T 0.0529
Rotation DS2 0.00841 0.501 0.1 0.511 T 0.0564
DS3 0.0123 0.382 0.1 0.395 T 0.414
DS4 0.0189 0.44 0.1 0.451 F 0.0145

7.5 Recommendations and Conclusions

This report presented the development of fragility functions for slender RC walls. Exper-

imental data from 66 wall specimens were used to associate damage states and methods

of repair with engineering demand parameters for walls. Rectangular, barbell, and flanged

walls with shear span ratios of two or greater were considered. It was determined that it
365

1 1
Likelihood of Achieving a DS

Likelihood of Achieving a DS
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 1 2 3 4 0 1 2 3 4
Specimen Drift, % Effective Drift, %

1
DS1: Empirical
Likelihood of Achieving a DS

DS1: Functional
0.8
DS2: Empirical
0.6 DS2: Functional
DS3: Empirical
0.4 DS3: Functional
DS4: Empirical
0.2
DS4: Functional

0
0 0.02 0.04 0.06 0.08 0.1
Rotation Demand

Figure 7.13: Fragility functions for cyclically loaded walls

was necessary to remove the damage data for monotonically loaded walls, and the final

fragilities were developed using 55 cyclic tests. One set of fragility functions was developed

as there was not sufficient data in individual categories to develop multiple sets of fragility

functions. The engineering demand parameters considered were experimentally reported

drift, drift at effective height of loading, and rotation demand. Recommendations to the

ATC-58 committee were made using rotation demand, and are summarized in Table 7.13

(with median values rounded to two significant figures and dispersions to the nearest 0.05),

however, effective drift values provided in Tables 7.12 could also be used.
366

Table 7.13: Recommended fragility function parameters

Method of Repair Damage State θ βd βu β


MOR-1: Cosmetic Repair DS1: 0.00087 0.90 0.10 0.90
Initial cracking
and yielding
MOR-2: Epoxy Injection and DS2: 0.0084 0.50 0.10 0.50
Patching Cover spalling,
vertical cracks
MOR-3: Replace Concrete DS3: 0.012 0.40 0.10 0.40
Exposed long.
reinforcement
MOR-4: Replace Steel and DS4: 0.019 0.45 0.10 0.45
Concrete Bar fracture or
buckling,
core damage,
bond slip,
crushed web,
shear failure
367

Chapter 8

EVALUATION OF BUILDINGS DAMAGED IN 2010 MAULE


EARTHQUAKE: OVERVIEW

An important aspect of the seismic performance of structural reinforced concrete walls

is the ability of engineers to identify, in existing buildings, potential deficiencies that are

susceptible to damage during a seismic event. One such method for evaluating the seismic

adequacy of a building component is the use of fragility functions such as those presented in

Chapter 7. Another option for the evaluation of existing structures is the use of standards

such as ASCE/SEI 31-03 and 41-06, published by the American Society of Civil Engineers

(ASCE).

Seismic Evaluation of Existing Buildings (ASCE/SEI 31-03) [7] provides guidance for

design professionals to evaluate the seismic vulnerability of an existing building. The evalu-

ation procedures are structured in a three-tiered approach. Each tier has an increasing level

of complexity with increased analysis, sophistication, and computation effort, but results in

a more thorough, detailed, and, ideally, realistic evaluation of how the structure would per-

form in an earthquake. Tier 1, or the Screening Phase, is intended to be a quick evaluation

to identify potential deficiencies. Tier 2, or the Evaluation Phase, uses results from linear

analyses. Tier 3, or the Detailed Evaluation Phase, requires a nonlinear analysis of either

the full structure or a portion of the structure. The Tier 2 and Tier evaluation criteria from

ASCE 31 were updated by Seismic Rehabilitation of Existing Buildings (ASCE/SEI 41-06)

[8, 9]. Consequently, both standards were considered in this study.

The 2010 Maule earthquake in Chile provided a unique opportunity to evaluate the

ASCE 31/41 standards. A large number of buildings affected by the earthquake were mid-

rise apartment or condominium buildings with reinforced concrete walls serving the dual

purpose of supporting gravity load and resisting earthquake induced lateral loads. Although
368

many of these buildings performed well in the earthquake, others sustained damage ranging

from minor to severe. For some buildings, damage was so extensive that the structures

were ultimately condemned. By studying several of these buildings and conducting the

three-tiered evaluations, the results of the ASCE 31/41 evaluations can be compared to the

actual performance of a building, providing insight into the benefits and shortcomings of

ASCE 31/41 evaluations previously not available.

The work presented in this and the following chapter was conducted as part of a col-

laborative project funded by the NEHRP Consultants Joint Venture (a partnership of the

Applied Technology Council (ATC) and Consortium of Universities for Research in Earth-

quake Engineering (CUREE)) and titled “Analysis of Seismic Performance of Reinforced

Concrete Buildings in the 2010 Chile Earthquake, Including Effects of Non-Seismic-Force-

Resisting Building Structural Elements”. The project objectives were to study buildings

damaged in the 2010 Chile earthquake and determine lessons learned that can be applied to

U.S. codes and standards. The author was a member of a subgroup tasked with investigation

of the response of four structural wall buildings. The objectives of the investigation included

evaluation of the provisions of ASCE/SEI 31-03 and 41-06 and providing recommendations

for improvements to the standards.

This chapter provides background for the study undertaken. Section 8.1 provides an

overview of the ASCE/SEI 31-03 and 41-06 standards. Section 8.2 provides details of

the models used to perform the evaluations. Section 8.3 provides a brief overview of the

buildings studied. The results of the study, as well as recommendations for modifications

to the ASCE 31/41 standards, are presented in Chapter 9.

8.1 Overview of Standards Used

Two standards were evaluated as a part of the study on damage to Chilean walled build-

ings, ASCE/SEI 31-03 and ASCE/SEI 41-06. Seismic Evaluation of Existing Buildings

(ASCE/SEI 31-03) [7] is a standard published by the American Society of Civil Engineers

(ASCE) to provide guidance for design professionals to determine if an existing building is


369

adequate to resist seismic forces. This standard is the source of the three tiered evaluation

approach briefly mentioned in the introduction to this chapter. The three tiers consist of (i)

quick evaluations to identify potential deficiencies, (ii) linear analysis of the structure, and

(iii) nonlinear analysis of the structure. This three-tiered approach was used as a guide-

line for conducting the research presented here; however, the details for conducting these

evaluations deviated from the ASCE/SEI 31-03 standard as a result of additional standards

published since publication of ASCE/SEI 31-03. Specifically, the provisions of Seismic Re-

habilitation of Existing Buildings (ASCE/SEI 41-06) [8, 9] were used, where applicable, in

place of those of ASCE/SEI 31-03.

Tier 1, or the Screening Phase, is intended to be a quick evaluation to identify potential

deficiencies, such as weak stories or insufficient shear strength, in a building. Such an

evaluation can be conducted with minimal time and computation effort on the part of the

design professional. ASCE/SEI 31-03 provides checklists for this purpose; these checklists

were used for this study. A brief summary of the requirements for this evaluation is presented

in Section 8.1.2.

ASCE/SEI 31-03 is structured such that if no deficiencies are identified by the Tier 1

evaluation, the building can be considered not having any seismic deficiencies and no further

analysis is required. If deficiencies are identified, a Tier 2 evaluation is required. Tier 2, or

the Evaluation Phase, requires the engineer to construct a linear model of the building and

evaluate the simulated demands against acceptable limits. ASCE/SEI 31-03 defines these

acceptable limits by evaluating simulated component demands against the strength of the

components. The specific procedures for these acceptable limits in ASCE/SEI 31-03 are

nearly identical to those for provided in ASCE/SEI 41-06 for evaluating a building using

linear procedures, thus, the discussion of Tier 2 in this document follows the ASCE/SEI

41-06 standard for linear procedures. Section 8.1.3 provides an overview of these procedures.

The Tier 3, or the Detailed Evaluation Phase, requires a nonlinear analysis of the build-

ing. ASCE/SEI 31-03 does not provide specific guidelines for this phase, stating that the
370

engineer should refer to typical design and/or evaluation procedures for nonlinear models.

The Tier 3 evaluations referred to in this study follow the guidelines of ASCE/SEI 41-06

for nonlinear time history analysis. Section 8.1.4 provides an overview of these procedures.

8.1.1 Evaluation Requirements

ASCE/SEI 31-03 and 41-06 provide details of the information necessary for evaluation of

a building, including the necessary levels of investigation, building type, level of seismicity,

and level of performance for evaluation. For the buildings evaluated in this study, the

data available was constrained to that collected by the reconnaissance teams. The building

type of all structures investigated was defined by ASCE 31 as Type 9, C2 (Concrete Shear

Walls with Stiff Diaphragms). High seismicity levels were assumed; additional details on

the seismicity levels assumed are provided in Section 8.2.1).

ASCE 31 Tier 1 checklists are based on two performance levels, Immediate Occupancy

(IO) and Life Safety (LS). ASCE 41 provides four target performance levels for the anal-

ysis of existing structures, namely, Collapse Prevention (CP), Life Safety (LS), Immediate

Occupancy (IO), and Operational. As a guideline for establishing the desired performance

level of a structure, a number of tables are provided in ASCE 41 relating performance levels

to overall building damage and damage to individual components. Table 9.5 provides those

relationships relevant to this study. ASCE 41 explicitly states that these relationships are

not to be used for post-earthquake evaluation of structures and for judging the safety or

level of repair of a building; however, as the goal of this study was to evaluate the adequacy

of the ASCE 41 standard, these relationships were used as guidelines for determining if spe-

cific performance levels can be accurately predicted for a structure or individual building

components.

Tier 2 evaluations were conducted to determine how the buildings were expected to

perform given a specific performance objective, which is defined as achieving a specified

performance level under a certain hazard level. The hazard levels considered were the
371

design and maximum considered events as defined by ASCE/SEI 7-10 [6]. The following

performance objectives were considered:

1. Achieving the Immediate Occupancy performance level under the design-level event

(IO-DBE).

2. Achieving the Life Safety performance level under the design-level event (LS-DBE).

3. Achieving the Collapse Prevention performance level under the maximum event (CP-

MCE).

8.1.2 Tier 1

The Tier 1 Evaluation (Screening Phase) was created as a means of quickly identifying

potential deficiencies of a building. When deficiencies are identified, further evaluation of

the structure, using Tier 2 and/or Tier 3 evaluation procedures, is required.

The steps involved in a Tier 1 analysis are as follows:

1. Identify if building is a benchmark building (see below for explanation).

2. If not a benchmark building, identify and complete appropriate checklists (basic and

supplemental) based on seismicity level, performance level, and building type.

3. Identify and complete appropriate foundation and non-structural component check-

lists.

4. Summarize deficiencies.

5. Determine if further evaluation (Tier 2 analysis) is required.

A benchmark building is one that is designed per specified design provisions (e.g. UBC

1997 or IBC 2000) and therefore does not require evaluation of the structural checklists.

Whether or not a building qualifies as a benchmark building is dependent on the design

professional verifying, through evaluation of construction documents and a site visit, that

the as-built condition of the building does indeed match the requirements of the bench-
372

mark provision. As the objective of this study is to investigate the appropriateness of the

structural checklists, all buildings evaluated were not considered to be benchmark buildings.

Checklists required are a function of level of seismicity, level of performance, and building

type. For the buildings evaluated in this study, the following checklists were required:

• Basic structural (ASCE/SEI 31-03 Section 3.7.9: Basic Structural Checklist for Build-

ing Type C2: Concrete Shear Walls with Stiff Diaphragms)

• Supplement structural (ASCE/SEI 31-03 Section 3.7.9S: Supplemental Structural

Checklist for Building Type C2: Concrete Shear Walls with Stiff Diaphragms)

• Geologic Site Hazard and Foundation (ASCE/SEI 31-03: Geologic Site Hazards and

Foundations Checklist)

• Basic, Intermediate, and Supplemental Nonstructural (ASCE/SEI 31-03 Sections 3.9.1

to 3.9.3)

Once checklists are completed, the design professional summarizes the deficiencies and

determines if further evaluation of the building is necessary. The factors that determine

what type of further analysis is needed include i) deficiencies present, ii) level of seismicity,

iii) level of performance, iv) building type and v) number of stories. For buildings that

exceed the story limits provided in ASCE 31 Table 3-3, a Tier 1 evaluation is considered

insufficient to adequately assess the building and a full building evaluation is required.

For buildings not exceeding these limits, it is permitted to do a Tier 2 deficiencies only

evaluation. The height of the buildings evaluated in this study exceed the limits provided

and thus a full building analysis is required.

Using the checklists listed above, Tier 1 evaluations were conducted for the four buildings

studied. The completed checklists are provided in Appendix E. A discussion of the most

critical checklist items for the buildings studied is provided in Section 9.1.
373

8.1.3 Tier 2

A Tier 2 Evaluation (Evaluation Phase) is intended to determine if building components

are adequate for a specific performance objective using a more in-depth evaluation than

provided by Tier 1 evaluation but without needing to conduct a full nonlinear analysis of

the building. A Tier 2 evaluation is necessary if deemed so based on the results of the Tier 1

analysis and/or the presence of specified building characteristics (i.e. number of stories). A

Tier 2 evaluation can consist of a full building analysis or a deficiency only analysis, in which

only aspects of the building deemed deficient (from the Tier 1 analysis) are analyzed. As

the objective of this study is to evaluate the provisions of ASCE/SEI 31-03 and ASCE/SEI

41-06, a full building analysis was conducted for the buildings studied.

A Tier 2 evaluation can be done using one or more of the following analysis types a)

linear static procedure (LSP), b) linear dynamic procedure (LDP), c) special procedure, or

d) procedures for nonstructural components. Special procedures are reserved for the analysis

of unreinforced masonry buildings with flexible diaphragms and nonstructural components,

which were not a focus of this study.

The LDP procedure is required for buildings taller than 100 ft (30.5 m); all of the

buildings studied exceed this limit. Thus, only the LDP procedures are described here. The

steps involved in an LDP are provided in the list below. The sections that follow discuss

each step in detail.

1. Develop a mathematical model of the building.

2. Develop a response spectrum for the site.

3. Perform a response spectrum analysis of the building.

4. Compute component actions.

5. Compare component actions with acceptance criteria.


374

8.1.3.1 Mathematical Model

This section summarizes the modeling requirements for a Tier 2 evaluation. The linear

mathematical models used in this study were created in SAP2000 [23]. Details of the model

are presented in Section 8.2.1.

ASCE 41 provides the effective stiffness that should be used for the building components.

These cross-sectional stiffness values are summarized in Table 8.1.

Table 8.1: Effective stiffness values for elastic building models.

Component Flexural Rigidity Shear Rigidity


Beams 0.30Ec Ig 0.40Ec Aw
Walls 0.50Ec Ig 0.40Ec Aw
Slabs 0.33Ec Ig 0.40Ec Aw

The model of the floor system depends on the type of diaphragm. Buildings with flexible

diaphragms can be modeled with vertical lines of the framing system acting independently.

Buildings with rigid diaphragms should explicitly account for the rigidity. Buildings with

stiff diaphragms should explicitly account for diaphragm flexibility. The buildings studied

here are shear wall buildings with stiff diaphragms.

ASCE 41 provides modeling criteria for modeling the foundation and soil-structure in-

teraction, however, these were not considered as a part of the study presented here.

8.1.3.2 Response Spectrum and Response Spectrum Analysis

The response spectrum used for the LDP may be either i) a mapped response spectrum or ii)

a site-specific response spectrum. A mapped response spectrum is based on mapped spectral

accelerations based on the soil conditions at the site of the building and is as specified by

ASCE/SEI 7 [6]. A site-specific response spectrum is based on geological, seismological,

and soil conditions at the building site and is defined as the mean spectra based on input

ground motions at the 2-percent/50-year probability of exceedance.


375

For the buildings studied, multi-directional effects were considered. ASCE 41 specifies

that ground motions be considered as 100% of the seismic forces in one horizontal directions

and 30% in the perpendicular horizontal direction, which can be achieved either by appling

both forces simultaneously or the effects of the two independent directions on an SRSS

basis; the later was used for the Tier 2 evaluations. Vertical ground motion, which must be

considered for horizontal cantilevers and prestressed elements, were not considered in this

study.

8.1.3.3 Component Actions, Deformations, and Acceptance

To evaluate the adequacy of the structure, component actions/demands are compared to

the component strength. The actions are determined from the LDP and must consider the

combined effects of gravity loads and seismic forces. Gravity loads, QG , are calculated as

1.1(QD + QL + QS ) and 0.9QD , where QD is the design dead load, QL is the design live

load, and QS is the snow load (assumed equal to zero for the buildings studied). Seismic

forces, QE , are determined from the response spectrum analysis.

QG = 1.1(QL + QD + QS ) (8.1)

QG = 0.9QD (8.2)

QE = from SAP RS analysis (8.3)

= ±X ± 0.3Y (8.4)

= ±Y ± 0.3X (8.5)

(8.6)

Actions are classified as either force- or deformation-controlled, with each type hav-

ing different criteria for calculating the component actions/demands and the component

strength. In reinforced concrete structural walls, shear and moment are deformation-
376

controlled components and axial is a forced-controlled component. The maximum actions

for each component are determined from an envelope of all possible load combinations,

including considerations for multi-directional effects.

The combined action of deformation-controlled actions, QU D , is calculated as QG ± QE .


QE
The combined action of force-controlled actions, QU F , is calculated as QG ± C1 C2 J . The

product of C1 and C2 was assumed to be equal to 1.25. The value of J is intended to account

for the contribution of additional components and is dependent on the level of seismicity,

the target performance level and whether or not actions introduced by adjacent components

are expected to remain elastic. J is equal to 1.0 for the Immediate Occupancy performance

level and 2.0 for all other performance levels.

QU D = QG ± QE (shear and moment) (8.7)


QE
QU F = QG ± (axial) (8.8)
C1 C2 J

Component strengths are classified as the nominal strength, QCL , for force-controlled

actions and the expected strength, QCE , for deformation-controlled actions. For reinforced

concrete components, the expected strength is equal to 1.25 times the nominal strength. The

nominal flexural strength was calculated from a fiber-section moment-curvature analysis.

The nominal shear strength of the walls was calculated using the provisions of ACI 318-11

Chapter 21. The nominal axial strength of the walls was taken as Ag fc0 .

QCE = expected strength (8.9)

QCL = nominal strength (8.10)

When considering deformation-controlled actions in reinforced concrete structural walls,

a wall must be classified as either flexure- or shear-controlled. This classification is used


377

in defining acceptance criteria for the deformation-controlled actions and is determined by

assuming a uniform lateral load distribution. Using this lateral load distribution, the shear

at the nominal moment strength is calculated. If this value is greater than the shear strength

of the component, the wall is considered shear-controlled. If the shear at nominal strength

is less than the shear strength of the component, the wall is considered flexure-controlled.

Two methods of comparing component actions to component strength are required by

ASCE 41. The first uses the demands from the SAP2000 modification, while the second

modifies these demands to account for the expected ductility of the components. The

unmodified demands are used to determined if the LDP procedure is appropriate for analysis

of the structure. The unmodified demand-capacity ratio, DCRU , is defined as the ratio

of the component demand (QU D or QU F ) to the component strength (QCE or QCL ). If

DCRu for all components and actions are less than 2.0, use of the LDP is appropriate.

If DCRU is greater than 2.0, than LDP is inappropriate and a Tier 3 evaluation, which

requires a nonlinear model, must be used to determine if building components are adequate.

Additionally, ASCE 41 indicates that the critical components in the building are those with

the largest DCRU values.

QU D
DCRu = (shear and moment) (8.11)
QCE
QU F
DCRu = (axial) (8.12)
QCL

The second method of comparing the component demands to component strength mod-

ifies the component strengths. Under the LDP provisions of ASCE 41, the model generates

reasonable displacements for the components but the forces necessary to generate these

displacements are larger than the forces the components are expected to experience. To

account for these larger forces, ASCE 41 increases the component strength of deformation-

controlled actions by multiplying the strength by an m-factor. The component strength


378

of force-controlled actions is not modified. The specific value of the m-factor is based

on if a component is shear- or flexure-controlled, the design shear demand, the axial de-

mand, boundary elements are confined, and the performance level considered. Table 8.2

provides the m-factors for reinforced concrete structural walls. The modified component

strengths are compared to the component actions/demands to determine if the components

are acceptable.The acceptability of components subjected to deformation-controlled and

force-controlled components, respectively, are determined based on the following equations:

QU D
QCE ≥ (Deformation-controlled actions) (8.13)
m
QCL ≥ QU F (Force-controlled actions) (8.14)

For simplicity, these relationships are defined as factored demand-capacity ratios, DCRf .

If DCRf is less than or equal to 1.0, the component is considered adequate. If DRF is

greater than 1.0, the component is considered inadequate.

QU D
DCRf = (shear and moment) (8.15)
mQCE
QU F
DCRf = (axial) (8.16)
QCL

8.1.4 Tier 3

A Tier 3 Evaluation (Detailed Evaluation Phase) provides the most in-depth evaluation

of a structure. While it is intended to provide the most accurate evaluation of a building

or of individual components, a Tier 3 Evaluation is also the most time-consuming and

expensive evaluation tool available for assessing the seismic adequacy of a structure. Thus,

it is typically used only when indicated as necessary by the Tier 2 analysis.

The decision of whether or not to conduct a Tier 3 Evaluation, the extent to which to

conduct the analysis, and the analysis methods employed require significantly more judg-
379

Table 8.2: m-factors for reinforced concrete shear walls (ASCE/SEI 41-06 [8, 9] Tables 6-20
and 6-21).

m-factors
Performance Level
Component Type
Primary Secondary
(As −A0s )fy +P 2 V√ 3
Description Confined?1 tw lw fc0 IO LS CP LS CP
tw lw fc0

Flexure controlled Y ≤ 0.1 ≤4 2 4 6 6 8


≥6 2 3 4 4 6
≥ 0.25 ≤4 1.5 3 4 4 6
≥6 1.25 2 2.5 2.5 4
N ≤ 0.1 ≤4 2 2.5 4 4 6
≥6 1.5 2 2.5 2.5 4
≥ 0.25 ≤4 1.25 1.5 2 2 3
≥6 1.25 1.5 1.75 1.75 2
Shear controlled N/A ≤ 0.05 N/A 2 2.5 3.0 4.5 6
> 0.05 N/A 1.5 2 3 3 4
1
Confined if the transverse reinforcement provides 75% of ACI 318 requirements
and spacing does not exceed 8db
2
P equal axial load due to gravity and earthquake ( (+) compression)
3
Design shear force

ment than do these same considerations for Tier 1 and 2 Evaluations. Consequently the

guidelines for Tier 3 evaluations provided in ASCE/SEI 31-03 are much more basic, indi-

cating that Tier 3 evaluations must be conducted using design procedures either intended

for new construction or the seismic rehabilitation of existing structures.

For this study, a Tier 3 evaluation was done using the nonlinear dynamic procedure

of the ASCE/SEI 41-06 Standard. For the ATC 94 project, models were developed using

both OpenSees [65] and Perform3D [24]. The Perform3D model was created to be a typical

model used in a design office and was developed by the practicing engineer members of

the project. Thus, the details and results of the model, as a well as a comparison to the

OpenSees model, are not provided here (see ATC-94 [75]). The OpenSees models, created

by the author, were created using methods common to evaluation of structures in a research
380

environment and are presented in this chapter.

Unlike the Tier 1 and 2 evaluation procedures, no specific procedures are available for a

Tier 3 evaluation. The following procedures were followed for this study:

1. Develop a mathematical building model.

2. Select a series of ground motions according to ASCE 7 provisions. Determine appro-

priate scale factors for DBE and MCE level events.

3. Perform time-history analysis for each ground motion.

4. Evaluate global demands.

5. Compute/extract rotation demands for comparison to acceptance criteria.

6. Identify deficient components.

Evaluation of the acceptability of the component performance was done using a) the

acceptance criteria from ASCE 41, b) fragility functions presented in Chapter 7, c) strains

corresponding to specific damage levels, and d) shear strength limits.

8.2 Overview of Models

This section provides a brief overview of the models used for Tier 2 and Tier 3 evaluation

of the buildings. The requirements of the models for each tier were presented in Sec-

tions 8.1.3 and 8.1.4, respectively, and are not repeated here. Section 8.2.1 provides details

on the Tier 2 models created in SAP2000 [23] for three of the four buildings studied (all but

Concepto Urbano). Section 8.2.2 provides details on the Tier 3 model created in OpenSees

for Plaza del Rio Building A.

8.2.1 Tier 2 Evaluation Models

Elastic models were used to determine modal properties of the buildings and to conduct Tier

2 evaluations using the linear dynamic procedure (LDP). Models were created in SAP2000

[23]. The models were developed following the guidelines provided in ASCE/SEI 41-06.
381

Thick shell elements were used to model the walls. Beams were modeled using elastic

beam column elements. Slabs were modeled with thick shell elements. Slab openings were

not considered and rigid diaphragm constraints were not used. The effective stiffness values

used are summarized in Table 8.1. To determine the optimal mesh size for the shell elements,

a mesh study was conducted on each building to determine the coarsest mesh size as which

the periods and distribution of shear force converged. The auto-meshing feature of SAP2000

was utilized and the maximum element dimensions were determined by dividing the typical

floor height in each building into a specified number of elements. The same meshing rules

were applied to all shell elements (walls and slabs) in the model. For the buildings studied, it

was determined that the optimal discretization was equal to four elements per floor height.

ASCE 41 provides modeling provisions for modeling the foundation. The impact of the

foundations was not accounted for in this study, rather a fixed-based was used, thus, these

provisions are not discussed here.

Table 8.3: Gravity loads used in creating the elastic building models

Building Dead Load Live Load


CM Self + 0.1 tonf/m2 (20.5 psf) 0.25 tonf/m2
PR-A Self + 0.1 tonf/m2 (20.5 psf) 0.25 tonf/m2
PR-B Self + 0.1 tonf/m2 (20.5 psf) 0.25 tonf/m2

Gravity loads were applied to the structures using the live and dead load values provided

in Table 8.3. Building masses were assigned in the horizontal directions using the total dead

load plus 25% of the live load. Masses were applied at the nodes and distributed based on

the tributary floor area associated with each node. Modal analysis of the structures were

conducted for a maximum of 50 modes, with the corresponding mode shape defined by the

corresponding eigen vector.

The fundamental periods were determined for each building using the gross section

properties (uncracked) and the effective section properties (cracked) provided in Table 8.1.
382

The uncracked periods were calculated for use in evaluating the Chilean design spectrum.

The cracked and uncracked periods for each building studied are presented in Section 9.2.3.

Tier 2 evaluation was done by conducting a response spectrum analysis of the models.

ASCE 41 permits the use of either a mapped response spectrum or a site-specific response

spectrum. As the goal of the this study was to evaluate the ASCE 41 evaluation criteria,

rather than use the ASCE 41 evaluation criteria to evaluate a particular structure, the

following spectra were used:

1. Mapped design response spectrum as specified (same as ASCE/SEI 7-10 spectrum)

2. Mapped design response spectrum as defined per Chilean code

3. Mapped design response spectrum as defined per proposed revisions to the Chilean

code

4. A site specific spectrum based not on ground motions for the 2-percent/50-year prob-

ability of exceedance, but rather on recorded ground motions from the event studied.

The sections that follow provide details on these spectra. Discussions of the spectra include

reference to the buildings studied, details of which can be found in Section 8.3, and the

fundamental periods of these buildings. The periods for each building were obtained from

modal analysis using SAP2000 [23], of which the details are provided above. A summary of

the results of the response spectra analyses are provided in Section 9.2.

8.2.1.1 ASCE Spectrum

The provisions of ASCE/SEI 41-06 state that the response spectrum to be used for the

LDP is that from ASCE/SEI 7-10 [6]. The design response spectrum used to study the

buildings was created assuming building occupancy category II, site class D, importance

factor of 1, and reduction factor of 5. Mapped spectral values from ASCE/SEI 7-010 were

not available. The buildings were considered to be located in a high seismicity region

similar to that found in California and spectral values selected on this basis. The short

period (0.2 seconds) spectral acceleration, Ss , used was equal to 2.0 and the one-second
383

spectral acceleration, S1 , was equal to 1.0. Figures 8.1 -8.3 show the ASCE 7 design spectra

(same for all buildings studied). Two spectra are shown in each figure. The elastic spectra

(unreduced spectra) is used for the response spectra analysis. The reduced spectra is shown

for comparison of the inelastic demands from the ASCE and Chilean spectra (see next

section).

In addition to considering the design spectrum, the spectrum for the the Maximum

Considered Earthquake (MCE) was considered. ASCE 7 requires that the MCE spectrum

be created using the design parameters multiplied by a factor of 1.5. The MCE spectra was

used in combination with the Collapse Prevention (CP) performance level when evaluating

the adequacy of components. The DBE spectra was used in combination with the Immediate

Occupancy (IO) and Life Safety (LS) performance levels.

ASCE ASCE
NCh NCh − Trans.
NCh Proposed NCh − Long.
Concepcion NCh Proposed − Trans.
San Pedro NCh Proposed − Long.
Mode 1: trans Mode 1: trans
Mode 2: torsion Mode 3: long
Mode 3: long

2.5 0.5
Spectral Acceleration, Sa (g)

Spectral Acceleration, Sa (g)

2 0.4

1.5 0.3

1 0.2

0.5 0.1

0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Period, T (sec) Period, T (sec)
(a) Elastic (b) Reduced

Figure 8.1: Centro Mayor (CM) response spectra. ASCE and NCh. spectra are design
spectra.
384

ASCE ASCE
NCh NCh − Trans.
NCh Proposed NCh − Long.
Concepcion NCh Proposed − Trans.
San Pedro NCh Proposed − Long.
Mode 1: trans Mode 1: trans
Mode 2: torsion Mode 3: long
Mode 3: long

2.5 0.5
Spectral Acceleration, Sa (g)

Spectral Acceleration, Sa (g)

2 0.4

1.5 0.3

1 0.2

0.5 0.1

0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Period, T (sec) Period, T (sec)
(a) Elastic (b) Reduced

Figure 8.2: Plaza del Rio Building A (PR-A) response spectra. ASCE and NCh. spectra
are design spectra.
385

ASCE ASCE
NCh NCh − Trans.
NCh Proposed NCh − Long.
Concepcion NCh Proposed − Trans.
San Pedro NCh Proposed − Long.
Mode 1: torsion Mode 2: trans
Mode 2: trans Mode 3: long
Mode 3: long

2.5 0.5
Spectral Acceleration, Sa (g)

Spectral Acceleration, Sa (g)

2 0.4

1.5 0.3

1 0.2

0.5 0.1

0 0
0 0.5 1 1.5 2 2.5 3 0 0.5 1 1.5 2 2.5 3
Period, T (sec) Period, T (sec)
(a) Elastic (b) Reduced

Figure 8.3: Plaza del Rio Building B (PR-B) response spectra. ASCE and NCh. spectra
are design spectra.
386

8.2.1.2 Chilean Design Spectrum

Although the primary objective of this study was to evaluate the provisions of ASCE 41,

which requires the use of the ASCE 7 response spectrum, the buildings studied were not

designed using the criteria of ASCE 7, but rather using the criteria of the Chilean building

code (NCh433 Of.96 [50]). Based on the performance of buildings during the 2010 Maule

earthquake, changes to the response spectrum in the Chilean building code were proposed.

Both the original and proposed Chilean design spectra were considered in this study.

Using NCH433 Of.96, the buildings studied were within building category C, seismic

zone 3, soil type III, and importance factor 1. In the Chilean building code reduction of

spectral accelerations is calculated using the factor R∗ which is calculated based on the

modal period corresponding to the direction of seismic analysis. This differs from ASCE 7

which uses the same R factor for all periods. The reduced spectral accelerations are defined

as:

IAo α
Sa = (8.17)
R∗  
p
1 + 4.5 TTno
α =  3 (8.18)
1 + TTno
T∗
R∗ = 1 + T∗
(8.19)
0.10To + Ro

where I is a coefficient relative to the importance, use and failure risk of the building;

Ao is the maximum effective acceleration and is determined based on the Chilean national

seismic zone; Tn is the vibration period of mode n; T ∗ is the period of the mode with highest

translational equivalent mass in the direction of analysis; To and p are parameters based on

soil type; Ro is the structural response modification factor for spectral analysis and is equal

to 11.

Figures 8.1 - 8.3 show the unreduced (R∗ = 1.0) and reduced design spectra for each

building. The reduced spectra are shown because there was concern as to the appropri-
387

Table 8.4: Comparison of code based spectral accelerations at fundamental periods.


ASCE ASCE
Sa Sa
Building Spectrum Mode SaASCE SaN Ch SaN ChP N Ch
Sa N ChP
Sa

CM Elastic Trans. 0.96 1.11 1.26 0.87 0.76


Long. 1.33 1.61 1.61 0.83 0.83
Reduced Trans. 0.19 0.18 0.20 1.10 0.96
Long. 0.27 0.30 0.30 0.89 0.89
PR-A Elastic Trans. 1.33 1.66 1.66 0.80 0.80
Long. 1.33 1.62 1.62 0.82 0.82
Reduced Trans. 0.27 0.33 0.33 0.81 0.81
Long. 0.27 0.42 0.42 0.63 0.63
PR-B Elastic Trans. 1.33 1.71 1.71 0.78 0.78
Long. 1.33 1.50 1.50 0.89 0.89
Reduced Trans. 0.27 0.36 0.36 0.74 0.74
Long. 0.27 0.42 0.42 0.64 0.64

ateness of comparing the elastic response spectra for the U.S. and Chilean design codes,

especially when evaluating results from the Chilean spectrum using ASCE 41 acceptance

criteria (m-factors) developed on the basis of the U.S. spectrum. For each building, the

periods of the fundamental mode shapes are shown to support this discussion. Of primary

concern was the spectral accelerations of the first mode. For Centro Mayor, these accelera-

tions were nearly identical, while the first mode spectral accelerations of both Plaza del Rio

buildings were slightly higher using the Chilean spectrum. These differences, summarized

in Table 8.4, were considered to be small enough that there was no need to adjust the

acceptance criteria when using the Chilean spectrum.

8.2.1.3 Spectra for Recorded Ground Motions

Figures 8.1a, 8.2a,and 8.3a show, in addition to the design spectra for the U.S. and Chilean

building codes, the response spectra generated using two ground motion accelerations that

were recorded during the Maule earthquake. The two motions were the Concepcion and San

Pedro de la Paz ground motions and were selected due to the proximity to the buildings

studied. Table 8.5 summarizes the location and soil types of the two recorded ground

motions. Table 8.6 provides the spectral accelerations for the translational and longitudinal
388

response modes for each ground motion, as well as a comparison the the ASCE 7 design

spectral values at these periods.

Table 8.5: Location and soil type of the ground motions used to generate site-specific
response spectra.

Location
Record Name Latitude Longitude Soil Description
Concepcion 36◦ 49’ 33.96” S 73◦ 3’ 16.92” W Silty Sand
San Pedro de la Paz 36◦ 50’ 39.12” S 73◦ 6’ 31.32” W Clay/gravel on rock

The Concepcion spectra was recorded in the same city as the buildings on silty sand and

had the highest spectral acceleration occur at a long period of approximately 1.5 seconds1 .

This long period response was one of the factors used in the proposed Chilean spectra mod-

ifications, for which an increase in spectral accelerations can be seen at long periods relative

to the spectra from Chilean code in affect at the time of the earthquake and the spectra

from U.S. code. The maximum spectral acceleration was lower than the maximum design

spectral accelerations for either the U.S. or Chilean codes, and at the first mode periods for

all three buildings studied, the Concepcion spectral accelerations were significantly lower

than the design spectral accelerations.

The San Pedro de la Paz record was recorded in the same city as the building sites on

clay and gravel. The spectrum generated from this ground motion had different charac-

teristics than did the shape of the Concepcion response spectrum. At long periods, the

San Pedro spectrum is similar to the ASCE and NCh design spectra. The maximum spec-

tral accelerations were found at a much lower period of approximately 0.2 seconds, which

is approximately the left-most point on the plateau in the ASCE spectrum. For Centro

Mayor, the San Pedro de la Paz spectral accelerations at the first, second and third modes

were nearly identical to those from the ASCE spectrum and slightly lower than the NCh

1
The acceleration records provided indicate that this may have been due to a shift in the instrumentation
during the earthquake
389

spectrum. For Plaza del Rio Building A, the first mode spectral acceleration was similar to

the ASCE value; all other values were significantly lower than the design accelerations. For

Plaza del Rio Building B, the San Pedro de la Paz spectral accelerations were lower than

the design spectral accelerations.

Table 8.6: Comparison of spectral accelerations from ground motion records at fundamental
periods.
ASCE ASCE
Sa Sa
Building Mode SaASCE SaConcepcion SaSanP edro Concepcion SanP edro
Sa
Sa

CM Trans. 0.96 1.26 0.78 2.29 1.23


Long. 1.33 0.63 1.45 2.13 0.92
PR-A Trans. 1.33 0.64 1.19 2.09 1.12
Long. 1.33 0.84 0.97 1.58 1.38
PR-B Trans. 1.33 0.66 0.96 2.03 1.38
Long. 1.33 0.97 1.05 1.37 1.27

8.2.2 Tier 3 Evaluation Models

Due to the larger amount of time and resources necessary for nonlinear analysis, a Tier

3 evaluation was conducted for only one building. Plaza del Rio Building A was selected

due to i) the rectangular plan with regular wall distribution, ii) well documented damage

information, and iii) the availability of nearby ground motions. Nonlinear analyses were

conducted in OpenSees [65] and Perform3D [24]. A discussion of the OpenSees modeling

techniques and results are presented here; details of the Perform3D model and a comparison

to the OpenSees results is presented in the ATC-94 project report [75].

The OpenSees model was created using nonlinear force-based beam-column elements

located at the the geometric centroid of the elements with one element used per floor of

each wall. For walls where the centroid was not at the same location from one floor to the

next, the continuity of the load path was ensured by connecting the top of the lower element

to the bottom of the upper element with an elastic beam with very large stiffness. Where

walls terminated, a continuous load path was ensured by providing very stiff transfer beams
390

connecting the bottom of the terminated wall to centroids of adjacent walls. The beams

and slabs were not explicitly modeled; instead, a rigid diaphragm constraint was used at

each floor to ensure that the walls acted as a system. Lumped masses were assigned at the

nodes of the walls at each floor level based on gravity loads. The gravity loads used were

the same as those used in the SAP2000 model for the Tier 2 LDP analyses.

Five integration points were used per element (i.e. five integration points for each floor

to floor segement), with each integration point consisting of an aggregated section consisting

of independent response of the wall to i) torsion, ii) shear, and iii) flexural and axial loading.

An elastic torsional stiffness of 0.5GJ was assumed.

The nonlinear shear response was simulated using a cyclic shear model with a trilinear

backbone curve, shown in Figure 8.4. The three parts of the backbone curve were as follows

i) a stiffness of GA to the approximate cracking strength2 , ii) a stiffness of 0.40GA to 0.60Vn

(chosen for consistency with parameters for shear controlled walls in ASCE 41), and iii) a

stiffness of 0.10GA thereafter. These stiffness values that define the backbone curve were

determined to be realistic on using the stiffness values presented in Section 6.3.2 for the

four UW-UIUC planar wall tests. The shear model was implemented using the Hysteretic

material model in OpenSees with pinching factors of 1.0 and 0.103 . The OpenSees material

model used prevented the use of strength limits, thus, this information was post-processed

from the model results.

The flexure and axial response was generated using fiber sections. Implementation of

the flexural component of the force-based beam column elements followed recommendations

by Pugh [89]. Pugh used a large dataset of experimental wall subassembly tests to validate

modeling techniques that accurately captured the stiffness, strength, and drift capacities of

flexural-controlled walls. The basis for Pugh’s recommendations is that distributed plasticity

elements have the potential for localization of damage at the critical section and that the

2 √
2 fc0 based on the experimental database discussed in Section 7.
3
Ideally a value of 0.0 would be used but the use of 0.10 instead eliminated a number of convergence
problems otherwise encountered
391

0.6Vn 0.1GA
0.4GA

p
2 fc0 Ag

GA

Shear strain
Figure 8.4: Nonlinear shear backbone curve used in Tier 3 OpenSees model.

global response is consequently sensitive to the number of integration points used. To

achieve an objective response regardless of mesh size, Pugh proposed recommendations

for post-peak regularization of the stress-strain response of the steel and concrete material

models and was subsequently able to significantly improve the prediction of slender wall drift

capacity using force-based elements. Regularization of the post-peak response for concrete

in compression used a fracture energy of 2.0fc0 (fc0 in MPa). The fracture energy for steel

was calculated assuming an ultimate strain of 16%.

8.2.2.1 Ground Motions

Ground motions were used to conduct time history analyses of the buildings with the objec-

tive of (a) understanding the response of the structure during the earthquake, including the

progression of damage in the structure, (b) evaluating ASCE 41 NDP acceptance criteria,

and (c) evaluating the fragility functions presented in Chapter 7.

Time history analyses were conducted using multiple ground motion records. Ideally, the

specific motion experienced by the building would be used to conduct an evaluation of the
392

building. Although no such records were available for the site studied, ground motions from

nearby sites were available. In selecting ground motions to use in studying the buildings,

only records from sites with similar soil types were considered. The sites closest to the

building site were Concepcion and San Pedro de la Paz (the response spectra for these

motions were also used for Tier 2 analysis). Five additional records were considered from

similar soil sites that were recorded during the Maule earthquake. All seven records, which

are summarized in Table 8.7, were corrected acceleration records available through the

ATC-94 project.

The Concepcion and San Pedro de la Paz records were both used to conduct time history

analyses of the building model to assess the response of the structure during the earthquake.

Time history analyses were run using these acceleration records without any scale factors

applied.

To assess the Tier 3 procedures, all seven ground motions acceleration records sum-

marized in Table 8.7 were scaled to the ASCE 7 response spectra for the building4 . The

response spectra is described in Section 8.2.1.1. The ground motions were scaled following

the amplitude scaling method of ASCE 7 Chapter 16. This method was chosen over spectral

scaling to increase the likeliehood of capturing unique behaviors excited by varied records.

The scaling was done over the period range of 0.2T to 1.5T , where T is the approximate

period of the first mode shape (T = 0.64 seconds for Plaza del Rio Building A). Scale factors

were determined for the ASCE 7 DBE spectrum; multiplying these factors by an additional

1.5 factor enabled study of the ASCE 7 MCE. The amplification factors are provided in

Table 8.8.

For each of the analysis runs described in the preceding two paragraphs (a total of 18),

a large number of recorders were used in the OpenSees models to capture the response

of the structure. To monitor the global response of the structure, nodal displacements,

accelerations, and damping forces were recorded at all nodes in the model. The distribution

4
Note: Scaling of the ground motion records was done by the ATC-94 project team, not by the author.
393

Table 8.7: Summary of ground motions studied.

Location Peak ground acceleration (g)


Record Name Latitude Longitude Dir. Value Dir. Value
Concepcion -36.8261◦ (S) -73.0547◦ (W) Long. 0.402 Trans. 0.284
Llolleo -33.6167◦ (S) -71.6176◦ (W) Long. 0.319 Trans. 0.564
Matanzas -33.9593◦ (S) -71.8727◦ (W) Long. 0.342 Trans. 0.308
San Pedro de la Paz -36.8442◦ (S) -73.1087◦ (W) E 0.605 N 0.65
Valparaiso Almendral -33.0458◦ (S) -71.6068◦ (W) Long. 0.224 Trans. 0.265
Vina del Mar Centro -33.0253◦ (S) -71.5508◦ (W) NS 0.219 EW 0.334
Vina del Mar El Salto -33.0469◦ (S) -71.51◦ (W) NS 0.351 EW 0.338

Table 8.8: Factors to scale ground motions to ASCE 7 DBE response spectrum

Record Name Scale Factor1


Concepcion 1.737
Llolleo 1.113
Matanzas 1.330
San Pedro de la Paz 1.495
Valparaiso Almendral 1.946
Vina del Mar Centro 1.516
Vina del Mar El Salto 1.339
1 Scale factors are for DBE. Multiply by 1.5 for MCE.
394

of shear forces throughout the building was enabled by recording element forces at each

integration point (five per element). The corresponding element deformations were recorded

to enable the use of rotation limits provided by the ASCE/SEI 41-06 evaluation procedures

and the fragility functions proposed in Chapter 7. For additional evaluation of performance

states, the stress-strain response of extreme steel and concrete fibers was recorded in the

lowest two floors of the structure where most nonlinear behavior was expected to occur5 .

8.2.2.2 Acceptance Criteria

ASCE 41-06 Section 3.4.3, which provides provisions for the evaluation of an existing struc-

ture analyzed using the nonlinear dynamic procedure (NDP), requires that the drifts from

all ground motions considered must not exceed the design drifts for the structure. Addition-

ally, ASCE 41-06 Chapter 6 provides acceptance criteria for flexure- and shear-controlled

walls that are used to determine the adequacy of components (see Section 8.1.3). For the

study undertaken, it was desired to evaluate all potential methods for evaluating a structure

and the following evaluation criteria were considered:

1. Acceptable rotation limits based in ASCE 41-06 for flexure-controlled walls

2. ASCE 41-06 usable strain limits

3. Strain values indicating first occurrence of reinforcement yield and concrete spalling

and crushing

4. Fragility functions presented in Chapter 7

5. Shear capacity

The ASCE 41-06 acceptance criteria for flexure-controlled walls are defined as plastic

hinge rotations, where the plastic hinge length is approximated as one-half of the wall length.

Acceptable hinge rotations, summarized in Table 8.9, are determined as a function of the

axial load, confinement, shear demand, and target performance level.

5
Limitations on the number of output files prevented the collection of strain data throughout the structure
395

Table 8.9: Acceptance criteria for reinforced concrete shear walls (ASCE/SEI 41-06 Tables
6-18 and 6-19).

m-factors
Performance Level
Component Type
Primary Secondary
(As −A0s )fy +P 1
Description Confined?1 tw lw fc0
V√
IO LS CP LS CP
tw lw fc0

Flexure controlled3 Y ≤ 0.1 ≤4 0.005 0.010 0.015 0.015 0.020


≥6 0.004 0.008 0.010 0.010 0.015
≥ 0.25 ≤4 0.003 0.006 0.009 0.009 0.012
≥6 0.0015 0.003 0.005 0.005 0.010
N ≤ 0.1 ≤4 0.002 0.004 0.008 0.008 0.015
≥6 0.002 0.004 0.006 0.006 0.010
≥ 0.25 ≤4 0.001 0.002 0.003 0.003 0.005
≥6 0.001 0.001 0.002 0.002 0.004
Shear controlled4 N/A ≤ 0.05 N/A 0.40 0.75 1.0 1.5 2.0
> 0.05 N/A 0.40 0.55 0.75 0.75 1.0
1 Confinement if transverse reinforcement provides 75% of ACI 318 requirements
2 P equal axial load due to gravity and earthquake ( (+) compression)
3 Flexure-controlled acceptance criteria are plastic rotations.
4 Shear-controlled acceptance criteria are drift values.

As with the acceptance criteria for Tier 2 evaluations, walls determined to be shear

controlled have separate acceptance criteria based on story drifts. Two walls6 in Plaza del

Rio Building A were deemed to be shear-controlled, however, because a separate evaluation

of walls reaching the shear capacity was conducted, these walls were assumed to be flexure-

controlled for simplicity as the shear evaluation flags them as shear critical.

ASCE 41 determines flexure- or shear-control based on an assumption of a uniform

lateral load distribution, however, studies by other researchers have found that the actual

load distribution can vary significantly. Given the possibility of a lower effective height and

consequently a higher shear demand, it was desired to add an additional check beyond that

of ASCE 41 to evaluate if the response of the individual components were shear critical.

6
The large L-shaped walls on the second floor and above
396

For simplicity, a wall was considered to be shear critical if the shear demand exceed the

shear capacity by more than 10%, where the shear capacity is defined as the expected shear

strength of the wall. A 10% overstrength was included based on the results by Pugh [89]

that indicates experimental tests of slender walls do not have shear demand capacity ratios

that exceed 1.10.

ASCE 41 Section 6.6.3.1 provides usable strain limits for reinforced concrete components.

These limits are very general (0.005 for concrete in compression and 0.05 for steel in tension),

and although they were initially evaluated as a part of this study, they were found to be poor

indicators of the acceptability of the performance of the building components and are not

presented. Instead, strain values corresponding to key damage states were evaulated. For

reinforcing steel, initial yield in tension and the maximum tensile strain were considered.

For concrete, spalling was considered to be the strain at which there was a 10% loss of

compressive strength and crushing was considered to be the strain at which there was an

80% loss of compressive strength.

In additional to the ASCE 41 acceptance criteria and the strains listed in the previous

paragraph, the rotation based fragility functions developed in Chapter 7 were evaluated.

The fragility function rotations, like those used specified by the ASCE 41-06 acceptance

criteria, are based on a lumped plasticity element with a plastic hinge length equal to one-

half the wall length. In order to ensure an accurate evaluation of rotation based criteria,

it is necessary to understand the characteristics of rotations from the distributed plasticity

and lumped plasticity models.

Figure 8.5a shows the load-drift for pushover analysis of three different models. The solid

blue line shows the results using a distributed plasticity element with the material regular-

ization recommended by Pugh [89]. The dashed red line shows the results using a lumped

plasticity model with material regularization taken into consideration. The magenta line

shows the results of the lumped plasticity model without material regularization. Although

the drift capacities are not identical for the two regularized models, the regularized lumped
397

plasticity model achieves a drift capacity much closer to the distributed plasticity than does

the non-regularized lumped plasticity model. Figure 8.5b shows rotations from the same

models. The dashed red and dotted magenta lines show the rotations from the regular-

ized and non-regularized rotations, respectively. The dashed green line shows the empirical

lumped plasticity rotations. These rotations were calculated in an identical manner to the

rotations used to develop the fragility functions with the drift from the distributed plasticity

models considered to be the true drift of the wall modeled. The empirical and regularized

lumped plasticity rotations, are very similar, however, they are much greater than the base

rotations from the distributed plasticity model, shown by the light blue solid line. The

base rotations were calculated as the curvature at the lowest section in an element times

the characteristic length. In order for rotations from the distributed plasticity model to be

comparable with the lumped plasticity rotations, it is necessary to consider the cumulative

rotation at multiple sections. The solid blue line shows the sum of all rotations at integra-

tion points within the length of the plastic hinge used by the lumped plasticity models. This

rotation, referred to as the equivalent plastic hinge rotation, has stiffness up to the peak

strength that closely matches that of the lumped plasticity rotations. The rotation capacity

using this method is larger than that for the regularized lumped plasticity model and the

empirical lumped plasticity model. Consequently, evaluation of the ASCE 41 acceptance

criteria and the fragility functions using the equivalent plastic hinge rotations is likely to

result in a conservative estimation of the number of walls exceeding acceptable limits.

Calculation of the equivalent plastic hinge rotations in the distributed plasticity model

required the use of curvatures from multiple integration points. For some walls, the plastic

hinge length of one-half the wall length was greater than one floor height and the use of

multiple elements was necessary.

8.3 Overview of Buildings Studied

Four mid-rise (12 − 24 stories) concrete buildings located in the city of Concepcion were

studied. The buildings were selected on the basis of available design drawings and the
398

0.8

0.6
V/Vmax

0.4

0.2 D.P (Reg.)


L.P. (Reg.)
L.P. (Unreg.)
0
0 0.5 1 1.5 2
Roof Drift, %
(a) Drifts

0.8

0.6
M/Mmax

0.4
D.P. Rotation
D.P. Equivalent Plastic Hinge Rotation
0.2 L.P. Rotation (Reg.)
L.P. Rotation (Unreg.)
Empirical L.P. Rotation
0
0 0.005 0.01 0.015 0.02
Rotation, radians
(b) Rotations

Figure 8.5: Generic wall model results to demonstrate the variation in rotation demands
resulting from lumped and distributed plasticity models.
399

availability of damage documentation. All four buildings were apartment or condominium

buildings with reinforced concrete bearing wall systems that served as both the gravity and

lateral systems. Two of the buildings were severely damaged, condemned, and eventually

demolished. A third building, immediately adjacent to one of the severely damaged build-

ings, was only moderately damaged in a few isolated locations. The forth building was not

reported to have sustained any damage.

The following sections provide a brief overview of all four buildings; more extensive

descriptions are available in the appendix. As a part of the building summaries, the most

critical floor plans in the buildings are shown with colors that indicate the level of damage

sustained by the walls on these floors. The discussion of damage is limited to that observed

in the walls as walls were the primary focus of this study. The damage levels used were

based on a report describing the damage to one of the buildings [27] and are summarized

by Table 8.10. These damage levels are displayed throughout this chapter using the color

scheme shown in Table 8.10. The damage levels were determined by photos and other

reconnaissance data available through the ATC 94 project team.

The following sections present an overview of the four buildings studied. Section 8.3.1

describes towers A and B of the Plaza del Rio apartment building. The two towers were

severely and moderately damaged, respectively. Tower A was the primary focus of the

work presented here. Section 8.3.2 describes the severely damaged Cenro Mayor building.

Section 8.3.3 describes the undamaged Concepto Urbano building.

8.3.1 Plaza del Rio (Buildings A & B)

Plaza del Rio Buildings A and B were adjacent to one another and separated by a con-

struction joint. They were both constructed in 2006. The two buildings form an L-shaped

configuration, with Building A (PRA) forming the web (oriented up and down on the figures

provided) and Building B (PRB) forming the flange (oriented left to right on the figures

provided). Figure 8.6 shows both buildings. The footprint of the buildings is shown in Fig-
400

ure 8.7. In both buildings, the walls were 15 cm (6 inches) thick and the slabs were 12 cm (5

inches). Building A was 12 stories tall and Building B was 13 stories tall. Neither building

had basement levels. Foundations were mat foundations 40 cm (16 inches) thick. The soil

type at the site was Type III, medium dense gravel/medium stiff clay, per the Chilean code

[50], or Type D-E per ASCE/SEI 7 [6]. The design concrete strength was 2500 T/m2 (3.6

ksi).

Figure 8.6: Photo of Plaza del Rio buildings.

Figure 8.8 shows the wall layout on typical floors for Plaza del Rio Building A, with

color indicating the level of damage. The footprint had approximate dimensions of 79 feet

(24.1 meters) by 47 feet (14.3 meters) (principal direction was oriented in approximately

the N-S direction). A number of damage patterns were observed in the building. Figure 8.9

shows damage to z-shaped walls at the south end of the building. The pair of T-walls

oriented east-west in the center of building were also significantly damaged. This region,

shown in in Figure 8.10, sustained greater damage in the second floor than in the first

floor, despite the same reinforcement layout. In the second floor, damage consisted of core

crushing and bar buckling in the boundary element regions, but also extended along a

horizontal plane nearly the entire length of the wall. At first floor, crushing occurred at

the toe of the right wall, but not in the left wall, where the boundary element was shorter.

Reconnaissance teams measured significant residual displacements following the earthquake


401

Table 8.10: Summary of damage levels used to describe damage to buildings.

Name (#) Color Description


None (0) Gray No visible damage
Minor (1) Cyan Minor cracks
Moderate (2) Yellow Initial spalling
Severe (3) Orange Exposed reinforcement; Ini-
tial bar buckling and/or con-
crete crushing
Total (4) Red Extensive bar buckling and
concrete crushing

Figure 8.7: Floor plan of Plaza del Rio.


402

None
None
Minor
Moderate
Moderate
Severe
Severe
Total
Total

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(a) Floor 1 (f1) (b) Floor 2 (f2) (c) Damage legend

Figure 8.8: Plaza del Rio Building A: Layout of walls in typical floors. Color is used to
indicate the level of observed damage.

and the building was ultimately condemned and destroyed. Although not visible from the

photographs available to the project team (and thus available for documentation here), there

was reported evidence of significant uplift of this region of the building at the foundation

level suggesting that rocking occurred during the earthquake.

Figure 8.11 shows the wall layout on typical floors for Plaza del Rio Building B, with

color indicating the level of damage. The footprint had approximate dimensions of 131 feet

(39.9 meters) by 43 feet (13.1 meters) (principal direction was oriented in approximately the

E-W direction). The wall damage in Building B was primarily restricted to minor cracking.

The difference in damage between the two very similar buildings was a primary moti-

vation for including the buildings in the study. Ultimately, only one building in the study

was studied using nonlinear analysis. For this purpose, Plaza del Rio Building A was cho-

sen because significant damaged occurred, and unlike the other damaged building (Centro
403

l$#ffi
fffil tu{
.H*
-==lq
Y?
tfi

Iiil
llLl
HE
Pts

Figure 8.9: Damage at south end of Plaza del Rio Building A.


*L-ftJ,L*EJ*#*LJ*-J, L-*--{#.'L
-
*ry-
ff

fi: rL *{

*8
ltt
38

${n
$i1
ll n
il$
il-
I
L'"-

Figure 8.10: Damage at center T-walls in Plaza del Rio Building A.


404

(a) Floor 1 (f1)

None
Minor
Moderate

(b) Floor 2 (f2)

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total
None

(c) Damage legend

Figure 8.11: Plaza del Rio Building B: Layout of walls in typical floors.
405

Mayor), the layout of the walls was relatively simple.

8.3.2 Centro Mayor

Centro Mayor (CM), is an 18-story building with one basement level. The building was

constructed in 2005. The overall configuration of the building is shown in Figure 8.12. The

footprint of the building is shown in Figure 8.13. Figures 8.14 - 8.15 show the wall layout

on typical floors, with color indicating the level of damage. The basement level (f0 ) has

approximate dimensions of 134 feet (40.8 meters) by 57 feet (17.4 meters) (note: this floor

had perimeter walls not shown on the floor plans). Floor slabs were 15 cm (6 inches) thick.

The wall thickness was constant throughout the building at 20 cm (8 inches) thick. The

foundation was a mat foundation. The soil type at the site was Type II, dense gravel or

clay, per the Chilean code [50], or Type D-E per ASCE/SEI 7 [6]. The design concrete

strength was 2500 T/m2 (3.6 ksi).

Figure 8.12: Photo of Centro Mayor.

The damage levels were determined from a post-reconnaissance report provided by a


406

Figure 8.13: Plans of typical floors of Centro Mayor.

(a) Floor 0 (f0) (b) Floor 1 (f1)

None 0: No Damage None


Minor 1: Minor (cracking) Minor
Moderate Moderate
Severe 2: Moderate Severe
Total Total
3: Severe
4: Total

(c) Damage legend

Figure 8.14: Concepto Mayor: Layout of walls in basement levels. Perimeter walls are not
shown. Color is used to indicate the level of observed damage.
407

(a) Floor 2 (f2) (b) Floor 3 (f3)

None 0: No Damage None


Minor 1: Minor (cracking) Minor
Moderate Moderate
Severe 2: Moderate Severe
Total Total
3: Severe
4: Total

(c) Damage legend

Figure 8.15: Concepto Mayor: Layout of walls in typical floors. Color is used to indicate
the level of observed damage.

Chilean engineering firm [27]. Damage was primarily associated with loading in the EW

direction (up-down as shown in floor plans). The level and characteristics of damage varied

greatly throughout the building. Some of the highlights of the damage are detailed here.

Many walls had confining reinforcement spacing exceeding that permitted by ACI 318 and

the upper limit of 8db needed to consider the walls confined per ASCE/SEI 41-06. The

damage to these regions of confinement was extensive, with significant amounts of core

crushing, bar buckling, and in some locations, bar fracture (post-buckling). There were

a number of locations throughout the building where heavy damage occurred in the lower

portion of a flag-shaped wall or in walls adjacent to a terminated wall. At a couple locations,

severe damage occurred where two adjacent, coupled walls merged in the lower stories, with

significant crushing occurring at the region where the two upper walls meet. Sketches of

this configuration and the associated damage are shown in Figure 8.16.
408

(a) (b)

Figure 8.16: Centro Mayor (CM): Damage at intersection of coupled wall piers [27].

8.3.3 Concepto Urbano

Concepto Urbano (CU), is a 24-story building, of which two are below grade. The building

was under construction from 2008-2010 and was not occupied at the time of the earth-

quake. Figure 8.17 shows a photograph of the building. The footprint of the building is

shown in Figure 8.18, with the different shades indicating the plan at various heights. Fig-

ures 8.19 - 8.21 show the wall layout on typical floors. The basement levels (f 1 and f0 )

have approximate dimensions of 170 feet (51.8 meters) by 94 feet (28.7 meters) (note: these

floors have perimeter reinforced concrete walls not shown on the floor plans). Floor slabs

were 16 − 17 cm (6.3 − 6.7 inches) thick with the exception of floor f10 which has a pool

and thus a 20 cm (7.9 inches) thick slab. The typical wall thickness varied throughout the

building, ranging from 30 cm (12 inches) thick at the lower floor to 15 cm (6 inches). At the

lower levels, column-like wall piers at the ends of the building that supported much larger
409

walls were 40 cm (16 inches) thick. The foundations was a 50 − 190 cm (20 − 75 inches)

thick mat foundation. The foundation was a mat foundation. The soil type at the site was

Type II, dense gravel or clay, per the Chilean code [50], or Type D-E per ASCE/SEI 7 [6].

The design concrete strength was 2500 T/m2 (3.6 ksi).

Figure 8.17: Sketch of Concepto Urbano


410

Figure 8.18: Plan of Concepto Urbano. The various shades of indicate the plan at different
floors.

(a) Floor -1 (f-1) (b) Floor 0 (f0)


None None

Figure 8.19: Concepto Urbano: Layout of walls in basement levels. Perimeter walls are not
shown. No damage was recorded.
411

(a) Floor 1 (f1) (b) Floor 2 (f2)


None None

(c) Floor 3 (f3) (d) Floor 11 (f11)


None None

Figure 8.20: Concepeto Urbano: Layout of walls in typical floors in lower portion of building.
No damage was recorded.

(a) Floor 12 (f12)


None

Figure 8.21: Concepto Urbano: Layout of walls in floors in the upper portion of the building.
No damage was recorded.
412

Chapter 9

EVALUATION OF BUILDING DAMAGED IN 2010 MAULE


EARTHQUAKE: RESULTS

The preceding chapter presented background for a study of buildings damaged in the

2010 Maule earthquake in Chile. The study focused on evaluating buildings using the

methods detailed in the ASCE/SEI 31-03 and 41-06 standards. Section 8.1 provides an

overview of these standards and the three tiered approach used for building analysis. The

second and third tiers of the evaluation process involve evaluation of the structures using

linear and nonlinear models, respectively. Section 8.2 provides an overview of the models.

Four buildings were included in the study with most of the research effort focused on one

severely damage building, Plaza del Rio Building A. Details of this and the other buildings

are presented in Section 8.3.

This chapter presents the results of the study and recommendations for future improve-

ments to the ASCE 31 and 41 standards. Section 9.1 presents a summary of the Tier 1

evaluations for the four buildings. A Tier 1 evaluation involves performing a series of quick

checks to identify potential deficiencies and determine if any further evaluation is required.

Section 9.2 presents the results of the Tier 2 evaluations of the structures. A Tier 2 evalu-

ation requires linear analysis of a structure. Section 9.3 presents the results of a Tier 3 of

Plaza del Rio Building A. A Tier 3 evaluation requires nonlinear analysis of a structure. In

addition to performing the three-tiered evaluation, recommendations for improvements to

the ASCE 31 and 41 standards were determined as a part of this study and are presented

in Section 9.4.
413

9.1 Summary of Tier 1 Evaluations

ASCE/SEI 31-06 Tier 1 is considered the Screening Phase, with the intention that the build-

ing be evaluated quickly with minimal calculations, to identify any potential deficiencies that

require a more in-depth analysis. For the Chilean study buildings, these evaluations were

performed by practicing engineers on the project team and are not discussed in this thesis.

The completed Tier 1 checklists are available in the project report [75]. A brief summary

of the results of the Tier 1 evaluations is provided here to serve as background for the Tier

2 and 3 evaluations as well as to support general conclusions about the Tier 1 evaluation

procedures.

The primary deficiencies identified by the Tier 1 evaluations were as follows:

1. Presence of walls with “vertical discontinuities”, defined as walls that were not con-

tinuous to the foundation. Such walls were found in all four buildings studied.

2. Weak stories, defined as a reduction in total wall length that results in a strength

discontinuity from one floor to an adjacent floor. This type of deficiency was identified

in three of the four buildings studied (Concepto Urbano, Centro Mayor, and Plaza del

Rio Building A).

3. Soft stories, defined as a reduction in total wall length that results in a stiffness

discontinuity from one floor to an adjacent floor.

4. Shear stress demands exceeding allowable levels was observed in all buildings.

Additional identified deficiencies included:

1. Longitudinal reinforcement ratios less than then the specified minimum (CU).

2. Inadequate detailing of coupling beams (all buildings).

3. Inadequate detailing at wall openings where stress concentrations are expected (all

buildings).

4. Plan irregularities and reentrant corners (CM and CU).


414

9.1.1 Additional Qualitative Analysis

A comprehensive list of deficiencies identified by the project team is presented in Ap-

pendix E.1. This list includes deficiencies flagged by ASCE 31 assessment criteria as well

as those identified through evaluation of construction drawings, post-earthquake reconnais-

sance reports, and photos of damage. A summary of deficiences identified by the team but

not flagged by ASCE 31 assessment follows.

1. Significant loss of flange area: The ASCE 31 check identifies as non-compliant the

termination of an entire wall, but not length of or flange of a wall.

2. Uncertainty in the location of “critical sections”: This can be a result of unclear load

paths, geometric irregularities, and/or reinforcement irregularities that result in either

an unclear location of where flexural yielding will initiate and/or how that plasticity

will spread.

3. Inadequate slab thickness: A minimum slab thickness is necessary to ensure adequate

diaphragm stength to allow transfer of seismic load to the walls.

4. Pounding: Damage due to pounding between Plaza del Rio Buildings A and B was

observed but the potential for this to cause damage is not addressed by the ASCE 31

Tier 1 checklists.

5. Coupling with non-structural components: Damage was observed in Plaza del Rio

where the connection of secondary components to primary components (inverted

beams supporting and forming a wall at balconies) was not adequately detailed.

6. Detailing of confining reinforcement: For transverse reinforcement to provide effec-

tive confinement for boundary element concrete, transverse reinforcement must be

adequately detailed with cross-ties and limited spacing. Many walls in the build-

ings studied had slender boundary elements confined by hoops spaced at dimensions

greater than or equal to the wall thickness and without cross-ties; significant bar

buckling and concrete crushing was observed in many of these walls. Poorly detailed

transverse reinforcement should be flagged by the Tier 1 evaluation as non-compliant.


415

7. Discontinuous wall reinforcement: Changes to or termination of longitudinal rein-

forcement can cause stress concentrations in the region of the discontinuity and result

in damage.

The most notable of the above deficiences were the “loss of significant flange area”

and the “uncertain critical sections”. The project team referred to these issues collectively

as “local vertical discontinuities”; these were considered to be different from the “vertical

discontinuities” check included in ASCE 31. Unlike the ASCE 31 check, which defines as

non-compliant the complete termination of a wall above the foundation, a “local vertical

discontinuity” defines a wall that is continuous to the foundation but, at some point along

the load path, had a change in geometry that could potentially affect the transfer of load

to the foundation. Recommendations for modifying ASCE 31 Tier 1 evaluation to include

a check for local vertical discontinuities is discussed in detail in Section 9.4.2.

9.2 Summary of Tier 2 Evaluations

A Tier 2 evaluation requires linear analysis of a building and comparing the simulated

demands to the component strengths to determine if any walls are inadequate. Tier 2

evaluations were conducted for Centro Mayor (CM), Plaza del Rio Building A (PRA), and

Plaza del Rio Building B (PRB) using the linear dynamic procedures (LDP) of ASCE/SEI

41-06. Details of the evaluation process are provided in Section 8.1.3. Multiple response

spectra were used to conduct the response spectra analyses (i) ASCE 7 response spectrum,

(ii) Chilean design response spectrum, and (iii) two spectra derived from ground motion

records near the building sites. Results of the evaluation are summarized here.

The Tier 2 evaluation is based on factored and unfactored demand-capacity ratios. These

ratios were found to be highest in the lowest typical floor of the buildings studied. Specifi-

cally, these were floor f1 for Plaza del Rio Buildings A (PR-A) and B (PR-B) and floor f3

for Centro Mayor. Figures showing all relevent demand-capacity data for the lower floors

of these buildings are presented in Appendix F.


416

To simplify the presentation of the Tier 2 results, the discussion here focuses on the

demand-capacity data for Plaza del Rio Building A as this is the building modeled for the

nonlinear Tier 3 evaluation. Also, to further simplify the discussion of the results, relatively

few figures are included; the reader is referred to Appendix F for a complete set of figures.

Section 9.2.1 provides a discussion of the unfactored demand capacity ratios. Sec-

tion 9.2.2 provides a discussion of the factored demand capacity ratios. Section 9.2.3 pro-

vides a discussion of the impact of building period on the results of the analysis. Section 9.2.4

provides an overview of the impact on demands of wall coupling due to beams and slabs.

9.2.1 Evaluation of Unfactored DCR Values

ASCE/SEI 41-06 defines unfactored demand-capacity ratios, DCRu , for the purpose of

evaluating the adequacy of using linear analysis procedures for assessment and identifying

critical components. Details of the Tier 2 procedures are provided in Section 8.1.3. The

primary purpose of evaluating the DCRu values is to determine if linear analysis procedures

are appropriate for use in assessing the structure. If any components have DCRu > 2.0,

ASCE 41 requires a nonlinear analysis of the structure.

As was mentioned previously, only the results for Plaza del Rio Building A are presented

here. Figure 9.1 shows the DCRu values for the lower two floors for results generated

using the ASCE response spectrum and the two ground motion spectra. For each wall,

the DCRu values for shear, moment, and axial demands were calculated following the

procedure outlined in Section 8.1.3. The maximum of these three values was considered to

be the controlling value and is shown on the floor plans. Figure 9.2 indicates the action

(shear, moment, or axial) that resulted in the largest DCRu value. The floor plans indicate

a large numbers of walls with DCRu > 2.0, therefore use of the linear dynamic procedure

was inappropriate per the ASCE 41 standard. This outcome was also the result of Tier 2

evaluation of the other buildings studied. Section F.2 in the appendix provides figures that

show the DCRu values for the lower floors of the other buildings.
417

(a) ASCE RS (b) Concepcion RS (c) San Pedro RS (d)

Figure 9.1: PR-A Floor 1: Unfactored demand capacity ratios (DCRu ).

(a) ASCE RS (b) Concepcion RS (c) San Pedro RS

Figure 9.2: PR-A Floor 1: Controlling action using unfactored demand capacity ratios
(DCRu ). Blue = Moment; Red = Shear; Black = Axial.
418

ASCE 41 also indicates that the DCRu values can also be used to identify critical

components in a building. For Plaza del Rio Building A, the most critical components (those

with the largest DCRu ) are the small wall-piers. These components are shear controlled per

the DCRu s (but axial controlled for the DCRf s). For completeness, all components were

considered primary components for this evaluation. However, if the ASCE 41 guidelines

were strictly followed, these small piers would likely be considered secondary components

due to their relavitely small contribution to the total base shear capacity and lateral stiffness.

If these small wall piers are not considered, the most critical component (critical action was

flexure) in Plaza del Rio Building A is the right T-wall in the center of the building, for which

significant damage was observed (see Figure 8.10). Other main walls (those carrying a large

percentage of the total shear) also had large DCRu values and would be considered critical

in the overall structure. The differences in values for walls that otherwise would be expected

to be the same (that is the pairs of mirrored T- and Z- shaped walls) is due to difference in

the component strengths resulting from different reinforcement (size, spacing and quantity)

in the boundary elements. When considering different response spectra, there is little change

in the which components are critical, nor in the critical action, however, the magnitude of

the DCRu values do vary. In the other buildings studied, the critical components were

also small wall piers and likely to be considered secondary components. As in PR-A, these

components were axial critical when considering the DCRf values. Again, if the small

wall piers were not considered, the critical components were some of the walls carrying the

largest percentage of total shear, with the critical actions throughout the building being a

mix of both shear and flexure. Despite the number of walls identified as shear critical by the

evaluation of the DCRu values, the damage in the buildings was predominately flexural.

9.2.2 Evaluation of Factored DCR Values

The Tier 2 evaluation procedures are based not on the DCRu values discussed in the

previous section, but rather on factored demand capacity ratios, DCRf . The DCRf values,
419

discussed in detail in Section 8.1.3, are calculated by applying a modification factor (m-

factor) to the computed strength of the component. The m-factor is intended to account

for the over-prediction of demands that occur when elastic models are used. For each

component, the “critical action” is that determined by the response mode (axial, flexure, or

shear) that has the largest DCRf . Ultimately, the Tier 2 procedure identifies a component

as adequate if, for a given earthquake demand, the DCRf value for the critical action is

less than 1.0. Appendix F.2 provides figures showing the critical actions and the value of

DCRf for each of the typical floors in the buildings studied.

When considering DCRf , the values vary as a function of the performance level and

earthquake hazard level considered. If these evaluations were carried out in a design office,

it is likely that only the performance objective of achieving the Life Safety (LS) performance

level under the design-level event would be considered. However, since the objective of this

study is to evaluate the Tier 2 procedures, three performance objectives were considered:

Immediate Occupancy for the design level event (IO-DBE), Life Safety for the design level

event (IO-DBE), and Collapse Prevention at the maximum event (CP-MCE)). The design

event used the ASCE 7 design spectrum. The maximum event used the ASCE design

spectrum scaled by a factor of 1.5. Additionally, all three performance levels (IO, LS, and

CP) were considered for demands determined from the unscaled spectra derived from the

measured ground motions.

The controlling actions identified using the DCRf values were generally the same as

the controlling actions identified using the DCRu values. For the unfactored DCRu values,

walls were either flexure critical or shear critical; however, when considering the factored

DCRf values, some shear-critical walls became axial-critical walls. In general, these walls

were small wall piers and had large axial load demands generated by the earthquake forces.

Figures 9.3 thru 9.4 show, for the ASCE 7 and ground motion response spectra, the walls

that are adequate for the various performance objectives in the lowest floor of Plaza del

Rio Building A (PR-A). Inadequate walls are indicated by dashed lines. Adequate walls are
420

shown as solid with the color (green through red) indicating the observed damage level; gray

indicates that no damage was observed. The most notable observation to make from these

figures is the difference between the results obtained using the two different response spectra.

Significantly more walls were identified as adequate when the Concepcion ground motion

spectrum was used. Such observations were expected given the lower spectral accelerations

for the ground motion spectrum at the fundamental periods of the buildings.

Figure 9.3 shows the adequate walls for Plaza del Rio Building A floor f1 when LDP

was performed using the ASCE response spectrum scaled to the DBE and MCE levels.

Figure 9.3a indicates that most walls were considered inadequate regardless of which per-

formance objective was considered, with the majority of these sustaining some level of dam-

age. Figure 9.3b indicates that a few walls with moderate to severe damage were considered

adequate for the collapse prevention performance objective. Tier 2 evaluation procedures

were successful though potentially conservative (that is, damage was expected because the

Tier 2 evaluations indicates the components is deficient), as indicated by the observation

of damage (of any type or severity) for walls that were determined to be inadequate for all

performance objectives or adequate only for collapse prevention under the MCE objective.

Concern with the Tier 2 procedure arises when walls identified as adequate for a specified

performance objective had observed damage in excess of the damage associated with that

performance objective (Table 9.4). For Plaza del Rio Building A floor f1, this occurs for the

IO-DBE performance objective (Figure 9.3d) for a few walls that are flag-shaped or could

be considered secondary components. These are the same walls identified as the critical

components in Section 9.2.1. For Plaza del Rio Building A floor f2 (not shown), no walls

were observed to have damage in excess of that expected per the Tier 2 evaluation.

The discussion in the proceeding paragraph was focused on the results of a Tier 2

evaluation for Plaza del Rio Building A using the response spectrum specified in ASCE/SEI

7-10. If the response spectrum derived from the recorded ground motion were used instead,

the results were significantly different. As shown in Figure 8.2a, using the ground motion
421

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-DBE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure 9.3: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under ASCE 7 response spectrum.
422

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-MCE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure 9.4: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under Concepcion response spectrum.
423

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-MCE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure 9.5: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under San Pedro response spectrum.
424

spectra the spectral accelerations at the fundamental period of the building is significantly

lower than if the ASCE DBE spectrum is used. This resulted in much lower demands and

consequently, much lower DCRf . Figure 9.4 shows the walls identified as adequate when

the Concepcion response spectrum was used. Whereas the ASCE spectrum resulted in

many inadequate walls and few walls that were acceptable at any performance level, the

Concepcion response spectrum resulted in only a few inadequate walls. Most walls were

identified as adequate per the collapse prevention performance level and observed damage

levels were consistent with that performance level. However, many walls were adequate

per the life safety and immediate occupancy performance objectives but damage was in

excess of that expected at these performance objectives, suggesting a failure of the Tier 2

procedures to adequately capture deficiencies under demands likely experienced during the

actual event.

Figure 9.3, Figure 9.4, and Figure 9.5 are summarized in Figure 9.6a, d, and e re-

spectively, which indicate for the various performand levels the number of walls that were

adequate (DCR < 1) and inadequate (DCR > 1), with the color indicating the observed

damage level. Figure 9.6 also provides this information for the other response spectra con-

sidered (Chilean design spectrum at the time of construction and design spectrum based

on changes proposed following the earthquake). Similar figures for all buildings studied are

found in Appendix F. For Plaza del Rio Building A, both the current and proposed Chilean

building code response spectra give results that are very similar to those for the ASCE 7

response spectra (the same observation can made for other buildings). This is because the

spectral accelerations, at the fundamental periods of the building, are very similar to the

different design spectra.

9.2.3 Impact of Building Period

One possible explanation for the low demands determined using the ground motion spectra

is that the periods in the model are significantly different from the actual periods of the
425

25 25 25

20 20 20
# of walls

# of walls

# of walls
15 15 15

10 10 10

5 5 5

0 0 0
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1
IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE

(a) ASCE response spectrum (b) Chilean response spectrum (c) Proposed Chilean response
spectrum

25 25

20 20
# of walls

# of walls

15 15

10 10
0: No Damage
1: Minor (cracking)
5 5
2: Moderate
0 0
3: Severe
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 4: Total
IO−GM LS−GM CP−GM IO−GM LS−GM CP−GM

(d) Concepcion response spec-(e) San Pedro de la Paz response (f) Damage color legend
trum spectrum

Figure 9.6: Plaza del Rio Building A Floor 1: # of damage state occurances in walls
achieving various performance objectives. Color indicates the observed damage.
426

building and that actual periods resulted in much higher spectral accelerations such as those

at approximately 1.7 seconds in the Concepcion ground motion record (see Figure 8.2). To

explore the possibility that the model was too stiff, the effective stiffness of the elements

were progressively reduced to increase the fundamental periods, however, even at very low

stiffness values, the periods were much lower that those associated with the peak accel-

erations. For Plaza del Rio Building A, reconnaissance teams noted significant uplift of

the mat foundation that suggested the building experienced significant rocking that could

have increased the period. This soil-structure interaction was not accounted for in the Tier

2 evaluation models, however, nonlinear Perform models conducted by other members of

the project team [75] included simple models for soil-structure interaction. These models

generated periods that were in the approximate range of the highest spectral accelerations.

9.2.4 Impact of Coupling Effects

One of the goals of the study was to explore the impact of overall building configuration

issues, including the impact of coupling due to coupling beams, slabs, and non-structural

components such as stairs. Some exploration of this issue was possible using the results of

the LDP analyses of the buildings conducted for the Tier 2 analyses.

To explore the overall impact of coupling, the axial loads in the walls in the lower floor

of the buildings were studied. For each response spectrum used in the Tier 2 evaluations,

the axial load from the earthquake forces, QU F , was determined according to the ASCE 41

provisions. This value was normalized by the axial load due to gravity to determine the

increase in the axial load due to earthquake effects. Figure 9.7 shows this index for the

first floor of Plaza del Rio Building A for the ASCE 7 design spectrum and the Concepcion

and San Pedro ground motion spectra. The factor increase in the axial load demand ranges

from approximately 1.5 in a few small wall piers to more than 5.0 in walls in the core

region at the top of the building plan and at the two T-shaped walls at the center of the

building. Similar results are obtained for the other buildings studied and are provided in
427

Appendix F.3. These ratios suggest that axial forces must be considered when evaluating

the adequacy of the structure. Given the significant impact of wall axial demand on wall

performance and ASCE 41 acceptance criteria, accurate prediction of axial load is critical.

Axial load due to coupling is explored further using the nonlinear analysis results (see

Section 9.3.3).

(a) (b) ASCE 7 (c) Concepcion GM (d) San Pedro GM

Figure 9.7: PR-A Floor 1: Ratio of maximum axial load due to earthquake loading to axial
load due to gravity loading from different response spectra.

9.3 Summary of Tier 3 Evaluations

A Tier 3 evaluation requires conducting a nonlinear dynamic analysis. Due to the time

required to conduct nonlinear analyses of a full building, Tier 3 evaluation was conducted

only for Plaza del Rio Building A. The ASCE/SEI 31-03 guidelines for a Tier 3 evaluation

are very general and simply indicate that evaluation using the results of any reasonable

nonlinear analysis procedure are acceptable. For this study, the Tier 3 evaluation followed

the ASCE/SEI 41-06 requirements for a nonlinear dynamic analysis and included additional

evaluations using fragility functions and strain limits. Details of the model and process

for nonlinear analysis are presented in Section 8.2.2; analysis results and assessment of the

performance based on these results are presented here. Section 9.3.1 summarizes the general
428

50

40

30

20

10

(a) (b) ASCE 7 (c) Concepcion GM (d) San Pedro GM

Figure 9.8: PR-A Floor 1: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial capacity of the
wall (Ag fc0 ).

response of the building model. Section 9.3.2 presents the evaluation of i) ASCE/SEI 41-06

acceptance criteria for walls, with modification to consider the shear strength of walls, ii) the

evaulation of the fragility functions presented in Chapter 7, and iii) evaluation using strain

limits associated with damage states. Section 9.3.3 discusses the progression of damage in

the model.

9.3.1 General Building Response

A total of sixteen time-history analyses were conducted for Plaza del Rio Building A (PRA).

Two ground motions, San Pedro de la Paz and Concepcion, were run without any scaling

to assess the response of the structure to the earthquake motions that represent the best

estimates of the actual ground motion experienced by the building during the 2010 Maule

earthquake. Analyses were done also for these two ground motions, plus an additional five

motions recorded during the Maule earthquake on similar soil sites, scaled to the ASCE 7

DBE and MCE hazard levels. Most time-history analyses were fully completed in OpenSees;

four analyses “failed to converge” prior to completion of the full history. The ground motions
429

Table 9.1: Summary of building drifts from Tier 3 analyses.

Record Maximum Residual


Name Scale Trans. (%) Long. (%) Trans. (%) Long. (%)
Concepcion 1 0.82 0.60 0.09 0.01
DBE 0.741 0.451 - -
MCE 2.011 0.941 - -
San Pedro 1 0.94 0.59 0.02 0.01
DBE 1.33 0.81 - -
MCE 2.49 1.03 - -
Llolleo DBE 0.54 0.47 0.00 0.00
MCE 0.82 0.60 0.00 0.00
Matanzas DBE 1.46 0.68 0.17 0.00
MCE 1.20 0.79 0.05 0.01
Valparaiso DBE 0.88 0.76 0.02 0.01
MCE 1.17 1.02 0.11 0.07
VdM - Centro DBE 0.82 0.70 0.00 0.01
MCE 1.13 0.95 0.04 0.02
VdM - El Salto DBE 1.05 1.05 0.00 0.01
MCE 1.27 1.38 0.02 0.04
1
Maximum displacement prior to “failure to converge”.

that failed to converge were: i) Concepcion scaled to the DBE (fail at 12 of 141 seconds),

ii) Concepcion scaled to the MCE (fail at 13 of 141 seconds), iii) San Pedro de la Paz scaled

to the DBE (fail at 31 of 202 seconds), and iv) San Pedro de la Paz scaled to the MCE (fail

at 27 of 202 seconds). These convergence failures occurred in local elements.

Tables 9.1 and 9.2 summarize the drift demands for each ground motion. All drifts were

determined from the displacement at the master node for the rigid diaphragm on each floor.

Roof drifts were calculated as the displacement, in the appropriate direction, divided by the

model height (30 meters). The story drifts were calculated as the differential displacement

at the top and bottom of the floor, divided by the floor height (2.5 m). The residual drifts

were calculated as the median drift over the last 10 seconds of each earthquake record.

Residual drifts are not included for those for which the analysis failed to converge prior to

completion of the record.


430

Table 9.2: Summary of story drifts from Tier 3 analyses.

Maximum
st
Record 1 Story 2nd Story
Name Scale Trans. (%) Long. (%) Trans. (%) Long. (%)
Concepcion 1 0.12 0.17 0.51 0.35
DBE 0.12 0.16 0.35 0.31
MCE 0.63 0.63 1.63 0.87
San Pedro 1 0.20 0.25 0.48 0.41
DBE 0.38 0.46 0.80 0.65
MCE 2.16 0.97 2.23 1.00
Llolleo DBE 0.16 0.17 0.30 0.25
MCE 0.22 0.22 0.39 0.32
Matanzas DBE 0.17 0.25 0.97 0.51
MCE 0.22 0.28 0.72 0.60
Valparaiso DBE 0.17 0.29 0.47 0.51
MCE 0.22 0.63 0.52 0.81
VdM - Centro DBE 0.14 0.25 0.41 0.43
MCE 0.19 0.31 0.59 0.51
VdM - El Salto DBE 0.18 0.36 0.52 0.70
MCE 0.23 0.40 0.66 0.73

Larger drifts were measured in the transverse direction of loading, which is consistent

with the larger spectral accelerations for motion in that direction and with the observed

damage. For the unscaled ground motions, the maximum roof drifts were very similar,

with a drift of 0.82 − 0.94% in the transverse direction and approximately 0.60% in the

longitudinal direction. When the seven scaled motions are considered, the drifts have a much

wider variation, ranging from 0.54 − 2.49% in the transverse direction and 0.47 − 1.38% in

the longitudinal direction. The largest drift demands were experienced using the San Pedro

ground motion. The maximum residual drifts were on very small and significantly less than

the observed residual drifts.

Table 9.3 summarizes the effective height of loading for the ground motions studied.

The effective height ratio was calculated as the ratio of the base moment to the base shear,

divided by the building height. The time steps corresponding to local (with respect to
431

Table 9.3: Summary of effective height ratios from Tier 3 analyses.

Record Effective Height1


Name Scale Trans. Long.
Concepcion 1 0.36 0.38
DBE 0.32 0.37
MCE 0.37 0.38
San Pedro 1 0.28 0.36
DBE 0.30 0.25
MCE 0.28 0.20
Llolleo DBE 0.30 0.36
MCE 0.33 0.40
Matanzas DBE 0.37 0.42
MCE 0.35 0.41
Valparaiso DBE 0.40 0.41
MCE 0.37 0.40
VdM - Centro DBE 0.41 0.41
MCE 0.40 0.40
VdM - El Salto DBE 0.35 0.36
MCE 0.35 0.34
1
Value listed is α where αHbldg is the effective
height of the lateral load.

time) peak base shear demands were used to determine the median effective heights. In

the longitudinal direction of loading, the effective height ranged from 0.28 − 0.41Hbldg , with

a median value of 0.35Hbldg . In the longitudinal direction of loading, the effective height

ranged from 0.20 − 0.42Hbldg , with a median value of 0.38Hbldg . These effective heights

are lower than that for the uniform load distribution (hef f ≈ 0.54Hbldg for PR-A) used by

ASCE/SEI 41-06 for calculation of the design shear forced used in determining i) if walls

are shear- or flexure-controlled, and ii) the shear demand used to determine acceptance

criteria for flexure controlled walls. Consequently, the ASCE 41 acceptance criteria likely

underestimate the impact of shear on the response of the wall and thus potentially fail to

capture shear failures. To consider the impact of shear demand in the study presented here,

an additional check beyond the ASCE 41 provisions was conducted for shear.
432

9.3.2 Evaluation of Acceptance Criteria

9.3.2.1 Presentation of Observed vs Predicted Damage Data

Figures 9.9 and 9.22 summarizes the results of application of the evaluation criteria, de-

scribed in Section 8.2.2.2, for the unscaled Concepcion and San Pedro ground motion.

Figure 9.9 present the results when the shear criterion is not applied. Figure 9.22 present

the results when the shear criterion is applied. The results are presented in three groups: i)

ASCE 41 acceptable rotations, ii) application of fragility functions, and iii) damage-based

strains. Each column in the bar charts indicates the total number of walls in the build-

ing that reach a particular level of demand. For ASCE rotations, walls are categorized by

highest performance level (IO, LS, or CP) for which the acceptable hinge rotation was ex-

ceeded. If the shear criterion was considered and the acceptable shear strength (1.1Vn ) was

exceeded, the wall is included in the ‘shear’ column. For fragility function rotations, walls

are categorized by the highest damage state (DS1, DS2, DS3, or DS4) for which the likeli-

hood of occurrance was at least 0.50. If the shear criteria was considered and the acceptable

shear strength was exceeded, the wall is included in the ‘shear’ column. For strains, walls

are categorized by the highest damage level (yield, spalling, or crushing) reached. If the

shear criterion was considered and the acceptable shear strength was exceeded, the wall is

included in the ‘shear’ column.

Each column in the bar charts consists of three colors indicating how the expected

damage relates to the observed damage. Green (medium gray) indicates the observed and

predicted damage levels were the same. Cyan (light gray) indicates the observed damage is

less severe than that predicted from the evaluation criteria. Red (dark gray) indicates that

the observed damage was more severe than that predicted by the model. The relationship

between the predicted damage levels, determined by the demands of the model, and the

observed damage are summarized in Table 9.4. The predicted damage levels for ASCE/SEI

41-06 are based on the expected damage for the different performance levels, summarized in
433

Table 9.51 . For fragility functions, the predicted damage is directly related to the damage

states used to develop the fragilities, which are described in Chapter 7. For evaluation of

strain demands, the predicted damage is determined based on the damage associated with

the strain limits of yielding, spalling, and crushing.

Figures 9.10 -9.21 and Figures 9.23 -9.34 show the walls that contribute to each column

in the bar charts in Figures 9.9 and 9.22, respectively. As with the bar charts, the wall color

indicates the relationship between predicted and observed damage.

9.3.2.2 Observations Based on Observed and Predicted Damage Data

Figure 9.9 summarizes the acceptance criteria without the shear strength limit. The left

four bars summarize the results of evaluation using the ASCE 41 acceptance criteria. The

ASCE 41 criteria are generally accurate or conservative in predicting the levels of the ob-

served damage, that is, most walls with maximum simulated rotations exceeding a particular

performance level have observed damaged levels equal to or less than that expected for that

performance level (the cyan (light gray) and green (medium gray) portions of the bars).

Concern with the ability of the evaluation procedure to predict damage is with walls that

are indicated by red (dark gray). Of particular concern are the number of walls that have

observed damage but are not flagged as exceeding any acceptable rotation limits (the red

(dark gray) portion of the ‘None’ column). Review of the data in Figures 9.10 and 9.11

show that all of these walls are small wall piers, mostly located on the first floor. For these

walls, a possible explanation for the inability to capture the damage in these walls is the

fact that the OpenSees model does not simulate increased axial load due to slab coupling.

Further discussion of this issue is provided in Section 9.3.3.

The fifth through ninth bars of Figure 9.9 summarize the results of evaluation of compo-

nent demands using the fragility functions from Chapter 7. In comparison with the ASCE

1
ASCE 41 explicitly states that these relationships are not to serve for post-earthquake evaluation of
structures and for the judging the safety or level of repair necessary for a structure; however, as the goal
of this study is to evaluate the adequacy of the ASCE 41 standard, these relationships were used as a
guideline for determining if specific performance levels are accurately predicted.
434

Table 9.4: Relationship between damage levels and target performance levels.

Acceptable damage levels


Performance Level None Minor Moderate Severe Total
None
IO
LS
CP
None
DS1
DS2
DS3
DS4
Shear
Yield
Spall
Crush
Observed damage < predicted damage
Observed damage = predicted damage
Observed damage > predicted damage
435

Table 9.5: ASCE/SEI 41-06 relationship between performance level and observed damage

Target Performance Levels


Element Collapse Prevention Life Safety Immediate Occupancy
Overall Damage Severe Moderate Light
General Little residual stiffness Some residual strength No permanent drift.
and strength, but load and stiffness left in all Structure substantially
bearing elements stories. retains original strength
function. Gravity load bearing el- and stiffness.
Large permanent drifts. ements function. No out Minor cracking of struc-
Some exits blocked. of plane failure of walls. tural elements. Eleva-
Building is near Some permanent drift. tors can be restarted.
collapse. Building may be beyond Fire protection
economical repair. operable.
Concrete walls Major flexural and shear Some boundary element Minor hairline cracking
cracks and voids. Slid- stress, including limited of walls, < 1/16 in. (1.6
ing at joints. Extensive bar buckling. mm) wide.
crushing and buckling of Some sliding at joints. Coupling beams experi-
reinforcement. Failure Damage at openings. ence cracking < 1/8 in.
around openings. Sever Some crushing and flex- (3.2 mm) wide.
boundary element ural cracking.
damage. Coupling beams: exten-
Coupling beams sive shear and flexural
shattered and virtually cracks; some crushing,
disintegrated. but concrete remains in
place.
2% transient or 1% transient drift; 0.5% transient drift;
permanent drift. 0.5% permanent. negligible permanent.
Foundations Major settlement and Total settlements < 6 in. Minor settlement and
tilting (15.24 cm) and negligible tilting.
differential settlements
< 1/2 in. (1.3 cm) in 30
ft. (9.1 m).
Diaphragms Extensive crushing and Extensive cracking Some minor cracking
observable offset across (< 1/4 in. (6.35 mm) along joints.
many cracks width).
Local crushing and
spalling.
436

41 criteria, assessment using the fragility functions indicates an overall expectation of more

severe damage. Unlike the ASCE 41 criteria, the fragility functions indicate that at least a

minor level of damage (DS1 - initial cracking or yielding) is expected in all walls. Unlike the

ASCE 41 results, which resulted in more severe damage for the Concepcion ground motion

record, the fragility functions indicate more severe damage states for the San Pedro de la

Paz ground motion record. The fragility functions produced results that are more conserva-

tive than the ASCE 41 criteria. This may be due in part to the development of the fragility

functions based on the initiation of damage states in the laboratory. In the field, with

extensively damaged walls, damage may not be “observed” until well past initiation. The

conservative results from the fragility functions may also be a result of the larger rotation

values predicted using the equivalent plastic hinge model (used for definition of fragility)

versus using the empirical lumped plasticity approach (used from processing analysis data).

A discussion of these different rotations is presented in Figure 8.5.

Use of the shear criterion (difference between Figure 9.9 and Figure 9.22) resulted in

many walls reaching the shear strength limit, many of which would otherwise have been

categorized as having high rotation or strains demands. The shear strength limit used is

likely a conservative value, therefore the two set of evaluations can be seen as an upper and

lower bound of the impact of shear on the results of the evaluation. The impact of shear

would be better explored with a model capable of capturing flexure and shear interaction,

and using a more precise shear strength, but the results nonetheless indicate that shear

should be a considered factor in evaluating the performance of what are typically considered

flexure-controlled walls.

The above discussion focuses on results generated using the unscaled Concepcion and

San Pedro de la Paz ground motions. Time-history analyses were also conducted using both

ground motions, as well as five additional ground motions, scaled to the ASCE 7 DBE and

MCE response spectra. For each of these analyses bar charts and floor plans similar to those

in Figures 9.9 - 9.34 are provided in Appendix G. A summary of these results are provided
437

in Tables 9.6 -9.8. Each table presents, for the three sets of limits considered (ASCE 41

rotation acceptance criteria, fragility function probabilities, and strains for initiation of

damage), the number of walls that fall within each performance level indicated in the bar

charts. Values are provided without and with the shear strength limit. Also provided are

the number of walls for which the simulated damage underpredicted the observed damage

(those walls that are red in the bar charts and floor plans).

As was mentioned previously, the analyses using the Concepcion and San Pedro de

la Paz ground motions failed to converge in local elements prior to all time steps being

completed. Consequently, the evaluation of the acceptance criteria using these results should

be considered with reservation as some components may have exceeded the specified limits

had the analysis been able to continue. In general, the scaled records produced a greater

number of walls with large rotation demands, resulting in more walls in the LS and CP

performance level categories when using the acceptance criteria and more walls in the DS3

and DS4 damage states when using the fragility functions for evaluation. The number of

walls exceeding the shear strength limit was mostly unaffected by the scaled ground motions

relative to the unscaled ground motions. The most notable result is that of the San Pedro

de la Paz ground motion scaled to the ASCE 7 MCE spectrum. Although this analysis

failed to converge prior to full analysis of the ground motion, the model simulated significant

demands (rotations exceeding the CP performance level and/or damage state DS4) in almost

all walls on the first floor, including the small wall piers that were significantly damaged

and identified as critical by the Tier 2 evaluation but did not have any simulated damage

using any other records.


438

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Concepcion

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(b) San Pedro de la Paz

Figure 9.9: Summary of walls exceeding acceptance criteria for unscaled ground motion
records.
439

(a) None (b) IO (c) LS (d) CP

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure
45
9.10: Floor 1 walls exceeding ASCE 41 rotation limits (Concepcion ground motion).
40

35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) IO (c) LS (d) CP

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure
45
9.11: Floor 2 walls exceeding ASCE 41 rotation limits (Concepcion ground motion).
40

35
# of Predicted walls

30

25

20

15

10

0
2

d
IO

LS

CP

1
e

ll

h
pa
DS

DS

DS

DS

iel
on

on

on

rus
440

(a) DS1 (b) DS2 (c) DS3 (d) DS4

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure
45
9.12: Floor 1 walls exceeding fragility function probabilities of 0.50 (Concepcion
ground
40
motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) DS1 (b) DS2 (c) DS3 (d) DS4

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure
45
9.13: Floor 2 walls exceeding fragility function probabilities of 0.50 (Concepcion
ground
40
motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
441

(a) None (b) Yield (c) Spall (d) Crush

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure 9.14:
45
Floor 1 walls exceeding strain limits (ConcepcionStrain
ground motion).
40

35
# of Predicted walls

30

25

20

15

10

0
IO

LS

CP

ld
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) Yield (c) Spall (d) Crush

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure 9.15:
45
Floor 2 walls exceeding strain limits (ConcepcionStrain
ground motion).
40

35
# of Predicted walls

30

25

20

15

10

0
1

d
IO

LS

CP
e

ll

h
pa
DS

DS

DS

DS

iel
on

on

on

rus
442

(a) None (b) IO (c) LS (d) CP

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.16: Floor 1 walls exceeding ASCE 41 rotation limits (San Strain
Pedro de la Paz ground
motion).
40

35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) IO (c) LS (d) CP

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.17: Floor 2 walls exceeding ASCE 41 rotation limits (San Strain
Pedro de la Paz ground
motion).
40

35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
443

(a) DS1 (b) DS2 (c) DS3 (d) DS4

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure45
9.18: Floor 1 walls exceeding fragility function probabilitiesStrain
of 0.50 (San Pedro de
la Paz40 ground motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) DS1 (b) DS2 (c) DS3 (d) DS4

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure45
9.19: Floor 2 walls exceeding fragility function probabilitiesStrain
of 0.50 (San Pedro de
la Paz40 ground motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
444

(a) None (b) Yield (c) Spall (d) Crush

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.20: Floor 1 walls exceeding strain limits (San Pedro de laStrain
Paz ground motion).
40

35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) Yield (c) Spall (d) Crush

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.21: Floor 2 walls exceeding strain limits (San Pedro de laStrain
Paz ground motion).
40

35
# of Predicted walls

30

25

20

15

10

0
2

d
IO

LS

CP

1
e

ll

h
pa
DS

DS

DS

DS

iel
on

on

on

rus
445

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr
(a) Concepcion

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) San Pedro de la Paz

Figure 9.22: Summary of walls exceeding acceptance criteria with consideration for shear
strength limit for unscaled ground motion records.
446

(a) None (b) IO (c) LS (d) CP (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.23: Floor 1 walls exceeding ASCE 41 rotation limits withStrain
consideration for shear
strength
40
limit (Concepcion ground motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) IO (c) LS (d) CP (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.24: Floor 2 walls exceeding ASCE 41 rotation limits withStrain
consideration for shear
strength
40
limit (Concepcion ground motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
447

(a) DS1 (b) DS2 (c) DS3 (d) DS4 (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure45
9.25: Floor 1 walls exceeding fragility function probabilities ofStrain
0.50 with consideration
for shear
40
strength limit (Concepcion ground motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) DS1 (b) DS2 (c) DS3 (d) DS4 (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure45
9.26: Floor 2 walls exceeding fragility function probabilities ofStrain
0.50 with consideration
for shear
40
strength limit (Concepcion ground motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
448

(a) None (b) Yield (c) Spall (d) Crush (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.27: Floor 1 walls exceeding strain limits with considerationStrain
for shear strength limit
(Concepcion
40
ground motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) Yield (c) Spall (d) Crush (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.28: Floor 2 walls exceeding strain limits with considerationStrain
for shear strength limit
(Concepcion
40
ground motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
449

(a) None (b) IO (c) LS (d) CP (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.29: Floor 1 walls exceeding ASCE 41 rotation limits withStrain
consideration for shear
strength
40
limit (San Pedro de la Paz ground motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) IO (c) LS (d) CP (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.30: Floor 2 walls exceeding ASCE 41 rotation limits withStrain
consideration for shear
strength
40
limit (San Pedro de la Paz ground motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
450

(a) DS1 (b) DS2 (c) DS3 (d) DS4 (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure45
9.31: Floor 1 walls exceeding fragility function probabilities ofStrain
0.50 with consideration
for shear
40
strength limit (San Pedro de la Paz ground motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) DS1 (b) DS2 (c) DS3 (d) DS4 (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure45
9.32: Floor 2 walls exceeding fragility function probabilities ofStrain
0.50 with consideration
for shear
40
strength limit (San Pedro de la Paz ground motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
451

(a) None (b) Yield (c) Spall (d) Crush (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.33: Floor 1 walls exceeding strain limits with considerationStrain
for shear strength limit
(San 40
Pedro de la Paz ground motion).
35
# of Predicted walls

30

25

20

15

10

0
ld
IO

LS

CP

4
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr

(a) None (b) Yield (c) Spall (d) Crush (e) Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure
45
9.34: Floor 2 walls exceeding strain limits with considerationStrain
for shear strength limit
(San 40
Pedro de la Paz ground motion).
35
# of Predicted walls

30

25

20

15

10

0
2

ld
IO

LS

CP

1
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
452

Table 9.6: Number of walls exceeding ASCE 41 rotation limits.

# Limits Exceeded1 # Under


Record Scale None IO LS CP Shear -predicted1,2
Concepcion 1 18 (17) 3 (1) 12 (3) 13 (4) 0 (21) 14 (8)
DBE 24 (21) 5 (1) 12 (6) 5 (3) 0 (13) 18 (13)
MCE 6 (6) 10 (8) 9 (6) 21 (4) 0 (22) 10 (7)
San Pedro 1 15 (15) 10 (8) 15 (5) 6 (1) 0 (17) 17 (10)
DBE 11 (11) 10 (9) 8 (4) 17 (4) 0 (18) 11 (8)
MCE 8 (8) 3 (1) 6 (2) 29 (15) 0 (20) 4 (2)
Llolleo DBE 23 (23) 10 (2) 12 (2) 1 (0) 0 (19) 19 (10)
MCE 15 (15) 16 (9) 12 (4) 3 (0) 0 (18) 18 (10)
Matanzas DBE 14 (13) 10 (9) 7 (2) 15 (5) 0 (17) 13 (9)
MCE 15 (15) 8 (8) 7 (1) 16 (5) 0 (17) 13 (9)
Valparaiso DBE 15 (15) 8 (8) 15 (10) 8 (3) 0 (10) 18 (14)
MCE 13 (13) 5 (5) 12 (9) 16 (4) 0 (15) 11 (9)
VdM - Centro DBE 17 (17) 10 (9) 16 (9) 3 (2) 0 (9) 20 (15)
MCE 15 (15) 8 (7) 13 (8) 10 (6) 0 (10) 17 (2)
VdM - El Salto DBE 13 (13) 9 (9) 8 (3) 16 (8) 0 (13) 14 (10)
MCE 14 (14) 8 (8) 4 (2) 20 (8) 0 (14) 12 (10)

Mean 1 16.5 (16) 6.5 (4.5) 13.5 (4) 9.5 (2.5) 0 (19) 15.5 (9)
DBE 16.7 (16.1) 8.9 (6.7) 11.1 (5.1) 9.3 (3.6) 0 (14.1) 16.1 (11.3)
MCE 12.3 (12.3) 8.3 (6.6) 9 (4.6) 16.4 (6) 0 (16.6) 12.1 (8.4)
1
# when shear criteria not considered (# when shear criteria considered)
2
Walls indicated as red (dark gray) in the associated bar charts and floor plans
453

Table 9.7: Number of walls exceeding fragility function probability of 0.50.

# Limits Exceeded1 # Under


Record Scale DS13 DS2 DS3 DS4 Shear -predicted1,2
Concepcion 1 9 (9) 7 (6) 3 (3) 27 (7) 0 (21) 8 (7)
DBE 7 (7) 7 (7) 10 (8) 22 (11) 0 (13) 8 (8)
MCE 1 (1) 3 (3) 2 (2) 40 (18) 0 (22) 1 (1)
San Pedro 1 5 (5) 6 (6) 7 (7) 28 (11) 0 (17) 5 (5)
DBE 3 (3) 2 (2) 8 (8) 33 (15) 0 (18) 3 (3)
MCE 3 (3) 3 (3) 0 (0) 40 (20) 0 (20) 0 (0)
Llolleo DBE 11 (11) 6 (6) 5 (4) 24 (6) 0 (19) 8 (8)
MCE 8 (8) 4 (4) 9 (8) 25 (8) 0 (18) 6 (6)
Matanzas DBE 3 (3) 7 (7) 7 (7) 29 (12) 0 (17) 6 (6)
MCE 5 (5) 3 (3) 11 (11) 27 (10) 0 (17) 6 (6)
Valparaiso DBE 7 (7) 5 (5) 9 (9) 25 (15) 0 (10) 8 (8)
MCE 3 (3) 4 (4) 5 (5) 34 (19) 0 (15) 5 (5)
VdM - Centro DBE 10 (10) 3 (3) 7 (7) 26 (17) 0 (9) 7 (7)
MCE 6 (6) 6 (6) 6 (6) 28 (18) 0 (10) 7 (7)
VdM - El Salto DBE 4 (4) 3 (3) 10 (10) 29 (16) 0 (13) 5 (5)
MCE 4 (4) 2 (2) 10 (10) 30 (16) 0 (14) 7 (7)

Mean 1 7 (7) 6.5 (6) 5 (5) 27.5 (9) 0 (19) 6.5 (6)
DBE 6.4 (6.4) 4.7 (4.7) 8 (7.6) 26.9 (13.1) 0 (14.1) 6.4 (6.4)
MCE 4.3 (4.3) 3.6 (3.6) 6.1 (6) 32 (15.6) 0 (16.6) 4.6 (4.6)
1
# when shear criteria not considered (# when shear criteria considered)
2
Walls indicated as red (dark gray) in the associated bar charts and floor plans
3
All walls exceeded damage state DS1 at minimum, thus, a ‘None’ column is not included.
454

Table 9.8: Number of walls exceeding strain limits.

# Limits Exceeded1 # Under


Record Scale None Yield Spall Crush Shear -predicted1,2
Concepcion 1 12 (12) 24 (11) 7 (2) 3 (0) 0 (21) 20 (0)
DBE 16 (16) 24 (13) 5 (3) 1 (1) 0 (13) 22 (13)
MCE 3 (3) 19 (16) 13 (4) 11 (1) 0 (22) 15 (7)
San Pedro 1 15 (15) 21 (9) 9 (5) 1 (0) 0 (17) 22 (11)
DBE 9 (9) 16 (13) 15 (5) 6 (1) 0 (18) 18 (9)
MCE 5 (5) 16 (11) 5 (4) 20 (6) 0 (20) 7 (3)
Llolleo DBE 15 (15) 27 (10) 3 (1) 1 (1) 0 (19) 21 (9)
MCE 14 (14) 25 (13) 6 (1) 1 (0) 0 (18) 21 (10)
Matanzas DBE 14 (14) 21 (11) 6 (1) 5 (3) 0 (17) 21 (11)
MCE 12 (12) 23 (13) 8 (3) 3 (1) 0 (17) 22 (10)
Valparaiso DBE 13 (13) 28 (19) 4 (3) 1 (1) 0 (10) 22 (15)
MCE 11 (11) 18 (14) 12 (4) 5 (2) 0 (15) 18 (10)
VdM - Centro DBE 15 (15) 27 (21) 4 (1) 0 (0) 0 (9) 22 (15)
MCE 13 (13) 24 (18) 7 (3) 2 (2) 0 (10) 21 (15)
VdM - El Salto DBE 10 (10) 22 (16) 12 (6) 2 (1) 0 (13) 21 (13)
MCE 9 (9) 20 (16) 15 (6) 2 (1) 0 (14) 21 (12)

Mean 1 13.5 (13.5) 22.5 (10) 8 (3.5) 2 (0) 0 (19) 21 (5.5)


DBE 13.1 (13.1) 23.6 (14.7) 7 (2.9) 2.3 (1.1) 0 (14.1) 21 (12.1)
MCE 9.6 (9.6) 20.7 (14.4) 9.4 (3.6) 6.3 (1.9) 0 (16.6) 17.9 (9.6)
1
# when shear criteria not considered (# when shear criteria considered)
2
Walls indicated as red (dark gray) in the associated bar charts and floor plans
455

9.3.3 Damage Progression

One advantage offered by the nonlinear time history analyses employed for the Tier 3 anal-

ysis, in comparison with the linear dynamic procedure employed for a Tier 2 analysis, is the

ability to evaluate the progression of damage and the sequence in which damage-state limits

were exceeded. The results of the time history analysis of Plaza del Rio Building A under

the unscaled Concepcion and San Pedro records were used to determine the progression

of damage in the building. Figures 9.14 - 9.15 and Figures 9.20 - 9.21 indicate the walls

for which a various damage states (yield, spalling, or crushing) were simulated in the first

two floors of Plaza del Rio Building A under the unscaled Concepcion and San Pedro de la

Paz ground motions, respectively. Appendix G provides tables summarizing the sequence

in which these damage states occurred. A brief discussion of the sequence of damage is

presented here.

Analyzing the structure using the unscaled Concepcion ground motion record (Fig-

ures 9.14 and 9.15), only three walls reached concrete strains associated with crushing.

These walls were the two z-shaped walls at the bottom of the cross-section and one of the

two t-walls located at the center of the structure. The crushing strain was first reached in

each of these walls in the second floor. In the T-shaped and Z-shaped walls on the right side

of the floor plans, these higher demands in the second floor were consistent with the ob-

served damage occurring at the bottom of the second floor (shown in Figures 8.9 and 8.10)2 .

It should be noted that the walls at these two E-W (left-right) gridlines carry the majority

of the shear associated with the E-W direction of loading which was the larger component

of the ground motion.

Analyzing the structure using the San Pedro de la Paz ground motion record (Fig-

ures 9.20 and 9.21), the strain demands were less than with the Concepcion record. Crush-

ing occurred only in the second floor of the lower right z-shaped wall. Unlike the Concepcion

2
It should be noted that the shear limits in the first floors of the z-shape was reached prior to crushing
in the second floors.
456

record, the San Pedro record did record spalling in both the first and second floor of the

z-shaped walls at the bottom of the plans. This is consistent with the damage observed in

both stories. However, as with the Concepcion record, the spalling demands in these walls

occurred first in the second floor and eventually progressed to crushing in one of the walls.

The differences in damage progression between the two unscaled ground motions indi-

cates the response of the structure is sensitive to the specific ground motion used. Table 9.8

summarizes the number of walls exceeding each strain limit for all ground motions and

hazard levels considered. Generally, scaling the ground motions to the DBE and MCE re-

sponse spectra had little affect on the number of walls reaching the strains corresponding

to initiation of spalling or crushing, with a mean of 9.4 and 6.3 walls reaching each initial

strain value, respectively, for the seven MCE ground motions. One exception to the lower

number of walls with large compressive strains is the San Pedro de la Paz ground motion

scaled to the MCE spectra. Use of this ground motion resulted in 20 walls reaching the

strain corresponding to initiation of crushing. Most of these walls were located in the first

floor and included small wall piers that were damaged but do not exceed acceptable limits

(ASCE 41 rotations, fragility function rotations, or strains) for any other ground motions.

Overall, the damage predicted using the strain measures is less than the observed dam-

age. Possible reasons for this stem from simplifications made in the modeling approach and

include i) a failure to capture strength loss due to shear failure, ii) reduced flexural capacity

due to shear, iii) failure to redistribute loads due to slab failures, iv) foundation flexibility,

and v) impact of axial load from coupling (see following section).

9.3.4 Impact of Axial Load

Figures 9.7 and 9.8 demonstrate, relative to the gravity load, the increased compressive axial

loads from coupling for the Tier 2 linear analysis. Such coupling affects cannot be captured

by the simplified OpenSees model, however, axial loads generated from nonlinear models

were availabe from a Perform3D model developed by practicing engineers collaborating on


457

the study of Plaza del Rio Building A. The details of these models (one with a fixed based

and one with soil-structure interaction (SSI)) are not discussed here and can be found in

the project report [63]. Figures 9.35 and 9.36 show the compressive and tensile axial load

ratios for walls in the first floor of the building. In the main T-walls at the center of the

building, compressive demands range from 15 − 35% of the gross capacity of the walls and

are approximately 5 or more times larger than the gravity loads, and would contribute to

the extent of spalling and crushing observed in the structure. In the small wall piers, large

compressive and tensile demands were simulated in the Perform3D model and likely played

a large role in the observed damage. The inability of the simplified OpenSees model to

include these axial demands likely impacts the inability to simulate the extent of damage

in many walls.

9.4 Recommendations for Changes to ASCE 31/41 Evaluations

Application and evaluation of the ASCE 31 and 41 Tier 1, Tier 2, and Tier 3 provisions for

assessment of Chilean buildings resulted in identification of a number of changes that could

be made to improve assessment results and the efficiency of the assessment process.

For the Tier 1 evaluation, which is intended to be a quick check to identify potential

deficiencies and determine if further evaluation is required, improvements are recommended,

based on the observed damage, to address the following issues for assessment of walled

buildings:

1. A check should be added to identify as “non-compliant” walls in which significant

local vertical discontinuities, such as a shift in the cross-section or reduction in wall

area, interrupt the load path.

2. Checks for basic seismic detailing such as minimum wall thickness, minimum spacing of

confining reinforcement, detailing of confining reinforcement, and detailing at potential

stress concentrations such as openings and flag-shaped walls are recommended.

3. Modification of soft- and weak-story checks to clarify that such a deficiency is only
458

50

40

30

20

10

(a) (b) SAP (c) Perform - Fixed (d) Perform - SSI

Figure 9.35: Ratio of maximum compressive axial demand to axial compressive capacity
(Ag fc0 ). Values shown are percents. Note: All walls are colored and apparent differences in
line thickness do not indicate different results.

50

40

30

20

10

(a) (b) Perform - Fixed (c) Perform - SSI

Figure 9.36: Ratio of minimum (most tensile) axial demand to axial tensile capacity (As fy ).
Values shown are a percent. Note: All walls are colored and apparent differences in line
thickness do not indicate different results.
459

critical when a soft- or weak-story occurs below another story.

For Tier 2 and 3 evaluations, which involve linear and nonlinear dynamic analysis of

part or all of the structure, respectively, the following issues were found to require further

evaluation or clarification for application to walled buildings:

1. Clarification on the use of ASCE 41 Tables 6-18 and 6-19, which define response

parameters for walls responding in flexure and shear, specifically as to the appropriate

method for modeling and evaluating the shear response of flexure controlled walls, or

the flexural response of shear-controlled walls.

2. Consideration of shear and flexure interaction for modeling the walls, for identifying

if walls are shear or flexure controlled, for evaluating the demand capacity ratios, and

for evaluating acceptable rotations. Currently, ASCE 41 specifies the use of a back-

bone curve for either shear or flexure response, however, discussions with practicing

engineers indicate that both are typically considered for modeling walls. ASCE 41

tables do account for the shear in selecting model parameters and acceptance criteria

for walls, but only based on a single value of the shear that does not account for the

variation in demands throughout a time-history analysis.

3. Evaluation/validation of m-factors and acceptance criteria using experimental data.

If needed, improvements should be made based on improved numerical modeling tech-

niques such as those provided by Pugh [89]. Consideration should be made for differ-

ent cross-section shape (i.e. planar, symmetric flanged, or asymmetric flanged) and

improved recommendations should include acceptance criteria for both rotation and

drift for flexure controlled walls. Documentation should be provided indicating how

improvements were developed and the potential limitations of the values provided.

Addressing all of these improvements is beyond the scope of the current study. Modifying

the Tier 1 checklist to include a check for local vertical discontinuities was addressed as

part of this study and is presented in Section 9.4.2. Also, ongoing research conducted
460

in collaboration with ACI Committee 369, is working to develop improved m-factors and

acceptance criteria for flexure-controlled walls. All other issues identified above are left to

future research efforts.

9.4.1 Comments on Usability of ASCE 31/41 Provisions

In addition to the recommendations for improvement presented above, there are a number

of issues in ASCE/SEI 31-03 and 41-06 that are somewhat confusing and/or are difficult to

implement given the standard methods for modeling a structure in the design office. These

include the following:

1. With respect to material strength modifiers, it is not clear for which response mech-

anisms and to which parameters of the strength model the 1.25 over-strength factor

should be applied. For example, for flexure, it is not clear if the factor should be

applied to the steel strength only or to the whole value, and for shear, it is not clear

if the factor should be applied at all.

2. Clarification is required as to the axial load that should be used to determine the

wall strength and the appropriate acceptance criteria. Currently in ASCE 41 it is

difficult to ascertain if the axial force should include forces resulting from earthquake

demands in calculating the nominal strength of wall. Additionally, use of this axial

load for definition of the response envelope is challenging as this value is dependent

on having created and executed a model and is not available in preliminary stages of

the analysis.

3. The C1 C2 J term for calculating the demands of force-controlled actions is inconvenient

because of how the demands are extracted from the results of the LDP. It would be

easier if a different m-factor were prescribed for the force-controlled actions.


461

9.4.2 Local Vertical Discontinuities

The ASCE 31 Tier 1 checklists consider vertical discontinuities in buildings by considering

the relative stiffness and strength of one floor relative to that above/below it. However,

study of mid-rise walled buildings damaged in the 2010 Chile earthquake suggest that local

changes in the stiffness/strength of individual building components may be correlated to the

observed damage. Thus, to explore this hypotheses, a study was undertaken to quantify the

local discontinuities and determine if there is a significant correlation to observed damage.

Recommendations are made for proposed additions to the current Tier 1 analysis procedures.

9.4.2.1 Characterization of Discontinuities

For the purpose of this study, a discontinuity was considered to be any location where the

cross-section of a wall at any given floor (referred to as the upper wall) differed from the

cross-section of the floor below it (referred to as the lower wall). In the four buildings

studied, such discontinuities were located at the building cores, interiors and perimeters,

and can be described as one of the following types of discontinuities:

• Termination of a wall from the upper floor to the lower floor.

• Openings present in the lower section of the wall. An example is shown in Figure 9.37b.

• Shift in the centroid of the cross-section, including a relocation of a flange(s) of a wall.

An example is shown in Figure 9.37c.

• Flag-shaped wall, such as the one shown in Figure 9.37d, where the upper wall is

larger than the lower wall.

When assessing the damage levels for the discontinuities, the difference between the

damage in the lower and upper floors was considered. Thus, for the wall in Figure 9.37d,

the damage level is equal to 4 (= 4 (lower floor) −0 (upper floor)) and for the wall in

Figure 9.37b, the damage level is equal to 3 (= 4 (lower floor) −1 (upper floor)). For

walls where there is no change, or the damage level is higher in the upper wall, such as the
462

wall in Figure 9.37c, the damage level is specified as zero (0). In figures referenced in the

following discussion, the discontinuity damage level is indicated by the corresponding color

in Figure 9.37a (e.g. red indicates the a damage level difference of 4).

To quantify the discontinuities, the following parameters were calculated for each dis-

continuity:

• ∆A/Aupper . The change in wall area, normalized area of upper wall and reported as

a percentage.

• ∆Lx /Lupper
x and ∆Ly /Lupper
y : The change in the maximum east-west and north-south

dimensions of the wall, normalized by the dimension of the upper wall and reported

as a percentage. This is calculated by the dimensions of a rectangle in which the

cross-section is inscribed. Ultimately this parameter was not found to be useful and

thus is not presented further.

• ∆Lcl /Lupper
cl , ∆Lcl,x /Lupper upper
cl,x , and ∆Lcl,y /Lcl,y : The change in the centerline length

of the wall normalized by that of the upper wall and reported as a percentage. The

centerline length, Lcl , of the wall is total length of wall as measured along the centerline

of the cross-section. Distinction is also made in the total length of wall in the x- and

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(a) Damage Legend (b) Opening (c) Shift (d) Flag


shaped

Figure 9.37: Examples of geometric discontinuities with damage level indicated.


463

y-directions, and are referred to as Lcl,x and Lcl,y , respectively.

• ∆xc /Lupper
x , ∆yc /Lupper
y : The change in the in-plane coordinates of the geometric

centroid of the wall, normalized by the length of the wall in the same direction and

reported as a percentage. This number is always presented as a negative number to

facilitate easy comparison with parameters that describe the loss of area of length.

In evaluating the discontinuities, the parameters described above were considered in

three groups. The first group, which includes change in area, ∆A/Aupper , and change in

total centerline length, ∆Lcl /Lupper


cl , are referred to as global properties. The second group,

referred to as directional properties, describe the discontinuities in terms of either the x

or ydimensions of the walls and may include the parameters ∆Lcl,x /Lupper upper
cl,x , ∆Lcl,y /Lcl,y ,

∆xc /Lupper
x , and ∆yc /Lupper
y . When using these parameters, only the relevant directions

were considered. For example, a flanged wall that losses area in the flange but not the

web is only described by the directional properties associated with the local axis parallel to

the flange. The third group of parameters considered is a reduction of the second group.

Whereas the second group consists of all directional parameters describing the discontinuity,

the third group consists of only the most severe directional parameter (that with the lowest,

or most negative, value).

The complexity of the discontinuities ranged from simple, isolated, flag-shaped walls

to those adjacent to other discontinuities or with damage believed to be associated with

a different deficiency of the component. Consequently, each discontinuity was assigned a

rank, or rating, intended to indicate the quality of the data for the purpose of exploring the

relationship between damage and local geometric discontinuities. The ranking, ranging from

0 for a high quality data point to 6 for a low quality data point, considered the following

factors i) simplicity of the geometry, ii) confidence that the observed damage is associated

with the discontinuity, iii) the presence of other local discontinuities associated with either

or both of the lower and upper walls, and iv) the importance of the discontinuity to the

overall performance of the building, considering factors such as location within the building
464

and the percentage of total floor shear carried by the wall. Figure 9.38 provides an indication

of the number of discontinuities of each ‘rank’ in each building studied. Table 9.9 provides

a summary of the number of discontinuities by each floor and whether or not damaged was

observed.

6 20
# of Discontinuities

# of Discontinuities
15
4
10
2
5

0 0
0 2 4 6 0 2 4 6
Rank Rank
(a) Plaza del Rio (A) (b) Plaza del Rio (B)

15 25
# of Discontinuities

# of Discontinuities

20
10
15

10
5
5

0 0
0 2 4 6 0 2 4 6
Rank Rank
(c) Centro Mayor (d) Concepto Urbano

Figure 9.38: Number of local geometric discontinuities, sorted by rank. A rank of 0 indicates
the highest quality data; a rank of 6 indicates the lowest quality of data.

9.4.2.2 Evaluation of Discontinuities in Severely Damaged Buildings

Figures 9.39 - 9.41 show, for severely damaged buildings (PR-A and CM), the global, direc-

tional, and controlling directional parameters, respectively; subfigures show the values for

a) all discontinuities, b) discontinuities with rank 0 or 1, c) discontinuities with rank 2 or


465

Table 9.9: Comparison of local and global vertical discontinuites

Bldg Floor (Upper/Lower) # Discontinuities Damaged?


PR-A f2/f1 12 Yes
PR-B f2/f1 21 No
CM f3/f2 10 Yes
f2/f1 18 Yes
f1/f0 14 Yes
CU f12/f11 6 No
f11/f10 5 No
f2/f1 4 No
f1/f0 3 No
f0/f-1 1 No

3, and d) discontinuities with rank ≥ 4.

In evaluating the data to establish if any correlation exists between observed damage

and local discontinuities, those considered to be “good” data points (rank 0 or 1) were

the primary focus as these discontinuities were clearly defined and unlikely to be affected

by other discontinuites. Figure 9.39b shows the global properties for these walls, with the

change in either area or centerline length ranging from a loss of over 50% to a gain of nearly

100%. Severe increases in damage were observed for both loss and gain of area or length,

indicating the global properties are inadequate for associating damage with discontinuities.

When considering only the directional parameters (Figure 9.40b), there is a greater associ-

ation of damage and negative values, however there is a number of discontinuities described

by both negative and positive changes in length or centroid. Based on the characteristics

of the individual discontinuities, either a loss of length or large shift in the centroid is the

value with the largest loss, therefore both should be considered. The controlling directional

parameter (that which results in the largest loss of length or shift in centroid) is shown in

Figure 9.41b. When the data set is considered in manner, there appears to be a clear trend

of severe damage occurring in walls with a controlling directional parameter of more than

30%, indicated by the dashed red line.

When considering the 30% limit for the discontinuities with rank ≥ 2 (Figures 9.41c and
466

9.41d), this trend is not as clear, however, it should be recalled that the higher rank was

assigned to these due to uncertainty in definition of the discontinuity and judgement as to

the likliehood of the damage resulting from a discontinuity and not another mechanism.

Regardless of the uncertainty of association with damage, the consistency of negative values

of the controlling directional parameters demonstrates the appropriateness of this method

for quantifying the discontinuities relative to the use of global parameters (Figures 9.39c and

9.39d) which demonstrate the same large variation in magnitude and value as was seen when

examining the discontinuities with lower rank.

9.4.2.3 Recommendation for Discontinuities Susceptible to Damage

Based on the observations presented, the following text is recommended as an additional

ASCE 31 Tier I checklist item for reinforced concrete structural wall buildings.

PARTIAL VERTICAL DISCONTINUITIES: All vertical wall elements in the lateral-

forceresisting system that are continuous to the foundation but have a change in cross-

sectional geometry shall not exceed i) a loss greater than 30% of centerline length (in either

orthogonal direction) from one floor to the floor below or ii) a change in centroid greater

than 30% of the corresponding wall length (largest of the two cross-sections).

The recommendation of 30% is consistent with the ‘GEOMETRY’ check in the ASCE 31

check lists, which indicates that there shall be no more than a 30% loss in the lateral force

resisting system. The difference between the ‘GEOMETRY’ checklist item and the proposed

one is that the proposed focuses on an individual wall. While the the ‘GEOMETRY’ check

is adequate for identifying potential deficiences of the system, the proposed ‘PARTIAL

VERTICAL DISCONTINUITIES’ check adds a method for identifying particaluar elements

particulary susseptible to damage.


467

400 400

350 350

300 300

250 250

200 200
∆, %

∆, %
150 150

100 100

50 50

0 0

−50 −50

−100 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
# #

(a) All (b) Rank: 0-1

400 400

350 350

300 300

250 250

200 200
∆, %

∆, %

150 150

100 100

50 50

0 0

−50 −50

−100 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
# #

(c) Rank: 2-3 (d) Rank: >= 4

Figure 9.39: Change in global parameters vs discontinuity ID.


468

300 300

250 250

200 200

150 150
∆, %

∆, %
100 100

50 50

0 0

−50 −50

−100 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
# #

(a) All (b) Rank: 0-1

300 300

250 250

200 200

150 150
∆, %

∆, %

100 100

50 50

0 0

−50 −50

−100 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
# #

(c) Rank: 2-3 (d) Rank: >= 4

Figure 9.40: Change in relevant directional parameters (severely damage buildings) vs dis-
continuity ID.
469

60 60

40 40

20 20

0 0
∆, %

∆, %
−20 −20

−40 −40

−60 −60

−80 −80

−100 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
# #

(a) All (b) Rank: 0-1

60 60

40 40

20 20

0 0
∆, %

∆, %

−20 −20

−40 −40

−60 −60

−80 −80

−100 −100
0 20 40 60 80 100 120 0 20 40 60 80 100 120
# #

(c) Rank: 2-3 (d) Rank: >= 4

Figure 9.41: Change in controlling directional parameter (severely damaged buildings) vs


discontinuity ID.
470

9.5 Conclusions

The Chilean buildings studied were complex and are not representative of modern construc-

tion in the United States. Thus, they are not ideal case studies for evaluating the ASCE

31/41 Standards. However, the study was undertaken as it presented a unique opportu-

nity to study, outside a laboratory environment, a physical structure that was subjected to

severe earthquake loading and did or did not sustain damage.

The Tier 1 procedures identified a number of deficiencies in the study buildings includ-

ing inadequate shear capacities for the expected shear strength demands, weak and soft

stories, and discontinuous vertical load paths. Comparing these deficiencies to the observed

performance of the buildings, it was noted that the presence of these deficiencies alone

was not a predictor of observed damage; thus, as is specified by ASCE/SEI 3103, further

evaluation of the structures was required. In addition to evaluating the existing provision,

recommendations for a new checklist item were made based on observations of the observed

damage in the buildings studied. This recommendation addresses walls that have partially

continuous load paths and reads as follows:

PARTIAL VERTICAL DISCONTINUITIES: All vertical elements in the lateralforce-

resisting system that are continuous to the foundation but have a change in crosssectional

geometry shall not exceed i) a loss greater than 30% of centerline length (in either orthogo-

nal direction) from one floor to the floor below or ii) a change in centroid greater than 30%

of the corresponding wall length (largest of the two cross-sections).

The next level of evaluation of a structure is a Tier 2 evaluation that involves linear

dynamic analysis of the structure. Models of three of the four study buildings were created

in SAP2000 and response spectra analysis was conducted using the following spectra:

• ASCE/SEI 7-10 [6] design spectrum for building sites,

• The Chilean Building code spectrum [50] in effect when the buildings were designed,
471

• Proposed changes to the Chilean building code spectrum following the 2010 Maule

earthquake,

• Spectra generated from two ground motions recorded near the building sites.

For all three buildings analyzed using all spectra, the amount by which demands exceed

capacities indicated that linear evaluation of the structures was inappropriate and that a

Tier 3 (nonlinear analysis) was required. Despite this, the Tier 2 acceptance criteria were

evaluated to identify potential areas of improvement for the procedure. Recommendations

for improvement of the Tier 2 evaluation procedures include:

1. Clarification of the method for classifying the walls as shear or flexure controlled and

modifying the assumed effective height of loading to better represent that observed in

the models.

2. Clarification of the axial load demands to be used to establish m-factors used in the

assessment and simplification of the calculation of force-control demands to better

reflect the way data are generated by commercial software.

3. Updating m-factors to reflect observed behavior of walls in the laboratory and to

include the impact of wall geometry on performance, including the difference between

planar and flanged walls and symmetric and asymmetric shapes.

4. Consideration for the interaction of shear, moment, and axial forces in evaluating the

adequacy of a component.

Tier 3 evaluation, the highest level of evaluation, involves creating a nonlinear model of a

structure. Due to the increased time and complexity of the analysis, only one of the buildings

studied was modeled for Tier 3 evaluation. Due to the relative simplicity of the wall layout

and the well documented damage to the building, Plaza del Rio Building A was chosen for

Tier 3 evaluation. In evaluating the model, the observed damage was compared with the

expected damage as indicated by i) shear strength of the walls, ii) the various performance

levels associated with ASCE/SEI 41-06 acceptance criteria for flexure controlled walls, iii)
472

strains at initial yield, initial spalling, and initial concrete crushing, and iv) damage states

associated with the fragility functions presented in Chapter 7. The evaluations results were

mixed; damage was successfully identified in some walls but not in other walls and/or was

predicted to occur in undamaged walls. Evaluation of the effective height and inclusion of

the shear strength limits indicated that the impact of shear demands needs to be thoroughly

addressed for flexure-controlled walls. The results of the models were able to provide some

insight into the potential response of the building during the earthquake by identifying the

order in which damage occurred in individual walls. As with the Tier 2 analysis, a number of

areas were identified that require clarification and/or improvements. The recommendations

for improvements to Tier 2 procedures generally apply to Tier 3 evaluation. The following

expands on these for Tier 3:

1. Clarification of shear response used for modeling flexure-controlled walls.

2. Updated backbone curve values for flexure controlled walls to reflect observed exper-

imental behavior and to include consideration for flanged and asymmetric sections.

Updated values should be available for both rotation and drift, with rotation values

clearly indicating the type of model they were generated for and how these values

might relate to the rotations generated from other commonly used modeling tech-

niques.
473

Chapter 10

CONCLUSIONS

The work presented in this dissertation is focused on enhancing the engineering commu-

nity’s i) ability to understand the seismic behavior of slender reinforced concrete walls, ii)

access to high-resolution data to validate and develop numerical models for performance-

based design of slender RC walls, iii) ability to predict potential damage to slender RC

walls, and iv) understanding of tools for the assessment of existing structures to identify

seismic deficiencies. Specifically, this included the following tasks:

1. A review of relevant literature (Chapter 2) to summarize i) current U.S. requirements

and recommendations for seismic design of structural walls, ii) design characteristics

for structural wall buildings designed for the West Coast using UBC 1991-1998, iii)

a review of structural wall damage observed following major earthquakes around the

world since 1957, and iv) previous experimental research investigating the seismic

performance of structural walls. Additionally, to assess the relevance of existing ex-

perimental data and to identify gaps in the knowledge base, design characteristics for

previously tested walls were compared with those of walls in the West Coast build-

ing inventory and with current ACI 318-08 [1] Code requirements. Damage observed

in laboratory test specimens was compared with that observed in the field following

recent earthquakes.

2. Documentation of experimental results from four planar (rectangular) reinforced con-

crete subassemblies (Chapters 3 and 4. These tests, part of the UW-UIUC-UCLA

NEESR project, investigated the impact of shear demand, longitudinal reinforcement

distribution, and lap splices within the expected plastic hinge region. Documentation

of experimental results included damage progression, failure mode, and drift capacity
474

of the wall specimens.

3. Analysis of experimental data (Chapters 5 and 6) collected from the four planar wall

tests to improve understanding of seismic behavior of planar walls. Data analysis in-

cluded calculation and evaluation of: i) displacement and rotation profiles, ii) effective

stiffness, iii) flexural, shear, and fixed-end deformations, iv) extent of reinforcement

yield, v) vertical, shear, and principal compressive strain fields, and vi) strain demands

at failure.

4. Development of fragility functions (Chapter 7) defining the probability of occurrence

of various damage states given a specific level of seismic demand defined by a specific

value of an engineering demand parameter. Fragilities were developed using data

from previous experimental tests by others as well as the specimens documented in

Chapters 3 through 6.

5. Evaluation of ASCE/SEI 31-03 and 41-06 evaluation procedures (Chapters 8 and 9)

for buildings damaged in the 2010 Maule earthquake. This consisted of quick checks

(Tier 1), linear dynamic analysis (Tier 2), and nonlinear dynamic analysis (Tier 3).

Results of these evaluations were discussed in comparison to the observed performance

of buildings. Recommendations for future improvements were made.

10.1 Conclusions from Literature Review

Review of the literature support the following conclusions:

1. Only 10 planar wall test specimens which meet ACI 318-08 Code requirements were

identified. These test specimens have characteristics that generally are representative

of modern buildings on the West Coast, however, some gaps were identified within

the overall set of data related to the inventory of buildings on the West Coast. These

gaps were i) few very slender (aspect ratio greater than or equal to 3.0) walls have

been tested, ii) boundary element length in experimental test specimens are generally

shorter relative to wall length than the boundary elements in modern walls, and iii)
475

no experimental tests of walls with slender cross-sections (wall length relative to wall

thickness) with axial loads have been conducted.

2. Based on a review of post-earthquake reconnaissance reports, the major damage type

(i.e. more extensive than cracking) observed most frequently (more than 20% of all

observed damage types) in U.S. structural walls is compressive damage (bar buckling

and concrete crushing) in the boundary element. By contrast, the major damage

types observed frequently (more than 20% of all observed damage types) in Chilean

construction included not only boundary element crushing but also large diagonal

shear cracks, web crushing, and horizontal planes of damage that span the length of

the wall (in addition to compressive damage in the boundary element).

3. Damage in the lab is primarily compressive, reflecting that documented in post-

earthquake reconnaissance reports. However, damage seen at discontinuities in the

field is a condition not observed in the lab.

4. Previous experimental research has indicated that wall drift capacity is affected by

i) axial load, ii) shear demand, iii) confining reinforcement ratio, and iv) shear span

ratio. Evaluation of all previous experimental research indicates that drift capacity

may also be affected by i) cross-sectional aspect ratio, ii) web reinforcement ratios,

and iii) specimen scale.

10.2 Conclusions for Experimental Test Program

Experimental observations and the data analysis for the four planar wall tests support the

following preliminary conclusions:

1. Shear demand appears to impact the drift capacity of the walls, with Specimen PW1
p
(shear of 2.8Acv fc0 ) exhibiting higher drift capacity than Specimen PW2 (identical
p
in design to Specimen PW1 but with shear demand of 4.5Acv fc0 ). The data indicate

that axial load, steel strain capacity, and the presence of a lap splice in the region of

maximum moment demand may also affect drift capacity.


476

2. In walls with lap splices in the region in which flexural yielding is expected:

(a) The increased strength and stiffness of the spliced region may result in a reloca-

tion of the critical section, from the expected region (base of the wall) to the top

of the splice.

(b) Strains calculated from displacements measured on the surface of the wall indicate

that compressive strains are transferred through the spliced region similar to the

unspliced region but tensile strains are not.

3. In walls with lap splices in the expected plastic hinge region and detailed with a ‘dog-

leg’ to represent modern construction, the increased stiffness of the spliced region

resulted in a relocation of the critical section of the wall from the base of the wall to

the top of the splice. Tensile strains were concentrated above the splice. Compressive

strains were distributed more evenly in and above the spliced region loaded first in

compression but more concentrated at at the top of the splice or above it when loaded

first in tension.

4. In walls with a uniform layout of longitudinal reinforcement, web crushing is observed

unlike in walls with longitudinal reinforcement concentrated in the boundary regions.

This web crushing appears to be the result of interaction between tensile strain and

shear strain.

10.3 Conclusions from Development of Fragility Functions

Evaluation of the damage progression from the 4 planar wall tests and from previous ex-

perimental test specimens support the following conclusions regarding the development of

fragility functions:

1. Damage progression in slender walls is best evaluated using the following groups of

damage states, listed with the associated method of repair:

(a) Initial cracking and initial yielding of longitudinal reinforcement. Cosmetic repair

only (e.g. painting).


477

(b) Spalling of cover concrete. Repaired by patching of cover concrete and epoxy

injection of cracks.

(c) Exposed longitudinal reinforcement. Repaired by replacement of concrete.

(d) Core crushing, bar buckling, bar fracture, and/or shear failure. Repaired by wall

replacement.

2. Evaluation of the median engineering demand parameters for each damage states

suggests the following may impact damage progression: i) load history, ii) cross-section

shape, iii) axial load, and iv) shear span ratio. However, the amount of data available

for some subsets of these parameters (e.g. flanged walls or axial load ratios greater

than 0.1Ag fc0 ) is limited and more data is needed to validate these observations and

to develop fragilities.

3. Recommendations for fragility function were made to the ATC-58 project. Fragility

functions were created using data from cyclically loaded test specimens only.

10.4 Conclusions from Study of Damaged Chilean Buildings

Evaluation of the ASCE/SEI 31-03 [7] and 41-06 [8, 9] procedures for buildings damaged in

the 2010 Maule earthquake resulted in the following conclusions:

1. The Tier 1 quick checks identified deficiencies for shear strength, soft-stories, and

weak-stories, however, these deficiencies did not consistently correlate with observed

damage.

2. Recommendations were made for expanding the Tier 1 checks to account for local

vertical discontinuities where load paths were interrupted. Current Tier 1 procedures

only account for termination of load paths.

3. The Tier 2 evaluation, which involved linear dynamic analysis of the buildings, indi-

cated that many small wall piers were axial critical and expected to have high demands

relative to the component strength. Some of these walls, especially those where cou-

pling effects may have been a factor, were significantly damaged in the buildings
478

studied. The Tier 2 evaluations also identified many walls as shear critical, however,

the damaged observed throughout the buildings was primarily flexural damage.

4. When buildings were modeled using linear dynamic analysis, the magnitude of the

component demands were sufficiently large relative to the component strength such

that Tier 2 evaluation alone was insufficient to determine the seismic adequacy of the

structure, thus requiring full nonlinear analysis of the structures.

5. Evaluation of the building using a nonlinear dynamic analysis (Tier 3 evaluation) in-

dicated drifts on the order of 0.80%-0.90%, which are much lower than would typically

be expected in a building with the level of observed damage. Residual drifts were much

smaller than measured in the structure following the earthquake, indicating that the

lack of inclusion of soil-structure interaction may have an impact on the results of the

model. The median effective height of the loading, using a suite of seven ground mo-

tions, ranged from 0.20-0.40 of the building height, which is lower than that assumed

by the ASCE 41 standard.

6. Results of the Tier 3 evaluation successfully indicated the level of damage in some of

the key walls (those carrying a large percentage of the shear) but did not identify any

damage in other key walls.

7. When the adequacy of the buildings was evaluated using the fragility functions pre-

sented in Chapter 7, the predicition of damage was more conservative than when the

acceptance criteria of ASCE 41 were used.

In addition to the conclusions above, the study of the ASCE/SEI 31-03 evaluation pro-

cedures resulted in a number of recommendations for improvements to future versions of

the standard and for additional research in other areas, including:

1. Improved acceptance criteria based on numerical models validated with experimental

data (e.g. recommendations by Pugh [89].) and that consider appropriate parameters

such as cross-section characteristics.

2. Clarification of use shear response models for flexure-controlled walls and modification
479

for acceptance criteria to allow failure in either flexure or shear in flexure-controlled

walls.

3. Improved effective height for determination of shear or flexure controlled response of

walls.

Overall, despite the any shortcomings or uncertainties in the modeling and evaluation

procedures used to study the damaged Chilean buildings, both the linear (Tier 2) and

nonlinear (Tier 3) evaluation procedures of ASCE/SEI 41-06 [8, 9] were able to, for a severely

damaged building, to identify the structure as inadequate under the hazard levels considered

and in general, the critical components identified by the evaluations were those building

components most severely damaged. Although recommendations for improvements of the

standards have been made within this document, the ASCE 41 procedures proved to be a

useful tool for determining whether or not the studied structures were seismically adequate.

Upon further development based on validation of acceptance criteria using experimental

data and revisions of procedures to reflect standard use in design offices, the ASCE 41

standard will continue to be a useful tool for engineers that when coupled with engineering

judgment and experience, can guide the process of evaluating the adequacy of existing

structures.
480

BIBLIOGRAPHY

[1] ACI. Building Code Requirements for Structural Concrete (ACI 318-08) and Com-
mentary. American Concrete Institute, 2008.

[2] ACI Committee 546. ACI 546R-96: Concrete Repair Guide. American Concrete
Institute, Farmington Hills, MI, 1996.

[3] A. Ali. Reinforced concrete structural walls with staggered opening configurations
under reversed cyclic loading. PhD thesis, University of Michigan, 1990.

[4] Applied Technology Council. FEMA 308: Repair of Earthquake Damaged Concrete
and Masonry Wall Buildings. FEMA, Washington, D.C., 1998.

[5] Applied Technology Council. Seismic Performance Assessment of Buildings Volume


1 - Methodology, ATC-58-1 75% Draft. Technical report, prepared for U.S. Dept. of
Homeland Security Federal Emergency Management Agency, May 2011.

[6] ASCE. Minimum Design Loads for Buildings and Other Structures (7-10). American
Society of Civil Engineers, Reston, Virginia, 2010.

[7] ASCE/SEI. Seismic Evaluation of Existing Buildings (ASCE/SEI Standard 3103).


Structural Engineering Institute, American Society of Civil Engineers, Reston, Vir-
ginia, 2003.

[8] ASCE/SEI. Seismic Rehabilitation of Existing Buildings (ASCE/SEI Standard 4106).


Structural Engineering Institute, American Society of Civil Engineers, 2007.

[9] ASCE/SEI. Supplement No. 1, Seismic Rehabilitation of Existing Buildings (ASCE/


SEI Standard 4106). Structural Engineering Institute, American Society of Civil
Engineers, Reston, Virginia, 2007.

[10] Glen V. Berg and James L. Stratta. Anchorage and the alaska earthquake of march
27, 1964. Technical report, American Iron and Steel Institute, 1964.

[11] K. Beyer, A. Dazio, and M.J.N. Priestley. Quasi-static cyclic tests of two U-shaped
reinforced concrete walls. Journal of Earthquake Engineering, 12:1023–1053, 2008.

[12] P. Brown. Probabilistic earthquake damage predictions for reinforced concrete build-
ing components. Master’s thesis, University of Washington, 2008.
481

[13] P. Brown, J. Ji, A. Sterns, D.E. Lehman, L.N. Lowes, D. Kuchma, and J. Zhang.
Investigation of the seismic behavior and analysis of reinforced concrete walls. In 8th
US National Conference on Earthquake Engineering, 2006.

[14] P.C. Brown and L.N. Lowes. Fragility functions for modern reinforced concrete beam-
column joints. Earthquake Spectra, 23(2):263–289, May 2007.

[15] B.L. Brueggen. Performance of T-shaped Reinforced Concrete Structural Walls under
Multi-Directional Loading. PhD thesis, University of Minnesota, August 2009.

[16] G. Butcher, D. Hopkins, R. Jury, W. Massey, G. McKay, and G. McVerry. The


September 1985 Mexico earthquakes: Final report of the New Zealand reconnaissance
team. Bulletin of the New Zealand National Society for Earthquake Engineering, 21(1),
March 1988.

[17] M.J.C. Caamano. Analisis de danos del edificio plaza del rio, provocados por el
terremoto del 27 de febrero del 2010. Master’s thesis, Universidad Technica Federico
Santa Maria, June 2011.

[18] A.E. Cardenas and D.D. Magura. Strength of high rise shear walls – rectangular cross
section. In Response of Multistory Concrete Structures to Lateral Forces, SP–36, pages
119–150. American Concrete Institute, 1973.

[19] R. (Editor) Chung. The january 17, 1995 nyogoken-nanbu (kobe) earthquake. Tech-
nical Report NIST Special Publication 901, National Institute of Standards and Tech-
nology, July 1996.

[20] C.D. Comartin, M. Greene, and S.K. Tubbesing. The hyogo-ken nanbu earthquake
january 17, 1995. Technical Report 95–04, Earthquake Engineering Research Institute,
February 1995.

[21] C.D. (Editor) Comartin. Guam earthquake reconnaissance report. Earthquake Spectra,
11(S3):63–137, April 1995.

[22] R.D. Cook, D.S. Malkus, M.E. Plesha, and R.J. Witt. Concepts and Applications of
Finite Element Analysis. John Wiley & Sons. Inc., fourth edition, 2002.

[23] CSI. SAP2000 Version 15.0.1. Computers and Structures, Inc., Berkeley, California,
http:\\www.csberkeley.com\sap2000, 2011.

[24] CSI. Perform3D. Computers and Structures, Inc., Berkeley, California,


http:\\www.csiberkeley.com\perform3d, 2012.
482

[25] A. Dazio, K. Beyer, and H. Bachmann. Quasi-static cyclic tests and plastic hinge
analysis of RC structural walls. Engineering Structures, 31:1556–1571, 2009.

[26] A. Dazio, T. Wenk, and H. Bachmann. Versuche an stahlbetontragwaden unter


zyklish-statischer Einwirkung (tests on RC walls under cyclic-static action). Tech-
nical Report 239, IBK, 1999. http://e–collection.ethbib.ethz.ch/view/eth:23296.

[27] DICTUC. Informe de Levantamiento de Daños Edificio Centro Mayor - Concepción.


Technical Report 906576/ 10-056-EE-01-R0, DICTUC S.A. Área de Ingenieria Es-
tructural, November 2010.

[28] DICTUC. Post-earthquake evaluation report, salas 1343 torre a (plaza del rio). Tech-
nical Report 879068, DICTUC, April 2010.

[29] Doğangün. Performance of reinforced concrete buildings during the May 1, 2003
Bingöl Earthquake in Turkey. Engineering Structures, 26:841–856, 2004.

[30] EERI. The nisqually, washington, earthquake: February 28, 2001. Technical report,
Earthquake Engineering Research Institute, March 2001.

[31] B.R. (Editor) Ellingwood. An investigation of the miyagi-ken-oki, japan, earthquake


of june 12, 1978. Technical report, U.S. Department of Commerce National Bureau
of Standards, October 1980.

[32] M. Elnady. Seismic Rehabilitation of RC Structural Walls. PhD thesis, McMaster


University, 2008.

[33] EQE Engineering. The october 17, 1989 loma prieta earthquake. Technical report,
EQE Engineering Inc., 1989.

[34] E. Fang. Failure Modes of RC Tall Shear Walls. In Concrete Shear in Earthquake,
pages 125–133. Taylor & Francis, 1992.

[35] FEMA 273. NEHRP Guidelines for the Seismic Rehabilitation of Buildings (FEMA
Publication 273). Building Seismic Safety Council for the Federal Emergency Man-
agement Association, Washington, D.C., October 1997.

[36] FEMA 356. Prestandard and Commentary for the Seismic Rehabilitation of Build-
ings. American Society of Civil Engineers for the Federal Emergency Management
Association, Washington, D.C., November 2000.

[37] C. Field. “RE: PCA Literature (Historic)”. to Anna Birely, March 2011.
483

[38] A.E. Fiorato, R.G. Oesterle, and W.G. Corley. Behavior of earthquake resistant
structural walls before and after repair. ACI Structural Journal, 80(5):403–413, 1983.

[39] I. Ghorbani-Renani, N. Velev, R. Tremblay, D. Palermo, B. Massicotte, and P. Leger.


Modeling and testing influence of scaling effects on inelastic response of shear walls.
ACI Structural Journal, pages 358–367, May 2009.

[40] C.K. Gulec, A.S. Whittaker, and J.D. Hooper. Fragility functions for low aspect ratio
reinforced concrete walls. Engineering Structures, 32:2894–2901, September 2010.

[41] K. Gulec, A. Whittaker, and J. Hooper. Damage states and fragility curves for low
aspect ratio reinforced concrete walls. Prepared for the Applied Technology Council
Project ATC–58, 2009.

[42] A. Halder and S. Mahadevan. Probability, Reliability and Statistical Methods in En-
gineering Design. John Wiley and Sons, Inc., New York, 2000.

[43] John F. (Editor) Hall. Northridge earthquake january 17, 1994 preliminary reconnais-
sance report. Technical report, Earthquake Engineering Research Institute, March
1994.

[44] C.R. Hart. Cracking of Reinforced Concrete Structural Walls Subjected to Cyclic
Loading. PhD thesis, University of Illinois Urbana-Champaign, 2012.

[45] M.S. Hoehler and J.F. Stanton. Simple phenomenological model for reinforcing steel
under arbitrary load. Journal of Structural Engineering, 132:1061–1069, 2006.

[46] J. Hooper. February 2010 Chilean Earthquake – Preliminary Observations from the
SEI/ASCE Chilean Earthquake Assessment Team. Powerpoint presentation, 2011.

[47] ICC. 2006 IBC Structural/Seismic Design Manual, volume 3. International Code
Council, 2007.

[48] IDIEM. Post-earthquake evaluation report, salas 1343 torre a (plaza del rio). Technical
Report 595657-12, IDIEM, March 2010.

[49] N. Ile and J.M. Reynouard. Behaviour of U-shaped walls subjected to uniaxial and
biaxial cyclic lateral loading. Journal of Earthquake Engineering, 9(1):67–94, 2005.

[50] INN. Earthquake Resistant Design of Buildings (NCh 433.Of96). Instituto Nacional
de Normalizacion (INN), Santiago, Chile, www.inn.cl, 2006.

[51] V.C. Kalevras. Earthquake damage evaluation of monolithic reinforced concrete build-
ings through detailed inspection and grading. In A. Koridze, editor, Earthquake Dam-
age Evaluation and Vulnerability Analysis of Building Structures, pages 19–56. 1990.
484

[52] H. Kaplan, S. Yilmaz, H. Binici, E. Yazar, and N. Cetinkaya. May 1, 2003 Turkey
– Bingöl earthquake: damage in reinforced concrete structures. Engineering Failure
Analysis, 11:279–291, 2004.

[53] A. Khalil and A. Ghobarah. Behaviour of rehabilitated structural walls. Journal of


Earthquake Engineering, 9(3):371–391, 2005.

[54] I. Lefas. Behaviour of Reinforced Concrete Structural Walls and Its Implication for
Ultimate Limit State Design. PhD thesis, University of London, February 1988.

[55] I.D. Lefas and M.D. Kotsovos. Strength and deformation characteristics of reinforced
concrete walls under load reversals. ACI Structural Journal, 87(6):716–726, 1990.

[56] I.D. Lefas, M.D. Kotsovos, and N.N. Ambraseys. Behavior of reinforced concrete
structural walls: Strength, deformation characteristics, and failure mechanism. ACI
Structural Journal, 87(1):23–31, 1990.

[57] H.S. Lew, E.V. Leyendecker, and R.D. Dikkers. Engineering aspects of the 1971 san
fernando earthquake. Technical Report NBS BSS–40, U.S. Department of Commerce
National Bureau of Standards, December 1971.

[58] H.W. Lilliefors. On the K-S test for normality with mean and variance unknown.
Journal of the American Statistical Association, 62:399–402, 1967.

[59] H. Liu. Effect of concrete strength on the response of ductile shear walls. Master’s
thesis, McGill University, 2004.

[60] L. Lowes. Chi–chi reconnaissance images. Personal Communication, 2001.

[61] L. Lowes, Birely A., K. Marley, C. Hart, D. Kuchma, and D. Lehman. Investigation
of the seismic response of slender planar concrete walls. In 9th US National and 10th
Canadian Conference on Earthquake Engineering, 2010.

[62] L. Lowes, J. Li, and ATC 58 SPP Team. Fragility functions for reinforced concrete
moment frames. Report to Applied Technology Council Project ATC–58, 2010.

[63] L.N. Lowes, D.E. Lehman, A.C. Birely, D.A. Kuchma, K.P. Marley, and C.R. Hart.
Behavior, analysis, and design of complex wall systems: Planar wall test program
summary document. http://nees.org/resources/3677, 2011.

[64] K. Marley. Title Unknown. PhD thesis, University of Illinois Urbana-Champaign,


2012 (expected).
485

[65] S. Mazzoni, F. McKenna, G.G. Fenves, and et al. Open system for earthquake engi-
neering simulation user manual. http://opensees.berkeley.edu, 2010.

[66] D. Mitchell, R.H. DeVall, K. Kobayashi, R. Tinawi, and W.K. Tso. Damage to
concrete structures due to the january 17, 1995, hyogo-ken nanbu (kobe) earthquake.
Canadian Journal of Civil Engineering, 23:757–770, 1996.

[67] D. Mitchell, R.H. DeVall, M. Saatcioglu, R. Simpson, R. Tinawi, and R. Tremblay.


Damage to concrete structures due to the 1994 northridge earthquake. Canadian
Journal of Civil Engineering, 22:361–377, 1995.

[68] S. Mobeen. Cyclic tests of shear walls confined with double head studs. Master’s
thesis, University of Alberta, 2002.

[69] J. Moehle. February 27, 2010 Chile Earthquake Reconnaissance Team Investigation:
Reinforced Concrete Buildings. Powerpoint presentation, 2011.

[70] J. Moehle. Photos of damage from 2010 maule earthquake. Provided through ATC94
project, 2011.

[71] Jack P. (Editor) Moehle. Preliminary report on the seismological and engineering
aspects of the january 17, 1994 northridge earthquake. Technical Report UCB/EERC-
94/01, Earthquake Engineering Research Center, 1994.

[72] B.J. Morgan, H. Hiraishi, and W.G. Corley. U.s.–Japan quasi-static test of isolated
wall planar reinforced concrete structure, volume I. Technical report, National Science
Foundation, 1986.

[73] B.J. Morgan, H. Hiraishi, and W.G. Corley. U.s.–Japan quasi-static test of isolated
wall planar reinforced concrete structure, volume II. Technical report, National Sci-
ence Foundation, 1986.

[74] NISEE (National Information Service for Earthquake Engineering). The Earthquake
Engineering Online Archive. http://nisee.berkeley.edu/elibrary/.

[75] NIST. Recommendations for Seismic Design of Reinforced Concrete Wall Buildings
Based on the 2010 Chile Earthquake. Technical Report GCR 13917-XX, prepared
by the NEHRP Consultants Joint Venture, a partnership of the Applied Technology
Council and the Consortium for Universities for Research in Earthquake Engineering,
for the National Institute of Standards and Technology, Gaithersburg, Maryland,
2013.

[76] NOAA–NGDC (National Oceanic and Atmospheric Administration – National Geo-


physical Data Center). The Significant Earthquake Database. http://www.ngdc.
noaa.gov/nndc/struts/form?t=101650&s=1&d=1.
486

[77] R.G. Oesterle. Inelastic Analysis for In-plane Strength of Reinforced Concrete Shear
Walls. PhD thesis, Northwestern University, June 1986.

[78] R.G. Oesterle, J.D. Aristizabal-Ochoa, A.E. Fiorato, H.G. Russell, and W.G. Corley.
Earthquake resistant structural walls – tests of isolated walls – phase II. Technical
report, National Science Foundation, 1979.

[79] R.G. Oesterle, J.D. Aristizabel-Ochoa, K.N. Shiu, and W.G. Corley. Web crushing of
reinforced concrete structural walls. ACI Structural Journal, 81(3):231–241, 1984.

[80] R.G. Oesterle, A.E. Fiorato, L.S. Johal, J.E. Carpenter, H.G. Russell, and W.G.
Corley. Earthquake resistant structural walls – tests of isolated walls. Technical
report, National Science Foundation, 1976.

[81] Y.-H. Oh, S.W. Han, and L.-H. Lee. Effect of boundary element details on the seis-
mic deformation capacity of structural walls. Earthquake Engineering and Structural
Dynamics, 31:1583–1602, 2002.

[82] C.A. Pagni and L.N. Lowes. Fragility functions for older reinforced concrete beam-
column joints. Earthquake Spectra, 22(1):215–238, February 2006.

[83] R. Park, M.N.J. Priestly, and W.D. Gill. Ductility of square-confined concrete
columns. Journal of the Structural Division, 108:135–137, 1982.

[84] J. Paterson. Seismic retrofit of reinforced concrete shear walls. Master’s thesis, McGill
University, month 2001.

[85] J. Paterson and D. Mitchell. Seismic retrofit of shear walls with headed bars and
carbon fiber wrap. Journal of Structural Engineering, pages 606–614, May 2003.

[86] K. Pilakoutas. Earthquake Resistant Design of Reinforced Concrete Walls. PhD thesis,
University of London, May 1990.

[87] K. Pilakoutas and A. Elnashai. Cyclic behavior of reinforced concrete cantilever walls,
part II: Discussions and theoretical comparisons. ACI Structural Journal, 92(4):425–
281, July–August 1995.

[88] K. Pilakoutas and A.S. Elnashai. Cyclic behavior of reinforced concrete cantilever
walls, part I: Experimental results. ACI Structural Journal, 92(3):271–281, May–June
1995.

[89] J. Pugh. Numerical Simulation of Walls and Capacity Design Recommendations for
Walled Buildings. PhD thesis, University of Washington, 2012.
487

[90] D.J. Raynor. Bond assessment of hybrid frame continuity reinforcement. Master’s
thesis, University of Washington, 2000.

[91] R. Riddell, S.L. Wood, and J.C. De La Llera. The 1985 Chile earthquake struc-
tural characteristics and damage statistics for the building inventory in Viña del Mar.
Technical report, University of Illinois at Urbana-Champaign, April 1987.

[92] P. Riva, A. Meda, and E. Giuriani. Cyclic behaviour of a full scale RC structural wall.
Engineering Structures, 25:835–845, 2003.

[93] M. Saatcioglu and M. Bruneau. The 1992 erzincan earthquake. Concrete Interna-
tional, pages 51–56, September 1994.

[94] M. Saatcioglu, N.J. Gardner, and A. Ghobarah. 1999 turkey earthquake performance
of rc structures. Concrete International, pages 46–56, March 2001.

[95] M. Saatcioglu, D. Mitchell, R. Tinawi, N.J. Gardner, A.G. Gillies, A. Ghobarah, D.L.
Anderson, and D. Lau. The August 17, 1999, Kocaeli (Turkey) earthquake – damage
to structures. Canadian Journal of Civil Engineering, 28:715–737, 2001.

[96] SEAW. Kobe earthquake reconnaissance report. Technical report, Structural Engi-
neers Association of Washington, 1995.

[97] H. Sezen, K.J. Elwood, A.S. Whittaker, K.M. Mosalam, J.W. Wallace, and J.F. Stan-
ton. Structural engineering reconnaissance of the August 17, 1999, Kocaeli (Izmit),
turkey, earthquake. Technical report, Pacific Earthquake Engineering Research Cen-
ter, December 2000.

[98] H. Sezen, A.S. Whittaker, K.J. Elwood, and K.M Mosalam. Performance of reinforced
concrete buildings during the august 17, 1999 kocaeli, turkey earthquake, and seismic
design and construction practise in turkey. Engineering Structures, 25:103–114, 2003.

[99] G.H. (Editor) Shea. Erzincan, turkey earthquake of march 13, 1992 reconnaissance
report. Earthquake Spectra, 9(S1):49–86, 1993.

[100] K.N. Shiu, J.I. Daniel, J.D. Aristizabal-Ochoa, A.E. Fiorato, and W.G. Corley. Earth-
quake resistant structural walls – tests of walls ith and without openings. Technical
report, National Science Foundation, 1981.

[101] R.K.L. Su and S.M. Wong. Seismic behaviour of slender reinforced concrete shear
walls under high axial load ratio. Engineering Structures, 29:1957–1965, 2007.

[102] A.A. Tasnimi. Strength and deformation of mid-rise shear walls under load reversal.
Engineering Structures, 22:311–322, 2000.
488

[103] J.H. Thomsen, IV and J.W. Wallace. Displacement-based design of reinforced con-
crete structural walls: Experimental studies of walls with rectangular and T–shaped
cross sections. Technical Report 95/06, Department of Civil and Environmental En-
gineering, Clarkson University, Potsdam, N.Y., 1995.

[104] J.H. Thomsen, IV and J.W. Wallace. Displacement-based design of slender reinforced
concrete structural walls – experimental verification. Journal of Structural Engineer-
ing, 130(4):618–630, April 2004.

[105] D. Todd, N. Carino, R. Chung, H.S. Lew, A.W. Taylor, W.D. Walton, J.D. Cooper,
and R. Nimis. 1994 northridge earthquake performance of structures, lifelines, and
fire protection systems. Technical Report NIST Special Publication 862, National
Institute of Standards and Technology, March 1994.

[106] B. Tupper. Seismic response of reinforced concrete walls with steel boundary elements.
Master’s thesis, McGill University, 1999.

[107] Unknown. Centro mayor structural drawings. Provided through ATC94 project.

[108] Unknown. Concepto urbano structural drawings. Provided through ATC94 project.

[109] Unknown. Plaza del rio building a etabs model. Provided through ATC94 project.

[110] Unknown. Plaza del rio structural drawings. Provided through ATC94 project.

[111] USGS (U.S. Geological Survey). Usgs earthquake hazards program. http://
earthquake.usgs.gov/learn/photos.php.

[112] USGS (U.S. Geological Survey). USGS Photographic Library. http://


libraryphoto.cr.usgs.gov/.

[113] J. Wallace and J. Moehle. Ductility and detailing requirements of bearing wall build-
ings. Journal of Structural Engineering, 118(6):1625–1644, 1992.

[114] B. Westenenk, J.C. de la Llera, J.J. Besa, R. Junemann, J. Moehle, C. Luders, J.A.
Inaudi, K.J. Elwood, and S.J. Hwang. Response of reinforced concrete buildings in
concepcion during the maule earthquake. Provided through ATC94 project, 2011.

[115] S.L. Wood. Observed Behavior of Slender Reinforced Concrete Walls Subjected to
Cyclic Loading. In Special Publication 127: Earthquake-Resistant Concrete Structures
Inelastic Response and Design, pages 453–477. American Concrete Institute, Detroit,
MI, 1991.
489

[116] S.L. Wood, R. Stark, and S.A. Greer. Collapse of eight-story RC building during 1985
Chile earthquake. Journal of Structural Engineering, 117(2), February 1991.

[117] S.L. Wood, J.K. Wight, and J.P. Moehle. The 1985 Chile earthquake observations
on earthquake-resistant construction in Viña del Mar. Technical report, University of
Illinois at Urbana-Champaign, February 1987.

[118] L.A. Jr. Wyllie. Lessons from the armenian earthquake. Concrete International, pages
21–26, August 1989.

[119] M. Yoshimine. Earthquake image archives (tokyo metropolitan university). http:


//geot.civil.metro-u.ac.jp/archives/eq/, 2008.

[120] Y. Zhang and Z. Wang. Seismic behavior of reinforced concrete shear walls subjected
to high axial loading. ACI Structural Journal, 97(5):739–750, September 2000.
490

LIST OF SYMBOLS

Acv Area of wall resisting shear

Ag Total cross-sectional area of wall


Total cross-sectional area of transverse reinforcement (including cross-ties)
Ash
within spacing s and perpendicular to dimension bc (ACI 318-08 Ch. 21)
D Dead load (ACI 318-08 Ch. 9)

E Load effects of earthquake (ACI 318-08 Ch. 9)

Ec Elastic modulus of concrete


Es Modulus of elasticity

AR Aspect ratio. Equal to the wall height, hw , divided by the wall length, `w

Esh Tangent stiffness at on-set of strain hardening

Ig Cross section moment of inertia


L Live load (ACI 318-08 Ch. 9)

M Base moment
M
V `w Shear span ratio

( VM`w )strong Shear span ratio for strong axis bending

( VM`w )weak Shear span ratio for weak axis bending

P Axial force
U Required strength (ACI 318-08 Ch. 9)

V Base shear
Vn ACI 318 nominal shear strength

Vu Maximum experimental shear strength

Vu /Vn Shear demand-capacity ratio


491

Cross-sectional dimension of column core measured center-to-center of outer

bc legs of the transverse reinforcement comprising area Ash (ACI 318-08 Ch.

21)
bf Width of boundary element (barbell wall); Flange width (flanged wall)

c distance from extreme compression fiber to neutral axis

fc0 Concrete compressive strength

fy Rebar yield stress

fu Rebar ultimate stress


h10 Height of 1/3-scale ten-story wall
Thickness of boundary element (barbell wall); Flange thickness (flanged
hf
wall)
hef f Effective height of load

hw Height of wall specimen


Maximum center-to-center horizontal spacing of crossties or hoop legs (ACI
hx
318-08 Ch. 21)
`be Boundary element length

`w Wall length

s center-to-center spacing of transverse reinforcement (ACI 318-08 Ch. 21)


center-to-center spacing requirement of transverse reinforcement (ACI 318-
so
08 Ch. 21)
tw Wall thickness
vn Normalized ACI 318 nominal shear strength

vu Normalized maximum experimental shear stress

αload Factor indicating height of effective load relative to specimen height hw

β Fragility function dispersion

βd Dispersion of experimental data

βu Dispersion due to uncertainty of experimental data

δu Design displacement (ACI 318-08 Ch. 21)

εy Yield strain
492

εsh Strain at on-set of strain hardening

εt Steel tensile strain at nominal strength (ACI 318-08 Ch. 9)

εu Ultimate strain
Factor accounting for lightweight concrete (equal to 1.0 for normalweight
λ
concrete)
λN Axial load ratio
φ Strength reduction factor (ACI 318-08 Ch. 9)

ρbe Boundary element longitudinal reinforcement ratio

ρcon Boundary element confining reinforcement ratio (volumetric)

ρh Horizontal reinforcement ratio


ρweb Web longitudinal reinforcement ratio

ρv Total vertical reinforcement ratio


σ Standard deviation of lognormal dispersion

µ Mean of lognormal distribution


493

Appendix A

SUMMARY OF EXPERIMENTAL DAMAGE

This appendix provides detailed summaries of the progression of damage during the

tests of the four planar wall specimens. Included are representative photos1 taken during

the tests and crack maps that indicate the crack patterns of the walls2 . For each specimen,

the damage discussion is provided a sequential order. Section headings indicate the drift

level and type(s) of damage that is discussed in the subsequent paragraph(s). Time-lapse

movies created from photos of the bottom story of the wall specimens can be found on the

project youTube Channel (http://www.youtube.com/user/NEESRWallProject).

A.1 PW1

Specimen PW1 was constructed with a boundary element distribution of longitudinal re-

inforcement. The longitudinal reinforcement was lap spliced at the base of the wall. The

wall was loaded to create base reactions representative of an ASCE 7 lateral load dis-

tribution on a 10-story wall. Section 4.1 provides an overall summary of the Specimen

PW1 test. A time-lapse movie of damage to the bottom story of PW1 can be found at

http://www.youtube.com/user/NEESRWallProject#p/u/1/KC7b-k3ZRjs. The following

sections provide a detailed summary of the wall damage.

1
All photos taken during the tests can be found in the NEES Project Warehouse https://nees.org/
warehouse/project/104/.
2
Crack maps were provided by Chris Hart at the University of Illinois [44].
494

0.10% Drift, Initial Horizontal and Diagonal Cracking

Horizontal cracks were first visible at 0.06% (Step 73) during the first ER+ cycle to 0.10%

drift. These cracks were less than 4 inches in length and evenly distributed along the edge

of the west boundary element. At the peak (0.101%, Step 75), a horizontal crack above the

splice had extended nearly the length of the boundary element and a diagonal crack were

visible in the upper half of the first floor web. In the same cycle in the WL- direction, a

diagonal crack was visible in the lower half of the first floor of the web next to the east

boundary element at −0.08% drift (Step 84). Following the next step at −0.101% drift,

horizontal cracks were visible in the east boundary element above the splice and the upper

half of the first floor in the adjacent web. Additional diagonal cracks were visible at the

ends of the horizontal cracks and in the upper quarter of the first floor in the east boundary

element. Figure A.1 shows the crack patterns in the first two floors of the wall following

three cycles at 0.10% drift.

Figure A.1: PW1 crack pattern in lower two stories following three cycles at 0.10% drift.
495

0.25% Drift, Extension of Cracking

Figure A.2 shows the first and second floor of the wall following three cycles at 0.25% drift.

Horizontal cracks in the upper half of the first floor have increased in density and extend

the length of the boundary element. A couple of these cracks extended into the web before

becoming or joining with diagonal cracks. In the lower half of the first floor, where the steel

in spliced, a few horizontal cracks had extended but there was no increase in the number of

cracks and a few diagonal cracks were visible. Diagonal cracks formed by displacement in

the ER+ and WL- directions cross and extend as far as 25 and 14 inches beyond the center

of the wall, respectively.

Figure A.2: PW1 crack pattern in lower two stories following three cycles at 0.25% drift.
496

0.347% and 0.50% Drift, Vertical Cracking

The next set of cycles displaced the wall to 0.347% drift. Figure A.3a shows crack pattern

in the bottom two stories following three cycles. At the ER+ peak of the first cycle at 0.50%

(Step 360), vertical cracks where visible on the east back corner of the wall above the splice,

as shown in Figure A.3b. At Step 456, a vertical crack formed at the front east corner of

the wall. At the final WL- peak at 0.50% (Step 460), a vertical crack was observed on the

front of the east boundary element near the top of the splice and a couple inches in from

the edge of the wall.

(a) Crack pattern in lower two stories following three (b) East side of wall at first ER+ peak
cycles at 0.35% drift. at 0.50% drift (Step 360)

Figure A.3: PW1 damage during 0.35% and 0.50% drift cycles.
497

0.75% Drift, East (Right) Side, Vertical Cracking and Cover Spalling

At a drift of 0.60% (Step 478) during the first 0.75% drift cycle, the vertical crack front east

(right) side of the wall extended to a few inches above the splice. At the top of this crack

the concrete cover was separated from the wall (see Figure A.4) and as the cycle continued,

this continued down the vertical crack. Spalling occurred at the east toe of the wall along

a crack about an half inch above the wall-foundation interface and extended about two

inches from the east edge of the wall. Following Step 560 (final ER+ peak at 0.75% drift),

a number of vertical cracks were observed in the east boundary element of the wall.

Figure A.4: PW1 vertical splitting at top of east splice at first ER+ peak at 0.75% drift.
498

0.75% and 1.0% Drift, West Side, Vertical Cracking, Cover Spalling and Exposed Reinforce-

ment

A vertical crack was observed in Step 480 (first E+ peak at 0.75% drift) in the west boundary

element above the top of the splice. This crack extended and the cover concrete separated in

steps 535 through 540 (second WL- peak @ −0.7856%). Unloading from this peak displace-

ment, the spalled concrete fell off the wall, but did not expose longitudinal reinforcement

(see Figure A.5a). More of the spalled concrete fell off the wall unloading from the final WL-

peak of the 0.75% drift level (Step 580). At Step 593 (0.30% drift) on the way to the first

ER+ peak at 1.0% drift, spalling above the splice on the west side of the wall progressed to

the point that longitudinal reinforcement was exposed, as shown in Figure A.5b.

A vertical crack formed at the bottom west corner of the wall at the first WL- peak at

0.75%3 . The cover at this location fell off the wall at Step 620 (first WL- peak at 1.0%).

Vertical cracks formed in west boundary element in the spliced region next to the web, in

the center of the boundary element in the spliced region and above the splice, and in the

upper quarter of the first floor in the web adjacent to the boundary element. Figure A.6

shows initial bar buckling on the west side of the wall, which occurred during the first cycle

to 1.0% drift.

3
Occurred at Step 500 but not marked until Step 540 in photos
499

(a) Second WL- peak at 0.75% drift (b) Exposed reinforcementon way
(Step 540) to first ER+ peak at 1.0% drift
(Step 593)

Figure A.5: PW1: Spalling on west side of wall during 0.75% and 1.0% drift cycles.

Figure A.6: PW1: Initial bar buckling on west side at first WL- peak at 1.0% drift (Step
620)
500

1.0% Drift, East Side

At a drift of 0.80% (Step 598), the vertical crack at the east edge of the wall extended

from above the splice to the base of the wall. Figure A.7 shows the crack, and spalling at

the top of the wall, at the first ER+ peak at 1.0% drift (Step 600). At the WL- peak of

0.9654% drift (Step 620) the cover defined by the vertical crack was separated from the wall

and spalling occurred at the toe of the wall. A vertical crack on the east side of the wall

also appeared and spanned the height of the splice. The separation of the cover concrete

from the wall continued throughout the test. At step 634 (0.368% in second ER+ cycle to

1.0%), an inclined crack formed on the east edge of the wall in the upper half of the first

floor and above the previously damaged region. In the remaining steps to the ER+ peak,

cover crushing was observed along this crack. Ultimately this fell off the wall as loading

continued.

Figure A.7: PW1: Vertical crack on east edge of wall at first ER+ peak at 1.0% drift (Step
600)
501

1.0% Drift, West Side, Bar Buckling

At the final WL- peak at 1.0% drift (Step 660), there was a region of concrete cover crushing

about approximately 10 inches above the splice in the west boundary element4 . Figure A.8

shows multiple bars buckled above the splice on the west side of the wall.

Figure A.8: PW1: Multiple bars buckled at final WL- peak at 1.0% drift (Step 620)

4
Note: this is very similar to what was seen for PW2 at the same place, but here the damage did not
continue to grow
502

1.5% Drift, East Side, Exposed Longitudinal Reinforcement, Bar Buckling

Longitudinal reinforcement was exposed on the east back corner of the wall (see Figure A.9a)

at 1.35% drift (Step 679 during first cycle to ER+ at 1.5%). Figure A.9b shows buckling

of the middle outer bar on east side above the splice at the first ER+ peak at 1.50% drift

(Step 680). When unloading from this step, loose concrete fell off the side of the wall and

exposed the front east corner bar the full length of the splice.

(a) Exposed reinforcement on way to first ER+ (b) Bar buckling on way to first ER+ peak at 1.5%
peak at 1.5% drift (Step 679) drift (Step 680)

Figure A.9: PW1: East side of wall during 1.5% drift cycles
503

0.75%-1.5% Drift, Localized Spalling Above Splice

During the cycles at 0.75% drift, spalling occurred along a horizontal crack above the splice

in the west boundary element. The extent of the spalling increased in the cycles at 1.0% and

1.5% drift. Similar spalling occurred along a horizontal crack in the east boundary element

beginning in the 1.0% drift cycles, although the extent of the damage was less than that in

the west boundary element. The spalling occurred in both elements while the regions were

in compression.5

5
This is not the primary damage for the PW1 test, but it is very similar to that seen above the splice in
other walls, and thus is noted.
504

1.5% Drift, West Side, Core Damage

During the first cycle at 1.5% drift, the spalled region above the west boundary element

splice grew but the concrete largely remained attached to the wall. Unloading from the

second ER+ peak, spalled concrete fell off the wall and revealed damage to the core above

the west boundary element splice. As loading the wall towards the second WL- peak, the

core concrete continued to deteriorate.6

6
Core damage did not progress as quickly or to the same extent as in other walls.
505

1.5% Drift, Loss of Load Carrying Capacity and Bar Fracture

In the steps leading up to the second WL- peak at −1.5% drift, there was i) significant

opening of a crack along the wall-foundation interface ii) increate loss of cover concrete

along the east side of the wall, iii) loss of cover concrete along the length of the boundary

element and iv) a shear crack extends into the base of the east boundary element. At

−1.379% (Step 739) the increase in the opening of the interface crack (see Figure A.10)

was substantially more than that seen in the previous steps and there was a drop in the

lateral force applied to the wall. Longitudinal reinforcement7 was exposed on the front and

back faces of the boundary element. Damage to the core concrete was visible. Figure A.11

shows the back east side of the wall, where the second longitudinal bar in from the edge is

fractured. Figure A.12 shows the front of the east boundary element where the third bar

in from the corner is fractured. At the peak (Step 740, −1.523%), uplift of the wall further

resulted in more core damage and bar fracture. Figures A.12 and A.13 show the fractured

rebar on the front and back faces of the east boundary element, respectively.

7
All seven bars along the front.
506

Figure A.10: PW1: Interface crack opening at drop in lateral load carrying capacity (−1.38%
drift, Step 739)

Figure A.11: PW1: Longitudinal bar fracture on back face of east boundary element at
drop in lateral load carrying capacity (−1.38% drift, Step 739)
507

Figure A.12: PW1: Longitudinal bar fracture on front face of east boundary element at
drop in lateral load carrying capacity (−1.38% drift, Step 739)

Figure A.13: PW1: Longitudinal bar fracture on front face of east boundary element at
final peak (−1.523% drift, Step 740)
508

Figure A.14: PW1: Longitudinal bar fracture on back face of east boundary element at
final peak (−1.523% drift, Step 740)
509

A.2 PW2

Specimen PW2 was constructed with a boundary element distribution of longitudinal re-

inforcement. The longitudinal reinforcement was lap spliced at the base of the wall. The

wall was loaded to create base reactions representative of a uniform lateral load distri-

bution on a 10-story wall. Section 4.2 provides an overall summary of the Specimen

PW2 test. A time-lapse movie of damage to the bottom story of PW2 can be found at

http://www.youtube.com/user/NEESRWallProject#p/u/2/Miepayt10Vk. The following

sections provide a detailed summary of the wall damage.


510

0.14% Drift, Initial Cracking

The first cracks were observed in the first cycle at the 0.139% drift level. When moving in

the ER+ direction, cracks were first observed and marked following Step 137 at 0.097% drift.

Horizontal cracks were observed in the boundary region, and diagonal cracks were observed

in the web, some of which appeared to be extensions of shrinkage cracks marked prior to the

beginning of testing. When moving in the WL- direction, small (approximately 2 inches)

horizontal and diagonal cracks were marked following step 153 at a drift of −0.041%. The

first large cracks, both horizontal and diagonal, were marked following Step 157 at −0.096%

drift. Upon completion of three cycles at ±0.139% drift (see Figure A.15 for crack pattern),

the inclined cracks extended between one-third and one-half the length of the wall8 .

Figure A.15: PW2 crack pattern in lower two stories following three cycles at 0.14% drift.

8
The cracks extend lower than in PW4. Cracks resulting from ER+ displacement have a lowest point
below that of those resulting from WL- displacement of the wall.
511

0.21% Drift, Extension of Cracking

After completion of the first cycle at ±0.208% drift (peak Steps 260 and 280), inclined

cracks had grown such that at least one in each direction passed the center of the wall and

a crack from each direction intersected9 . Following the third cycle at 0.208% drift (peak

steps 340 and 360), inclined cracks extend to the third rod in from the east and west of

the application of the first floor shear force. The crack pattern in the lower two floors after

three cycles at 0.21% drift is shown in Figure A.16).

Figure A.16: PW2 crack pattern in lower two stories following three cycles at 0.21% drift.

9
Cracks initial crossed mid-point of wall at 0.190% (ER+ step 259) and −0.184% (WL- step 279)
512

0.347% Drift, Vertical Cracking

Vertical cracks were first observed at the peak displacement steps of the second cycles to

±0.347% drift (Steps 470 and 510). Cracks formed just below the top of the splice region in

the east boundary element and just below and above the top of the spliced region in the west

boundary element. The vertical cracks were formed while the boundary elements were in

tension. The crack pattern following three cycles at ±0.347% drift is shown in Figure A.17).

Figure A.17: PW2 crack pattern in lower two stories following three cycles at 0.35% drift.
513

0.50% Drift, Extension of Cracks

Three cycles were completed at ±0.50% drift. During these cycles, existing cracks extended

and new cracks formed Inclined cracks extended to approximately (give or take an inch or

two) the eastern and western most rods used to apply the first floor shear force. The crack

pattern in the lower two floors following three cycles to 0.50% drift is shown in Figure A.18.

Figure A.18: PW2 crack pattern in lower two stories following three cycles at 0.14% drift.
514

0.75% Drift, Cover Spalling, Wide Horizontal Crack,

Spalling was first observed during the first cycle at 0.75% drift, spalling was first observed.

At the first ER+ peak (0.757% drift), minor flaking of the concrete was observed at the

very bottom at the east end of the wall. At this same step, a horizontal crack in the

west boundary element, just above the top (aprroximately 1 inch) of the splice, measured

0.1 inches wide (see Figure A.19). After unloading to approximately zero lateral force,

this crack was still very noticeable while surrounding cracks were hairline cracks if visible

at all. With increasing lateral load in the WL- direction, this crack continued to close,

reaching its smallest width at −0.1416% (Step 894). At this same step, vertical splitting

and spalling of the concrete10 ocurred above the closed crack at the west edge of the wall as

shown in Figure A.20 (Figure A.21 shows the west toe of the wall at the same step). The

spalled concrete pushed out over the next couple steps but the amount of spalled concrete

did not increase as the wall was displaced to the cycle peak. At the WL- direction peak

(−0.739%) drift, additional spalling was noticed at closed cracks (horizontal and inclined)

at the top of the splice in the west boundary element. This damage is marked by dotted

lines in Figure A.22. At this same step, a flexure-shear crack just above the splice in the

east boundary element measured 0.075 inches. At lateral force of approximately zero after

unloading from this step, the crack was still visible11 while the surrounding cracks were not

visible. Concrete fell of the bottom east corner at Step 1031, 0.718% drift, (see Figure A.23),

but it did not expose longitudinal reinforcement. Figure A.24 shows spalling along vertical

cracks at the top of the east boundary element splice.

10
The crack formed under compressive loading. Vertical cracks had been observed previously, but were
formed while the concrete was in tension.
11
The crack was visible, but visual inspection indicated it was not wide as the residual crack in the west
boundary element was previously
515

Figure A.19: PW2: West boundary element at first ER+ peak at 0.75% drift, with 0.1 inch
wide crack above splice (Step 870)

Figure A.20: PW2: Vertical splitting above west boundary element splice at WL- peak at
−0.75% drift (Step 910)
516

Figure A.21: PW2: West toe of wall at WL- peak at −0.75% drift (Step 910)

Figure A.22: PW2: Spalling along cracks above splice in west boundary element at WL-
peak at −0.75% drift (Step 910)
517

Figure A.23: PW2: East toe of wall following third ER+ peak at 0.75% drift (Step 1030)

Figure A.24: PW2: Vertical cracks at top of splice in east boundary element following third
ER+ peak at 0.75% drift (Step 1030)
518

0.75% Drift, Exposed Reinforcement and Bar Buckling (West B.E.)

As loading continued, concrete continued to spall along the west edge of the wall at and

above the top of the splice, ultimately exposing the longitudinal reinforcement at −0.179%

drift (Step 1055) while moving to the third WL- peak of the 0.75% drift cycles. When

the final peak was reached at this drift level (Step 1070 @ −0.741%), longitudinal buckling

was observed in one of the outer most longitudinal bars in the west boundary element (see

Figure A.25).
519

Figure A.25: PW2: Bar buckling in west boundary element following third WL- peak at
−0.75% drift (Step 1070)
520

1.0% Drift, Exposed Reinforcement and Bar Buckling (East B.E.)

During the first cycle to 1.0% drift (Step 1102 @ 0.61% drift), more spalling occurred above

the east boundary element splice was seen when a small piece (approximately 1 cm x 3

cm) fell off the corner of the wall. Throughout the remainder of the 1.0% drift cycles,

the extent of the cover spalling at the top of the splice in the boundary element grew to a

region of approximately 2 inches x 6 inches, revealing the confining ties and the longitudinal

reinforcing bars. At the final ER+ peak (Step 1190 @ 1.007% drift), bar buckling was

observed in the east boundary element (see Figure A.26).

Figure A.26: PW2: Bar buckling above east boundary element splice at third ER+ peak at
1.0% drift (Step 1190)
521

1.0% Drift, Progression of Spalling in West B.E.

The damage at the top of the west boundary element splice grew more rapidly during the

1.0% drift cycles than that of the east boundary element. When the first WL- peak was

reached at a drift of −0.987% (Step 1150), the three outer most longitudinal bars had

buckled (Figure A.27). In earlier cycles, spalling was detected at a horizontal crack at

the top of the splice; the spalling spread out vertically about an inch from this crack and

extended approximately 2/3 into the boundary element. In the second and final cycle at

1.0% drift, the spalled concrete region expanded to a square region about 18x10 inches, with

the majority of this being the concrete cover seperated from the core, but not yet having

fallen off the wall. A 3x10 inch region of the core was visible on the front of the wall and

core damage was not present.

Figure A.27: PW2: Extensive bar buckling above west boundary element splice at first Wl-
peak at 1.0% drift (Step 1150)
522

1.5% Drift, Core Damage (West B.E.), Loss of Lateral Load Carry Capacity

The first cycle beyond the 1.0% drift level saw the wall pushed to a drift of 1.5% in the

ER+ direction. Figure A.28 shows spalling in the east boundary element spanning from the

bottom of the wall to above the splice. Unloading from this point, the west boundary element

experienced core crushing and bar buckling to the extent that, when pushing the wall in the

WL- direction, the load-displacement curve fell below that of the previous cycle. The first

step beyond the previous maximum drift in the WL- direction was completed (Step 1304

@ −1.046% drift). Damage at this step is shown in Figure A.29. While loading to the next

step, the axial load dropped from 560 kips to 220 kips, longitudinal bars buckled in the entire

west boundary element, the region of crushed concrete expanded from the boundary element

into the web, and longitudinal and horizontal #2 bars in the web buckled. To complete

the step, a reduced axial load and a reduced moment were necessary. Figures A.30 to A.33

show the damage following completion of Step 1305 to a drift of −1.090%. At this point, the

prescribed loading protocol was discarded and the wall was manually unloaded laterally by

reducing the displacement and moment. The axial load was reapplied during this process.

During this process, web crushing extended to two-thirds of the web and additional #2 bars

buckled. Figure A.34 shows the wall after the test was completed.
523

Figure A.28: PW2: Spalling in east boundary element at 1.5% drift (Step 1250)

Figure A.29: PW2: Damage to west boundary element at −1.05% drift (Step 1304). One
step prior to loss of load carrying capacity.
524

Figure A.30: PW2: Web crushing and bar buckling following loss of load carrying capacity
(Step 1305)

Figure A.31: PW2: West boundary element and web following loss of load carrying capacity
(Step 1305)
525

Figure A.32: PW2: Damage to back side of wall following loss of load carrying capacity
(Step 1305)

Figure A.33: PW2: East boundary element following loss of load carrying capacity (Step
1305)
526

Figure A.34: PW2: Damage to the wall in its final resting place
527

A.3 PW3

Specimen PW3 was constructed with a uniform distribution of longitudinal reinforcement.

The longitudinal reinforcement was lap spliced at the base of the wall. The wall was loaded

to create base reactions representative of a uniform lateral load distribution on a 10-story

wall. Section 4.3 provides an overall summary of the Specimen PW2 test. A time-lapse

movie of damage to the bottom story of PW3 can be found at http://www.youtube.com/

user/NEESRWallProject#p/u/3/WVtKmeC0FPU. The following sections provide a detailed

summary of the wall damage.


528

0.14% Drift, Initial Cracking (Horizontal, Diagonal and Vertical)

Cracks were first visible during the first cycle to 0.139% drift. Horizontal cracks appeared in

the third quarter (above the splice) of the first floor in both boundary elements at ±0.055%

drift (Step 134 and 154). In the west boundary element, two horizontal cracks spanned the

length of the boundary region. In the east boundary element, cracks spanned the length of

the boundary region just above the top of the splice and another crack was visible in the

center of the upper half of the splice.

At 0.066% drift in the ER+ direction (Step 135), an inclined crack was visible in the

bottom quarter of the first floor web next to the boundary element. Inclined cracking was

first visible at −0.082% when loading in the WL- direction (Step 156). This crack appeared

in the third quarter of the first floor, originating at the web-boundary region interface and

extending downward into the web.

Following the first cycle to 0.139% drift, at least one inclined crack from each direction

of loading had extended to, or past, the center of the wall and had interesected each other.

Although initial horizontal cracks appeared in the boundary region, horizontal cracks also

occurred in the web, the most noticable being in the eastern half of the wall, approximately

13 inches above the foundation. At the first ER+ peak (Step 140, 0.136% drift), a vertical

crack appeared in the lower quarter of the first floor at the interface of the boundary region

and the web. Figure A.35 shows the first floor of the wall following three cycles at 0.139%

drift.
529

Figure A.35: PW3: Crack pattern following 3 cycles at ±0.14% drift (Step 250)
530

0.21% Drift, Extension of Cracks

Figure A.36 shows the damage following three cycles at 0.208% drift. During these cycles,

the progression of damage was extention of existing cracks and formation of new cracks.

Most notably, another vertical crack appeared in the west half of the web12 , and an inclined

crack formed at the west edge of the wall a few inches above the splice.

Figure A.36: PW3: Crack pattern following 3 cycles at ±0.21% drift (Step 370)

12
In the top half of the splice
531

0.347% Drift

Figure A.37 shows the damage in the first floor following three cycles at 0.347% drift. Above

the splice in the west boundary region additional inclined cracks formed under tension,

including one slope opposite that of the other cracks. Vertical cracks formed on the west

side of the wall above the splice following the second ER+ peak (Step 470). Following the

second WL- peak (Step 510), additional vertical cracks had formed in the spliced region

at the interface of the boundary region and web. In the east boundary element, a vertical

crack formed above the splice at −0.291% drift (Step 427). At the final WL- peak (Step

590), a vertical crack formed in tension near the interface of the web and the east boundary

region in the top half of the spliced region.

Figure A.37: PW3: Crack pattern following 3 cycles at ±0.35% drift (Step 610)
532

0.50% Drift, Cover Spalling (West B.E.), Residual Crack Widths

On the way to the first 0.50% WL- peak (−0.396% drift, Step 666), spalling of the cover

concrete in the west boundary region was observed. The spalling occurred near the top of

the splice and was not visible from a distance. At the peak (Step 670), a crack on the east

side of the wall along the wall-foundation interface measured 0.03 inches wide13 . At the

second ER+ peak (0.50% drift, Step 710), the horizontal crack above the splice in the west

boundary region measured 0.075 inches in width14 . After unloading to zero lateral force,

no cracks measured wider than 0.005 inches. At the second WL- peak (−0.495% drift, Step

750), the widest crack was the second horizontal crack above the top of the splice in the east

boundary region. This crack measured 0.035 inches wide15 at the peak and 0.02 inches at

zero lateral force. Also at the second WL- peak, a vertical crack was observed in the center

of the lower half of the spliced region of the west boundary element, and an inclined crack

oriented in the opposite direction expected formed above the east boundary region splice.

During the final cycle to ±0.50% drift, the width of the previously mentioned horizontal

crack in the west boundary region did not increase, but the widest inclined crack increased

to a width of 0.025 inches. At the final WL- peak, the crack at the wall-foundation interface

had increased in width16 . The widest crack in the east boundary region had increased in

width to 0.040 inches. Figure A.39 shows the region at the top of the west boundary element

splice, where additional spalling of the cover was marked with dashed lines17 . Figure A.38

shows the damage to the first floor following completion of the the cycles.

13
Extended approximately 3 inches into boundary element from corner of wall
14
This was clearly the widest crack. Most other cracks, horizontal and inclined, measured 0.01-0.02 inches
wide
15
This is smaller than the similar crack in the west boundary region, but width of the other cracks at this
step, 0.015-0.025 inches, is greater those formed by the opposite direction
16
Based on visual observations of photographs; a measured width was not recorded
17
The spalling itself, which occurred along closed cracks, cannot be seen in the photograph, and therefore
the damage was marked with the dashed lines
533

Figure A.38: PW3: Crack pattern following 3 cycles at ±0.50% drift (Step 850)

Figure A.39: PW3: West boundary element at the third WL- peak at 0.50% drift (Step
830)
534

0.75% Drift Cycle A, Cover Spalling (East B.E.),

Three cycles were completed to ±0.75% drift. During the first cycle at 0.524% (Step 864),

moderate cover spalling occurred in the east boundary region above the splice18 At the first

ER+ peak, the most noticable spalling was the seperation of the cover concrete along a

vertical crack (formed in compression) at the east edge of the wall at the top of the third

quarter of the first floor, as shown in Figure A.40. Figure A.41 shows the damage above

the west boundary region splice at the same step. The wide horizontal crack measured 0.10

inches wide19 . The surrounding horizontal cracks were no greater that 0.025 inches and the

inclined cracks in the web were no greater that 0.02 inches. Spalling was indicated by dashed

lines in the west web of the wall above the splice. At the first WL- peak (Step 910, −0.748%

drift), the widest horizontal crack in the east boundary region above the splice measured

0.075 inches. Other horizontal cracks measured 0.025-0.035 inches wide and inclined cracks

measured 0.015-0.030 inches wide20 . In the step leading to the WL- peak, additional regions

of spalling were marked in the west half of the wall and vertical cracks formed in the west

boundary region in the upper portion of the third quarter of the first floor.

18
Marked with dash line.
19
When the lateral force unloading from this step was approximately zero, the residual width of this crack
was 0.05 inches. The residual width of the surrounding cracks were no greater that 0.015 inches. Inclined
cracks in the second floor of the wall had a width no greater that 0.005 inches.
20
Most residual cracks width were no greater 0.02 inches in the first floor and 0.005 inches for inclined
cracks in the second floor. The widest horizontal crack width at the peak had a residual width of 0.03
inches
535

Figure A.40: PW3: Spalling in region above splice in east boundary element at first ER+
peak at 0.75% drift (Step 870)

Figure A.41: PW3: Damage above splice in west boundary element at first ER+ peak at
0.75% drift (Step 870)
536

0.75% Drift Cycle B and C, Progression of Spalling, Exposed Reinforcement

After reaching the second ER+ peak (0.754% drift, Step 950) spalling of the cover above the

splice spanned the length of the wall, extending as high as a foot above the splice21 . After

reaching the second WL- peak (−0.744% drift, Step 990), new inclined cracks formed nearly

as far west as the west boundary element and vretical cracks formed in the spliced region

at the center of the west boundary element and at the interface of the boundary element

and the web. At the east side and corner of the wall above the splice, where spalling

was first observed, a six inch tall region of concrete fell off the wall, exposing transverse

reinforcement on the edge of the wall (see Figure A.42). At the third ER+ peak (Step

1030) and a small amount of the back corner longitudinal bar was exposed22 . A region

of concrete was marked with a blue dashed line where the concrete cover was ‘bulging’

out of plane nearly an half-inch23 . The bulging region extended from the west edge of the

wall and across the entire web, ending next to the east boundary element. The height of

the region varied from approximately six inches at the west edge (located directly above

the splice), to a maximum height of nearly two feet in the web (one foot below the top

of the splice to nearly one foot above the top of the splice). The crack widths had not

increased from that observed at the first ER+ peak at 0.75% drift. Spalled concrete fell

of the west side of the wall above the splice and exposed transverse reinforcement but not

longitudinal reinforcement. After the final WL- peak at 0.75% drift, the region of ‘bulging’

extended higher in the west half of the web. During the final cycle at 0.75% drift, residual

crack widths (outside the bulging region) typically ranged from 0.005-0.02 inches, with the

largest measuring 0.04 inches above the east boundary region splice. Figure A.43 shows the

21
It is important to note that the cover is spalling in this region but the concrete has not been damaged
enough to fall away from the wall and thus the damage is not necessarily apparent in the photographs
aside from the dashed lines indicating the spalling
22
The amount of longitudinal reinforcement exposed is so small it was difficult to tell it was exposed until
more was exposed at Step 1108, visible in picture DCS0165 on the roaming camera.
23
The region where the bulging was occuring was identified by tapping the concrete and listening for a
hollow sound for the bulging region and a solid sound for the non bulging region
537

Figure A.42: PW3: Spalling above splice on east side of wall at second WL- peak at 0.75%
drift (Step 990)

first floor following completion of the three cycles.


538

Figure A.43: PW3: Crack pattern following 3 cycles at ±0.75% drift (Step 1090)
539

1.0% Drift (Cycle A), Bar Buckling

In the first half-cycle at 1.0% drift, concrete fell off the wall on the east edge of the web

above the splice and the area of bulging in the web extended higher and lower in the web.

Concrete fell off the east corner above the splice and exposed longitudinal reinforcement at

the front corner of the wall and further exposed the bar at the back corner. At the cycle

ER+ peak (Step 1110), bar buckling was observed on the east side of the wall just above the

splice. Unloading from the peak, the crack above the west boundary region splice measured

0.05 inches at approximately zero lateral load, with other cracks were no larger than 0.02

inches and most were equal to or smaller than 0.005 inches. At −0.0743% (Step 1145) drift

on the way to the first WL- peak at 1.0% drift, transverse reinforcement pushed out from

the walls indicated lonitudinal bar buckling on the west side of the wall above the splice

(Figure A.44). By the time the peak was reached at 0.993% drift (Step 1150), the buckled

bar was clearly visible (Figure A.45). Concrete in the ‘bulging’ region in eastern most part

of the web had fallen off the wall exposing horizontal web reinforcement. More concrete fell

off this area when unloading from this peak, and at −0.298% drift (Step 1164), longitudinal

reinforcment was exposed in the web next to the east boundary region.
540

Figure A.44: PW3: Bar buckling on west side of wall above splice at −0.743% drift on way
to first WL- peak at 1.0% drift (Step 1145).

Figure A.45: PW3: Bar buckling on west side of wall above splice at first WL- peak at 1.0%
drift (Step 1150).
541

1.0% Drift (Cycle B), Bar Buckling, Core Damage

At the second ER+ peak (1.001% drift, Step 1190), the web reinforcement adjacent to the

boundary region was buckled, as shown in Figure A.46. Figure A.47 shows the east side

of the wall, where all three outer bars were buckled above the splice. On the way to the

peak, concrete fell off the wall over the entire length of web and longitudinal reinforcement

was exposed on the west half of the web. Figure A.48 shows the first two floors of the wall.

Figure A.49 shows the wall following the second (and final) WL- peak of the 1.0% drift

cycles (Step 1230). The extent of damage to the web grew, exposing more reinforcement

and revealing damage to the web concrete other than the cover concrete24 Initial damage

to the core concrete in the west boundary region also occurred.

Figure A.46: PW3: Bar buckling in east web at second ER+ peak at 1.0% drift (Step 1190)

24
Step 1230 is recorded as initial web crushing since it was the first time damage to concrete that was not
the cover concrete was visible.
542

Figure A.47: PW3: Bar buckling on east side at second ER+ peak at 1.0% drift (Step 1190)

Figure A.48: PW3: Damage to first two floors at second ER+ peak at 1.0% drift (Step
1190)
543

Figure A.49: PW3: Damage to first two floors at second WL- peak at 1.0% drift (Step
1230)
544

1.25% Drift, Failure

Following completion of the two cycles at ±1.0% drift, the loading protocol called for two

cycles at ±1.5% drift. While loading to the first ER+ peak, at 1.28% drift (Step 1267),

the axial load applied to the wall dropped to 190 kips. This drop was accompanied by

additional damage to the web concrete, spalling revealing transverse reinforcement the full

length of the east boundary element, and increased severity of bar buckling in the web and

the outer most boundary element bars. Leading up to this drop in axial load, the increase in

noticable damage was minimal from the damage in the previous cycles, but at the prior step

the lateral load carrying capacity dropped. On the front of the wall, the damage (spalling,

buckling and crushing) was concentrated at above the splice along the full length of the

wall, but Figure A.50 reveals the back of the wall was damaged above the splice in the web,

with the boundary region damaged in the top half of the splice and a few inches above the

splice.

Figure A.50: PW3: Back of wall at loss of axial load capacity (1.28% drift, Step 1267)
545

A.4 PW4

Specimen PW4 was constructed with a boundary element distribution of longitudinal re-

inforcement. The longitudinal reinforcement was continuous (no lap splice). The wall was

loaded to create base reactions representative of a uniform lateral load distribution on a 10-

story wall. Section 4.4 provides an overall summary of the Specimen PW4 test. A time-lapse

movie of damage to the bottom story of PW4 can be found at http://www.youtube.com/

user/NEESRWallProject#p/u/4/auU03q70EVg. The following sections provide a detailed

summary of the wall damage.


546

0.14% Drift, Initial Horizontal and Diagonal Cracks

Cracks were first visible in the first cycle to 0.139% drift. At a drift of 0.055% (Step

134) in the ER+ direction, a horizontal crack appeared in the upper quarter of the first

floor along the west edge of the wall. The next step, at 0.069% drift, this crack extended

further and became inclined halfway along the length of the boundary element. Additional

horizontal and inclined cracks appeared as well. At the ER+ peak (0.138%, Step 140),

two inclined cracks25 extended to the center of the wall. The lowest horizontal crack was

located approximately 2 inches from the top of the second-quarter of the first floor. When

the wall was displaced in the WL- direction, inclined cracks were first visible at −0.070%

drift (Step 155), and horizontal cracks one step later at −0.083% drift. As with displacement

in the ER+ direction, at the WL- peak (−0.138, Step 160) a shear crack extended from

the web/east boundary element interface in the lower quarter of the second floor to the

center of the web in the first floor (intersects crack for ER+ loading). Only two horizontal

cracks were visible in the first floor and were found in the upper third of the floor, while

the second floor had evenly spaced cracks nearly the entire floor height. Once three cycles

at the 0.139% drift level had been completed, horizontal cracks were visible in the second

quarter of the first floor, with greater density in the west boundary element than in the east

boundary element. The crack patterns are visible in Figure A.51.

25
One of the inclined cracks started in the boundary element as the previously mentioned horizontal crack
and then became inclined. The other started in the lower quarted on the second floor at the boundary
element-web interface.
547

Figure A.51: PW4: Crack pattern following 3 cycles at ±0.14% drift (Step 250)
548

0.21% Drift, Extension of Cracks

During the three cycles to 0.21% drift, additional cracking was observed (see Figure A.52).

Horizontal cracks appeared as low as one foot from the base of the wall in the west boundary

element. In the east boundary element, a horizontal crack formed approximately 4 inches

from the base of the wall at a region of patched concrete26 .

Figure A.52: PW4: Crack pattern following 3 cycles at ±0.21% drift (Step 370)

26
This was a ‘key’ crack to observe during the test. The next horizontal crack was approximately 15 inches
higher, and the region inbetween these two cracks remained uncracked for a large portion of the test, while
the equivalent region in the west boundary elment saw evenely distributed cracks
549

0.35% Drift

Damage progression was again extension of previous cracks and formation of new cracks

during the next three cycles, in which the wall was pushed to drifts of 0.351% in the ER+

direction and −0.341% in the WL- direction (see Figure A.53). During the ER+ cycles,

a horizontal crack at the foundation-wall interface was measured as extending four inches

into the boundary element from the west edge of the wall.

Figure A.53: PW4: Crack pattern following 3 cycles at ±0.35% drift (Step 610)
550

0.50% Drift, Vertical Cracks

During the first cycle to 0.50% drift, at the ER+ peak of 0.503% (Step 630), vertical cracks

and spalling of the cover27 concrete was first noticed at on the east side of the wall in the

lower 6 inches of the wall (see Figure A.54). At −0.394% (Step -0.394%) drift, a vertical

crack was first noticed on the west side of the wall at the base. Following the second cycle

ER+ peak (Step 710), the crack at the foundation-wall interface extended from the west

edge of the wall to the center of the wall, and spalling was observed at the base of the

west boundary element28 . Figure A.55 shows damage to the east toe of the wall at the

second WL- peak (Step 750). Following the final ER+ peak (Step 790), vertical cracks were

marked at the inner edge of the west boundary element in the third quarter of the first floor.

The damage following the 0.50% drift cycles is best summarized by the spalling in the west

boundary element shown by Figure A.56 and by the spalling at the patched region in the

east boundary element by Figure A.57.

27
Confining tie was visible, but longitudinal reinforcement was not; this damage is visible on photos with
the roaming camera (picture 92), and is only barely visible on the EastBottom camera (and may have
been visible at earlier steps but it is difficult to tell from photos)
28
Not noticable on the WestBottom camera because blocked by instrumentation
551

Figure A.54: PW4: Spalling at east toe of wall (Step 630)

Figure A.55: PW4: Spalling at east toe of wall (Step 750)


552

Figure A.56: PW4: Spalling at west side of wall following ±0.50% drift cycles (Step 830)

Figure A.57: PW4: Spalling of patched region of east boundary element following ±0.50%
drift cycles (Step 830)
553

0.75% Drift, Exposed Reinforcement, Bar Buckling, Core Damage

The next set of cycles was three cycles to ±0.75% drift. On the way to the first ER+

peak (0.754% @ Step 870), the following was observed: i) spalling on the front face of

the east boundary element extended higher (Figure A.58), ii) an increase in the length of

cracks iii) an increase in the density of horizontal cracks, iv) additional vertical cracks at the

interior edge of the west boundary element, including one approximately two inches long

above the wall-foundation interface (see Figure A.59), and v) longitudinal reinforcement

exposed on the back corner of the east boundary element (see Figure A.60). A substantial

increase in vertical cracks, shown in Figure A.61, formed along the west side toe of the wall

−0.521% drift (Step 904). At the first WL- peak at −0.744% drift (Step 910), removal of

loose, spalled cover revealed longitudinal reinforcement at the base of the front corner of

the east boundary element. At this step, a horizontal crack formed, other than that at the

patched region, in the lower 10 inches east boundary element29 . At the second ER+ peak

(Step 950), bar buckling30 was observed in the east boundary element (see Figure A.62).

At the corresponding WL- peak (Step 990), the density of the horizontal cracks in the

east boundary element had increased. At the final ER+ peak at 0.753% drift (Step 1030),

longitudinal reinforcement is exposed in the west boundary element (Figure A.63), but there

is no evidence of core crushing. At the final WL- peak at −0.744% drift (Step 1070), there

is damage to the core concrete in the east boundary element (see Figure A.64). Figure A.65

shows a vertical crack extending the lower 13 inches of the west side of the wall. Following

the three cycles at ±0.75% drift, only one longitudinal reinforcing bar was visible on the

front of the east boundary element.

29
Up until this step, there had been only the one crack in the east boundary element, contrasted to the
much denser horizontal cracks in the west boundary element
30
Middle outer bar
554

Figure A.58: PW4: Extension of spalling in east boundary element at first ER+ peak at
0.75% drift (Step 870).

Figure A.59: PW4: Vertical crack in west boundary element at 0.602% drift (Step 866).
555

Figure A.60: PW4: Exposed longitudinal reinforcement in east boundary element at first
ER+ peak at 0.75% drift (Step 870).

Figure A.61: PW4: Increase in vertical cracks on west toe of wall at −0.521% drift (Step
904).
556

Figure A.62: PW4: Bar buckling in east boundary element at second ER+ peak at 0.75%
drift (Step 950).

Figure A.63: PW4: Exposed longitudinal reinforcement in west boundary element at third
ER+ peak at 0.75% drift (Step 1030).
557

Figure A.64: PW4: Damage to east boundary element core concrete at third WL- peak at
0.75% drift (Step 1070).

Figure A.65: PW4: Damage to west boundary element at third WL- peak at 0.75% drift
(Step 1070).
558

1.0% Drift Cycle A

The next cycle was to ±1.0% drift. On the way to the ER+ peak (Step 1110), the east

boundary element core continued to deteriorate and multiple bars were buckled on the side

of the wall (see Figure A.66). Two longitudinal bars were visible on the front of the wall,

as seen in Figure A.67. The the extent of spalling of the cover concrete increased at the

base of the west boundary element, exposing the longitudinal reinforcement at the front of

the wall31 . At −0.895% (Step 1149) in the WL- direction, vertical cracks fromed along the

east boundary element/web interface. At the peak (Step 1150, −0.995% drift), a horizontal

crack formed in the web, next to the east boundary element, in the lower part of the third

quarter of the wall. An inclined crack extended into the west boundary element32 . Spalling

of the cover on the west boundary element base, shown in Figure A.68, extends five inches

up from the wall-foundation interface on the west side of the wall. Approximately three

inches from the wall-foundation interface, crushing of the cover concrete occurred along a

horizontal crack33

31
previously the longitudinal reinforcement was visible only on the side of the wall
32
The inclined crack ends approximately 10 inches from the base of the wall, and 2 inches into the boundary
element from the web.
33
This location is about the same as the patched region in the east boundary element.
559

Figure A.66: PW4: Bar buckling and core damage in east boundary element at first ER+
peak at 1.0% drift (Step 1110).

Figure A.67: PW4: Face of east boundary element at the first ER+ peak at 1.0% drift
(Step 1110).
560

Figure A.68: PW4: Spalling in west boundary at the first WL- peak at 1.0% drift (Step
1150).
561

1.0% Drift Cycle B, Loss of Load Capacity

While loading to the second ER+ peak at 1.0% drift, there was a drop in the axial load at

0.854% drift (Step 1187). Damage to the first floor is shown by the East Bottom and West

Bottom cameras in Figure A.69 and A.70. Six of the seven longitudinal bars at the front of

the east boundary element were visible34 (see Figure A.71). These bars were buckled, with

increasing severity from the inner side to outer side of the boundary element. Figure A.72

shows bar buckling on the east side of the wall. Damage to the core concrete extended

most of the length of the boundary element. An inclined crack that originated about six

inches below the eastern most rod applying the first floor shear force extended into the

boundary element, terminated at the spalled region a few inches from the foundation-wall

interface and approximately one inch into the boundary element. A vertical crack appeared

east of the center of the wall. While documenting the damage associated with the drop in

axial load at Step 187, loose concrete was removed from the backside (north face) of the

wall to protect the instruments mounted to the back of the wall. Removal of this concrete

revealed buckled longitudinal and horizontal reinforcement in the web35 (see Figure A.73).

In the west boundary element, cover spalling spread further along the length of the wall as

is shown in Figure A.74. At the cycle ER+ peak (1.001% drift, Step 1190), spalling occured

at the location of this vertical crack in the web (see Figure A.75). Figure A.76 shows the

extent of out-of-plane movement of the wall36 .

As the wall was loaded to the second WL- peak (Step 1230), spalling occurred at the

location of the previously ‘crushed’ cover concrete. The east boundary element core concrete

continued to deteriorate, and all longitudinal bars were visible. Once the wall was unloaded

34
Three bars were visible in the previous step
35
Spalled concrete extended three feet in from the east end of the wall. The first two longitudinal web
reinforcing bars had buckled. Although spalling extended further into the wall on the backside, the damage
to the boundary element core concrete did not appear as severe from the backside of the wall as it did
from the frontside.
36
This movement was documented at this step, but out-of-plane motion did occur previously, this is just
the extreme at the end of the test.
562

to zero drift (Step 1250), the test was officially concluded. At this point, the core of

the west boundary element remained undamaged. The east boundary element core was

almost completely deteriorated in the lower few inches and the reinforcing bars were severely

buckled. Web reinforcement was visible on the front of the wall and buckled on the back

side of the wall.

Figure A.69: PW4: East Bottom camera at loss of axial load capacity (Step 1187).
563

Figure A.70: PW4: West Bottom camera at loss of axial load capacity (Step 1187).

Figure A.71: PW4: East boundary element damage at loss of axial capacity (Step 1187).
564

Figure A.72: PW4: Loss of axial capacity: bar buckling at the east side of the wall (Step
1187).

Figure A.73: PW4: Loss of axial capacity: spalling and bar buckling on the back side of
web next to the east boundary element (Step 1187).
565

Figure A.74: PW4: Base of the west boundary element at the second ER+ peak at 1.0%
drift (Step 1190).

Figure A.75: PW4: East half of first floor at second ER+ peak at 1.0% drift (Step 1190).
566

Figure A.76: PW4: Out-of-plane displacement of wall as seen from the east side view of the
first floor at the second ER+ peak at 1.0% drift (Step 1190).
567

Appendix B

ADDITIONAL DATA ANALYSIS RESULTS

This appendix is provided as an “overflow” for data analysis figures.


568

−0 5% A
A

−0.34 A

.1 A
0. % A

50 A
.2 7%

0. % A

0. 7%A

A
0%

0%

5%

−0 0%

34%

%
0
.5

.0

.7

.5

10
0.25

75

00

50
−1

−1

−0

−0

0.

1.

1.
12

10
Height, feet

0
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
Displacement, in.

Figure B.1: Specimen PW1 displacement profiles at first peaks for each drift level.
AA
A

−0 % A
.5 A
−0 % A

0. % A
.247%

0.25 A
7 A

A
0%

.05%

5%

0%

%
34%

%
0

0
.5

−1.7

−0.3

.1

10

50

75

00

50
−1

−0

−0

0.

0.

0.

1.

1.
12

10
Height, feet

0
−0.015 −0.01 −0.005 0 0.005 0.01 0.015 0.02
Rotation, rad.

Figure B.2: Specimen PW1 rotation profiles at first peaks for each drift level.
Height, feet Height, feet

10
12

0
2
4
6
8
10
12

0
2
4
6
8

−2

−0.01
−1
−1.50 −1
−0.0 % .

−1.5
.70% A −110%
5% A .0
−0 A 0%
.5

−0.005
−0 A
−0 0%

−1
.3 A .7
−0 47% 5%
−0.20 A −0 A
.1 8% .5
0

0
−0 39% A −0 %

−0.5
0. .0
022 A .3 A
0. 8% 8% −0 47%
0.139 AA −0.2
20 % .0 A
8

0
−0 1389%
0. % A 0. .0 % A
34 A 0 28 A
7 0. 28% %

0.005
0. % 0.139 AA
50 A 20 %
%

Rotation, rad.
0.5
A 0. 8% A

Displacement, in.
34 A
0. 0. 7%
75 50 A
% %
A
1

0.01
0.
75
%
A
1. 1.
1.5

00 00
% %
A A

0.015
2

1.
50 1.
% 50
A %
A
2.5

0.02

Figure B.4: Specimen PW2 rotation profiles at first peaks for each drift level.
Figure B.3: Specimen PW2 displacement profiles at first peaks for each drift level.
569
Height, feet Height, feet

10
12

0
2
4
6
8
10
12

0
2
4
6
8

−1.5

−0.015
−1
.0
0%
A
−0

−1
.7
−1 5%
.0 A

−0.01
0%
A −0
.5
−0 0
.7 −0 % A

−0.5
5% .3
A 4
−0 7%
−0 −0.20 A
.5 .1 8
0

−0.005
−0 39%

0
−0 %
.3 A 0. .0 % A
4 0228 A
−0 7% 0. 8%%
.
−0 0 A2 0.139 AA
.1 8% 20 %
8

0
−0 39 A 0. % A

0.5
0. .0 % 34 A
022 A
0. 8% 8% 7

Rotation, rad.
0. % A
0.139 AA 50
Displacement, in.
20 % %
8 A
1
0. % A
34 A 0.
7 75
%

0.005
0. % A
50
% A
A 1.
00
1.5

0. %
75
% A
A

0.01
1.
00
2

%
A
1. 1.
50 50
% %
A A
2.5

0.015

Figure B.6: Specimen PW3 rotation profiles at first peaks for each drift level.
Figure B.5: Specimen PW3 displacement profiles at first peaks for each drift level.
570
571

−0 % A

−0 9% A
0. 02 A

0. % A
A

13 A

8% A
A

A
.1 8%

8 %

A
0%

5%

0%

0. 9%

7%
47

02 8

%
−0.20
3
.0

.7

.5

.3

20

34

50

75

00
.
−1

−0

−0

−0

0.

0.

0.

1.
12

10
Height, feet

0
−1.5 −1 −0.5 0 0.5 1 1.5
Displacement, in.

Figure B.7: Specimen PW4 displacement profiles at first peaks for each drift level.
−0.20 A
−0 39% A
022 A

0.139 AA
A

0. % A
34 A

A
−0 47%

0. .0 %
8%

A
0%

5%

0%

0. 8%
20 %

7%
.1 8

%
8
.0

.7

.5

.3

50

75

00
−1

−0

−0

−0

0.

0.

1.

12

10
Height, feet

0
−0.015 −0.01 −0.005 0 0.005 0.01 0.015
Rotation, rad.

Figure B.8: Specimen PW4 rotation profiles at first peaks for each drift level.
572

1.0 1.0
1st Flr (Fixed) 1st Flr (Fixed)
1st Flr 1st Flr
2nd Flr 2nd Flr
0.8 0.8

0.6 0.6
EIeff /EIg

EIeff /EIg
0.4 0.4

0.2 0.2

0 0
0 0.5 1.0 1.5 0 0.5 1.0 1.5
Drift, % Drift, %
(a) PW1 (b) PW2

1.0 1.0
1st Flr (Fixed) 1st Flr (Fixed)
1st Flr 1st Flr
2nd Flr 2nd Flr
0.8 0.8

0.6 0.6
EIeff /EIg

EIeff /EIg

0.4 0.4

0.2 0.2

0 0
0 0.5 1.0 1.5 0 0.5 1.0 1.5
Drift, % Drift, %
(c) PW3 (d) PW4

Figure B.9: Comparison of effective flexural stiffness values for the first floor (base fixed and
allowed to rotate) and the second floor. Values calculated using calculated using Timoshenko
beam theory.
573

1.0 1.0
1st Flr (Fixed) 1st Flr (Fixed)
1st Flr 1st Flr
2nd Flr 2nd Flr
0.8 0.8
AGeff /AGg

AGeff /AGg
0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 0.5 1.0 1.5 0 0.5 1.0 1.5
Drift, % Drift, %
(a) PW1 (b) PW2

1.0 1.0
1st Flr (Fixed) 1st Flr (Fixed)
1st Flr 1st Flr
2nd Flr 2nd Flr
0.8 0.8
AGeff /AGg

AGeff /AGg

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 0.5 1.0 1.5 0 0.5 1.0 1.5
Drift, % Drift, %
(c) PW3 (d) PW4

Figure B.10: Comparison of effective shears stiffness values for the first floor (base fixed and
allowed to rotate) and the second floor. Values calculated using Timoshenko beam theory.
574

120 120

100 100
% Total Deformation

% Total Deformation
80 80

60 60

40 40

20 20

0 0
%

%
0.101%
25 %
0. 1%
7%

0.101%
25 %
0. 1%
7%
%

%
50

50
0.050

0.050
50

75

00

50

75

00
34

34
1.

1.
0.

0.

1.

0.

0.

1.
0.

(a) 1st Floor ER+ Peaks 0. (b) 1st Floor WL- Peaks

3rd story flexure 3rd story shear


2nd story flexure 2nd story shear
1st story flexure 1st story shear
Base slip Base rotation
(c)

120 120

100 100
% Total Deformation

% Total Deformation

80 80

60 60

40 40

20 20

0 0
%

%
0.101%
25 %
0. 1%
7%

0.101%
25 %
0. 1%
7%
%

%
50

50
0.050

0.050
50

75

00

50

75

00
34

34
1.

1.
0.

0.

1.

0.

0.

1.
0.

0.

(d) 3rd Floor ER+ Peaks (e) 3rd Floor WL- Peaks

Figure B.11: Specimen PW1 deformation contributions.


575

120 120

100 100
% Total Deformation

% Total Deformation
80 80

60 60

40 40

20 20

0 0
%

%
0.028%
13 %
0. %
0. %
7%

0.028%
13 %
0. %
0. %
7%
0%

5%

0%

5%

%
50

50
0.014

9
8

0.014

9
8
00

00
20
34

20
34
5

7
1.

1.
0.

0.

1.

0.

0.

1.
0.

0.

(a) 1st Floor ER+ Peaks (b) 1st Floor WL- Peaks

3rd story flexure 3rd story shear


2nd story flexure 2nd story shear
1st story flexure 1st story shear
Base slip Base rotation
(c)

120 120

100 100
% Total Deformation

% Total Deformation

80 80

60 60

40 40

20 20

0 0
%

%
0.028%
13 %
0. %
0. %
7%

0.028%
13 %
0. %
0. %
7%
%

%
50

50
0.014

9
8

0.014

9
8
50

75

00

50

75

00
20
34

20
34
1.

1.
0.

0.

1.

0.

0.

1.
0.

0.

(d) 3rd Floor ER+ Peaks (e) 3rd Floor WL- Peaks

Figure B.12: Specimen PW2 deformation contributions.


576

120 120

100 100
% Total Deformation

% Total Deformation
80 80

60 60

40 40

20 20

0 0
%

%
0.028%
13 %
0. %
0. %
7%

0.028%
13 %
0. %
0. %
7%
0%

5%

0%

5%

%
50

50
0.014

9
8

0.014

9
8
00

00
20
34

20
34
5

7
1.

1.
0.

0.

1.

0.

0.

1.
0.

0.
(a) 1st Floor ER+ Peaks (b) 1st Floor WL- Peaks

3rd story flexure 3rd story shear


2nd story flexure 2nd story shear
1st story flexure 1st story shear
Base slip Base rotation
(c)

120 120

100 100
% Total Deformation

% Total Deformation

80 80

60 60

40 40

20 20

0 0
%

%
0.028%
13 %
0. %
0. %
7%

0.028%
13 %
0. %
0. %
7%
%

%
50

50
0.014

9
8

0.014

9
8
50

75

00

50

75

00
20
34

20
34
1.

1.
0.

0.

1.

0.

0.

1.
0.

0.

(d) 3rd Floor ER+ Peaks (e) 3rd Floor WL- Peaks

Figure B.13: Specimen PW3 deformation contributions.


577

120 120

100 100
% Total Deformation

% Total Deformation
80 80

60 60

40 40

20 20

0 0
%

%
0.028%
13 %
0. %
0. %
7%

0.028%
13 %
0. %
0. %
7%
0%

5%

0%

5%

%
50

50
0.014

9
8

0.014

9
8
00

00
20
34

20
34
5

7
1.

1.
0.

0.

1.

0.

0.

1.
0.

0.
(a) 1st Floor ER+ Peaks (b) 1st Floor WL- Peaks

3rd story flexure 3rd story shear


2nd story flexure 2nd story shear
1st story flexure 1st story shear
Base slip Base rotation
(c)

120 120

100 100
% Total Deformation

% Total Deformation

80 80

60 60

40 40

20 20

0 0
%

%
0.028%
13 %
0. %
0. %
7%

0.028%
13 %
0. %
0. %
7%
%

%
50

50
0.014

9
8

0.014

9
8
50

75

00

50

75

00
20
34

20
34
1.

1.
0.

0.

1.

0.

0.

1.
0.

0.

(d) 3rd Floor ER+ Peaks (e) 3rd Floor WL- Peaks

Figure B.14: Specimen PW4 deformation contributions.


578

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.15: Vertical strain fields at first and last peaks at 0.50% drift.
579

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.16: Vertical strain fields at first and last peaks at 0.75% drift.
580

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (B) (h) PW2 WL- (B)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (B) (l) PW3 WL- (B)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (B) (p) PW4 WL- (B)

Figure B.17: Vertical strain fields at first and last peaks at 1.00% drift.
581

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) (h)

(i) PW3 ER+ (A) (j) (k) (l)

(m) (n) (o) (p)

Figure B.18: Vertical strain fields at first and last peaks at 1.50% drift.
582

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.19: Shear strain fields at first and last peaks at 0.50% drift.
583

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.20: Shear strain fields at first and last peaks at 0.75% drift.
584

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (B) (h) PW2 WL- (B)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (B) (l) PW3 WL- (B)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (B) (p) PW4 WL- (B)

Figure B.21: Shear strain fields at first and last peaks at 1.00% drift.
585

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) (h)

(i) PW3 ER+ (A) (j) (k) (l)

(m) (n) (o) (p)

Figure B.22: Shear strain fields at first and last peaks at 1.50% drift.
586

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.23: 2nd principal strain fields at first and last peaks at 0.50% drift.
587

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.24: 2nd principal strain fields at first and last peaks at 0.75% drift.
588

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (B) (h) PW2 WL- (B)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (B) (l) PW3 WL- (B)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (B) (p) PW4 WL- (B)

Figure B.25: 2nd principal strain fields at first and last peaks at 1.00% drift.
589

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) (h)

(i) PW3 ER+ (A) (j) (k) (l)

(m) (n) (o) (p)

Figure B.26: 2nd principal strain fields at first and last peaks at 1.50% drift.
590

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.27: 1st principal strain fields at first and last peaks at 0.50% drift.
591

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (C) (d) PW1 WL- (C)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (C) (h) PW2 WL- (C)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (C) (l) PW3 WL- (C)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (C) (p) PW4 WL- (C)

Figure B.28: 1st principal strain fields at first and last peaks at 0.75% drift.
592

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) PW2 ER+ (B) (h) PW2 WL- (B)

(i) PW3 ER+ (A) (j) PW3 WL- (A) (k) PW3 ER+ (B) (l) PW3 WL- (B)

(m) PW4 ER+ (A) (n) PW4 WL- (A) (o) PW4 ER+ (B) (p) PW4 WL- (B)

Figure B.29: 1st principal strain fields at first and last peaks at 1.00% drift.
593

(a) PW1 ER+ (A) (b) PW1 WL- (A) (c) PW1 ER+ (B) (d) PW1 WL- (B)

(e) PW2 ER+ (A) (f) PW2 WL- (A) (g) (h)

(i) PW3 ER+ (A) (j) (k) (l)

(m) (n) (o) (p)

Figure B.30: 1st principal strain fields at first and last peaks at 1.50% drift.
594

Appendix C

EXPERIMENTAL DATA SUPPORTING FRAGILITY FUNCTION


DEVELOPMENT

Tables C.1 and C.2 contain a summary of the walls used in this study. Tables C.3 and C.4

provide a summary of the reinforcing steel properties, including i) the location in the wall

where there steel is used (B = boundary element longitudinal steel, W = web longitudinal

steel, H = horizontal steel, and C = confining steel) ii) the material model used for OpenSees

fiber models, ReinforcingSteel (reinfStl) or Steel01 (Steel01), and iii) estimated material

properties (see Section 7.1.1.5), which are denoted by a ’*’.

Section C.3 provides summaries of the individual wall tests. For each specimen, the

following is provided:

• A table listing the recorded damage states. The damage states are sorted in alpha-

numeric order. When developing the fragility functions, only one damage state for

each method of repair was used and was selected to be the damage state with the

lowest magnitude of drift, which is not necessarily the first damage state listed for

each damage state. The damage states, which were listed in the main body of the

report in Table 7.2 is reiterated here in Table C.5 and contains damage states that

were entered into the database but which were not used in developing the fragility

functions. These damage states begin with the the letters ‘DSr’, are primarily related

to crack width measurements, and are included in the appendix for completeness.

• A table listing recorded crack width measurements. Any crack width measurements

entered into the slender wall database were included here. The tables indicate the i)

associated damage state, ii) the drift and force at the measurement, iii) if the mea-

surement corresponds to the exact drift or the maximum cycle drift for the occurrence
595

of the damage, iv) if the crack measurement is for an inclined crack (shear), a hor-

izontal crack (flexure), or an unspecified crack (any), and v) the maximum and/or

residual crack widths. Crack widths that were reported but not associated with a

particular damage state are indicated by the letter ’C’ and a number corresponding

to the number of the reading.

• A figure showing the positive and negative load-drift envelopes (and the full hysteresis

if available), with markers indicating the occurrences of the damage states. The

markers shown in the legends correspond to the cycle peak where the damage occurred.

Smaller markers of the same shape and color indicate the exact point on the load

displacement envelope where the damage occurred, if available.

• Any figures (photos, sketches, crack maps) collected and added to the slender wall

database to provide a visual reference for the damage data. Figure captions contain

the name of the specimen, the damage state the photo is associated with, and a

description of the figure.


596

C.1 Database Properties

Table C.1: Geometric properties for slender walls included in fragility function database

Specimen Database
Source Name Name Shape Loading AR M/(V `w ) Scale
Pilakoutas SW4 PilaSW4 Rect. cyclic 2.00 2.00 0.20
and Elnashai SW5 PilaSW5 Rect. cyclic 2.00 2.00 0.20
SW6 PilaSW6 Rect. cyclic 2.00 2.00 0.20
SW7 PilaSW7 Rect. cyclic 2.00 2.00 0.20
SW8 PilaSW8 Rect. cyclic 2.00 2.00 0.20
SW9 PilaSW9 Rect. cyclic 2.00 2.00 0.20
Tasnimi SHW1 TasSHW1 Rect. cyclic 2.20 2.20 0.16
SHW2 TasSHW2 Rect. cyclic 2.20 2.20 0.16
SHW3 TasSHW3 Rect. cyclic 2.20 2.20 0.16
SHW4 TasSHW4 Rect. cyclic 2.20 2.20 0.16
Thomsen and RW1 ThomRW1 Rect. cyclic 3.00 3.00 0.33
Wallace RW2 ThomRW2 Rect. cyclic 3.00 3.00 0.33
TW1 ThomTW1 T-wall cyclic 3.00 3.00 0.33
TW2 ThomTW2 T-wall cyclic 3.00 3.00 0.33
Dazio et al. WSH1 DazioWSH1 Rect. cyclic 2.28 2.28 0.49
WSH2 DazioWSH2 Rect. cyclic 2.28 2.28 0.49
WSH3 DazioWSH3 Rect. cyclic 2.28 2.28 0.49
WSH4 DazioWSH4 Rect. cyclic 2.28 2.28 0.49
WSH5 DazioWSH5 Rect. cyclic 2.28 2.28 0.49
WSH6 DazioWSH6 Rect. cyclic 2.26 2.26 0.49
Lefas et al. SW21 LefasSW21 Rect. monotonic 2.00 2.00 0.21
SW22 LefasSW22 Rect. monotonic 2.00 2.00 0.21
SW23 LefasSW23 Rect. monotonic 2.00 2.00 0.21
SW24 LefasSW24 Rect. monotonic 2.00 2.00 0.21
SW25 LefasSW25 Rect. monotonic 2.00 2.00 0.21
SW26 LefasSW26 Rect. monotonic 2.00 2.00 0.21
SW30 LefasSW30 Rect. monotonic 2.00 2.00 0.21
SW31 LefasSW31 Rect. monotonic 2.00 2.00 0.21
SW32 LefasSW32 Rect. monotonic 2.00 2.00 0.21
SW33 LefasSW33 Rect. monotonic 2.00 2.00 0.21
597

Table C.1: Geometric properties for slender walls included in fragility function database
(con’t)

Specimen Database
Source Name Name Shape Loading AR M/(V `w ) Scale

Oesterle et al. R1 OestR1 Rect. cyclic 2.40 2.40 0.33


R2 OestR2 Rect. cyclic 2.40 2.40 0.33
B1 OestB1 Barbell cyclic 2.40 2.40 0.33
B2 OestB2 Barbell cyclic 2.40 2.40 0.33
B3 OestB3 Barbell cyclic 2.40 2.40 0.33
B4 OestB4 Barbell monotonic 2.40 2.40 0.33
B5 OestB5 Barbell cyclic 2.40 2.40 0.33
B6 OestB6 Barbell cyclic 2.40 2.40 0.33
B7 OestB7 Barbell cyclic 2.40 2.40 0.33
B8 OestB8 Barbell cyclic 2.40 2.40 0.33
B9 OestB9 Barbell cyclic 2.40 2.40 0.33
B10 OestB10 Barbell cyclic 2.40 2.40 0.33
F1 OestF1 H-wall cyclic 2.40 2.40 0.33
F2 OestF2 H-wall cyclic 2.40 2.40 0.33
Morgan et al. W1 Morgan Barbell cyclic 2.86 2.79 0.19
Liu W1 LiuW1 Rect. cyclic 3.08 3.13 0.66
W2 LiuW2 Rect. cyclic 3.08 3.13 0.66
Ali, Wight W1 AliW1 Barbell cyclic 2.92 3.04 0.25
Tupper W3 TupperW3 Rect. cyclic 3.90 3.75 0.50
Mobeen W1 MobeenW1 Barbell cyclic 2.98 2.74 0.66
Riva et al. W1 RivaW1 Rect. cyclic 4.11 3.17 0.98
Shiu et al. C1 ShiuC1 Rect. cyclic 2.88 2.88 0.33
Khalil C1 KhaC1 Rect. cyclic 1.10 2.25 0.39
Elnady and CW2 ElnCW2 Rect. cyclic 1.10 5.00 0.39
Ghobarah CW3 ElnCW3 Rect. cyclic 1.10 2.25 0.39
Ile et al. X IleX C-wall cyclic 2.40 2.40 0.82
Y IleY C-wall cyclic 2.40 2.40 0.82
XY IleXY C-wall cyclic 2.40 2.40 0.82
Beyer et al. TUA TUA C-wall cyclic 2.58 2.58 0.49
TUB TUB C-wall cyclic 2.58 2.58 0.33
Brueggen NTW1 NTW1 T-wall cyclic 3.20 3.47 0.50
NTW2 NTW2 T-wall cyclic 1.60 3.47 0.50
Lowes et al. PW1 PW1 Rect. cyclic 1.20 2.85 0.50
PW2 PW2 Rect. cyclic 1.20 2.15 0.50
PW3 PW3 Rect. cyclic 1.20 2.00 0.50
PW4 PW4 Rect. cyclic 1.20 1.99 0.50
598

Table C.2: Loading and reinforcement properties for slender walls included in fragility
function database

Database fc0 ρbe ρweb ρh ρcon


Name λn vn vu Vu /Vn ksi (MPa) % % % %
PilaSW4 0.00 6.2 5.7 0.92 5.4 (37) 6.85 0.75 0.39 1.54
PilaSW5 0.00 5.0 7.0 1.40 4.6 (32) 11.41 0.79 0.35 0.88
PilaSW6 0.00 4.7 5.8 1.24 5.6 (39) 6.85 0.75 0.35 0.71
PilaSW7 0.00 6.6 7.5 1.15 4.6 (32) 11.08 0.79 0.39 1.85
PilaSW8 0.00 5.0 4.7 0.94 6.6 (46) 7.14 0.75 0.42 0.85
PilaSW9 0.00 6.6 5.3 0.79 5.6 (39) 7.14 0.75 0.60 1.22
TasSHW1 0.00 3.6 1.6 0.45 3.1 (22) 3.23 0.28 0.28 0.00
TasSHW2 0.00 3.6 2.1 0.57 3.1 (22) 3.23 0.28 0.28 0.00
TasSHW3 0.00 3.6 1.7 0.47 3.3 (22) 3.23 0.28 0.28 0.00
TasSHW4 0.00 3.5 2.3 0.66 3.4 (23) 3.23 0.28 0.28 0.00
ThomRW1 0.10 5.1 2.6 0.50 4.6 (32) 2.93 0.33 0.33 1.06
ThomRW2 0.07 5.0 2.6 0.53 4.9 (34) 2.93 0.33 0.33 1.21
ThomTW1 0.07 4.7 4.3 0.92 6.3 (44) 3.27 0.33 0.33 1.06
ThomTW2 0.07 4.7 5.5 1.16 6.0 (42) 2.51 0.45 0.33 0.89
DazioWSH1 0.05 4.6 2.0 0.44 6.5 (45) 1.57 0.30 0.25 1.08
DazioWSH2 0.06 4.3 2.3 0.53 5.9 (41) 1.57 0.30 0.25 1.12
DazioWSH3 0.06 4.4 2.9 0.67 5.7 (39) 1.74 0.54 0.25 1.04
DazioWSH4 0.06 4.5 2.8 0.63 5.9 (41) 1.74 0.54 0.25 0.00
DazioWSH5 0.13 4.5 2.9 0.63 5.6 (38) 0.77 0.27 0.25 1.00
DazioWSH6 0.11 4.3 3.5 0.81 6.6 (46) 1.74 0.54 0.25 1.68
LefasSW21 0.00 10.8 6.2 0.57 5.0 (34) 2.99 2.49 0.82 0.56
LefasSW22 0.12 10.7 7.2 0.68 5.1 (35) 2.99 2.49 0.82 0.56
LefasSW23 0.21 10.3 8.3 0.80 5.6 (39) 2.99 2.49 0.82 0.56
LefasSW24 0.00 10.7 5.8 0.54 5.1 (35) 2.99 2.49 0.82 0.56
LefasSW25 0.21 10.5 7.1 0.67 5.3 (37) 2.99 2.49 0.82 0.56
LefasSW26 0.00 7.2 7.1 0.99 3.5 (24) 2.99 2.49 0.41 0.56
LefasSW30 0.00 6.6 6.8 1.03 3.5 (24) 2.99 1.55 0.36 0.49
LefasSW31 0.00 6.3 6.2 0.99 4.1 (28) 2.99 1.55 0.36 0.49
LefasSW32 0.00 5.5 4.9 0.89 6.0 (42) 2.99 1.55 0.36 0.49
LefasSW33 0.00 5.7 5.2 0.91 5.4 (37) 2.99 1.55 0.36 0.49
OestR1 0.00 4.9 1.1 0.22 6.5 (45) 1.47 0.28 0.31 0.00
OestR2 0.00 5.0 2.0 0.40 6.7 (46) 4.00 0.28 0.31 3.99
OestB1 0.00 4.7 2.3 0.49 7.7 (53) 1.11 0.28 0.31 0.48
OestB2 0.00 7.5 5.8 0.77 7.8 (54) 3.67 0.28 0.63 0.42
OestB3 0.00 7.2 2.5 0.34 6.9 (47) 1.11 0.28 0.63 4.75
OestB4 0.00 7.7 3.1 0.41 6.5 (45) 1.11 0.28 0.63 4.75
OestB5 0.00 7.6 7.0 0.93 6.6 (45) 3.67 0.28 0.63 5.91
OestB6 0.13 10.2 11.0 1.07 3.2 (22) 3.67 0.28 0.63 3.57
OestB7 0.08 7.2 8.7 1.20 7.2 (49) 3.67 0.28 0.63 5.91
OestB8 0.09 14.3 9.4 0.66 6.1 (42) 3.67 0.28 1.38 5.91
OestB9 0.09 7.2 9.2 1.27 6.4 (44) 3.67 0.28 0.63 5.91
OestB10 0.08 7.3 9.0 1.24 6.6 (46) 1.97 0.28 0.63 5.93
OestF1 0.00 9.3 8.4 0.90 5.6 (38) 3.89 0.28 0.71 1.11
OestF2 0.07 6.5 8.2 1.25 6.6 (46) 5.01 0.24 0.55 3.15
599

Table C.2: Loading and reinforcement properties for slender walls included in fragility
function database (con’t)

Database fc0 ρbe ρweb ρh ρcon


Name λn vn vu Vu /Vn ksi (MPa) % % % %

Morgan 0.05 6.8 2.5 0.37 4.6 (32) 1.11 0.42 0.44 1.20
LiuW1 0.08 5.9 2.3 0.39 4.8 (33) 3.00 0.33 0.40 2.32
LiuW2 0.04 5.1 1.7 0.33 10.3 (71) 3.00 0.33 0.47 2.32
AliW1 0.08 5.3 3.7 0.69 4.7 (32) 3.20 0.28 0.28 2.67
TupperW3 0.10 8.9 4.3 0.48 5.6 (39) 5.36 0.44 0.73 0.96
MobeenW1 0.15 11.4 3.8 0.34 3.5 (24) 1.60 0.67 1.00 2.15
RivaW1 0.00 4.0 2.0 0.51 4.7 (33) 2.30 0.17 0.17 0.96
ShiuC1 0.00 6.3 4.3 0.69 3.4 (23) 5.58 0.24 0.37 2.78
KhaC1 0.07 5.2 5.9 1.14 5.5 (38) 10.00 0.83 0.27 0.00
ElnCW2 0.08 4.9 1.6 0.33 5.4 (37) 6.41 1.67 0.26 0.00
ElnCW3 0.08 4.9 3.5 0.71 5.4 (38) 6.41 1.67 0.26 0.00
IleX 0.10 9.4 6.0 0.64 3.4 (24) 0.98 0.25 0.54 1.09
IleY 0.10 9.4 3.0 0.32 3.4 (24) 0.98 0.25 0.54 1.09
IleXY 0.11 9.9 6.0 0.60 3.0 (21) 0.98 0.25 0.54 1.09
TUA 0.02 4.1 3.2 0.79 11.3 (78) 1.29 0.25 0.30 1.64
TUB 0.04 5.8 5.7 0.99 7.9 (55) 2.17 0.38 0.45 2.44
NTW1 0.03 4.5 4.9 1.09 7.3 (50) 3.78 0.28 0.29 1.13
NTW2 0.03 7.5 5.4 0.72 6.6 (45) 2.64 0.86 0.61 1.12
PW1 0.10 4.9 3.6 0.73 5.2 (36) 3.33 0.29 0.28 2.10
PW2 0.13 4.8 5.0 1.05 5.8 (40) 3.50 0.29 0.28 1.86
PW3 0.10 5.0 4.5 0.89 5.0 (34) 2.01 1.59 0.28 1.86
PW4 0.12 5.2 4.6 0.89 4.3 (29) 3.50 0.29 0.28 1.86
600

Table C.3: Rebar Properties (* denotes assumed value) - Part A

Area db
Name Bar Size Location in2 (mm2 ) in (mm) Material Model
PilaSW4 d12 B 0.18 (113) 0.47 (12.0) Steel01
d6 WHC 0.04 (28) 0.24 (6.0) Steel01
PilaSW5 d16 B 0.31 (201) 0.63 (16.0) Steel01
d6 BW 0.04 (28) 0.24 (6.0) Steel01
d4 HC 0.02 (13) 0.16 (4.0) Steel01
PilaSW6 d12 B 0.18 (113) 0.47 (12.0) Steel01
d6 W 0.04 (28) 0.24 (6.0) Steel01
d4 HC 0.02 (13) 0.16 (4.0) Steel01
PilaSW7 d16 B 0.31 (201) 0.63 (16.0) Steel01
d6 BWHC 0.04 (28) 0.24 (6.0) Steel01
PilaSW8 d10 B 0.12 (79) 0.39 (10.0) Steel01
d6 W 0.04 (28) 0.24 (6.0) Steel01
d4 HC 0.02 (13) 0.16 (4.0) Steel01
PilaSW9 d10 B 0.12 (79) 0.39 (10.0) Steel01
d6 W 0.04 (28) 0.24 (6.0) Steel01
d4 HC 0.02 (13) 0.16 (4.0) Steel01
TasSHW1 d6 B 0.04 (28) 0.24 (6.0) Steel01
d3 WH 0.01 (7) 0.12 (3.0) Steel01
TasSHW2 d6 B 0.04 (28) 0.24 (6.0) Steel01
d3 WH 0.01 (7) 0.12 (3.0) Steel01
TasSHW3 d6 B 0.04 (28) 0.24 (6.0) Steel01
d3 WH 0.01 (7) 0.12 (3.0) Steel01
TasSHW4 d6 B 0.04 (28) 0.24 (6.0) Steel01
d3 WH 0.01 (7) 0.12 (3.0) Steel01
ThomRW1 no3 B 0.11 (71) 0.38 (9.5) reinfStl
no2 WH 0.05 (32) 0.25 (6.3) Steel01
ThomRW2 no3 B 0.11 (71) 0.38 (9.5) reinfStl
no2 WH 0.05 (32) 0.25 (6.3) Steel01
ThomTW1 no3 B 0.11 (71) 0.38 (9.5) reinfStl
no2 BWH 0.05 (32) 0.25 (6.3) Steel01
ThomTW2 no3 B 0.11 (71) 0.38 (9.5) reinfStl
no2 BW 0.05 (32) 0.25 (6.3) Steel01
DazioWSH1 d10 B 0.12 (79) 0.39 (10.0) Steel01
d6 WHC 0.04 (28) 0.24 (6.0) Steel01
DazioWSH2 d10 B 0.12 (79) 0.39 (10.0) Steel01
d6 WHC 0.04 (28) 0.24 (6.0) Steel01
DazioWSH3 d12 B 0.18 (113) 0.47 (12.0) Steel01
d8 W 0.08 (50) 0.31 (8.0) Steel01
d6 HC 0.04 (28) 0.24 (6.0) Steel01
DazioWSH4 d12 B 0.18 (113) 0.47 (12.0) Steel01
d8 W 0.08 (50) 0.31 (8.0) Steel01
d6 H 0.04 (28) 0.24 (6.0) Steel01
DazioWSH5 d8 B 0.08 (50) 0.31 (8.0) Steel01
d6 WH 0.04 (28) 0.24 (6.0) Steel01
601

Table C.3: Rebar Properties (* denotes assumed value) - Part A (con’t)

Area db
Name Bar Size Location in2 (mm2 ) in (mm) Material Model

DazioWSH6 d12 B 0.18 (113) 0.47 (12.0) Steel01


d8 W 0.08 (50) 0.31 (8.0) Steel01
d6 HC 0.04 (28) 0.24 (6.0) Steel01
LefasSW21 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW22 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW23 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW24 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW25 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW26 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW30 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW31 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW32 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
LefasSW33 d8 BW 0.08 (50) 0.31 (8.0) Steel01
d6.25 H 0.05 (31) 0.25 (6.3) Steel01
OestR1 no3 B 0.11 (71) 0.38 (9.5) Steel01
d6 WH 0.05 (32) 0.25 (6.3) Steel01
OestR2 no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestB1 no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 WH 0.05 (32) 0.25 (6.3) Steel01
OestB2 no6 B 0.44 (284) 0.75 (19.1) Steel01
d6 WH 0.05 (32) 0.25 (6.3) Steel01
OestB3 no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestB4 no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestB5 no6 B 0.44 (284) 0.75 (19.1) Steel01
d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestB6 no6 B 0.44 (284) 0.75 (19.1) Steel01
d6 WH 0.05 (32) 0.25 (6.3) Steel01
OestB7 no6 B 0.44 (284) 0.75 (19.1) Steel01
d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestB8 no6 B 0.44 (284) 0.75 (19.1) Steel01
d6 WC 0.05 (32) 0.25 (6.3) Steel01
no3 H 0.11 (71) 0.38 (9.5) Steel01
602

Table C.3: Rebar Properties (* denotes assumed value) - Part A (con’t)

Area db
Name Bar Size Location in2 (mm2 ) in (mm) Material Model

OestB9 no6 B 0.44 (284) 0.75 (19.1) Steel01


d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestB10 no5 B 0.31 (200) 0.63 (15.9) Steel01
no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestF1 no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 WHC 0.05 (32) 0.25 (6.3) Steel01
OestF2 no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 BWHC 0.04 (28) 0.24 (6.0) Steel01
Morgan d6 B 0.04 (28) 0.24 (6.0) Steel01
D3 WH 0.03 (19) 0.20 (5.0) Steel01
LiuW1 20M B 0.47 (300) 0.77 (19.5) reinfStl
10M WHC 0.16 (100) 0.44 (11.3) reinfStl
LiuW2 20M B 0.47 (300) 0.77 (19.5) reinfStl
10M WHC 0.16 (100) 0.44 (11.3) reinfStl
AliW1 no4 B 0.20 (129) 0.50 (12.7) reinfStl
no2 WH 0.05 (32) 0.25 (6.3) Steel01
TupperW3 20M B 0.47 (300) 0.77 (19.5) reinfStl
10M WH 0.16 (100) 0.44 (11.3) reinfStl
MobeenW1 15M B 0.31 (200) 0.63 (16.0) reinfStl
10M WHC 0.16 (100) 0.44 (11.3) reinfStl
RivaW1 d20 B 0.49 (314) 0.79 (20.0) Steel01
d8 BWHC 0.08 (50) 0.31 (8.0) Steel01
ShiuC1 no4 B 0.20 (129) 0.50 (12.7) Steel01
d6 WHC 0.04 (28) 0.24 (6.0) Steel01
KhaC1 20M B 0.47 (300) 0.77 (19.5) Steel01
10M W 0.16 (100) 0.44 (11.3) Steel01
6M H 0.05 (29) 0.24 (6.1) Steel01
ElnCW2 15M BW 0.31 (200) 0.63 (16.0) Steel01
15M Web 0.31 (200) 0.63 (16.0) Steel01
d6 H 0.04 (28) 0.24 (6.0) Steel01
ElnCW3 15M BW 0.31 (200) 0.63 (16.0) Steel01
d6 H 0.04 (28) 0.24 (6.0) Steel01
IleX d12 B 0.18 (113) 0.47 (12.0) Steel01
d10 W 0.12 (79) 0.39 (10.0) Steel01
d8 HC 0.08 (50) 0.31 (8.0) Steel01
IleY d12 B 0.18 (113) 0.47 (12.0) Steel01
d10 W 0.12 (79) 0.39 (10.0) Steel01
d8 HC 0.08 (50) 0.31 (8.0) Steel01
IleXY d12 B 0.18 (113) 0.47 (12.0) Steel01
d10 W 0.12 (79) 0.39 (10.0) Steel01
d8 HC 0.08 (50) 0.31 (8.0) Steel01
TUA d12 B 0.18 (113) 0.47 (12.0) reinfStl
d6 BWHC 0.04 (28) 0.24 (6.0) Steel01
603

Table C.3: Rebar Properties (* denotes assumed value) - Part A (con’t)

Area db
Name Bar Size Location in2 (mm2 ) in (mm) Material Model

TUB d12 B 0.18 (113) 0.47 (12.0) reinfStl


d6 BWHC 0.04 (28) 0.24 (6.0) Steel01
NTW1 no6 B 0.44 (284) 0.75 (19.1) Steel01
no5 B 0.31 (200) 0.63 (15.9) reinfStl
no3 BWH 0.11 (71) 0.38 (9.5) Steel01
NTW2 no6 B 0.44 (284) 0.75 (19.1) Steel01
no5 BW 0.31 (200) 0.63 (15.9) reinfStl
no4 BW 0.20 (129) 0.50 (12.7) Steel01
no3 BWH 0.11 (71) 0.38 (9.5) Steel01
PW1 no4 B 0.20 (129) 0.50 (12.7) reinfStl
no2 WHC 0.05 (32) 0.25 (6.3) reinfStl
PW2 no4 B 0.20 (129) 0.50 (12.7) reinfStl
no2 WHC 0.05 (32) 0.25 (6.3) reinfStl
PW3 no4 BW 0.20 (129) 0.50 (12.7) reinfStl
no2 HC 0.05 (32) 0.25 (6.3) reinfStl
PW4 no4 B 0.20 (129) 0.50 (12.7) reinfStl
no2 WHC 0.05 (32) 0.25 (6.3) reinfStl
604

Table C.4: Rebar Properties (* denotes assumed value) - Part B

E Esh fy fu
Name Bar Size ksi (GPa) ksi (MPa) ksi (MPa) εy εsh ksi (MPa) εu
PilaSW4 d12 29028 (200) 228 (1573) 68 (470) 0.0024 - 87 (600) 0.085
d6 29028 (200) 378 (2605) 79 (545) 0.0027 - 86 (590) 0.02
PilaSW5 d16 29753 (205) 110 (760) 78 (535) 0.0026 - 86 (590) 0.075
d6 29028 (200) 378 (2605) 79 (545) 0.0027 - 86 (590) 0.02
d4 29028 (200) 150 (1034) 58 (400) 0.002 - 67 (460) 0.06
PilaSW6 d12 29028 (200) 228 (1573) 68 (470) 0.0024 - 87 (600) 0.085
d6 29028 (200) 378 (2605) 79 (545) 0.0027 - 86 (590) 0.02
d4 29028 (200) 150 (1034) 58 (400) 0.002 - 67 (460) 0.06
PilaSW7 d16 29753 (205) 110 (760) 78 (535) 0.0026 - 86 (590) 0.075
d6 29028 (200) 378 (2605) 79 (545) 0.0027 - 86 (590) 0.02
PilaSW8 d10 29028 (200) 838 (5772) 62 (430) 0.0022 - 96 (660) 0.042
d6 29028 (200) 378 (2605) 79 (545) 0.0027 - 86 (590) 0.02
d4 29028 (200) 150 (1034) 58 (400) 0.002 - 67 (460) 0.06
PilaSW9 d10 29028 (200) 838 (5772) 62 (430) 0.0022 - 96 (660) 0.042
d6 29028 (200) 378 (2605) 79 (545) 0.0027 - 86 (590) 0.02
d4 29028 (200) 150 (1034) 58 (400) 0.002 - 67 (460) 0.06
TasSHW1 d6 29000 (145)* 293 (2018) 40 (276) 0.0014 - 69 (475) 0.10*
d3 29000 (145)* 148 (1021) 31 (216) 0.0011 - 46 (317) 0.10*
TasSHW2 d6 29000 (145)* 293 (2018) 40 (276) 0.0014 - 69 (475) 0.10*
d3 29000 (145)* 148 (1021) 31 (216) 0.0011 - 46 (317) 0.10*
TasSHW3 d6 29000 (145)* 293 (2018) 40 (276) 0.0014 - 69 (475) 0.10*
d3 29000 (145)* 131 (900) 31 (216) 0.0011 - 44 (305) 0.10*
TasSHW4 d6 29000 (145)* 293 (2018) 40 (276) 0.0014 - 69 (475) 0.10*
d3 29000 (145)* 131 (900) 31 (216) 0.0011 - 44 (305) 0.10*
ThomRW1 no3 29000 (200) 1450 (9991) 63 (433) 0.0022 0.017 93 (639) 0.094
no2 29000 (200) 397 (2733) 64 (438) 0.0022 - 85 (588) 0.057
ThomRW2 no3 29000 (200) 1450 (9991) 63 (433) 0.0022 0.017 93 (639) 0.094
no2 29000 (200) 397 (2733) 64 (438) 0.0022 - 85 (588) 0.057
ThomTW1 no3 29000 (200) 1450 (9991) 63 (433) 0.0022 0.017 93 (639) 0.094
no2 29000 (200) 397 (2733) 64 (438) 0.0022 - 85 (588) 0.057
ThomTW2 no3 29000 (200) 1450 (9991) 63 (433) 0.0022 0.017 93 (639) 0.094
no2 29000 (200) 397 (2733) 64 (438) 0.0022 - 85 (588) 0.057
DazioWSH1 d10 29000 (200) 178 (1228) 79 (545) 0.0027 - 89 (612) 0.057
d6 29000 (200) 154 (1060) 83 (569) 0.0028 - 86 (590) 0.023
DazioWSH2 d10 29000 (200) 318 (2190) 84 (580) 0.0029 - 108 (746) 0.079
d6 29000 (200) 117 (808) 70 (483) 0.0024 - 78 (534) 0.065
DazioWSH3 d12 29000 (200) 218 (1504) 87 (599) 0.003 - 105 (723) 0.085
d8 29000 (200) 240 (1655) 83 (569) 0.0028 - 101 (699) 0.082
d6 29000 (200) 143 (985) 71 (488) 0.0024 - 80 (552) 0.068
DazioWSH4 d12 29000 (200) 163 (1125) 83 (574) 0.0029 - 98 (672) 0.09
d8 29000 (200) 262 (1806) 85 (583) 0.0029 - 103 (713) 0.075
d6 29000 (200) 108 (741) 75 (518) 0.0026 - 81 (558) 0.058
DazioWSH5 d8 29000 (200) 262 (1806) 85 (583) 0.0029 - 103 (713) 0.075
d6 29000 (200) 108 (741) 75 (518) 0.0026 - 81 (558) 0.058
605

Table C.4: Rebar Properties (* denotes assumed value) - Part B (con’t)

E Esh fy fu
Name Bar Size ksi (GPa) ksi (MPa) ksi (MPa) εy εsh ksi (MPa) εu

DazioWSH6 d12 29000 (200) 163 (1125) 83 (574) 0.0029 - 98 (672) 0.09
d8 29000 (200) 262 (1806) 85 (583) 0.0029 - 103 (713) 0.075
d6 29000 (200) 108 (741) 75 (518) 0.0026 - 81 (558) 0.058
LefasSW21 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW22 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW23 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW24 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW25 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW26 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW30 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW31 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW32 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
LefasSW33 d8 23077 (159) 142 (979) 68 (470) 0.003 - 82 (565) 0.10*
d6.25 21771 (150) 135 (932) 75 (520) 0.0035 - 89 (610) 0.10*
OestR1 no3 27800 (192) 386 (2660) 74 (511) 0.0027 - 111 (765) 0.098
d6 31400 (216) 216 (1486) 76 (522) 0.0024 - 102 (699) 0.12
OestR2 no4 26900 (185) 310 (2137) 65 (450) 0.0024 - 103 (708) 0.12
d6 32600 (225) 186 (1280) 78 (535) 0.0024 - 100 (690) 0.12
OestB1 no4 28300 (195) 327 (2253) 65 (449) 0.0023 - 103 (708) 0.12
d6 32500 (224) 242 (1665) 76 (520) 0.0023 - 101 (695) 0.11
OestB2 no6 30200 (208) 315 (2172) 60 (410) 0.002 - 101 (695) 0.13
d6 32100 (221) 245 (1688) 77 (532) 0.0024 - 102 (700) 0.1
OestB3 no4 25900 (178) 352 (2425) 64 (438) 0.0025 - 101 (696) 0.11
d6 30400 (209) 228 (1568) 69 (478) 0.0023 - 96 (658) 0.12
OestB4 no4 27500 (189) 322 (2217) 65 (450) 0.0024 - 103 (706) 0.12
d6 31900 (220) 207 (1426) 73 (504) 0.0023 - 99 (681) 0.13
OestB5 no6 29500 (203) 324 (2229) 64 (444) 0.0022 - 106 (733) 0.13
d6 31400 (216) 207 (1428) 73 (502) 0.0023 - 97 (671) 0.12
OestB6 no6 28500 (196) 383 (2638) 64 (440) 0.0022 - 106 (732) 0.11
d6 30400 (209) 266 (1831) 74 (511) 0.0024 - 98 (675) 0.092
OestB7 no6 28400 (196) 401 (2765) 66 (457) 0.0023 - 109 (750) 0.11
d6 28500 (196) 296 (2036) 71 (489) 0.0025 - 101 (696) 0.1
OestB8 no6 27500 (189) 410 (2824) 65 (447) 0.0024 - 108 (745) 0.11
d6 28200 (194) 220 (1518) 66 (453) 0.0023 - 89 (615) 0.11
no3 25000 (172) 405 (2789) 70 (482) 0.0028 - 106 (728) 0.091
606

Table C.4: Rebar Properties (* denotes assumed value) - Part B (con’t)

E Esh fy fu
Name Bar Size ksi (GPa) ksi (MPa) ksi (MPa) εy εsh ksi (MPa) εu

OestB9 no6 27600 (190) 414 (2853) 62 (429) 0.0023 - 107 (734) 0.11
d6 28600 (197) 219 (1506) 67 (461) 0.0023 - 89 (613) 0.1
OestB10 no5 27100 (187) 477 (3286) 55 (378) 0.002 - 108 (746) 0.11
no4 27000 (186) 328 (2259) 64 (438) 0.0024 - 102 (706) 0.12
d6 31300 (216) 210 (1444) 69 (475) 0.0022 - 92 (632) 0.11
OestF1 no4 28100 (194) 338 (2329) 65 (444) 0.0023 - 103 (707) 0.12
d6 31300 (216) 256 (1764) 76 (525) 0.0024 - 102 (704) 0.1
OestF2 no4 28100 (194) 338 (2329) 65 (444) 0.0023 - 103 (707) 0.12
d6 29400 (203) 190 (1306) 67 (464) 0.0023 - 88 (607) 0.11
Morgan d6 29000 (145)* 107 (739) 65 (448) 0.0022 - 86 (593) 0.2
D3 29000 (145)* 134 (925) 74 (510) 0.0026 - 84 (579) 0.077
LiuW1 20M 28029 (193) 1401 (9656) 66 (458) 0.0024 0.015 88 (606) 0.18
10M 23485 (162) 1174 (8091) 67 (464) 0.0029 0.022 80 (552) 0.17
LiuW2 20M 28029 (193) 1401 (9656) 66 (458) 0.0024 0.015 88 (606) 0.18
10M 23485 (162) 1174 (8091) 67 (464) 0.0029 0.022 80 (552) 0.17
AliW1 no4 29037 (200) 1452 (10003) 78 (540) 0.0027 0.01 110 (756) 0.07
no2 29107 (201) 2911 (20055)* 82 (562) 0.0028 - 364 (2511)* 0.10*
TupperW3 20M 26556 (183) 1328 (9148) 65 (450) 0.0025 0.015 87 (600) 0.17
10M 24841 (171) 1242 (8558) 71 (488) 0.0029 0.017 79 (547) 0.23
MobeenW1 15M 27576 (190) 1379 (9500) 57 (395) 0.0021 0.01 80 (553) 0.094
10M 27431 (189) 1372 (9450) 56 (384) 0.002 0.015 77 (527) 0.16
RivaW1 d20 29000 (145)* 143 (985) 81 (560) 0.0028 - 93 (640) 0.084
d8 29000 (145)* 143 (985) 81 (560) 0.0028 - 93 (640) 0.084
ShiuC1 no4 26100 (180) 425 (2930) 69 (476) 0.0026 - 111 (761) 0.10*
d6 29500 (203) 263 (1813) 69 (473) 0.0023 - 94 (650) 0.10*
KhaC1 20M 29000 (145)* 2900 (19981)* 68 (470) 0.0024 - 351 (2421)* 0.10*
10M 29000 (145)* 2900 (19981)* 68 (470) 0.0024 - 351 (2421)* 0.10*
6M 29000 (145)* 2900 (19981)* 87 (600) 0.003 - 368 (2538)* 0.10*
ElnCW2 15M 29000 (145)* 460 (3171) 65 (450) 0.0023 - 110 (760) 0.10*
15M 29000 (145)* 460 (3171) 65 (450) 0.0023 - 110 (760) 0.10*
d6 29000 (145)* 69 (474) 83 (570) 0.0029 - 89 (616) 0.10*
ElnCW3 15M 29000 (145)* 460 (3171) 65 (450) 0.0023 - 110 (760) 0.10*
d6 29000 (145)* 69 (474) 83 (570) 0.0029 - 89 (616) 0.10*
IleX d12 29000 (145)* 59 (403) 75 (516) 0.0026 - 89 (615) 0.25
d10 29000 (145)* 56 (384) 76 (525) 0.0026 - 90 (617) 0.24
d8 29000 (145)* 50 (344) 81 (557) 0.0028 - 93 (642) 0.25
IleY d12 29000 (145)* 59 (403) 75 (516) 0.0026 - 89 (615) 0.25
d10 29000 (145)* 56 (384) 76 (525) 0.0026 - 90 (617) 0.24
d8 29000 (145)* 50 (344) 81 (557) 0.0028 - 93 (642) 0.25
IleXY d12 29000 (145)* 59 (403) 75 (516) 0.0026 - 89 (615) 0.25
d10 29000 (145)* 56 (384) 76 (525) 0.0026 - 90 (617) 0.24
d8 29000 (145)* 50 (344) 81 (557) 0.0028 - 93 (642) 0.25
TUA d12 29000 (145)* 1450 (10)* 71 (488) 0.0024 0.024 86 (595) 0.13
d6 29000 (145)* 291 (2002) 75 (518) 0.0026 - 99 (681) 0.084
607

Table C.4: Rebar Properties (* denotes assumed value) - Part B (con’t)

E Esh fy fu
Name Bar Size ksi (GPa) ksi (MPa) ksi (MPa) εy εsh ksi (MPa) εu

TUB d12 29000 (145)* 1450 (10)* 68 (471) 0.0024 0.032 83 (574) 0.13
d6 29000 (145)* 291 (2002) 75 (518) 0.0026 - 99 (681) 0.084
NTW1 no6 28700 (198) 194 (1338) 64 (438) 0.0022 - 92 (636) 0.15
no5 28000 (193) 1400 (9646) 63 (434) 0.0022 0.0046 92 (630) 0.15
no3 29000 (200) 237 (1634) 73 (502) 0.0025 - 106 (727) 0.14
NTW2 no6 29600 (204) 238 (1641) 67 (460) 0.0023 - 101 (695) 0.14
no5 27700 (191) 1385 (9543) 66 (456) 0.0024 0.0069 103 (706) 0.14
no4 29000 (145)* 222 (1532) 75 (513) 0.0026 - 106 (727) 0.14
no3 29000 (145)* 229 (1577) 73 (504) 0.0025 - 105 (723) 0.14
PW1 no4 29000 (200) 1450 (9991) 84 (579) 0.0029 0.015 91 (628) 0.12
no2 29000 (200) 1450 (9991) 76 (522) 0.0026 0.015 76 (526) 0.058
PW2 no4 29000 (200) 1450 (9991) 84 (579) 0.0029 0.015 91 (628) 0.12
no2 29000 (200) 1450 (9991) 76 (522) 0.0026 0.015 76 (526) 0.058
PW3 no4 29000 (200) 1450 (9991) 51 (353) 0.0018 0.012 70 (484) 0.2
no2 29000 (200) 1450 (9991) 76 (522) 0.0026 0.015 76 (526) 0.058
PW4 no4 29000 (200) 1450 (9991) 67 (462) 0.0023 0.0075 99 (682) 0.13
no2 29000 (200) 1450 (9991) 76 (522) 0.0026 0.015 76 (526) 0.058
608

C.2 Damage States

Table C.5: Damage States and Methods of Repair

Method of Repair Damage State Description


DS1a Initial cracking
DS1b Initial flexural cracking
DS1c Initial shear cracking
Cosmetic repair
DS1d Tensile yield of extreme longitudinal steel
DS1e Compression yield of longitudinal steel
DS1f Yield of horizontal reinforcement
DS2a Spalling of boundary region cover concrete
Epoxy injection
(not revealing longitudinal reinforcement)
and
DS2b Spalling of patched concrete
concrete patching
DS2c Spalling of web concrete
DS2d Vertical cracks/splitting
Replace concrete DS3a Spalling revealing longitudinal reinforcement
DS4a Core crushing (boundary element)
DS4b Bar buckling
DS4c Compressive failure of boundary element
DS4d Failure by core crushing (boundary element)
DS4e Bar fracture
DS4f Failure due to bar buckling
Replace wall
DS4g Failure due to bar fracture
DS4i Shear failure
DS4k Web crushing
DS4m Failure due to web crushing
DS4o Failure by bond slip
DS4p Core crushing in confined boundary element of
flange tips (bi-directional tests only)
DS4q Confining reinforcement open or fractured
DSr1 First yield of flexural reinforcement
DSr2 Residual cracks ≤ 1/16 inches
DSr3 Wall fully cracked
DSr4 Large shear cracks
Not used for DSr5 Shear cracks extending most of wall length
development of DSr9 Cracks ≥ 1/16 inches
fragility functions DSr10 Cracks ≥ 1/8 inches
DSr11 Cracks ≥ 1/4 inches
DSr12 Cracks ≥ 1/2 inches
DSr13 Cracks ≥ 1 inch
DSr14 Stirrup yield
609

C.3 Damage Data for Individual Specimens


610

C.3.1 PilaSW4

Table C.6: PilaSW4 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.08% 4.2 (18.8) Occurred before 1mm disp, halfway Fig. C.2
to first maximum (p. 102)
DS1d -0.56% -19.4 (-86.2) As determined from
momentcurvature analysis
2 DS2d 1.33% 22.7 (100.9) Vertical cracks in boundary element Fig. C.3
at base, at location of long. reinf.
(p. 102); Identical other than de-
scription to large shear crack width
damage state for same wall
4 DS4a 1.83% 22.4 (99.8) Concrete confined by lowest 2 B.E. Fig. C.4
hoops spalling (p. 102)
DS4d -2.00% -20.3 (-90.3) Failure by core crushing in bound- Fig. C.5
ary elements (p. 102); No load his-
tory reported at this drift (LVDT
maxed out). Load is extrapolated
from the negative envelope, as the
envelope and figure of damage sug-
gests the left side (compression un-
der (-) drift) is the side that saw
compressive failure
DSr1 0.46% 17.1 (76.0) Occurred before 6mm disp (cycle Fig. C.6
peak) (p. 102); exact values re-
ported in Table 7.2 (p. 193)
DSr4 1.33% 22.7 (100.9) Lower web crack widths larger than Fig. C.7
others at end of 16mm cycle, as
noted by researchers (p. 102)
611

PilaSW4
25

20

15

10

5
Vbase, kips

−5

−10
Env.
DS1a
−15
DS1d
DS2d
−20 DS4a
DS4d
−25
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.1: Envelope for PilaSW4

Figure C.2: PilaSW4: DS1a - cracks at MDL-2 (peak after initial cracking)

4.(4).2 Crack pattern of wall SW4 at MDL-2 and MDL-4

First yield occurred just before MDL-6, but MDL-8 will be pre
as the first post-yield displacement in figure 4.(4).3. Following yield
612

4.(4).3 Crack pattern of wall SW4 at MDL-8 and MDL-1

Figure C.3: PilaSW4: DS2d - cracks at MDL-16 (larger shear crack widths)
4.(4).3 Crack pattern of wall SW4 at MDL-8 and MDL-16

Figure C.4: PilaSW4: DS4a - cracks at MDL-22 (core crushing)

Figure 4.(4).4 Crack pattern of wall SW4 at MDL-22 and failur

Figure 4.(4).4 Crack pattern of wall SW4 at MDL-22 and failure


4.(4).3 Crack pattern of wall SW4 at MDL-8 and MDL-16
613

Figure C.5: PilaSW4: DS4d - cracks at MDL-22 (core crushing)

igure 4.(4).4 Crack pattern of wall SW4 at MDL-22 and failure

103

Figure C.6: PilaSW4: DSr1 - cracks at MDL-8 (peak of cycle following yield cycle)
4.(4).3 Crack pattern of wall SW4 at MDL-8 and MDL-1
614

Figure C.7: PilaSW4: DSr4 - cracks at MDL-16 (larger shear crack widths)
4.(4).3 Crack pattern of wall SW4 at MDL-8 and MDL-16
615

C.3.2 PilaSW5

Table C.7: PilaSW5 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.08% 4.8 (21.1) Occurred before 1mm disp, halfway Fig. C.9
to first maximum; cracks 3/4 of wall
at end of cycle; more cracks in B.E.
than web; web cracks 30 degrees (p.
105)
DS1d 0.75% 25.5 (113.5) As determined from
momentcurvature analysis
DS1f 0.83% 26.0 (115.8) Shear reinf. yielded (flex. not Fig. C.10
yielded) (p. 106)
4 DS4i -0.83% -23.0 (-102.4) Loss of shear capacity on way to neg. Fig. C.11
peak; shear cracks signficantly open;
researcher noted failure due to lower
shear crack extending into boundary
element in compression (extending
from nearly top corner to lower cor-
ner) (p. 106)
DSr1 0.69% 24.7 (110.0) Yield of flexural reinforcement, from
force reported in Table 7.2 (p. 193)
DSr5 0.67% 24.4 (108.7) Shear cracks extend most of wall Fig. C.10
length (p. 107)
616

PilaSW5
30

20

10
Vbase, kips

−10

Env.
−20 DS1a
DS1d
DS1f
DS4i
−30
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
% Drift

Figure C.8: Envelope for PilaSW5

Figure C.9: PilaSW5: DS1a - cracks at MDL-2 (peak after initial cracking) and MDL-4
(wall fully 4.(5).2
cracked) Crack pattern of wall SW5 at MDL-2 and MDL-4

By MDL-8 the frequency of cracking in the boundaries increased


considerably as shown in figure 4.(5).3. The main web cracks started
joining with cracks originating higher in the boundary than before and
hence, increased their apparent inclination.The most precarious crack
617

3 Crack pattern of wall SW5 at MDL-8 and MDL-1O

Figure C.10: PilaSW5:


Following DS1f
failure- in
cracks at MDL-8
one direction, (shear
it was cracksto extend
decided lengththe
avoid cycling of wall) and
MDL-10 wall
(shear reinf.
twice yield)
at the same MDL and failure in the other direction was
immediate after exceeding MDL-10 at a load of about 108 KM. The last MDL
imposed in both directions was MDL-12. For the sake of demonstrating the
effect of cycling on a wall that failed in shear, displacements were
increased by 2 mm in each direction as shown in figure 4.(5).1, until the
limits of the control transducers were reached. The state of the wall at
MDL-14 is shown in figure 4.(5).4, demonstrating failure in the forward
direction which was caused by the crack originating at the top left corner.
Again two main cracks have opened considerably. After failure, several
hoops and stirrups opened up and considerable opening was noticed in the
middle sections of the wall. Additionally, it could be seen that the wall was
displacing in a rigid body mode above the main cracks. The considerable
degradation of the wall at MDL. 26 is shown in figure 4.(5).4.

3 Crack pattern of wall SW5 at MDL-8 and MDL-1O

Figure C.11: PilaSW5:


Following DS4i
failure- in
cracks at MDL-8
one direction, (shear
it was cracksto extend
decided length the
avoid cycling of wall) and
MDL-10 wall
(shear reinf. yield)
twice at the same MDL and failure in the other direction was
immediate after exceeding MDL-10 at a load of about 108 KM. The last MDL 107
imposed in both directions was MDL-12. For the sake of demonstrating the
effect of cycling on a wall that failed in shear, displacements were
increased by 2 mm in each direction as shown in figure 4.(5).1, until the
limits of the control transducers were reached. The state of the wall at
618

C.3.3 PilaSW6

Table C.8: PilaSW6 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.08% 4.5 (20.0) Occurred before 1mm disp, halfway Fig. C.13
to first maximum; flexure and shear
(extending into b.e.) (p. 109)
DS1d 0.58% 18.7 (83.2) As determined from
momentcurvature analysis
2 DS2a 1.33% 23.9 (106.2) Spalling at base of wall; diagonal Fig. C.14
cracks extending into boundary el-
ement; maximum load occurred (p.
110);
4 DS4a 1.83% 20.3 (90.2) Crushing initiated (p. 111)
DS4d 1.83% 20.2 (89.7) Wide crack in B.E. (not specified as
shear and flexure); Cycle wall failed
at (p. 111)
DS4q 1.50% 24.1 (107.4) Stirrup opened; cracks extending Fig. C.15
into compressed region (p. 111)
DSr1 0.47% 16.6 (74.0) Occurred just before 6mm disp (cy- Fig. C.14
cle peak) (p. 109); exact reported in
Table 7.2 (p. 193)
619

PilaSW6
25

20

15

10

5
Vbase, kips

−5

−10 Env.
DS1a
DS1d
−15
DS2a
DS4q
−20 DS4a
DS4d
−25
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.12: Envelope for PilaSW6

4.(6).2 Crack pattern of wall SW6 at MDL-2 and MDL-4

Figure C.13: PilaSW6:


After yield,DS1a - cracks
the extent at MDL-2
of cracking in (peak
the webafter initial considerably
increased cracking) and MDL-4
(fully cracked)
as shown in figure 4.(6).3. The main web cracks started joining with cracks
originating higher in the boundary, hence increased more their apparent
inclination than in SW4. Following MDL-8, the width of cracking started to
open considerably more than in SW4. However, by MDL-16 (figure 4.(6).3)
the maximum load achieved was 107 KN which was higher than the
ultimate for SW4. Diagonal cracking propagated from the web into the
opposite boundaries and seemed to have pushed the neutral axis, at least in
620

4.(6).3 Crack pattern of wall SW6 at MDL-8 and MDL-16


Figure C.14: PilaSW6: DS2a
By MDL-18 - cracks
(figure at the
4.(6).4), MDL-8 (yield)
concrete and MDL-16
deterioration (spalling/maximum
started affecting
load) the strength and even though the maximum load was achieved during the
BHS MDL, less load was resisted when the displacement was reversed. An
opening of the stirrup in the lower level was observed at this MDL as well as
progression of the cracks through the compressed area. Crushing of
concrete was initiated just before achieving MDL-20 in the LHS. The loss of
strength continued in the subsequent cycles and by MDL-22 the load was
below 75% of ultimate in the RHS direction as well. In this direction, a wide
crack crossing through the bottom boundary area was observed. MDL-22
(figure 4.(6).4) was the last displacement to be imposed since the wall was
considered to have failed.

Figure 4.(6).4 Crack pattern of wall SW6 at MDL-18 and MDL-22

Figure C.15: PilaSW6: DS4q - cracks at MDL-18 (open stirrup) and Failure

4.6.2 Load-displacement curves 111


The load-displacement curves are given in figures A.(6).1 - A.(6).1O
for displacements 3 to 12. Base vertical and out-of-plane displacements are
shown in figures A.(6).11 and A.(6).12, respectively. The wall extensions at
621

C.3.4 PilaSW7

Table C.9: PilaSW7 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.08% 4.5 (20.1) Occurred before 1mm disp, halfway Fig. C.17
to first maximum; flexure and shear
cracks 3/4 of wall height (p. 113)
DS1d -0.62% -25.5 (-113.5) As determined from
momentcurvature analysis
4 DS4e -1.83% -22.8 (-101.5) Fracture of 6mm reinf. bar (2 frac- Fig. C.18
tured by end of cycle to LHS) (p.
115)
DSr1 0.68% 24.1 (107.0) Yield of flexural reinforcement, from
force reported in Table 7.2 (p. 193)
DSr5 1.17% 27.6 (123.0) Main web cracks extend into com- Fig. C.19
pressive area (figure validates clas-
sification as extending nearly full
length of wall) (p. 115)

PilaSW7
30

20

10
Vbase, kips

−10

−20 Env.
DS1a
DS1d
DS4e
−30
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.16: Envelope for PilaSW7


-12
-16

-24
0246 810121416182O24
CycleNo 622
4.(7).1 Loading history for SW7

4.(7).2 Crack pattern of wall SW7 at MDL-2 and MDL-4

Figure C.17: PilaSW7:


At MDL-8,DS1a - cracks
the extent at MDL-2
of cracking (peak
at the after initial
boundaries cracking) and MDL-4
increased
(fully cracked)
considerably as shown in figure 4.(7).3. The main web cracks started
joining with cracks originating higher in the boundary, hence increased
114

Figure C.18: PilaSW7: DS4e - cracks at MDL-22 (failure due to bar fracture)
4.(7).4 Crack pattern of wall SW7 at MDL-22

4.7.2 Load-displacement curves


623

their apparent inclination but not to the extent seen for SW5. Following
MDL-8, cracking continued to become denser at the boundaries with an
average inclination of more than 45°, 'millie for wails SW4 and SW6 where
the boundaries were wider. By MDL-14 (figure 4.(7).3), the miin web cracks
were wider than before and seemed to have penetrated the compressive
area, at least on the surface.

4.(7).3 Crack pattern of wall SW7 at MDL-8 and MDL-14

Figure C.19:
ThePilaSW7: DSr5
load capacity - cracks
peaked at MDL-8
at MDL-18 and MDL-4to(shear
corresponding a load cracks
of 127.3in b.e.)
KM. Following thatthe widening of the web cracks increased and there was
some loss of strength at MDL-22 in the RHS direction. However, failure
occurred at MDL-22 by fracturing of a 6 mm reinforcement bar on the LHS
virgin cycle. A second bar snapped during the completion of this MDL and
the strength was reduced to less than half .The considerable degradation of
the wail at MDL-22 is shown in figure 4.(7).4.
624

C.3.5 PilaSW8

Table C.10: PilaSW8 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.08% 4.1 (18.3) Occurred before 1mm disp, halfway Fig. C.21
to first maximum; flexure and shear
cracks 1/2 of wall height (p. 117)
DS1d -0.46% -17.2 (-76.5) As determined from
momentcurvature analysis
2 DS2a 1.50% 21.3 (94.8) Spalling at interior of B.E. (p. 119) Fig. C.22
DS2d 1.17% 20.8 (92.5) Vertical cracks at bottom end main
reinforcement (p. 119)
3 DS3a 2.00% 21.2 (94.4) Flexural reinforcement exposed at
extreme bottom parts of wall (p.
119)
4 DS4c 2.17% 19.8 (87.9) Assummed failure mode. Loss of
lateral load capacity dropped to
75% of max
DSr1 0.44% 16.0 (71.0) Yield of flexural reinforcement, from
force reported in Table 7.2 (p. 193)
DSr4 1.00% 21.1 (93.8) Wider web cracks without change in Fig. C.23
crack pattern (p. 118)
625

PilaSW8
25

20

15

10

5
Vbase, kips

−5

−10 Env.
DS1a
DS1d
−15
DS2d
DS2a
−20 DS3a
DS4c
−25
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
% Drift

Figure C.20: Envelope for PilaSW8

4.(8).2 Crack pattern of wall SW8 at MDL2 and MDL-4

Figure C.21: PilaSW8:


At MDL-6,DS1a - cracks
a crack at MDL-2
originating (peak
from the top after
cornerinitial cracking) and MDL-4
propagated
(fully cracked)
through the web at an angle slightly higher then 45° to meet the web crack
below, as shown in figure 4.(8).3. After MDL-6, craiking continued to
become denser at the boundaries but the number of opening msin cracks in
the web stabilised to about four to five. By MDL-12 (figure 4.(8).3), the same
pattern of cracking continued, but the main web cracks were wider than
before.
626
Crack pattern of wall SW8 at MDLI and MDL-12

4.(8).4 Crack pattern of wall SW8 at MDL-18 and failure

Figure C.22: PilaSW8: DS2a - cracks at MDL-18 (spalling) and Failure

119

Crack pattern of wall SW8 at MDLI and MDL-12

Figure C.23: PilaSW8: DSr4 - cracks at MDL-6 and MDL-12


627

C.3.6 PilaSW9

Table C.11: PilaSW9 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.08% 4.1 (18.1) Occurred before 1mm disp, halfway Fig. C.25
to first maximum; flexure and shear
cracks 1/2 of wall height (p. 120)
DS1d -0.42% -17.0 (-75.7) As determined from
momentcurvature analysis
2 DS2a 1.50% 20.9 (93.0) Spalling at intersection of main Fig. C.26
cracks in web (p. 122)
4 DS4c 2.20% 12.2 (54.1) Assummed failure mode. Loss of
lateral load capacity dropped
to 75% of max
DSr1 0.41% 15.7 (70.0) Yield of flexural reinforcement, from
force reported in Table 7.2 (p. 193)
DSr4 1.17% 20.5 (91.0) Wider web cracks and opening fast Fig. C.27
(p. 122)

PilaSW9
25

20

15

10

5
Vbase, kips

−5

−10

Env.
−15
DS1a
DS1d
−20 DS2a
DS4c
−25
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
% Drift

Figure C.24: Envelope for PilaSW9


-10
-14
-18
-22
-26
0246810 12 14 16 18 )22 24 26
CydeNo 628

1 hi for SW9

4.(9).2 Crack pattern of wall SW9 at MDL-2 and MDL-4

Figure C.25: PilaSW9:


By MDL-6,DS1a - cracks at
the frequency MDL-2 (peak
of cracking in the after initial increased
boundaries cracking) and MDL-4
(fully cracked)
considerably as shown in figure 4.(9).3. However, the top crack, which also
appeared in SW8 at this stage, had a shallower angle. After MDL-6,
121

Crack pattern of wall SW9 at MDL-18 and failure

Figure C.26: PilaSW9: DS2a - cracks at MDL-18 (spalling) and Failure


4.9.2 Load-disDiacement curves

The load-displacement curves are given in figures A.(9).1 - A.(9).1O


for displacements 3 to 12. Base vertical and out-of-plane displacements are
shown in figures A.(9).11 and A.(9).12 respectively. The wall extensions at
top, half and quarter height locations is shown in figures A.(9).13 - A.(9).15,
629

cracking continued to increase in density as observed for all previous walls.


By MDL-14, (figure 4.(9).3) the main lower web cracks were wider than
before and seemed to be opening faster than the top ones.

The cracking pattern until MDL. 18 was similar to SW8 as shown in


figure 4.(9).4. Subsequently the upper cracks of SW9 were not opening as
much as the corresponding ones for SW8. Spalling of concrete in the web at
the intersections of main cracks was only confined to the lower quarter of
the wall. The wall survived both cycles at MDL-24 and lost strength during
the virgin cycle reverse and at the second cycle RHS MDL-26. MDL-26 is
considered to be failure as shown in figure 4.(9).4.

4.(9).3 Crack vattern of wall SW9 at MDL-6 and MDL-14

Figure C.27: PilaSW9: DSr4 - cracks at MDL-6 and MDL-14

122
630

C.3.7 TasSHW1

Table C.12: TasSHW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 0.3 (1.5) flexural cracks initiate at 10% of
wall capacity
DS1c 0.38% 2.0 (9.1) significant inclined cracking
initiated at 62% of maximum load
DS1d -0.52% -2.6 (-11.4) As determined from
momentcurvature analysis
2 DS2a 1.56% 3.3 (14.5) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DS2d 1.56% 3.3 (14.5) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DSr1 -0.41% -2.3 (-10.3) Yield displacement reported in
Table 8 of paper

TasSHW1
4

1
Vbase, kips

−1

−2 Env.
DS1b
DS1c
−3 DS1d
DS2a
DS2d
−4
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.28: Envelope for TasSHW1


631

C.3.8 TasSHW2

Table C.13: TasSHW2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 0.4 (1.8) flexural cracks initiate at 10% of
wall capacity
DS1c 0.34% 2.5 (11.3) significant inclined cracking
initiated at 62% of maximum load
DS1d 0.35% 2.6 (11.4) As determined from
momentcurvature analysis
2 DS2a 0.86% 3.9 (17.3) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DS2d 0.86% 3.9 (17.3) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DSr1 0.31% 2.4 (10.6) Yield displacement reported in
Table 8 of paper

TasSHW2
5

1
Vbase, kips

−1

−2
Env.
DS1b
−3
DS1c
DS1d
−4 DS2a
DS2d
−5
−1.5 −1 −0.5 0 0.5 1 1.5
% Drift

Figure C.29: Envelope for TasSHW2


632

C.3.9 TasSHW3

Table C.14: TasSHW3 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.03% 0.3 (1.5) flexural cracks initiate at 10% of
wall capacity
DS1c 0.39% 2.1 (9.4) significant inclined cracking
initiated at 62% of maximum load
DS1d -0.42% -2.6 (-11.5) As determined from
momentcurvature analysis
2 DS2a -1.03% -3.7 (-16.4) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DS2d -1.03% -3.7 (-16.4) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DSr1 -0.46% -2.7 (-12.1) Yield displacement reported in
Table 8 of paper

TasSHW3
4

1
Vbase, kips

−1

−2 Env.
DS1b
DS1c
−3 DS1d
DS2a
DS2d
−4
−1.5 −1 −0.5 0 0.5 1
% Drift

Figure C.30: Envelope for TasSHW3


633

C.3.10 TasSHW4

Table C.15: TasSHW4 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 0.4 (2.0) flexural cracks initiate at 10% of
wall capacity
DS1c 0.56% 2.8 (12.3) significant inclined cracking
initiated at 62% of maximum load
DS1d 0.50% 2.6 (11.5) As determined from
momentcurvature analysis
2 DS2a -1.70% -5.3 (-23.5) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DS2d -1.70% -5.3 (-23.5) A few semi-vertical cracks and
spalling in compressive zone just
prior to failure
DSr1 0.43% 2.4 (10.7) Yield displacement reported in
Table 8 of paper

TasSHW4
6

2
Vbase, kips

−2

Env.
DS1b
−4 DS1d
DS1c
DS2a
DS2d
−6
−2 −1.5 −1 −0.5 0 0.5 1 1.5
% Drift

Figure C.31: Envelope for TasSHW4


634

C.3.11 ThomRW1

Table C.16: ThomRW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.25% -14.7 (-65.3) Flexural cracking over first floor at
approximately 0.25% drift; also hor-
izontal cracks at construction joints
DS1c -0.50% -26.8 (-119.4) Flexural cracks extend to become
shear-flexure cracks
DS1d -0.74% -30.2 (-134.2) As determined from
momentcurvature analysis
2 DS2a -1.00% -31.7 (-140.9) Vertical splitting and minor crush-
ing at wall edge at approximately
1.0% drift (first cycle)
DS2d -1.00% -31.7 (-140.9) Vertical splitting and minor crush-
ing at wall edge at approximately
1.0% drift (first cycle)
4 DS4b -2.00% -33.4 (-148.3) Slight buckling of edge boundary Fig. C.33
bars observed in first cycle at 2.0%
drift; extensive spalling & crushing
in B.E.
DS4f 2.22% 30.6 (135.9) Loss of lateral load carrying capac-
ity due to buckling at 2.5% drift
(first excursion)
DSr1 -0.75% -30.3 (-135.0) Yielding of reinforcement at approx-
imately 0.75% drift
DSr10 -1.50% -32.8 (-145.7) Crack widths measured at 3/32”

Table C.17: ThomRW1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr10 -1.50% -32.8 (-145.7) Max Any 0.094 (2.4)
DS4b -2.00% -33.4 (-148.3) Max Any 0.094 (2.4)
635

ThomRW1
40

30

20

10
Vbase, kips

−10
Env.
DS1b
−20 DS1c
DS1d
DS2a
−30 DS2d
DS4b
DS4f
−40
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5 3
% Drift

Figure C.32: Envelope for ThomRW1

Figure C.33: ThomRW1: DS4b - Slight buckling of RW1 bars at 2% drift


trains—Specimen TW2

is used in ACI 318-99 to


636

C.3.12 ThomRW2

Table C.18: ThomRW2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.25% -17.5 (-77.8) Flexural cracking over first floor at
approximately 0.25% drift
DS1c -0.50% -25.3 (-112.5) Flexural cracks extend to become
shear-flexure cracks
DS1d -0.63% -27.7 (-123.1) As determined from
momentcurvature analysis
DS1e -0.75% -29.9 (-132.8) Compression yielding of reinforce-
ment at approximately 0.75% drift
(tension yielding occurs in same cy-
cle)
2 DS2a -1.50% -32.2 (-143.4) Minor crushing and spalling of cover
concrete during first cycle to 1.5%
drift (vertical cracking was extensive
in boundary regions)
DS2d -1.00% -31.0 (-137.9) Vertical splitting during first cycle
to 1.0% drift
4 DS4f -2.50% -35.2 (-156.4) Loss of lateral load carrying capac- Fig. C.35
ity due to buckling at 2.5% drift
(second excursion); multiple bars
(8) buckled simultaneously; core
crushing was also occuring
DSr1 -0.75% -29.9 (-132.8) Tensile yielding of reinforcement at
approximately 0.75% drift
(compression yielding occurs
in same cycle)
DSr10 -2.00% -34.0 (-151.1) Crack widths measured at 3/32”
(2.0% drift)

Table C.19: ThomRW2 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr10 -2.00% -34.0 (-151.1) Max Any 0.094 (2.4)
637

ThomRW2
40

30

20

10
Vbase, kips

−10
Env.
DS1b
−20 DS1c
DS1d
DS1e
−30 DS2d
DS2a
DS4f
−40
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
% Drift

Figure C.34: Envelope for ThomRW2

Figure C.35: ThomRW2: DS4f - RW2 damage at 2.5% drift


638

C.3.13 ThomTW1

Table C.20: ThomTW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.25% 19.8 (88.2) Initial flexural cracking during sec-
ond drift level
DS1c 0.50% 30.0 (133.6) Initial shear cracking seen in south
web boundary element during cycle
to 0.5% drift
DS1d -0.39% -33.5 (-148.9) As determined from
momentcurvature analysis
2 DS2d -0.75% -48.9 (-217.6) Onset of vertical splitting of web
boundary element
4 DS4f -1.25% -65.4 (-290.8) Loss of lateral load carrying capac- Fig. C.37
ity due to buckling of web and B.E.
reinforcement on way to 1.25% drift
DSr1 0.75% 38.0 (169.1) Longitudinal steel in web b.e. yield-
ing in tension

ThomTW1
60

40

20

0
Vbase, kips

−20

−40
Env.
DS1b
DS1d
−60 DS1c
DS2d
DS4f
−80
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.36: Envelope for ThomTW1


639

W2

-99 to
ndary
RW1
3d b ),

pends
7 and
ior of
city of
2, and
Figure C.37: ThomTW1: DS4f - TW1 damage at 1.25% drift
of the
te the
d sud- Fig. 16. Specimen 共a兲 RW1 damage—2% drift; 共b兲
he lat- 2.5% drift; 共c兲 TW1 damage—1.25% drift; and 共d兲
N兲, is 2.5% drift
s, 276
using
l load ture during the first cycle to 2.5% lateral drift. A
640

C.3.14 ThomTW2

Table C.21: ThomTW2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1d -0.24% -33.2 (-147.5) As determined from
momentcurvature analysis
4 DS4f -2.00% -81.3 (-361.5) Loss of lateral load carrying capac- Fig. C.39
ity due to buckling over several hoop
spacing in web boundary compres-
sion zone

ThomTW2
50

0
Vbase, kips

−50

Env.
DS1d
DS4f
−100
−3 −2 −1 0 1 2 3
% Drift

Figure C.38: Envelope for ThomTW2


641

Figure C.39: ThomTW2: DS4f - TW2 damage at 2.5% drift

g. 16. Specimen 共a兲 RW1 damage—2% drift; 共b兲 RW2 damage—


5% drift; 共c兲 TW1 damage—1.25% drift; and 共d兲 TW2 damage—
5% drift

re during the first cycle to 2.5% lateral drift. As expected, the


havior of specimen RW2 was slightly better then specimen
W1 due to the closer hoop spacing. Fig. 16共b兲 reveals that, at
lure, all eight longitudinal bars of Specimen RW2 buckled si-
ultaneously. The brittle failure of Specimen TW1 is shown in
642

C.3.15 DazioWSH1

Table C.22: DazioWSH1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.17% 52.4 (233.1) First cracking (from data
downloaded for project)
DS1d -0.21% -58.4 (-259.7) As determined from
momentcurvature analysis
2 DS2d 0.69% 74.7 (332.4) Vertical cracks in cover at base of
wall (same time as web reinf. rup-
ture)
4 DS4e 0.69% 74.7 (332.4) Bar rupture in web reinforcement
(vertical cracking beginning to ap-
pear)
DSr1 0.18% 53.6 (238.3) First yield as reported in paper Ta-
ble 4 (∆ y b )
DSr10 0.69% 74.7 (332.4) Cracks exceeding or equal to 1/8”
(determined from download crack
data)
DSr9 0.69% 74.7 (332.4) Cracks exceeding or equal to 1/16”
(determined from download crack
data)
643

Table C.23: DazioWSH1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS1a 0.17% 52.4 (233.1) Max Shear 0.0039 (0.1)
Flexure 0.0039 (0.1)
Any 0.0039 (0.1)
DSr1 0.47% 74.7 (332.4) Max Shear 0.016 (0.4)
Flexure 0.024 (0.6)
Any 0.024 (0.6)
DSr9 0.69% 74.7 (332.4) Max Shear 0.071 (1.8)
Flexure 0.14 (3.5)
Any 0.14 (3.5)
DSr10 0.69% 74.7 (332.4) Max Shear 0.071 (1.8)
Flexure 0.14 (3.5)
Any 0.14 (3.5)
DS4e 0.69% 74.7 (332.4) Max Shear 0.071 (1.8)
Flexure 0.14 (3.5)
Any 0.14 (3.5)
DS2d 0.69% 74.7 (332.4) Max Shear 0.071 (1.8)
Flexure 0.14 (3.5)
Any 0.14 (3.5)

DazioWSH1
80

60

40

20
Vbase, kips

−20

−40 Env.
DS1a
DS1d
−60 DS2e
DS4e
DS2d
−80
−0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1 1.2
% Drift

Figure C.40: Envelope for DazioWSH1


644

C.3.16 DazioWSH2

Table C.24: DazioWSH2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.16% 54.3 (241.7) First cracking (from data
downloaded for project)
DS1d 0.23% 59.5 (264.5) As determined from
momentcurvature analysis
2 DS2d 1.15% 77.6 (345.3) Vertical cracking beginning to ap-
pear (web bar fracture also occurs)
3 DS3a -1.15% -72.8 (-323.7) Cover spalled fell off, revealing buck-
led bars (buckling recorded as inde-
pendent damage state)
4 DS4b -1.15% -72.8 (-323.7) Initial buckling seen; revealed by
cover spalling (independent damage
states)
DS4e 1.15% 77.6 (345.3) Bar rupture in web reinforcement
(vertical cracking beginning to ap-
pear)
DS4g 1.40% 66.0 (293.5) Previously buckled bars ruptured
causing loss of load carrying capac-
ity
DSr1 0.17% 55.4 (246.3) First yield as reported in paper Ta-
ble 4 (∆ y b )
645

Table C.25: DazioWSH2 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS1a 0.16% 54.3 (241.7) Max Shear 0.0039 (0.1)
Flexure 0.0059 (0.15)
Any 0.0059 (0.15)
DSr1 0.46% 75.7 (336.7) Max Shear 0.018 (0.45)
Flexure 0.024 (0.6)
Any 0.024 (0.6)
DS2d 1.15% 77.6 (345.3) Max Shear 0.043 (1.1)
Flexure 0.24 (6)
Any 0.24 (6)
DS4e 1.15% 77.6 (345.3) Max Shear 0.043 (1.1)
Flexure 0.24 (6)
Any 0.24 (6)
DS3a -1.15% -72.8 (-323.7) Max Shear 0.35 (9)
Flexure 0.098 (2.5)
Any 0.35 (9)
DS4b -1.15% -72.8 (-323.7) Max Shear 0.35 (9)
Flexure 0.098 (2.5)
Any 0.35 (9)

DazioWSH2
100

80

60

40

20
Vbase, kips

−20
Env.
−40 DS1a
DS1d
−60 DS2d
DS4e
DS3a
−80 DS4b
DS4g
−100
−2 −1.5 −1 −0.5 0 0.5 1 1.5
% Drift

Figure C.41: Envelope for DazioWSH2


646

C.3.17 DazioWSH3

Table C.26: DazioWSH3 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.26% 67.9 (302.2) First cracking (from data
downloaded for project)
DS1d -0.37% -75.1 (-334.1) As determined from
momentcurvature analysis
2 DS2a 1.02% 100.9 (448.9) Initial spalling, but long. reinf.
NOT seen
3 DS3a 1.70% 101.9 (453.2) Cover continued to spall, revealing
long. reinf. for first time (initial
buckling also seen)
4 DS4b 1.70% 101.9 (453.2) Initial buckling seen when spalling
first revealed long. reinf. (indep.
damage state)
DS4g 2.04% 101.9 (453.2) Bar rupture in corner bar (previ-
ously buckled) shortly before cycle
peak
DSr1 0.25% 64.5 (286.8) First yield as reported in paper Ta-
ble 4 (∆ y b )
DSr9 1.36% 101.9 (453.2) Cracks exceeding or equal to 1/16”
(determined from download crack
data)
647

Table C.27: DazioWSH3 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.26% 67.9 (302.2) Max Shear 0.002 (0.05)
Flexure 0.0079 (0.2)
Any 0.0079 (0.2)
DS1a 0.26% 67.9 (302.2) Max Shear 0.002 (0.05)
Flexure 0.0079 (0.2)
Any 0.0079 (0.2)
DS2a 1.02% 100.9 (448.9) Max Shear 0.031 (0.8)
Flexure 0.043 (1.1)
Any 0.043 (1.1)
DSr9 1.36% 101.9 (453.2) Max Shear 0.043 (1.1)
Flexure 0.071 (1.8)
Any 0.071 (1.8)
DS3a 1.70% 101.9 (453.2) Max Shear 0.071 (1.8)
Flexure 0.1 (2.6)
Any 0.1 (2.6)
DS4b 1.70% 101.9 (453.2) Max Shear 0.071 (1.8)
Flexure 0.1 (2.6)
Any 0.1 (2.6)

DazioWSH3
150

100

50
Vbase, kips

0
Env.
DS1a
DS1d
−50 DS2a
DS2e
DS3a
DS4b
DS4g
−100
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
% Drift

Figure C.42: Envelope for DazioWSH3


648

C.3.18 DazioWSH4

Table C.28: DazioWSH4 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.25% 68.9 (306.5) First cracking (from data
downloaded for project)
DS1d 0.31% 73.5 (327.1) As determined from
momentcurvature analysis
2 DS2a -1.01% -97.0 (-431.7) Initial spalling, but long. reinf. not
seen (initial buckling also occurs)
3 DS3a 1.01% 100.0 (444.6) Cover continued to spall, revealing
long. reinf. for first time
4 DS4b -1.01% -97.0 (-431.7) Initial buckling seen at same time as
initial spalling
DS4d 1.56% 90.2 (401.4) Failure due to crushing of core con-
crete
DSr1 0.25% 69.1 (307.5) First yield as reported in paper Ta-
ble 4 (∆ y b )
DSr9 1.01% 100.0 (444.6) Cracks exceeding or equal to 1/16”
(determined from download crack
data)
649

Table C.29: DazioWSH4 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS1a 0.25% 68.9 (306.5) Max Shear 0.0059 (0.15)
Flexure 0.012 (0.3)
Any 0.012 (0.3)
DSr1 0.67% 98.0 (436.0) Max Shear 0.02 (0.5)
Flexure 0.047 (1.2)
Any 0.047 (1.2)
DSr9 1.01% 100.0 (444.6) Max Shear 0.035 (0.9)
Flexure 0.073 (1.9)
Any 0.073 (1.9)
DS2a -1.01% -97.0 (-431.7) Max Shear 0.043 (1.1)
Flexure 0.087 (2.2)
Any 0.087 (2.2)
DS3a 1.01% 100.0 (444.6) Max Shear 0.035 (0.9)
Flexure 0.073 (1.9)
Any 0.073 (1.9)
DS4b -1.01% -97.0 (-431.7) Max Shear 0.043 (1.1)
Flexure 0.087 (2.2)
Any 0.087 (2.2)

DazioWSH4
100

80

60

40

20
Vbase, kips

−20
Env.
−40 DS1a
DS1d
−60 DS2e
DS2a
DS3a
−80 DS4b
DS4d
−100
−1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.43: Envelope for DazioWSH4


650

C.3.19 DazioWSH5

Table C.30: DazioWSH5 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.10% 66.0 (293.5) First cracking (from data
downloaded for project)
DS1d 0.19% 77.3 (344.0) As determined from
momentcurvature analysis
2 DS2a 0.55% 95.1 (423.0) Initial spalling
3 DS3a 1.08% 94.1 (418.7) Cover continued to spall, revealing
long. reinf. for first time (B.E bars
buckled and web bars fractured)
4 DS4b 1.08% 94.1 (418.7) Initial buckling of corner bars
(spalling reveals reinf. for first time
and web bars rupture)
DS4e 1.08% 94.1 (418.7) Rupture of web bars (spalling re-
veals reinf. for first time and corner
bars have initial buckling)
DS4g 1.54% 78.6 (349.6) Failure due to rupture of bar (other
had previously buckled/fractured)
DSr1 0.17% 74.7 (332.5) First yield as reported in paper Ta-
ble 4 (∆ y b )
DSr9 0.55% 95.1 (423.0) Cracks exceeding or equal to 1/16”
(determined from download crack
data)
651

Table C.31: DazioWSH5 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS1a 0.10% 66.0 (293.5) Max Shear 0.002 (0.05)
Flexure 0.0059 (0.15)
Any 0.0059 (0.15)
DSr1 0.27% 88.3 (392.8) Max Shear 0.014 (0.35)
Flexure 0.016 (0.4)
Any 0.016 (0.4)
DSr9 0.55% 95.1 (423.0) Max Shear 0.051 (1.3)
Flexure 0.067 (1.7)
Any 0.067 (1.7)
DS2a 0.55% 95.1 (423.0) Max Shear 0.051 (1.3)
Flexure 0.067 (1.7)
Any 0.067 (1.7)
DS3a 1.08% 94.1 (418.7) Max Shear 0.14 (3.5)
DS4b 1.08% 94.1 (418.7) Max Shear 0.14 (3.5)
DS4e 1.08% 94.1 (418.7) Max Shear 0.14 (3.5)

DazioWSH5
100

80

60

40

20
Vbase, kips

−20 Env.
DS1a
−40 DS1d
DS2e
−60 DS2a
DS3a
DS4b
−80 DS4e
DS4g
−100
−1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.44: Envelope for DazioWSH5


652

C.3.20 DazioWSH6

Table C.32: DazioWSH6 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.22% 90.2 (401.4) First cracking (from data
downloaded for project)
DS1d -0.32% -103.4 (-459.9) As determined from
momentcurvature analysis
2 DS2a 0.57% 125.2 (556.8) Initial spalling
3 DS3a 0.84% 131.0 (582.7) Spalling revealing corner bars but
no bar buckling
4 DS4b -1.73% -132.0 (-587.1) Initial buckling of corner bars
(spalling reveals reinf. for first time
and web bars rupture)
DS4d 1.42% 132.0 (587.1) Failure due to crushing of core con-
crete (occurred in negative direction
after positive direction reached ac-
tuator displacement limit)
DSr1 0.22% 88.3 (392.7) First yield as reported in paper Ta-
ble 4 (∆ y b )
DSr9 -0.87% -132.9 (-591.4) Cracks exceeding or equal to 1/16”
(determined from download crack
data)

Table C.33: DazioWSH6 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS2a 0.57% 125.2 (556.8) Max Flexure 0.024 (0.6)
DS3a 0.84% 131.0 (582.7) Max Flexure 0.045 (1.1)
DSr9 -0.87% -132.9 (-591.4) Max Shear 0.043 (1.1)
Flexure 0.063 (1.6)
Any 0.063 (1.6)
DS4b -1.73% -132.0 (-587.1) Max Shear 0.079 (2)
653

DazioWSH6
150

100

50
Vbase, kips

−50 Env.
DS1a
DS1d
DS2a
−100 DS3a
DS2e
DS4d
DS4b
−150
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
% Drift

Figure C.45: Envelope for DazioWSH6


654

C.3.21 LefasSW21

Table C.34: LefasSW21 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.02% 2.2 (10.0) First flexural cracking reported
DS1c 0.45% 18.0 (80.0) First shear cracking reported
DS1d 0.38% 15.9 (70.5) As determined from
momentcurvature analysis
2 DS2a 0.92% 25.7 (114.3) Spalling at 90% of ultimate horizon-
tal load (accompanied by large ver-
tical crack in compressive B.E.)
4 DS4c 1.59% 28.6 (127.0) Compressive failure at ultimate load
due to vertical cracking in compres-
sion zone
DSr1 0.45% 18.0 (80.0) First flexural steel yielding reproted

Table C.35: LefasSW21 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS4c 1.59% 28.6 (127.0) Max Flexure 0.059 (1.5)
655

LefasSW21
30

25

20
Vbase, kips

15

10

Env.
DS1b
5 DS1d
DS1c
DS2a
DS4c
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
% Drift

Figure C.46: Envelope for LefasSW21


656

C.3.22 LefasSW22

Table C.36: LefasSW22 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.03% 3.1 (14.0) First flexural cracking reported
DS1c 0.38% 24.7 (110.0) First shear cracking reported
DS1d 0.31% 21.8 (96.9) As determined from
momentcurvature analysis
4 DS4c 1.18% 33.7 (150.0) Compressive failure at ultimate load
due to vertical cracking in compres-
sion zone
DSr1 0.38% 22.5 (100.0) First flexural steel yielding reported
DSr5 0.59% 30.3 (135.0) Inclined crack reaches within
200mm of wall edge (650mm wall
length)

LefasSW22
35

30

25

20
Vbase, kips

15

10
Env.
DS1b
5 DS1d
DS1c
DS4c
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
% Drift

Figure C.47: Envelope for LefasSW22


657

C.3.23 LefasSW23

Table C.37: LefasSW23 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 4.5 (20.0) First flexural cracking reported
DS1c 0.40% 27.0 (120.0) First shear cracking reported
DS1d 0.41% 26.7 (118.6) As determined from
momentcurvature analysis
4 DS4c 1.01% 40.5 (180.0) Compressive failure at ultimate load
due to inclined crack becoming a
vertical crack in compression zone,
followed by immediate cruching
DSr1 0.40% 27.0 (120.0) First flexural steel yielding

LefasSW23
45

40

35

30
Vbase, kips

25

20

15

10 Env.
DS1b
DS1c
5
DS1d
DS4c
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
% Drift

Figure C.48: Envelope for LefasSW23


658

C.3.24 LefasSW24

Table C.38: LefasSW24 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.02% 2.2 (10.0) First flexural cracking reported
DS1c 0.48% 18.0 (80.0) First shear cracking reported
DS1d 0.41% 15.7 (70.0) As determined from
momentcurvature analysis
2 DS2a 0.94% 24.3 (108.0) Spalling at 90% of ultimate horizon-
tal load (accompanied by large ver-
tical crack in compressive B.E.)
4 DS4c 1.39% 27.0 (120.0) Compressive failure at ultimate load
due to vertical cracking in compres-
sion zone
DSr1 0.48% 18.0 (80.0) First flexural steel yielding reported

Table C.39: LefasSW24 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS4c 1.39% 27.0 (120.0) Max Flexure 0.059 (1.5)
659

LefasSW24
30

25

20
Vbase, kips

15

10

Env.
DS1b
5 DS1d
DS1c
DS2a
DS4c
0
0 0.5 1 1.5
% Drift

Figure C.49: Envelope for LefasSW24


660

C.3.25 LefasSW25

Table C.40: LefasSW25 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 5.6 (25.0) First flexural cracking reported
DS1c 0.45% 29.2 (130.0) First shear cracking reported
DS1d 0.38% 26.0 (115.6) As determined from
momentcurvature analysis
DSr1 0.45% 29.2 (130.0) First flexural steel yielding reported

LefasSW25
35

30

25

20
Vbase, kips

15

10

Env.
5 DS1b
DS1d
DS1c
0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8
% Drift

Figure C.50: Envelope for LefasSW25


661

C.3.26 LefasSW26

Table C.41: LefasSW26 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.03% 2.2 (10.0) First flexural cracking reported
DS1c 0.42% 15.3 (68.0) First shear cracking reported
DS1d 0.44% 15.1 (67.3) As determined from
momentcurvature analysis
2 DS2a 1.02% 24.9 (110.7) Spalling at 90% of ultimate horizon-
tal load (accompanied by large ver-
tical crack in compressive B.E.)
4 DS4c 1.61% 27.7 (123.0) Compressive failure at ultimate load
due to vertical cracking in compres-
sion zone
DSr1 0.42% 15.3 (68.0) First flexural steel yielding reported

Table C.42: LefasSW26 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS4c 1.61% 27.7 (123.0) Max Flexure 0.059 (1.5)
662

LefasSW26
30

25

20
Vbase, kips

15

10

Env.
DS1b
5 DS1c
DS1d
DS2a
DS4c
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
% Drift

Figure C.51: Envelope for LefasSW26


663

C.3.27 LefasSW30

Table C.43: LefasSW30 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.06% 4.0 (17.7) First flexural cracking occurred at
15% of maximum load (p. 722)
DS1c 0.06% 4.0 (17.7) First shear cracking occurred at 60%
of maximum load
DS1d 0.38% 13.8 (61.2) As determined from
momentcurvature analysis
4 DS4c 1.61% 26.5 (117.7) Spalling and vertical cracks in com-
pressive b.e. ultimately led to loss
of lateral load carrying capacity (p.
722-723)

Table C.44: LefasSW30 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS4c 1.61% 26.5 (117.7) Max Flexure 0.047 (1.2)
664

LefasSW30
30

25

20

15
Vbase, kips

10

5
Env.
DS1b
0 DS1c
DS1d
DS4c
−5
−0.2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
% Drift

Figure C.52: Envelope for LefasSW30


665

C.3.28 LefasSW31

Table C.45: LefasSW31 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 4.5 (20.0) First flexural cracking (Table 8.3)
DS1c 0.32% 14.6 (64.9) First shear cracking (Table 8.3)
DS1d 0.30% 14.0 (62.5) As determined from
momentcurvature analysis
4 DS4c 1.71% 26.0 (115.8) Spalling and vertical cracks in com-
pressive b.e. ultimately led to loss
of lateral load carrying capacity (p.
216)
DSr1 0.32% 14.6 (64.9) First yield reported (Table 8.3)

Table C.46: LefasSW31 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS4c 1.71% 26.0 (115.8) Max Flexure 0.016 (0.4)
666

LefasSW31
30

25

20

15
Vbase, kips

10

−5 Env.
DS1b
DS1d
−10
DS1c
DS4c
−15
−0.5 0 0.5 1 1.5 2
% Drift

Figure C.53: Envelope for LefasSW31


667

C.3.29 LefasSW32

Table C.47: LefasSW32 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 4.5 (20.0) First flexural cracking (Table 8.3)
DS1c 0.34% 14.6 (64.9) First shear cracking (Table 8.3)
DS1d 0.32% 14.4 (64.3) As determined from
momentcurvature analysis
4 DS4c 1.89% 25.0 (111.0) Spalling and vertical cracks in com-
pressive b.e. ultimately led to loss
of lateral load carrying capacity (p.
216)
DSr1 0.34% 14.6 (64.9) First yield reported (Table 8.3)

Table C.48: LefasSW32 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS4c 1.89% 25.0 (111.0) Max Flexure 0.031 (0.8)
668

LefasSW32
25

20

15

10
Vbase, kips

−5

−10 Env.
DS1b
DS1d
−15
DS1c
DS4c
−20
−1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.54: Envelope for LefasSW32


669

C.3.30 LefasSW33

Table C.49: LefasSW33 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 3.7 (16.4) First flexural cracking (Table 8.3)
DS1c 0.44% 16.1 (71.5) First shear cracking (Table 8.3)
DS1d 0.44% 16.1 (71.5) Experimental reporting of yield
used in place of theoretical
calculation; First yield reported
(Table 8.3)
4 DS4c 1.92% 25.1 (111.5) Spalling and vertical cracks in com-
pressive b.e. ultimately led to loss
of lateral load carrying capacity (p.
216)
DSr1 0.44% 16.1 (71.5) First yield reported (Table 8.3)

Table C.50: LefasSW33 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS4c 1.92% 25.1 (111.5) Max Flexure 0.031 (0.8)
670

LefasSW33
30

25

20

15

10
Vbase, kips

−5

Env.
−10
DS1b
DS1c
−15 DS1d
DS4c
−20
−1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.55: Envelope for LefasSW33


671

C.3.31 OestR1

Table C.51: OestR1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 12.3 (54.8) Initial flexural cracking; data taken
from force-disp. figure of initial cy-
cles
DS1d -0.15% -17.2 (-76.3) As determined from
momentcurvature analysis
2 DS2a -1.09% -27.1 (-120.7) Minor spalling and flaking along
cracks observed in cycle 14 (data for
cycle 13 used)
4 DS4b -2.20% -27.2 (-121.0) First buckling of main flex. reinf.:
outer 2 bars in compression region
buckled 15 inches above base (cycle
20 reported, info from cycle 19 on
envelope is used)
DS4e -2.75% -20.0 (-89.0) First bar fracture of bars that were
first to buckle (cycle 26 is reported,
info from cycle 25 on envelope is
used)
DSr1 0.24% 20.1 (89.4) Reported tensile yield

Table C.52: OestR1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 -0.57% -25.3 (-112.7) Max Any 0.018 (0.46)
DS4b -2.20% -27.2 (-121.0) Max Any 0.05 (1.3)
C1 -1.66% -27.2 (-121.2) Max Flexure 0.2 (5.1) 0.02 (0.51)
672

OestR1
30

20

10
Vbase, kips

−10

Env.
DS1b
−20 DS1d
DS2a
DS4b
DS4e
−30
−3 −2 −1 0 1 2 3
% Drift

Figure C.56: Envelope for OestR1


673

C.3.32 OestR2

Table C.53: OestR2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.06% 13.6 (60.5) Initial flexural cracking; data taken
from force-disp of initial cycles;
cracking reported as occurring at 15
kips
DS1d 0.30% 32.6 (144.8) As determined from
momentcurvature analysis
2 DS2a -0.56% -42.9 (-191.0) Minor spalling and flaking along
cracks observed in cycle 19 along
horizontal web crack in lower 3ft
4 DS4e -3.33% -29.3 (-130.2) Several bars fractuctured during cy-
cle 37
DSr1 0.33% 37.0 (164.6) Reported tensile yield
DSr2 -1.67% -49.4 (-219.9) Residual cracks of 0.003 in

Table C.54: OestR2 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 -0.56% -42.9 (-191.0) Max Shear 0.019 (0.48)
Flexure 0.012 (0.3)
DSr2 -1.67% -49.4 (-219.9) Max Any 0.003 (0.076)
674

OestR2
60

40

20
Vbase, kips

−20

Env.
−40 DS1b
DS1d
DS2a
DS4e
−60
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.57: Envelope for OestR2


675

C.3.33 OestB1

Table C.55: OestB1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 28.6 (127.4) Initial flexural cracks; data taken
from force-disp of initial cycles
DS1d -0.17% -42.1 (-187.1) As determined from
momentcurvature analysis
2 DS2a 0.57% 54.1 (240.7) Minor spalling and flaking along
cracks observed in cycle 14 along
web cracks (cycle 13 used for enve-
lope data)
4 DS4b -1.67% -65.0 (-289.0) Corner bar near outer face of b.e. Fig. C.59
bowed out between ties
DS4e 2.79% 56.6 (252.0) First bar to fracture was first bar to
buckle
DSr1 -0.27% -45.1 (-200.6) Reported tensile yield (only nega-
tive direction for cycle 10 consid-
ered; yield occurred in positive di-
rection in cycle 13)
DSr2 1.11% 60.5 (269.0) Residual cracks of 0.005 in

Table C.56: OestB1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 -0.25% -49.9 (-222.2) Max Shear 0.014 (0.36)
Flexure 0.009 (0.23)
DSr2 1.11% 60.5 (269.0) Max Any 0.005 (0.13)
676

OestB1
80

60

40

20
Vbase, kips

−20

−40 Env.
DS1b
DS1d
−60 DS2a
DS4b
DS4e
−80
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.58: Envelope for OestB1

Figure C.59: OestB1: DS4b - Buckling of reinforcing bars


677

C.3.34 OestB2

Table C.57: OestB2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 31.1 (138.4) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d -0.38% -112.0 (-498.4) As determined from
momentcurvature analysis
2 DS2a 0.54% 121.0 (538.0) Minor spalling and flaking along
cracks observed in cycle 14 along di-
agonal cracks (cycle 13 used for en-
velope data)
3 DS3a 2.23% 153.1 (681.2) Considerable flaking and spalling in
web during cycles 25 thru 27 (cycle
25 data used)
4 DS4b 2.23% 153.1 (681.2) Two bars buckled in lower 1 ft of
wall in compression column under
positive load () cycle 26
DS4k -2.71% -151.3 (-673.0) Web failure occurred on way to -5
inch displacement in cycle 28
DSr1 0.47% 119.7 (532.5) Reported tensile yield (only nega-
tive direction for cycle 10 consid-
ered; yield occurred in positive di-
rection in cycle 13)

Table C.58: OestB2 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.54% 121.0 (538.0) Max Shear 0.017 (0.43)
Flexure 0.005 (0.13)
C1 -1.12% -149.3 (-664.1) Max Flexure 0.003 (0.076)
678

OestB2
200

150

100

50
Vbase, kips

−50

Env.
−100 DS1b
DS1d
DS2a
−150 DS4b
DS3a
DS4k
−200
−3 −2 −1 0 1 2 3
% Drift

Figure C.60: Envelope for OestB2


679

C.3.35 OestB3

Table C.59: OestB3 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 29.5 (131.2) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d -0.23% -41.0 (-182.6) As determined from
momentcurvature analysis
4 DS4b -3.89% -62.2 (-276.9) First buckling in cycle 38
DS4e 3.92% 57.6 (256.2) Vertical bar fractured at base (cover
still present)
DSr1 0.31% 45.2 (201.1) Reported tensile yield

Table C.60: OestB3 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.31% 45.6 (202.7) Max Shear 0.025 (0.64)
Flexure 0.012 (0.3)
680

OestB3
80

60

40

20
Vbase, kips

−20

−40
Env.
DS1b
−60 DS1d
DS4b
DS4e
−80
−5 −4 −3 −2 −1 0 1 2 3 4 5
% Drift

Figure C.61: Envelope for OestB3


681

C.3.36 OestB4

Table C.61: OestB4 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 26.4 (117.6) First cracking between steps 3 and
4 (called step 3.5 in envelope)
DS1d 0.23% 40.9 (182.1) As determined from
momentcurvature analysis
2 DS2a 1.11% 64.7 (287.7) Crushing of outer shell of compres-
sion face at load step 9
4 DS4e 4.97% 78.6 (349.8) Bar fracture of longitudinal bar in
web at max load (point named 15.5)
DS4g 7.40% 15.6 (69.2) All column bars (at diagonal crack)
and 4 additional web bars fractured,
causing drop in strength
DSr1 0.28% 45.2 (201.2) First yield at load step 7
DSr13 6.97% 70.7 (314.6) Largest shear crack 1.0”

Table C.62: OestB4 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.28% 45.2 (201.2) Max Shear 0.02 (0.51)
Flexure 0.01 (0.25)
DS4e 4.97% 78.6 (349.8) Max Shear 0.44 (11)
Flexure 0.1 (2.5)
DSr13 6.97% 70.7 (314.6) Max Shear 1 (25)
DS4g 7.40% 15.6 (69.2) Max Shear 0.44 (11)
Flexure 0.1 (2.5)
682

OestB4
80

60

40

20
Vbase, kips

−20

−40 Env.
DS1b
DS1d
−60 DS2a
DS4e
DS4g
−80
−4 −2 0 2 4 6 8
% Drift

Figure C.62: Envelope for OestB4


683

C.3.37 OestB5

Table C.63: OestB5 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.03% -30.6 (-135.9) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d -0.42% -117.1 (-520.7) As determined from
momentcurvature analysis
2 DS2a 0.53% 125.8 (559.5) Minor spalling and flaking along di-
agonal cracks observed in cycle 16
along web cracks (cycle 13 used be-
cause this was maximum historic
drift)
3 DS3a -1.67% -165.6 (-736.7) Significant spalling and crushing
along right half of construction joint
at 3 ft level
4 DS4k -2.62% -165.1 (-734.3) Web crushing in cycle 29 in negative
direction
DSr1 0.47% 112.3 (499.5) Reported tensile yield

Table C.64: OestB5 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.53% 125.8 (559.5) Max Shear 0.025 (0.64)
Flexure 0.007 (0.18)
DS3a -1.67% -165.6 (-736.7) Max Flexure 0.005 (0.13)
684

OestB5
200

150

100

50
Vbase, kips

−50

−100 Env.
DS1b
DS1d
−150 DS2a
DS3a
DS4k
−200
−3 −2 −1 0 1 2 3
% Drift

Figure C.63: Envelope for OestB5


685

C.3.38 OestB6

Table C.65: OestB6 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.11% -59.6 (-265.2) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d -0.54% -145.1 (-645.4) As determined from
momentcurvature analysis
2 DS2a 1.11% 186.1 (827.8) Crushing of concrete cover (cycle
22); Spalling & flaking along diag-
onal cracks (cycle 23)
DS2d 0.60% 145.7 (648.1) Splitting of concrete cover in outer
compression face
4 DS4k 1.11% 186.1 (827.8) Initial indication of crushing of com-
pression strut in right side of web 18
in above base in cycle 24 (use cycle
22)
DS4m 1.41% 166.4 (740.4) Several compression struts crushed
simultaneously
DSr1 0.54% 144.2 (641.4) Reported tensile yield

Table C.66: OestB6 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.60% 145.7 (648.1) Max Shear 0.014 (0.36)
Flexure 0.005 (0.13)
686

OestB6
200

150

100

50
Vbase, kips

−50

Env.
−100 DS1b
DS1d
DS2d
−150 DS2a
DS4k
DS4m
−200
−2 −1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.64: Envelope for OestB6


687

C.3.39 OestB7

Table C.67: OestB7 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.15% 98.5 (438.0) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d 0.58% 165.7 (736.9) As determined from
momentcurvature analysis
2 DS2a 0.28% 123.4 (549.0) Spalling & flaking along diagonal
cracks (cycle 10); Crushing of con-
crete cover (cycle 19)
4 DS4k 2.19% 214.9 (955.9) Initial indication of crushing of com-
pression strut in right and left side
of web
DS4m 2.71% 192.2 (854.8) Several compression struts crushed
simultaneously
DSr1 0.54% 161.2 (717.1) Reported tensile yield

Table C.68: OestB7 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.63% 174.2 (775.1) Max Shear 0.017 (0.43)
Flexure 0.007 (0.18)
688

OestB7
250

200

150

100

50
Vbase, kips

−50

−100
Env.
DS1b
−150
DS2a
DS1d
−200 DS4k
DS4m
−250
−4 −3 −2 −1 0 1 2 3
% Drift

Figure C.65: Envelope for OestB7


689

C.3.40 OestB8

Table C.69: OestB8 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.23% -85.0 (-378.1) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d -0.52% -162.1 (-721.1) As determined from
momentcurvature analysis
2 DS2a -0.53% -163.9 (-729.2) Spalling & flaking along diagonal
cracks (cycle 20); Crushing of con-
crete cover (cycle 14; use 13)
4 DS4k 2.17% 211.2 (939.6) Initial indication of crushing on
both ends of horizontal crack formed
along length of web 9 inches above
the base (cycle 26; use 25)
DS4m -3.21% -235.5 (-1047.6) Initial indication of crushing on
both ends of horizontal crack formed
along length of web 9 inches above
the base (cycle 26; use 25)
DSr1 0.49% 155.4 (691.3) Reported tensile yield
DSr9 -2.79% -238.6 (-1061.5) Diagonal cracks of 0.086 inches fol-
lowing cycle 28 displacement level

Table C.70: OestB8 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.53% 158.0 (703.0) Max Shear 0.011 (0.28)
Flexure 0.01 (0.25)
DSr9 -2.79% -238.6 (-1061.5) Max Shear 0.086 (2.2)
690

OestB8
250

200

150

100

50
Vbase, kips

−50

−100 Env.
DS1b
DS1d
−150
DS2a
DS4k
−200 DS2e
DS4m
−250
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.66: Envelope for OestB8


691

C.3.41 OestB9

Table C.71: OestB9 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.16% 75.0 (333.6) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d 0.45% 156.1 (694.4) As determined from
momentcurvature analysis
2 DS2a 1.24% 209.8 (933.3) First indication of cover crushing on
outer compression face (cycle 20)
4 DS4m -2.99% -220.5 (-981.0) Crushing of compression structs in
lower left region; immediately
followed by development of failure
plane along diagonal crack (used
maximum displacement of test due
to unique disp. history)
DSr1 0.26% 158.0 (702.8) Reported tensile yield

Table C.72: OestB9 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.45% 157.5 (700.5) Max Shear 0.015 (0.38)
Flexure 0.011 (0.28)
692

OestB9
250

200

150

100

50
Vbase, kips

−50

−100

Env.
−150
DS1b
DS1d
−200 DS2a
DS4m
−250
−4 −3 −2 −1 0 1 2 3
% Drift

Figure C.67: Envelope for OestB9


693

C.3.42 OestB10

Table C.73: OestB10 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.14% 75.0 (333.6) Initial flexural cracking occurred on
way to cycle 2 stage 13 at 75 kips
DS1d 0.27% 102.9 (457.8) As determined from
momentcurvature analysis
2 DS2a 0.63% 147.6 (656.7) Crushing of concrete cover (cycle 2
load stage 15)
DS2b -1.13% -147.0 (-654.0) Spalling of patched concrete (cycle
2 load stage 21)
4 DS4f -3.37% -110.6 (-491.9) Failure ultimately occurred due to
damage of boundary element, in-
cluding core crushing, bar buckling
and bar fracture
DSr1 0.22% 120.0 (533.8) Reported tensile yield

Table C.74: OestB10 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.22% 94.3 (419.4) Max Shear 0.019 (0.48)
Flexure 0.011 (0.28)
C1 0.37% 117.9 (524.4) Max Shear 0.019 (0.48)
Flexure 0.011 (0.28)
C2 2.72% 169.8 (755.4) Max Shear 0.095 (2.4)
Flexure 0.036 (0.91)
694

OestB10
200

150

100

50
Vbase, kips

−50

−100 Env.
DS1b
DS1d
−150 DS2a
DS2b
DS4f
−200
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.68: Envelope for OestB10


695

C.3.43 OestF1

Table C.75: OestF1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.05% 40.1 (178.2) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d 0.56% 147.6 (656.5) As determined from
momentcurvature analysis
2 DS2a -1.11% -173.4 (-771.2) Minor spalling and flaking along
construction joint and diagonal
cracks observed in cycle 20 (cycle 19
used because this was maximum his-
toric drift)
DS2d -1.11% -173.4 (-771.2) Vertical cracking in outer face of
flange in cycle 20, possibly due to
bowing caused by web expansion
(cycle 19 used because this was max-
imum historic drift)
4 DS4k -1.86% -173.5 (-771.8) Steepest diagonal strut in web
crushed and slipped (flange hinged
to allow)
DSr1 0.56% 150.6 (669.9) Reported tensile yield

Table C.76: OestF1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 -0.62% -152.4 (-678.0) Max Shear 0.018 (0.46)
696

OestF1
200

150

100

50
Vbase, kips

−50

−100 Env.
DS1b
DS1d
−150 DS2a
DS2d
DS4k
−200
−2.5 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2 2.5
% Drift

Figure C.69: Envelope for OestF1


697

C.3.44 OestF2

Table C.77: OestF2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.09% 78.7 (350.0) Initial flexural cracking; data taken
from force-disp of initial cycles
DS1d 0.47% 159.9 (711.1) As determined from
momentcurvature analysis
2 DS2a -1.11% -188.3 (-837.4) Minor spalling and flaking along di-
agonal cracks
DS2d -2.21% -204.3 (-908.7) Vertical cracking in lower 2.5 ft of
flange
4 DS4k 2.77% 205.7 (915.1) Steepest diagonal strut in web
crushed and slipped (flange hinged
to allow this)
DS4m -2.63% -202.1 (-899.1) Several struts in unconfined region
crushed at once, immediately fol-
lowed by shearing of compression
flange and fracture of horizontal re-
inforcing bar
DSr1 0.46% 156.0 (693.9) Reported tensile yield

Table C.78: OestF2 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.52% 174.2 (774.8) Max Shear 0.012 (0.3)
Flexure 0.012 (0.3)
698

OestF2
250

200

150

100

50
Vbase, kips

−50

−100 Env.
DS1b
DS1d
−150
DS2a
DS2d
−200 DS4m
DS4k
−250
−3 −2 −1 0 1 2 3 4
% Drift

Figure C.70: Envelope for OestF2


699

C.3.45 Morgan

Table C.79: Morgan damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a -0.11% -12.2 (-54.3) Crack map for first displacement Fig. C.72
level shows cracking occurred
DS1d 0.10% 14.7 (65.3) As determined from
momentcurvature analysis
2 DS2a -1.10% -20.9 (-92.8) Crack map shows minor spalling oc- Fig. C.73
curred on way to steps 108/111; oc-
curred approximately 1/3 way up
first floor
DS2d 0.72% 22.9 (102.1) Crack map shows vertical cracking Fig. C.74
occurred on way to steps 88/91
3 DS3a -1.46% -17.6 (-78.2) Crack map shows exposed Fig. C.75
reinforcement ccurred on way
to steps 128/131
4 DS4c 1.79% 16.0 (71.0) Assummed failure mode. Loss of
lateral load capacity dropped less
than 80% of max
DSr1 0.30% 19.7 (87.7) Initial yield occurred in boundary Fig. C.76
element at 83% of maximum load.
Data taken from positive envelope
since 83% occurs on this in cycle
sooner than on negative envelope
700

Morgan
25

20

15

10

5
Vbase, kips

−5

−10 Env.
DS1d
DS1a
−15
DS2d
DS2a
−20 DS3a
DS4c
−25
−1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.71: Envelope for Morgan

Figure C.72: Morgan: DS1a - crack map after steps 19/22


701

Figure C.73: Morgan: DS2a - crack map after steps 108/111

Figure C.74: Morgan: DS2d - crack map after steps 88/91


702

Figure C.75: Morgan: DS3a - crack map after steps 128/131

Figure C.76: Morgan: DSr1 - crack map after steps 65/68


703

C.3.46 LiuW1

Table C.80: LiuW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.09% -17.8 (-79.4) Hairline crack at 150mm from base Fig. C.78
of wall at location of confining rein-
forcement; cycle 4a/4b (p. 39)
DS1c -0.28% -34.4 (-153.2) Shear crack developed from initial Fig. C.79
bottom flexural crack, inclined 45
degrees towards base; cycle 7a/7b
(p. 39)
DS1d 1.06% 58.1 (258.3) As determined from
momentcurvature analysis
DS1e -1.16% -57.6 (-256.1) Compression yield ; cycle 20a (use Fig. C.80
data from 19a/b) (p. 41)
DS1f -3.02% -55.5 (-246.8) Shear yield; cycle 37a (p. 42)
2 DS2a 1.50% 58.5 (260.3) 50mm above base in boundary re-
gion; Described as crushing, but in-
terpretted to be spalling of cover
and reinforcement not exposted; cy-
cle 25a (p. 40)
DS2c 2.61% 56.9 (253.2) Spalling in web; cycle 34a (p. 40) Fig. C.81
4 DS4b -2.64% -57.3 (-254.7) Longitudinal buckling of bars in Fig. C.82
boundary elements and spalling
revealing confinement hoops; cycle
36b (p. 40)
DS4f 2.98% 57.1 (253.8) Loss of lateral load carrying capac- Fig. C.83
ity due to severe buckling of rein-
forcement and severe spalling and
crushing of the concrete; cycle 40a
(p. 41) but cycle 37a used for
recording damage state b/c larger
drift
DSr1 0.54% 48.4 (215.2) First yield in long. reinf. at base of Fig. C.80
wall; cycle 10a (p. 41)
704

Table C.81: LiuW1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS1c -0.28% -34.4 (-153.2) Max Flexure 0.0039 (0.1)
DSr1 0.54% 48.4 (215.2) Max Shear 0.016 (0.4)
Flexure 0.024 (0.6)
C1 -0.28% -34.4 (-153.2) Max Flexure 0.0039 (0.1)
C2 -0.28% -34.4 (-153.2) Max Flexure 0.0079 (0.2)
C3 -0.58% -48.0 (-213.6) Max Shear 0.016 (0.4)
Flexure 0.024 (0.6)
C4 -0.79% -56.7 (-252.0) Max Shear 0.022 (0.55)
Flexure 0.026 (0.65)

LiuW1
80

60

40

20
Vbase, kips

0 Env.
DS1b
DS1c
−20 DS1d
DS1e
DS2a
DS2c
−40 DS4b
DS4f
DS1f
−60
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.77: Envelope for LiuW1


705

Figure C.78: LiuW1: DS1b - photo after initial cracking

Figure C.79: LiuW1: DS1c - photo after 2.0Mcr cycles


706

Figure C.80: LiuW1: DS1e - cracking of first yield

Figure C.81: LiuW1: DS2c - cracking at first cycle of 3.80 Delta y; spalling in web
707

Figure C.82: LiuW1: DS4b - cracking after final cycle of 3.80 Delta y; longitudinal buckling

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.

Figure C.83: LiuW1: DS4f - photo at failure


708

C.3.47 LiuW2

Table C.82: LiuW2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.09% 22.4 (99.6) Hairline crack at 50mm from base Fig. C.85
of wall at location of confining rein-
forcement; cycle 4a/4b (p. 44)
DS1c -0.40% -44.5 (-197.8) Shear crack developed from initial Fig. C.86
bottom flexural crack, inclined 45
degrees towards base; cycle 7a/7b
(p. 45)
DS1d 1.24% 60.8 (270.6) As determined from
momentcurvature analysis
DS1f 1.95% 62.9 (279.7) Shear yield; cycle 31a (p. 46) Fig. C.89
2 DS2a 1.30% 61.1 (271.8) 150mm above base in boundary re- Fig. C.88
gion; Described as crushing, but in-
terpretted to be spalling of cover
and reinforcement not exposted; cy-
cle 25a (p. 45)
DS2c 1.95% 62.9 (279.7) Spalling in web; cycle 31a (p. 45) Fig. C.89
4 DS4b 2.92% 57.4 (255.4) Longitudinal buckling of bars in Fig. C.90
boundary elements and spalling
revealing confinement hoops; cycle
40a (p. 45)
DS4f 2.97% 46.9 (208.8) Loss of lateral load carrying capac- Fig. C.91
ity due to severe buckling of rein-
forcement and severe spalling and
crushing of the concrete; cycle 40a
(p. 45-46)
DSr1 0.66% 57.9 (257.7) First yield in long. reinf. at base of Fig. C.92
wall; cycle 13a/b (p. 46)
709

Table C.83: LiuW2 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS1c -0.40% -44.5 (-197.8) Max Flexure 0.0079 (0.2)
DSr1 0.66% 57.9 (257.7) Max Shear 0.016 (0.4)
Flexure 0.024 (0.6)
C1 -0.40% -44.5 (-197.8) Max Flexure 0.0079 (0.2)
C2 -0.40% -44.5 (-197.8) Max Flexure 0.014 (0.35)
C3 -0.45% -46.5 (-207.1) Max Shear 0.016 (0.4)
Flexure 0.016 (0.4)
C4 0.66% 57.9 (257.7) Max Shear 0.022 (0.55)
Flexure 0.024 (0.6)

LiuW2
80

60

40

20
Vbase, kips

0
Env.
DS1b
−20 DS1c
DS1d
DS2a
DS2c
−40 DS1f
DS4b
DS4f
−60
−3 −2 −1 0 1 2 3
% Drift

Figure C.84: Envelope for LiuW2


710

Figure C.85: LiuW2: DS1b - photo after initial cracking

Figure C.86: LiuW2: DS1c - photo after 2.0Mcr cycles


711

Figure C.87: LiuW2: DS1f - cracking of 3.38 Deltay

Figure C.88: LiuW2: DS2a - cracking of 2.25*Deltay


712

Figure C.89: LiuW2: DS2c - cracking at first cycle of 3.80 Delta y; spalling in web

Figure C.90: LiuW2: DS4b - cracking after cycle 40a


Figure C.91: LiuW2: DS4f - photo at failure

Figure C.92: LiuW2: DSr1 - cracking of first yield


713

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
714

C.3.48 AliW1

Table C.84: AliW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.26% -16.5 (-73.4) Cycles 1 and 2 flexural cracking in
b.e. lower two stories grew into
shear cracks (p. 32)
DS1c -0.26% -16.5 (-73.4) Cycles 1 and 2 flexural cracking in
b.e. lower two stories grew into
shear cracks (p. 32)
DS1d 0.75% 31.0 (137.7) As determined from
momentcurvature analysis
2 DS2a -1.54% -31.7 (-141.0) Cycle 9: minor crushing in east Fig. C.94
boundary element that spalled of in
next cycle (p. 34-35)
DSr10 -2.57% -35.2 (-156.6) Cycle 13/14: largest crack measur-
ing 0.2 inches (p. 35)
DSr9 -2.06% -33.3 (-148.1) Cycle 11/12: largest crack measur-
ing 0.12 inches (p. 35)

Table C.85: AliW1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr9 -2.06% -33.3 (-148.1) Max Shear 0.12 (3)
DSr10 -2.57% -35.2 (-156.6) Max Shear 0.2 (5.1)
C1 -0.77% -29.3 (-130.3) Max Shear 0.025 (0.64)
C2 -0.77% -29.3 (-130.3) Max Shear 0.04 (1)
C3 -1.03% -31.3 (-139.2) Max Shear 0.06 (1.5)
C4 -2.06% -33.3 (-148.1) Max Shear 0.12 (3)
C5 -2.57% -35.2 (-156.6) Max Shear 0.2 (5.1)
715

AliW1
40

30

20

10
Vbase, kips

−10

−20 Env.
DS1b
DS1c
−30 DS1d
DS2a
DS2e
−40
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.93: Envelope for AliW1

Figure C.94: AliW1: DS2a - photo of cycle 9 damage


716

C.3.49 TupperW3

Table C.86: TupperW3 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.15% -16.4 (-73.0) Small hairline flexural cracks; cycle
2(p. 45)
DS1c 0.76% 58.9 (262.0) cycle 4(p. 45)
DS1d 0.72% 56.9 (253.2) As determined from
momentcurvature analysis
DS1e 0.76% 58.9 (262.0) Compression yield of b.e. reinforce-
ment; cycle 4(p. 45)
2 DS2a 1.60% 75.3 (334.9) Stated as crushing in Fig. C.96
compression zone, interpretted to be
cover spalling; cycle 7(p. 46)
4 DS4f 2.92% 71.3 (317.1) ”it is clear that failure occurred due Fig. C.97
to severe distress in the compression
zone, with concrete crushing, rup-
turing of one of the confining hoops
and local buckling of the longitudi-
nal bars”; cycle 10(p. 46-47)
DSr1 0.72% 56.5 (251.2) Reported tensile yield; value given
in report is said to be estimate; cycle
4(p. 45)

Table C.87: TupperW3 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS2a 1.60% 75.3 (334.9) Max Shear 0.059 (1.5)
C1 1.09% 72.4 (322.1) Max Flexure 0.016 (0.4)
C2 1.99% 74.0 (329.0) Max Flexure 0.059 (1.5)
717

TupperW3
80

60

40

20
Vbase, kips

−20

Env.
−40 DS1b
DS1d
DS1c
−60 DS1e
DS2a
DS4f
−80
−3 −2 −1 0 1 2 3
% Drift

Figure C.95: Envelope for TupperW3

Figure C.96: TupperW3: DS2a - deflection of approximately 1.75 Deltay


718

Figure C.97: TupperW3: DS4f - final damage of wall


719

C.3.50 MobeenW1

Table C.88: MobeenW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.08% 34.2 (152.0) Initial cracking; second cycle at
3mm drift level; 238mm above base
of wall (p. 101)
DS1d -0.51% -74.3 (-330.6) As determined from
momentcurvature analysis
DS1e 0.67% 81.8 (363.7) Reported compression yield; 24mm
displacement level, nearly all ex-
treme vertical bars yielded in ten-
sion and compression (p. 101)
2 DS2a 1.17% 84.1 (374.3) Spalling started at 42mm disp level,
not stated if exposes reinforcement
(p. 102)
4 DS4b 1.85% 85.5 (380.4) Large chunks of concrete spalled off
revealing buckled visible bars at ap-
prox. 125 mm above base (p. 102)
DS4e -2.18% -68.1 (-303.0) Moving towards negative 78mm, bar
fracture in southwest corner caused
drop in lateral load (p. 102)
DSr1 0.35% 59.5 (264.7) Reported tensile yield; 12mm dis-
placement level vertical reinforce-
ment just beyond yield strain (p.
101)
DSr11 -3.85% -67.8 (-301.6) Maximum crack width at end of test Fig. C.99
is 8.5mm (p. 108)
DSr14 -1.51% -82.8 (-368.5) Stirrups first yielded on north side
near base at 54mm disp (p. 102)

Table C.89: MobeenW1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 0.35% 59.5 (264.7) Max Any 0.016 (0.4)
DS1e 0.67% 81.8 (363.7) Max Any 0.024 (0.6)
DS2a 1.17% 84.1 (374.3) Max Any 0.059 (1.5)
DSr11 -3.85% -67.8 (-301.6) Max Any 0.33 (8.5)
720

MobeenW1
100

80

60

40

20
Vbase, kips

−20

−40 Env.
DS1a
DS1d
−60
DS1e
DS2a
−80 DS4b
DS4e
−100
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.98: Envelope for MobeenW1

Figure C.99: MobeenW1: DSr11 - Final damage of wall (close up of boundary element)

Reproduced with permission of the copyright owner. Further reproduction prohibited without permission.
721

C.3.51 RivaW1

Table C.90: RivaW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1d -0.48% -100.5 (-447.2) As determined from
momentcurvature analysis
2 DS2d -2.09% -181.2 (-806.0) Vertical crack close to the ground
floor
4 DS4e 3.14% 150.2 (667.9) Bar fracture of longitudinal web re-
inforcement when unloading from
positive peak of cycle 30 to the nega-
tive peak; fracture was described as
a tensile failure with necking
DSr1 -1.06% -131.7 (-585.8) Yield assummed to occur in cycles
25 due to presentation of results
DSr11 -3.17% -173.6 (-772.4) 10mm wide crack in boundary ele-
ment

Table C.91: RivaW1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DSr1 -1.06% -131.7 (-585.8) Max Any 0.031 (0.8)
DSr11 -3.17% -173.6 (-772.4) Max Shear 2 (50)
Flexure 0.39 (10)
722

RivaW1
200

150

100

50
Vbase, kips

−50

−100

Env.
−150 DS1d
DS2d
DS4e
−200
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.100: Envelope for RivaW1


723

C.3.52 ShiuC1

Table C.92: ShiuC1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.14% 15.1 (67.2) Initial cracking occurred at 15.1 kips Fig. C.102
(p. B4)
DS1d 0.52% 47.4 (211.0) As determined from
momentcurvature analysis
4 DS4b 2.86% 74.8 (332.5) Buckling of flexural reinforcement in Fig. C.103
spalled regions at base of wall at end
of test( (p. B6)
DS4i 2.86% 74.8 (332.5) Loss of lateral load carrying capac- Fig. C.104
ity due to sliding shear failure (se-
vere distress in boundary element at
main horizontal crack) (p. 23); ex-
act point not included because oc-
curs at lower drift than max historic
DSr1 0.97% 62.4 (277.6) Initial yielding occurred at 15.1 kips
(p. B4)

ShiuC1
80

60

40

20
Vbase, kips

−20

−40
Env.
DS1a
−60 DS1d
DS4i
DS4b
−80
−3 −2 −1 0 1 2 3
% Drift

Figure C.101: Envelope for ShiuC1


724

Figure C.102: ShiuC1: DS1a - crack pattern following first phase of testing

Figure C.103: ShiuC1: DS4b - final damage of wall


725

Figure C.104: ShiuC1: DS4i - final damage of wall


726

C.3.53 KhaC1

Table C.93: KhaC1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.10% 6.7 (30.0) Tension cracks at bottom and mid-
height of wall
DS1c 1.48% 67.7 (301.0) Pairs of diagonal cracks formed
DS1d NaN% NaN (NaN) As determined from
momentcurvature analysis
4 DS4i -2.57% -67.6 (-300.8) Shear failure occurred at 363 kN,
but recorded as max neg. peak
as this is the largest drift experi-
enced; failure accompanied by large
diagonal cracks from corner to cor-
ner, crushing of concrete at toe, and
buckling vertical steel (no yielding)

KhaC1
100

80

60

40
Vbase, kips

20

−20

−40 Env.
DS1b
DS1c
−60
DS4i
DS1d
−80
−3 −2 −1 0 1 2 3 4 5 6
% Drift

Figure C.105: Envelope for KhaC1


727

C.3.54 ElnCW2

Table C.94: ElnCW2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.06% 10.0 (44.7) Tension cracks developed near base
of wall
DS1d NaN% NaN (NaN) As determined from
momentcurvature analysis
4 DS4o 0.21% 22.1 (98.2) Premature failure of wall due to
bond slip of lap splice

ElnCW2
25

20

15

10
Vbase, kips

−5

Env.
−10 DS1b
DS4o
DS1d
−15
−0.05 0 0.05 0.1 0.15 0.2 0.25 0.3
% Drift

Figure C.106: Envelope for ElnCW2


728

C.3.55 ElnCW3

Table C.95: ElnCW3 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 15.1 (67.0) Horizontal cracks observed at base
of wall
DS1c 0.10% 34.8 (155.0) Diagonal cracks observed in both di-
rections
DS1d NaN% NaN (NaN) As determined from
momentcurvature analysis
4 DS4o 0.30% 44.5 (198.0) Wall failed due to bond slip of lap
splice
DSr5 0.24% 43.8 (195.0) Diagonal cracks extended from cor-
ner to corner in both diagonal direc-
tions

ElnCW3
50

40

30

20

10
Vbase, kips

−10

−20

Env.
−30
DS1b
DS1c
−40 DS4o
DS1d
−50
−0.3 −0.2 −0.1 0 0.1 0.2 0.3 0.4 0.5
% Drift

Figure C.107: Envelope for ElnCW3


729

C.3.56 IleX

Table C.96: IleX damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1a 0.28% 150.1 (667.7) First cracking
DS1d 0.43% 169.7 (754.9) As determined from
momentcurvature analysis
4 DS4b 2.05% 205.5 (914.1) Buckling at ends of flanges
DS4g -3.08% -189.6 (-843.4) Rupture of previously buckled bars
(some stirrups also ruptured)

IleX
250

200

150

100

50
Vbase, kips

−50

−100

Env.
−150
DS1a
DS1d
−200 DS4b
DS4g
−250
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.108: Envelope for IleX


730

C.3.57 IleY

Table C.97: IleY damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.56% -156.2 (-694.6) Horizontal web cracks
DS1c 0.56% 137.8 (612.8) First inclined cracking at base of
flanges
DS1d -0.23% -123.2 (-548.2) As determined from
momentcurvature analysis
4 DS4b -1.02% -170.9 (-760.3) Steel bars at edges of flanges tended
to buckle
DS4f -3.07% -142.4 (-633.4) Drop in lateral load carrying capac-
ity dictated by bar buckling at base
of one flange extremity where one
stirrup was missing (pos direction);
bars buckle on corner of u-shape
wall web (neg direction)
DSr1 0.36% 118.3 (526.3) Initial yielding at 1.3cm

IleY
200

150

100

50
Vbase, kips

−50

−100 Env.
DS1d
DS1c
−150 DS1b
DS4b
DS4f
−200
−4 −3 −2 −1 0 1 2 3 4
% Drift

Figure C.109: Envelope for IleY


731

C.3.58 IleXY

Table C.98: IleXY damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1d 0.58% -170.2 (-757.0) As determined from
momentcurvature analysis
4 DS4e 2.03% -101.8 (-452.6) During final ’fly’ of 8cm displace- Fig. C.111
ment cycle, previously buckled bars
fractured when straightening out; a
shear failure of the flange was said
to follow

IleXY: strong IleXY: weak


200 150

150
100

100

50
50
Vbase, kips

Vbase, kips

0 0

−50
−50

−100

−100
−150
Env. DS4e
DS4e DS1d
−200 −150
−2 −1 0 1 2 3 −3 −2 −1 0 1 2
% Drift % Drift

Figure C.110: Envelope for IleXY


732

Figure C.111: IleXY: DS4e - Shear failure of flange


733

C.3.59 TUA

Table C.99: TUA damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1d 0.49% -96.8 (-430.5) As determined from
momentcurvature analysis
2 DS2a 1.17% 87.7 (390.1) Limited spalling following cycles at
ductility=3
4 DS4b 1.89% -104.3 (-463.8) Bar buckling first during diagonal
cycle to ductility of -6
DS4e 2.51% 99.9 (444.2) Bar fracture of D6 bars in west
flange at ductility of 8 in positive
strong axis bending (other cycles en-
tered at ductility of 6)
DS4k 3.10% 77.7 (345.7) Unconfined concrete in web and
flanges ‘decomposed’

TUA: strong TUA: weak


150 100

100 50
50
Vbase, kips

Vbase, kips

0
0
−50
−50
Env. −100
−100
DS2a DS4b
−150 −150
−4 −2 0 2 4 −4 −2 0 2 4
% Drift % Drift
TUA: diagStrong TUA: diagWeak
100 100

50 50
Vbase, kips

Vbase, kips

0 0

−50 DS4e −50


DS4k DS1d
−100 −100
−2 −1 0 1 2 −2 −1 0 1 2
% Drift % Drift

Figure C.112: Envelope for TUA


734

C.3.60 TUB

Table C.100: TUB damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1d 0.66% -94.4 (-419.8) As determined from
momentcurvature analysis
2 DS2a 0.91% 78.9 (350.9) Initial spalling (no exposed rebar) at
ductility=2
3 DS3a 1.80% 81.1 (360.7) Spalling revealing longitudinal rein-
forcement at ductility=4 (NS cycles)
4 DS4b 2.72% 83.3 (370.6) Buckling during diagonal cycle to
position E; two D12 bars in flange
end buckled
DS4m 2.72% 83.3 (370.6) Crushing of diagonal compression
strut in web during sweeping cycle

TUB: strong TUB: weak


150 100

100
50
50
Vbase, kips

Vbase, kips

0 0

−50
Env. −50
−100
DS2a DS3a
−150 −100
−4 −2 0 2 4 −4 −2 0 2 4
% Drift % Drift
TUB: diagStrong TUB: diagWeak
100 100

50 50
Vbase, kips

Vbase, kips

0 0

−50 DS4b −50


DS4m DS1d
−100 −100
−2 −1 0 1 2 −2 −1 0 1 2
% Drift % Drift

Figure C.113: Envelope for TUB


735

C.3.61 NTW1

Table C.101: NTW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.10% 37.4 (166.4) Flex. cracks in web tip when flange
in compression (flex. cracks in
flange begin at next drift level and
thus not entered)
DS1c 0.14% -67.7 (-301.1) Shear cracks in web under FC/FT
loading
DS1d 1.93% -72.5 (-322.5) As determined from
momentcurvature analysis
2 DS2a 1.63% -224.3 (-997.9) Spalling of cover at base of web tip
when flange is in tension (cracks
equal to 1/8 inch at same step)
4 DS4a 2.22% -207.2 (-921.7) Core crushing at third cycle in drift
level
DS4b 2.22% -207.2 (-921.7) Bars buckled in web tip in first ramp
of hourglass load cycle
DS4e 4.45% -36.9 (-163.9) Bar fracture and buckling in flange
tips after first peak at drift level
(weak axis bending); Bar fracture in
web tip on next strong axis bending
cycle
DS4m 0.99% -29.5 (-131.4) Web crushing while attempting to
reach peak, unable to sustain grav-
ity load
DS4p 4.44% 58.6 (260.5) Core crushing flange tips under
weak axis bending
DSr10 1.63% -224.3 (-997.9) cracks equal to 1/8 inch in middle
two feet of flange (same point at
which spalling occurs)
736

Table C.102: NTW1 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS2a -1.63% -224.3 (-997.9) Max Shear 0.06 (1.5)
Flexure 0.13 (3.2)
DSr10 -1.63% -224.3 (-997.9) Max Shear 0.06 (1.5)
Flexure 0.13 (3.2)
C1 -0.38% -117.0 (-520.4) Max Shear 0.02 (0.51)
C2 0.28% 62.3 (277.1) Max Shear 0.016 (0.41)
C4 -0.10% -26.1 (-116.2) Max Shear 0.01 (0.25)
C5 -0.07% -8.8 (-39.3) Max Shear 0.013 (0.33)
C6 0.09% 29.6 (131.6) Max Shear 0.005 (0.13)
C7 -0.07% -8.8 (-39.3) Max Shear 0.013 (0.33)
C8 -0.38% -110.5 (-491.4) Max Shear 0.025 (0.64)
C9 0.28% 62.3 (277.1) Max Shear 0.016 (0.41)
C10 0.31% 71.4 (317.7) Max Shear 0.02 (0.51)
C11 0.17% 50.4 (224.2) Max Shear 0.02 (0.51)
C12 -0.07% -8.8 (-39.3) Max Shear 0.013 (0.33)
C13 -0.55% -144.6 (-643.3) Max Shear 0.025 (0.64)
C14 0.41% 80.9 (360.0) Max Shear 0.03 (0.76)
Flexure 0.007 (0.18)
C15 -0.73% -175.0 (-778.4) Max Shear 0.035 (0.89)
Flexure 0.035 (0.89)
C16 0.54% 95.0 (422.6) Max Shear 0.025 (0.64)
Flexure 0.01 (0.25)
C17 -1.13% -212.5 (-945.4) Max Shear 0.06 (1.5)
Flexure 0.013 (0.33)
C18 0.82% 108.4 (482.0) Max Shear 0.05 (1.3)
Flexure 0.025 (0.64)
C19 -1.13% -212.5 (-945.4) Max Shear 0.06 (1.5)
Flexure 0.13 (3.2)
C20 0.82% 106.1 (472.1) Max Shear 0.03 (0.76)
Flexure 0.03 (0.76)
737

NTW1: strong NTW1: weak


200 100

100 50
Vbase, kips

0 Vbase, kips 0

−100 −50
Env.
−200 DS1b −100 DS2a
DS1c DS4a
−300 −150
−4 −2 0 2 4 −5 0 5
% Drift % Drift
NTW1: diagStrong NTW1: diagWeak
200 100

100 50
Vbase, kips

Vbase, kips

0 0

−100 −50
DS4b
−200 DS4p −100 DS4m
DS4e DS1d
−300 −150
−4 −2 0 2 4 −5 0 5
% Drift % Drift

Figure C.114: Envelope for NTW1


738

C.3.62 NTW2

Table C.103: NTW2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b 0.04% 30.1 (133.9) Flex. cracks in web tip when flange
in compression)
DS1c 0.05% -44.0 (-195.8) Shear cracks in web under FC/FT
loading
DS1d 0.24% -103.5 (-460.3) As determined from
momentcurvature analysis
2 DS2a 1.67% -238.1 (-1058.9) Spalling of cover at base of web tip
when flange is in tension
DS2d 2.19% -208.0 (-925.1) Vertical cracks in flange tips dur-
ing figure-eight cycle (not specified
when)
4 DS4b 2.19% -208.0 (-925.1) Bars buckled in web tip (4 end-
most); accompanied by stirrup frac-
ture
DS4e 2.70% -212.2 (-944.0) Previously buckled bars fractured in
web tip after cycling
DS4q 2.19% -208.0 (-925.1) Fractured confining hoops 8 inches
above foundation in web tip (and
bars buckled)
DSr10 2.19% -218.3 (-971.1) shr cracks equal to 1/8 in. when
flange is in tension (flex. cracks
nearly 1/16, but not quite )
DSr9 1.71% 130.2 (579.0) shr cracks equal to 0.10 in. when
flange is in compression
739

Table C.104: NTW2 crack width information

Crack Width, in (mm)


DS % Drift Force, kips (kN) Exact/Max Crack Type Max. Resid.
DS2a -1.67% -238.1 (-1058.9) Max Shear 0.06 (1.5)
Flexure 0.04 (1)
DSr9 1.71% 130.2 (579.0) Max Shear 0.1 (2.5)
Flexure 0.035 (0.89)
DS4b 1.93% 132.1 (587.6) Max Shear 0.13 (3.2)
Flexure 0.06 (1.5)
DS4q 1.93% 132.1 (587.6) Max Shear 0.13 (3.2)
Flexure 0.06 (1.5)
C1 0.04% 21.2 (94.2) Max Shear 0.02 (0.51)
C2 -0.35% -138.5 (-615.9) Max Shear 0.025 (0.64)
C3 0.21% 63.4 (281.8) Max Shear 0.013 (0.33)
Flexure 0.005 (0.13)
C4 -0.51% -177.8 (-791.0) Max Shear 0.03 (0.76)
Flexure 0.0013 (0.033)
C5 0.43% 106.3 (472.6) Max Shear 0.04 (1)
Flexure 0.035 (0.89)
C7 1.09% 123.6 (549.7) Max Shear 0.035 (0.89)
Flexure 0.013 (0.33)
C9 0.49% 91.8 (408.2) Max Shear 0.035 (0.89)
Flexure 0.04 (1)
C10 -2.19% -232.3 (-1033.4) Max Shear 0.13 (3.2)
Flexure 0.06 (1.5)
C11 1.71% 131.7 (585.7) Max Shear 0.1 (2.5)
Flexure 0.035 (0.89)
C12 1.93% 132.1 (587.6) Max Shear 0.13 (3.2)
Flexure 0.06 (1.5)
740

NTW2: strong NTW2: weak


200 100

100
50
Vbase, kips

0 Vbase, kips
0
−100
Env.
DS1b −50 DS2a
−200
DS1c DS2e
−300 −100
−3 −2 −1 0 1 2 −5 0 5
% Drift % Drift
NTW2: diagStrong NTW2: diagWeak
200 100

100
50
Vbase, kips

Vbase, kips

0
0
−100
DS2d
DS4b −50 DS4e
−200
DS4q DS1d
−300 −100
−3 −2 −1 0 1 2 −5 0 5
% Drift % Drift

Figure C.115: Envelope for NTW2


741

C.3.63 PW1

Table C.105: PW1 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.11% -69.5 (-309.4)
DS1c -0.11% -69.5 (-309.4)
DS1d 0.62% 167.9 (746.7) As determined from
momentcurvature analysis
2 DS2a 0.56% 170.6 (758.9)
DS2d -0.54% -157.0 (-698.2)
3 DS3a -0.79% -173.4 (-771.4)
4 DS4a 1.47% 171.9 (764.8)
DS4b 1.47% 187.7 (834.9)
DS4e -1.53% -83.8 (-372.7)

PW1
200

150

100

50
Vbase, kips

Env.
−50 DS1b
DS1c
DS2d
−100 DS2a
DS1d
DS3a
−150 DS4b
DS4a
DS4e
−200
−2 −1.5 −1 −0.5 0 0.5 1 1.5
% Drift

Figure C.116: Envelope for PW1


742

C.3.64 PW2

Table C.106: PW2 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.10% -107.0 (-476.0)
DS1c -0.10% -107.0 (-476.0)
DS1d 0.75% 254.1 (1130.2) As determined from
momentcurvature analysis
2 DS2a 0.76% 260.9 (1160.5)
DS2d 0.35% 194.7 (866.2)
3 DS3a -0.76% -260.2 (-1157.5)
4 DS4a 1.50% 253.7 (1128.6)
DS4b -1.10% -3.7 (-16.5)
DS4k -1.10% -3.7 (-16.5)

PW2
300

200

100
Vbase, kips

Env.
DS1b
−100 DS1c
DS2d
DS1d
DS3a
−200 DS2a
DS4k
DS4b
DS4a
−300
−1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.117: Envelope for PW2


743

C.3.65 PW3

Table C.107: PW3 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.06% -77.6 (-345.0)
DS1c -0.08% -93.3 (-415.0)
DS1d -0.37% -185.8 (-826.3) As determined from
momentcurvature analysis
2 DS2a -0.40% -188.1 (-836.9)
DS2d -0.30% -170.8 (-759.7)
3 DS3a 0.75% 210.2 (934.9)
4 DS4a -1.00% -205.9 (-915.8)
DS4b 1.00% 211.1 (939.1)
DS4k -1.00% -205.9 (-915.8)
DSr9 0.50% 201.0 (894.2) From testing notes

PW3
250

200

150

100

50
Vbase, kips

0
Env.
DS1b
−50 DS1c
DS2d
−100 DS1d
DS2a
−150 DS2e
DS3a
DS4b
−200 DS4a
DS4k
−250
−1.5 −1 −0.5 0 0.5 1 1.5 2
% Drift

Figure C.118: Envelope for PW3


744

C.3.66 PW4

Table C.108: PW4 damage information

MOR DS % Drift Force, kips (kN) Description Figure


1 DS1b -0.09% -80.2 (-356.7)
DS1c -0.07% -72.2 (-321.4)
DS1d NaN% NaN (NaN) As determined from
momentcurvature analysis
2 DS2a 0.51% 199.2 (886.2)
DS2d 0.51% 199.2 (886.2)
3 DS3a -0.76% -213.5 (-949.9)
4 DS4a -0.75% -206.7 (-919.6)
DS4b 1.00% 75.4 (335.4)
DS4k 1.00% 75.4 (335.4)
DSr9 -1.01% -217.8 (-968.6) From testing notes

PW4
250

200

150

100

50
Vbase, kips

0
Env.
DS1c
−50 DS1b
DS2d
−100 DS2a
DS4a
−150 DS3a
DS4k
DS4b
−200 DS2e
DS1d
−250
−1.5 −1 −0.5 0 0.5 1 1.5
% Drift

Figure C.119: Envelope for PW4


745

Appendix D

BUILDING SUMMARIES

Chapters 8 and 9 presented a study of buildings damaged in the 2010 Chile earthquake.

The work was conducted as a part of the ATC-94 project [75] and involved the collaboration

of many academic and practicing engineers. Building plans and damage documentation were

made available through many of these project team members. The pages that follow contain

slides providing a detailed description of the buildings studied. The slides were assembled

by Ady Aviram and Dominic Kelly of Simpson Gumpertz & Heger using information from

references made available to the ATC-94 project [17, 48, 28, 114, 70, 110, 109, 107, 27, 108].
746

D.1 Plaza del Rio (Buildings A & B)

PR: Building Overview


Summary:
• Name: Plaza del Rio, Towers A & B.
• Location: Concepcion (Salas St. 1343).
• Construction: 2006.
• Use: Residential (1st level parking and storage).
• Layout: L-shape building with 2.5 cm (1”) expansion joint. Bldg. A – web, Bldg. B –
flange.
• Geometry: Bldg. A – 12 stories (basement below elevator/stairway shaft), Bldg. B –
13 stories (no basement).
• Footprint: Bldg. A – 24.2 x 14.2 m (79.4 x 46.7 ft), Bldg. B – 39.8 x 13.2 m (130.5 x
43.3 ft).
• Total height: Bldg. A - 30.1 m (98.8 ft), Bldg. B - 32.5 m (106.6 ft).
• Typical story height: 2.47 m (8.10 ft).
• Construction area: Bldg. A – 4,100 m2 (44,500 ft2), Bldg. B – 6,800 m2 (73,500 ft2).
• Construction type: RC bearing wall system.
• Soil type (NSF Rapid): III (Medium dense gravel/ medium stiff clay w/ vs>180
m/s)~Soil Type D-E per ASCE 7.
• Foundation type: Mat foundation over natural soil or fill.

(D.1.1)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B.
747

PR: Building Overview


Summary:
• Longitudinal axis orientation of Bldg. B aligns with Concepcion Station (Azimuth 60o)
• Geographic Coordinates: 36o49’14.42”S, 73o03’42.90”W

Bldg. B

E
W

Bldg. A

Figures: Plaza del Rio location and orientation


per Google Earth/Maps

(D.1.2)

PR: Building Overview


Bldg. A S
W N
Bldg. B Bldg. B
E E W
W
N Bldg. A Bldg. B
S

W
2 W
1
4
N Bldg. A Bldg. B
S E
Bldg. B

W E

1 3
Bldg. A

S
3
2

(D.1.3)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
748

PR: Building Overview

1
Expansion O Q S U V
joint
2 N P
1 1

2
5

(D.1.4)

PR: Building Overview


Structure:
• Wall thickness: 15 cm (6”)
• Slab thickness: 12 cm (5”)
• Beam depth: Varies from 42 cm (16“) to 181 cm (71“)
• Concrete: f’c=2500 Ton/m2 (3.6 ksi)
• Steel: Fy=60 ksi

Detailing:
• Wall horizontal reinforcement with 90o at ends on each side.
• Mesh reinforcement w/o boundary elements in some long walls or wall
intersections (e.g., line 28, 5-P, 29-P).
• Coupling beams with conventional orientation reinforcement and 135o
stirrups at spacing <d/2.
• Wall piers without cross ties or overlapping hoops.

(D.1.5)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
749

Floor Plans:
Bldg. A
13th Floor 14th Floor North

Strong
coupling

Additional
wall

Weak
coupling

Loss of wall flange


Strong
coupling

Typical floor plan


1st Floor (2nd-12th)

(D.1.6)

Damage Overview
• Building A (web, contains elevator & stairway shaft) severely damaged.
• Building B (flange) pounding damage at seismic joint at roof, light damage
at 1st level.
• Damage concentrated on 1st and 2nd levels in Building A.
• Severe wall failure (shear, flexural, compression-tension) in the EW
(transverse) walls, local/global buckling of wall piers.
• Different damage states in shear walls from minor 1 mm cracks to spalling,
bar buckling, rebar fracture, and loss of concrete core.
• Heaved soils around the edges of the mat foundation suggesting rocking
response.
• Severe damage in elevator shaft, and coupling beams at building center.
• Severe cracking in 2nd floor diaphragm.
• Building condemned, out-of-plumb to EW direction (transverse).

(D.1.7)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
750

Plaza del Rio- A

(D.1.8)

24’ 28 34’
Damage Overview
• Out-of-plumbness to EW
(Transverse) direction: 46 cm in 31 m
(18” in 102 ft) = L/67 (1.5% drift)
W E

Structural damage scale:


Total destruction
Severe damage
Deterioration
Cracking
No damage (unapparent)

Elevation at South Wall


(Lines V-W)
10

(D.1.9)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
751

V
U
Damage Overview (PR-A) S
Q P
O
W V U S N S
S N

Eastern facade
Eastern facade
S U
S
N

West facade West facade

(D.1.10)

Damage Overview (PR-A)

Southern façade, West side Southern facade, East side

Soil heaving

Bar buckling, fracture

12

(D.1.11)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
752

Damage Overview: Floor Plans

1st Floor Cracking due to 2nd Floor Typical damage


overturning and 2nd-12th Floor
shear transfer 13

(D.1.12)

Damage: Roof
• 1” construction joint coinciding with
escalator/elevator shaft on line N. The
floors of the two buildings were not
connected. Evidence of pounding during
EQ at all levels.
Typical floor plan
• 12 cm (5”) slab at roof level severely
damaged during EQ.
Construction joint

Bldg. A Bldg. B

13th Floor

Eastern side

14th Floor

(D.1.13)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
753

Transverse Walls

15

(D.1.14)

Line N

Bldg. B Bldg. A

Bldg. A

Western facade

• Damage at line N due to pounding at all levels.


16

(D.1.15)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
754

Damage to Stairs
• Horizontal tension cracks and
diagonal shear cracks in shear
walls.
• Offset of stairway shaft wrt’
Bldg. B in EW and NS
directions: 15 cm (6”).
Bldg. A
moved E
wrt’ Bld. B

Change in construction
detail. Pounding
damage.
N

Bldg. A

Bldg. B
17

(D.1.16)

Lines V-W

24 27 28 31 34

18

(D.1.17)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
755

W & 31
28 V & 34

Lines V-W

Wall at V moving
towards N

Is this a reasonable
damage pattern?

Wall at W & 31 shearing towards N Symmetrical diagonal


cracks in spandrel, line W

Damage details
at East side

Coupling between lines


V and U?

(D.1.18)

Lines V-W
• Flexural overturning and rocking in Long. NS direction
(evidence of foundation pull-out /soil heaving).
• Diagonal cracks in spandrel beam on line W appear to be
from relative differences in overturning in NS direction of
edges of line V and 24 and 34 compared to line 28.
• The longitudinal wall on line 28 possibly prevented yielding
of the pier on line W due to motion in transverse EW
direction.
• Wall on line V and W displaced North between 8 and 10
cm (4”).
BE at wall V on West side
• Coupling between walls on lines V and U and between P near line 27
and Q and similar damage lines V and P.

Soil heaving
adjacent to
Line V and 34
Wall at V on West
side near line 27

(D.1.19)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
756

Lines V-W

Compression
failures

Diagonal cracking Tension


cracks

Compression
Diagonal cracking
failures

Walls at lines V & W


near line 27 (West)

Walls at lines V & W


near line 31 (East)

21

(D.1.20)

Line U

East side looking N

West side looking N


• Compression and tension failures in East and West wall piers
• Coupling with wall on line V for NS motion.
• Wall buckling on West side? 22

(D.1.21)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
757

Line S

BE at Line 24’

BE at Line 34’

Damage details West wall

(D.1.22)

Line S
• The boundary elements at lines 24 and 34 have different vertical
reinforcement amounts and different boundary element lengths.
• Rebar congestions of longitudinal reinforcement around conduits at line 24’.
• Diagonal cracks throughout wall height on West side.
• Failure extends from BE to the wall interior with significant loss of concrete
core.
• Hypothesis: Compression failure/loss of load bearing capacity at line S
connecting 2 “rigid bodies” consisting of N-S and S-W, which overturned in NS
direction wrt’ line S (building center).
• Reason why the damage to the wall on line S occurred at the 2nd level instead
of the 1st level?

24

(D.1.23)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
758

Line P Q
P

Eastern side

(D.1.24)

Line P Is this a reasonable plastic hinge


zone to expect?

Is this a reasonable damage


pattern?

• A clear region in which yield can extend over a reasonable height of the wall
does not exist.
• The damage is consistent with compression damage due to overturning in
NS direction in which the wall acts as a flange for the P-Q coupled wall
(discontinuous).

(D.1.25)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
759

Line Q

• Strong EW directionality (East side


presents compression failure and
West side tension cracking).
• Coupling with wall on P for NS
motion.

(D.1.26)

Line Q
• Loss of flange wall
• Coupling between walls on lines Q and P?
• Bar buckling and complete loss of concrete core
• Wall global buckling?

28

(D.1.27)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
760

Line O

29

(D.1.28)

Longitudinal Walls

30

(D.1.29)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
761

Line 28

Wall Line T
Wall Line W

• Little redundancy of shear walls in Longitudinal


(NS) direction.
• Main walls near bldg. centerline, no perimeter walls.
• No transverse reinforcement for confinement of
boundary element in wall 28.

31

(D.1.30)

Lines 5, 29

32

(D.1.31)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
762

ASCE 31- Tier 1 Assessment


Findings:
• Weak and soft stories identified according to ASCE 31 definition in
longitudinal direction (however, severe damage in transverse walls).
• Discontinuities in vertical elements (primarily loss of flange for transverse
walls).
• Shear stresses exceed allowable limits.
• Large openings in diaphragm near few shear walls (stairway/elevator shaft).
• Wall mesh reinforcement stops at foundation level. Only boundary element
reinforcement doweled to foundation.
• Remaining checklist items for LS performance level are not triggered.

33

(D.1.32)

ASCE 41 Assessment TT=0.47 sec


TL=0.30 sec
Tq=0.37 sec
• ETABS/SAP model: T1,ASCE-41=0.64 sec
– Effective cross-section properties
– Dead and live load per NCh433Of.96
– Nominal material properties in model
• Response Spectra Analysis
– Elastic spectra per NCh433Of.96
– Concepcion GM spectra

SAP2000
model of PR

34

(D.1.33)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
763

Plaza del Rio- B

35

(D.1.34)

PR-B: Overview
N
E

S N

Light damage, 1st Floor

36

(D.1.35)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
764

PR-B: Overview
• Diagonal (hairline) cracks throughout 1st story.
• Element dimension and reinforcement details similar to Bldg. A

W
E

Northern facade

Southern facade

West facade 37

(D.1.36)

PR-B: Damage Overview

Initiation of
flexural failure

Cracks in slab

Hairline
Diagonal cracks diagonal cracks

38

(D.1.37)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
765

Loss of wall flange/


PR-B: Floor Plans discontinuity

Additional
wall

1st Floor

Typical floor plan, 2nd-13th Floor


39

(D.1.38)

PR-B: Observations
Configuration:

• Wall vertical irregularity throughout building. However, new walls are added
near discontinued walls at 1st level.
• Irregularity present in both longitudinal and transverse directions.
• Building configuration does not result in strong coupling of façade walls.
• Walls distributed around building interior (corridor) as well as perimeter.

40

(D.1.39)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
766

PR-B: Reinforcement Details


PR-A PR-B

Example: Wall reinforcement at 1st level

Example: Coupling beam reinforcement at 1st floor


41

(D.1.40)

PR-B: Observations
Reinforcement:

• Minimum wall reinforcement and dimensions same in both buildings.


• Confinement details and boundary elements same in both buildings.
• Slab dimensions and reinforcement similar in both buildings.
• Coupling beam dimensions same in both buildings.
• PR-B wall reinforcement: 18mm bars used more commonly than 16mm at
1st level. Higher wall reinforcement ratios in PR-B.
• Coupling beams more heavily reinforced in PR-B: 18mm bars used more
commonly than 16mm at lower levels. Higher number of bars used near
vertical irregularities.

42

(D.1.41)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
767

ASCE 31- Tier 1 Assessment


Findings:
• Only soft story identified according to ASCE 31 definition in both longitudinal
and transverse directions.
• Discontinuities in vertical elements: Walls continuous from 2nd to 13th levels
interrupted at 1st level below, new walls on 1st level interrupted on 2nd level
above. Similar observation in both longitudinal and transverse directions.
• Shear stresses exceed allowable limits (worse than PR-A: Vj=243
psi>2(f’c)1/2).
• Wall mesh reinforcement stops at foundation level. Only boundary element
reinforcement doweled to foundation.
• Remaining checklist items for LS performance level are not triggered.

43

(D.1.42)

Sources of Information
• Maria Jose Cerda Caamano’s thesis, Universidad Tecnica Federico Santa
Maria, 06/2011
• Post-Earthquake Evaluation Report No. 595657-12, Salas 1343 Torre A
(Plaza del Rio) by IDIEM, 03/23/2010
• Post-Earthquake Evaluation Report No. 879068, Salas 1343 Torre A (Plaza
del Rio) by DICTUC, 04/12/2010
• NSF Rapid Report (Westenenk et al. 2011)
• Damage photos (from Moehle) and survey by UW

(D.1.43)

Figure D.1: Presentation providing overview of Plaza del Rio Buildings A & B (con’t).
768

D.2 Centro Mayor

Centro Mayor: Building Overview


Summary:
• Construction: 2005
• Location: Concepcion
• Use: Residential
• Geometry: 18 stories and 2 basement levels
• Footprint: Varies from 51 x 28 m (167 x 92 ft) in 1st level to 40.85 x 17.25 m
(134 x 57 ft) in floors 2-17
• Total height: 47.15 m (155 ft)
• Total construction area: 12,000 m2 (130,000 ft2)
• Longitudinal axia orientation: 62oW
• Soil type (dwgs): III (Medium dense gravel/ medium stiff clay w/ vs>180 m/s)
• Soil type (NSF Rapid): II (Dense gravel or clay w/ vs>400 m/s)
• Construction type: RC bearing wall system
• Foundation (geotech. report): Mat foundation over natural soil or refill
• Structural Engineer: JMS Ingenieros Consultores

(D.2.1)

Figure D.2: Presentation providing overview of Centro Mayor.


769

Centro Mayor: Building Overview


Y

(D.2.2)

Centro Mayor: Building Overview


Structure:
• Wall thickness: 20 cm (8”)
• Slab thickness: 15 cm (6”)
• Beam depth: Varies from 48 cm (19“) to 206 cm (81“)
• Concrete: f’c=2500 Ton/m2 (3.6 ksi)
• Steel: Fy=60 ksi
• Wall density: 2.5% ratio of wall area in each direction to floor area

Detailing:
• Coupling beams with conventional orientation reinforcement and 135o
stirrups at spacing <d/2
• Wall horizontal reinforcement with 135o at ends on each side (construction?)
• At wall vertical discontinuity the longitudinal rebar extends only min
development length.

(D.2.3)

Figure D.2: Presentation providing overview of Centro Mayor.


770

Centro Mayor: Building Overview


Description of damage:
• Multiple failures in EW-oriented walls in the first 2-3 elevated stories and
basement
• Shear failure through multiple piers separated by openings
• Compression or flexural-compression failures
• Shear failure of horizontal wall segments/coupling beams
• T-shape, L-shape, C-shape wall failures, slab damage

(D.2.4)

SAP Model: Linear Response Spectrum Analysis

Floor 1

View from
NW corner

Floor 2-17
6

(D.2.5)

Figure D.2: Presentation providing overview of Centro Mayor.


771

SAP Model: Elevations


• Transverse (Y-direction, EW):

Line M Line F Line E 7

(D.2.6)

Damage: Elevations

Line M Line F Line E


8

(D.2.7)

Figure D.2: Presentation providing overview of Centro Mayor.


772

SAP Model: Elevations


• Longitudinal (X-direction, NS):

Line 9 Line 8 Line 15 9

(D.2.8)

Damage: Elevations

Line 9 Line 8 Line 15


10

(D.2.9)

Figure D.2: Presentation providing overview of Centro Mayor.


773

ASCE 31- Tier 1 Assessment


Preliminary findings:
• Major setback from 1st level to 2nd level (mass irregularity)
• Weak and soft stories identified according to ASCE 31 definition
• High aspect ratio of maximum to minimum horizontal dimension of lateral
force-resisting system
• Discontinuities in vertical elements (walls)
• Torsion and permanent offsets in reports
• No evidence of deterioration prior to EQ according to reports
• Wall thickness complies with requirement L/25 but wall shear stresses could
exceed limits (pending)
• Inconsistent reinforcement in reentrant corners
• Large openings in diaphragm
• Strength of coupling beams strength is non-compliant with wall uplift
requirement for IO level
• Wall piers without cross ties or overlapping hoops

11

(D.2.10)

Figure D.2: Presentation providing overview of Centro Mayor.


774

D.3 Concepto Urbano

CU: Building Overview


Summary:
• Name: Concepto Urbano.
• Location: Concepcion (Orompello 129).
• Construction: 2008-2010 (in construction during EQ).
• Use: Residential (Parking and storage at basements and level 1, pool at 10th level).
• Geometry: 22 stories + 2 basements.
• Layout: L-shape floor plan with rectangular tower.
• Footprint: Basement to 1st level: 51.92 x 28.64 m (170.3 x 94.0 ft). Levels 2-10: 48.59
x 22.85 m (159.4 x 75.0 ft), Tower levels 11-22: 26.34 x 17.40 m (86.4 x 57.1 ft).
• Total height: Base: 25.49 m (83.6 ft). Tower: 55.69 m (182.7 ft).
• Typical story height: 2.50 m (8.20 ft).
• Construction area: 14,582 m2 (X ft2)
• Construction type: RC bearing wall system.
• Soil type (Dwgs): III (Medium dense gravel/ medium stiff clay w/ vs>180 m/s)~Soil
Type D-E per ASCE 7.
• Foundation type: Mat foundation over natural soil or fill.

(D.3.1)

Figure D.3: Presentation providing overview of Concepto Urbano.


775

CU: Building Overview


Summary:
• Longitudinal axis orientation aligns with Concepcion Station (Azimuth 60o)
• Geographic Coordinates: 36o49’40.68”S, 73o02’34.42”W

Figures: Plaza del Rio location and orientation


per Google Earth/Maps

(D.3.2)

CU: Building Overview 13


11 6
B D I L
5
N 1 1

Eastern Northern
facade facade

2
Southern facade
3
3 2 1
2

Entrance at
Eastern facade
3
4

(D.3.3)

Figure D.3: Presentation providing overview of Concepto Urbano.


776

CU: Building Overview


Northern
facade
Western facade

5 6

4
Eastern
facade

SW corner

3
5

(D.3.4)

• Negligible damage reported


CU: Damage Overview • Spalling observed in Southern
façade, slabs
N

(D.3.5)

Figure D.3: Presentation providing overview of Concepto Urbano.


777

CU: Building Overview


• Geometric configuration

28.64 m
51.92 m

Figure: Basement and 1st level plan


7

(D.3.6)

CU: Building Overview

48.59 m

22.85 m

Figure: Floors 2-10 plan

(D.3.7)

Figure D.3: Presentation providing overview of Concepto Urbano.


778

CU: Building Overview

26.34 m

17.40 m
Figure: Floors 11-22 plan

(D.3.8)

CU: Building Overview


Dimensions:
• Wall thickness: Varies.
o Basement-Level 1: 25-45 cm (10-18”).
o Levels 2-10: 20-30 cm (8-12”).
o Levels 11-22: 20-25 cm (8-10”).
• Slab thickness: Varies.
o Basement - Floor 1: 17 cm (6.7”)
o Floors 2-9, 11-22: 16 cm (6.3”)
o Floor 10: 20 cm (7.9”)
• Coupling beam width: Matches adjacent walls.
• Beam depth: Varies from 50 cm (20“) to 190 cm (75“)
• Mat foundation: Thickness 80-120 cm (2.6-4’).

Materials:
• Concrete: f’c=2500 Ton/m2 (3.6 ksi) throughout.
• Steel: Fy=60 ksi throughout.

10

(D.3.9)

Figure D.3: Presentation providing overview of Concepto Urbano.


779

CU: Building Overview


Detailing:
• Wall horizontal reinforcement with 135o at ends on each side or closed ties with 135o
bend.
• Typical: Mesh reinforcement w/o boundary elements. Selected walls with confined
boundary elements (closed ties at 10-20 cm (4-8”) spacing).
• Minimum wall reinforcement: 0.25%. Typical wall w/ higher ratios.
• Cross ties and overlapping hoops in selected walls.
• Coupling beams with conventional orientation reinforcement and 135o stirrups at
spacing <d/2.
• Vertical wall discontinuities reinforced using detailed coupling beams, and/or diagonal
bars. However, majority of wall lines are continuous.

11

(D.3.10)

CU: Building Overview


20 cm
• Selected elevations and details

25 cm

30 cm

40 cm

45 cm

12

(D.3.11)

Figure D.3: Presentation providing overview of Concepto Urbano.


780

CU: Building Overview


• Selected elevations and details

Line 1

13

(D.3.12)

CU: Building Overview


• Selected elevations and details

Line M

14

(D.3.13)

Figure D.3: Presentation providing overview of Concepto Urbano.


781

CU: Building Overview


• Selected elevations and details

15

(D.3.14)

ASCE 31- Tier 1 Assessment


Findings:
• Major setbacks at 2nd and 10th levels in both directions.
• Weak and soft stories triggered due to these two setbacks. However, this
corresponds to weak/soft story above a strong floor.
• Geometric irregularities: Length of lateral-force-resisting system reduces by
more than 30%. Torsion and mass check OK.
• Discontinuities in vertical elements: A few wall discontinuities at Levels 2, 3,
and 11 in both longitudinal and transverse directions. In 11th level: walls
above interrupted. At 2nd and 3rd levels: walls below interrupted.
• Shear stresses OK (Vj=110 psi <2(f’c)1/2).
• Mesh reinforcement below minimum requirements in few locations.
• Remaining checklist items for LS performance level are not triggered.

16

(D.3.15)

Figure D.3: Presentation providing overview of Concepto Urbano.


782

Appendix E

TIER 1 EVALUATION OF CHILEAN BUILDINGS

E.1 Comprehensive List of Deficiencies

Section 9.1 presented a overview of deficiencies identified by the ASCE/SEI 31-03 Tier 1

evaluation as well as those identified by the project team that were not flagged by the Tier 1

evaluations. Here, a comprehensive list of deficiencies is presented, including an indication

of which building(s) in which the deficiencies were observed and whether or not they can

be identified by the Tier 1 (or Tier 2) evaluation procedures. The list of deficiencies was

developed as a part of the ATC-94 project [75] and was compiled by Ady Aviram and

Dominic Kelly of Simpson Gumpertz & Heger.

Table E.1: List of wall and building characteristics resulting in structural damage.

Deficiency Description Building(s) Identified


by Tier 1
or 2?
Geometric
Weak story Reduction in wall length resulting in CM, CU, PR-A T1
strength discontinuity.
Soft story Reduction in wall length resulting in CM, CU, PR-A, PR- T1
stiffness discontinuity in building B
Loss of flange area Discontinuity in T-, C-, or Z-shape PR-A, PR-B -
wall flanges reduces strength of walls
Wall geometric irreg- Change in wall geometry and discon- CM, CU, PR-A, PR- T1
ularity/discontinuity tinuity to foundation B
Overall building geo- Significant horizontal or vertical set- CM, CU T1
metric irregularity backs, change in mass, which cause
stress concentration at the setback
Lack of redundancy Multiple and complete load paths are - T1
needed to ensure adequate seismic and
gravity load transfer to foundation
Unclear location of Unclear load path, geometric and re- PR-A, PR-B, -
wall plastic hinges inforcement irregularities result in un-
predictable locations and spreading of
damage
783

Table E.1: List of wall and building characteristics resulting in structural damage. (con’t)

Deficiency Description Building(s) Identified


by Tier 1
or 2?
Torsion Overall building torsion results in - T1
shear stress increase in walls
Inadequate wall Bearing walls with minimum thick- - T1
thickness ness requirement related to unsup-
ported wall height to ensure gravity-
load carrying capacity.
Inadequate slab Minimum slab thickness to ensure PR-A T2
thickness adequate diaphragm strength and
seismic load transfer to walls.
Diaphragm Diaphragm split levels and expansion - T1
discontinuity joints prevents adequate seismic load
transfer at each level.
Diaphragm openings Openings near shear walls or consist- CM, PR-A, PR-B T1
ing of significant portion of building
dimensions cause stress concentration
in diaphragms and inadequate shear
load transfer.
Overturning Wall aspect ratio controlled for in IO CM, PR-A, PR-B T1
level only
Pounding or coupling Between adjacent buildings due to PR-A, PR-B -
with non-structural lack of adequate expansion joints or
components proper anchorage of secondary com-
ponents.
Inadequate Reinforcement Details
Wall minimum Minimum wall reinforcement ratio CU T1
reinforcement and spacing required for ductile wall
behavior
Confinement Boundary elements required for shear CM, CU, PR-A, PR- T1
reinforcement walls with 2:1 aspect ratio with ade- B
quate confining reinforcement spacing
Confinement Inadequate spacing of confining rein- CU, PR-A, PR-B T1
spacing forcement (IO level only)
Confinement Lack of cross-ties in slender CM, PR-A, PR-B T1
cross-ties boundary elements or inadequate
spacing/orientation
Wall reinforcement Proper anchorage and development CM -
anchorage lengths of wall reinforcement at lower
levels to ensure load transfer
Wall reinforcement Change in wall reinforcement can CM -
discontinuity cause stress concentration at that lo-
cation
Wall reinforcement Wall dowels able to develop the PR-A, PR-B T1
doweled to strength of the walls or uplift capacity
foundation of foundations
Diaphragm-shear Diaphragm connectivity to shear walls - T1
wall connection is required for adequate seismic load
transfer. Details not specified.
784

Table E.1: List of wall and building characteristics resulting in structural damage. (con’t)

Deficiency Description Building(s) Identified


by Tier 1
or 2?
Coupling beams Reinforcement details with proper PR-A, PR-B T1
hooks and at maximum spacing to
ensure ductile behavior and coupling
of adjacent walls.
Reinforcement at Trim reinforcement at wall openings CU, CM, PR-A, PR- T1
openings to prevent damage at locations of B
stress concentration (corners).
Diaphragm reinforce- Reinforcement needed around large - T1
ment at openings diaphragm openings to ensure ade-
quate seismic load transfer.
Beam/spandrel Adequate reinforcement anchorage - -
reinforcement and development length.
anchorage
Insufficient Strength
Wall shear stress Shear capacity in shear walls must be CM, PR-A, PR-B T1, T2
exceeded adequate to prevent loss of concrete
area at all building levels.
Coupling beams Strength required to develop the uplift CM, PR-A, PR-B T1, T2
capacity of adjacent walls.
Plan irregularity and Tensile capacity needed to develop the CM T1
reentrant corners strength of the diaphragm at
reentrant corners of floor.
Intersection of Z-, C-, Demand from different directions - T2
T-shape walls results in blow-out of boundary ele-
ments at wall intersections.
Demand-to-capacity Adequate strength of all elements of - T2
ratio lateral-force-resisting system (beams,
spandrels, walls, piers, columns, slabs,
foundations)
Strength of short Short columns with high shear - -
column demands can results in shear failure.
785

Appendix F

TIER 2 EVALUATION OF CHILEAN BUILDING

F.1 Summary of Adequate Walls

The following pages present figures indicating the walls that are adequate for various perfor-

mance objectives using the ASCE/SEI 41-06 [8, 9] standard. For each building, the lowest

floor above grade is included. An explanation of the method of evaluation and a summary

of the results are presented in Section 9.2.2.

F.1.1 Plaza del Rio Building A (PR-A)

25 25 25

20 20 20
# of walls

# of walls

# of walls

15 15 15

10 10 10

5 5 5

0 0 0
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1
IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE

(a) ASCE response spectrum (b) Chilean response spectrum (c) Proposed Chilean response
spectrum

25 25

20 20
# of walls

# of walls

15 15

10 10
0: No Damage
1: Minor (cracking)
5 5
2: Moderate
0 0
3: Severe
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 4: Total
IO−GM LS−GM CP−GM IO−GM LS−GM CP−GM

(d) Concepcion response spec-(e) San Pedro de la Paz response (f) Damage color legend
trum spectrum

Figure F.1: PR-A Floor 1: # of damage state occurances in walls achieving various perfor-
mance objectives. Color indicates the observed damage.
786

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-DBE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.2: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under ASCE 7 response spectrum.
787

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-DBE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.3: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under Chilean response spectrum.
788

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-DBE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.4: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under proposed Chilean response spectrum.
789

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-DBE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.5: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under Concepcion response spectrum.
790

(a) Inadequate all (b) Adequate for (c) Adequate for (d) Adequate for
performance CP-DBE. LS-DBE. IO-DBE.
objectives.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.6: PR-A Floor 1: Wall damage for walls achieving various performance objectives
under San Pedro de la Paz response spectrum.
791

F.1.2 Plaza del Rio Building B (PR-B)

45 45 45
40 40 40
35 35 35
30 30 30
# of walls

# of walls

# of walls
25 25 25
20 20 20
15 15 15
10 10 10
5 5 5
0 0 0
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1
IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE

(a) ASCE response spectrum (b) Chilean response spectrum (c) Proposed Chilean response
spectrum

45 45
40 40
35 35
30 30
# of walls

# of walls

25 25
20 20
15 15 0: No Damage
10 10
1: Minor (cracking)
5 5
2: Moderate
0 0
3: Severe
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 4: Total
IO−GM LS−GM CP−GM IO−GM LS−GM CP−GM

(d) Concepcion response spec-(e) San Pedro de la Paz response (f) Damage color legend
trum spectrum

Figure F.7: PR-B Floor 1: # of damage state occurances in walls achieving various perfor-
mance objectives. Color indicates the observed damage.
792

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.8: PR-B Floor 1: Wall damage for walls achieving various performance objectives
under ASCE 7 response spectrum.
793

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.9: PR-B Floor 1: Wall damage for walls achieving various performance objectives
under Chilean response spectrum.
794

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.10: PR-B Floor 1: Wall damage for walls achieving various performance objectives
under proposed Chilean response spectrum.
795

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.11: PR-B Floor 1: Wall damage for walls achieving various performance objectives
under Concepcion response spectrum.
796

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.12: PR-B Floor 1: Wall damage for walls achieving various performance objectives
under San Pedro de la Paz response spectrum.
797

F.1.3 Centro Mayor (CM)

40 40 40

35 35 35

30 30 30

25 25 25
# of walls

# of walls

# of walls
20 20 20

15 15 15

10 10 10

5 5 5

0 0 0
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1
IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE IO−DE LS−DE CP−MCE

(a) ASCE response spectrum (b) Chilean response spectrum (c) Proposed Chilean response
spectrum

40 40

35 35

30 30

25 25
# of walls

# of walls

20 20

15 15 0: No Damage
10 10 1: Minor (cracking)
5 5 2: Moderate
0 0
3: Severe
DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 DCR < 1 DCR > 1 4: Total
IO−GM LS−GM CP−GM IO−GM LS−GM CP−GM

(d) Concepcion response spec-(e) San Pedro de la Paz response (f) Damage color legend
trum spectrum

Figure F.13: CM Floor 2: # of damage state occurances in walls achieving various perfor-
mance objectives. Color indicates the observed damage.
798

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.14: CM Floor 2: Wall damage for walls achieving various performance objectives
under ASCE 7 response spectrum.
799

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.15: CM Floor 2: Wall damage for walls achieving various performance objectives
under Chilean response spectrum.
800

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.16: CM Floor 2: Wall damage for walls achieving various performance objectives
under proposed Chilean response spectrum.
801

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.17: CM Floor 2: Wall damage for walls achieving various performance objectives
under Concepcion response spectrum.
802

(a) Inadequate all performance objectives. (b) Adequate for CP-DBE.

(c) Adequate for LS-DBE. (d) Adequate for IO-DBE.

0: No Damage
1: Minor (cracking)
2: Moderate
3: Severe
4: Total

(e) Legend

Figure F.18: CM Floor 2: Wall damage for walls achieving various performance objectives
under San Pedro de la Paz response spectrum.
803

F.2 Demand Capacity Ratios

The figures on the following pages provided, for the lowest above grade floor in each building

the unfactored, DCRu , and factored, DCRf , demand-capacity ratios, as well as the con-

trolling action (shear, moment or axial) for each wall. Discussion of the demand-capacity

ratios and controlling actions is provided in Section 9.2.


804

(a) ASCE RS (b) Concepcion RS (c) San Pedro RS (d)

Figure F.19: PR-A Floor 1: Unfactored demand capacity ratios (DCRu ).

(a) ASCE RS (b) Concepcion RS (c) San Pedro RS

Figure F.20: PR-A Floor 1: Action resulting in largest DCRu value. Blue: Moment; Red:
Shear; Black: Axial
805

(a) ASCE RS (b) Concepcion RS (c) San Pedro RS (d)

Figure F.21: PR-A Floor 1: Factored demand capacity ratios (DCRf ) for immediate occu-
pancy performance level.

(a) ASCE RS (b) Concepcion RS (c) San Pedro RS

Figure F.22: PR-A Floor 1: Action resulting in largest DCRf value for immediate occu-
pancy performance level. Blue: Moment; Red: Shear; Black: Axial
806

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

0
0 1 2 3 4 5 6 7 8

(d)

Figure F.23: PR-B Floor 1: Unfactored demand capacity ratios (DCRu ).


807

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

Figure F.24: PR-B Floor 1: Action resulting in largest DCRu value. Blue: Moment; Red:
Shear; Black: Axial
808

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

0
0 0.5 1 1.5 2 2.5 3 3.5 4

(d)

Figure F.25: PR-B Floor 1: Factored demand capacity ratios (DCRf ) for immediate occu-
pancy performance level.
809

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

Figure F.26: PR-B Floor 1: Action resulting in largest DCRf value for immediate occu-
pancy performance level. Blue: Moment; Red: Shear; Black: Axial
810

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

0
0 1 2 3 4 5 6 7 8

(d)

Figure F.27: CM Floor 2: Unfactored demand capacity ratios (DCRu ).


811

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

Figure F.28: CM Floor 2: Action resulting in largest DCRu value. Blue: Moment; Red:
Shear; Black: Axial
812

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

0
0 0.5 1 1.5 2 2.5 3 3.5 4

(d)

Figure F.29: CM Floor 2: Factored demand capacity ratios (DCRf ) for immediate occu-
pancy performance level.
813

(a) ASCE RS

(b) Concepcion RS

(c) San Pedro RS

Figure F.30: CM Floor 2: Action resulting in largest DCRf value for immediate occupancy
performance level. Blue: Moment; Red: Shear; Black: Axial
814

F.3 Axial Demands

To explore the overall impact of coupling, the axial loads in the walls in the lower floor of the

buildings were studied. For each response spectrum used in the Tier 2 evaluations, the axial

load from the earthquake forces, QU F , was determined according to the ASCE 41 provisions.

This value was normalized by the axial load due to gravity to determine the increase in the

axial load due to earthquake effects. The figures on the follow pages show this index for the

first floor of the three buildings modeled with SAP2000. Also included are figures showing

the ratio of the maximum compressive axial demand to the axial compressive strength of

the walls, Ag fc0 .

(a) (b) ASCE 7 (c) Concepcion GM (d) San Pedro GM

Figure F.31: PR-A Floor 1: Ratio of maximum axial load due to earthquake loading to
axial load due to gravity loading from different response spectra.
815

50

40

30

20

10

(a) (b) ASCE 7 (c) Concepcion GM (d) San Pedro GM

Figure F.32: PR-A Floor 1: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial capacity of the
wall (Ag fc0 ).
816

1 1.5 2 2.5 3 3.5 4 4.5 5


(a)

(b) ASCE 7

(c) Concepcion GM

(d) San Pedro GM

Figure F.33: PR-B Floor 1: Ratio of maximum axial load due to earthquake loading to
axial load due to gravity loading from different response spectra.
817

0
0 10 20 30 40 50
(a)

(b) ASCE 7

(c) Concepcion GM

(d) San Pedro GM

Figure F.34: PR-B Floor 1: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial capacity of the
wall (Ag fc0 ).
818

1 1.5 2 2.5 3 3.5 4 4.5 5


(a)

(b) ASCE 7

(c) Concepcion GM

(d) San Pedro GM

Figure F.35: CM Floor 2: Ratio of maximum axial load due to earthquake loading to axial
load due to gravity loading from different response spectra.
819

0
0 10 20 30 40 50
(a)

(b) ASCE 7

(c) Concepcion GM

(d) San Pedro GM

Figure F.36: CM Floor 2: Axial load ratio, λN , due to earthquake loading resulting from
different response spectra. Values shown are a percentage of the gross axial capacity of the
wall (Ag fc0 ).
820

Appendix G

TIER 3 EVALUATION OF CHILEAN BUILDING

Section 9.3.2 provided a discussion of Tier 3 evaluation results, which included evaluation

of ASCE/SEI 41-06 [8, 9] acceptance criteria, fragility function probabilities, and strain

limits for steel yield and concrete spalling and crushing. A full discussion of the methods

of evaluation are presented in Section 9.3.2. Discussion of these results included figures for

only a couple of ground motions studied. For completeness, all figures are provided here,

including bar charts summarizing the number of walls exceeding each evaluation criterion

and floor plans indicating where these walls are located in the building. Unlike in the

main body, the figures depicting the evaluation criteria with and without use of the shear

criterion are combined to reduce the total number of figures included. For each set of floor

plans, subfigures (a)-(d) and (f)-(i) are applicable to the evaluation without use of the shear

criterion. If it is desired to consider the impact of the shear criterion, subfigures (e) and (j)

must be considered, with any colored wall in these subfigures removed from consideration in

subfigures (a)-(d) and (f)-(i), respectively. Following the figures for each individual ground

motion are tables documenting the order in which the acceptance criteria, fragility function

probability, or strain limits were exceeded. The time during the ground motion record is

given to provide a sense of how rapidly the demands occurred relative to one another.
821

G.1 Concepcion (Unscaled)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.1: Summary of walls exceeding acceptance criteria (Concepcion ground motion).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure
45
G.2:ASCE Rotation
Walls Fragility Functions
exceeding ASCE 41 Strain
rotation limits (Concepcion ground motion).
40

822
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure G.3:
45
WallsASCE Rotation
exceeding Fragility Functions
fragility function Strain
probabilities of 0.50 (Concepcion ground motion).
40

823
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


45
Figure G.4: Walls exceeding strain limits (Concepcion ground motion).

824
40

35
# of Predicted walls

30

25

20

15
825

Table G.1: Sequence of exceedence of ASCE 41 acceptance criteria (Concepcion ground


motion). Shading indicates that the wall also exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wV-34-f1 9.1 wV-34-f1 10.3 wV-24-f2 13.8 wP-5-f1 11.0
2 wN-5-f2 9.4 wV-24-f1 11.0 wV-34-f2 13.8 wO-5-f1 11.1
3 wV-24-f1 9.6 wS-29-f2 11.0 w34-V1-f2 14.7 wO-29-f1 12.0
4 wN-5-f1 9.8 wS-5-f2 11.1 wN-5-f2 15.0 wV-24-f1 12.2
5 wP-5-f1 9.8 wV-34-f2 11.1 wV-24-f1 20.8 w28-W-f1 12.7
6 wV-24-f2 10.4 wS-5-f1 11.1 wS-29-f2 20.9 wV-34-f1 12.8
7 wS-29-f1 10.5 wV-24-f2 11.2 wN-5-f1 21.0 wP-29-f1 12.9
8 wS-5-f2 10.5 wN-5-f2 11.6 wS-5-f2 21.9 wO-29-f2 14.5
9 wS-29-f2 10.5 wN-5-f1 11.6 wS-29-f1 22.0 wS-5-f2 20.1
10 wS-5-f1 10.5 w34-V1-f2 12.8 w24-V1-f2 22.0 wP-23-f2 20.8
11 wV-34-f2 10.5 w24-V1-f2 13.8 wS-5-f1 22.1 wV-34-f2 21.1
12 wO-5-f1 10.6 wO-29-f2 14.6 wU-34-f2 23.2 wS-5-f1 21.1
13 wP-23-f2 11.0 wS-29-f1 14.6 wU-24-f2 23.2 w5-O-f2 21.4
14 wP-35-f2 11.0 w28-W-f1 15.1 wN-5-f2 22.6
15 wP-29-f1 11.1 wP-5-f1 15.2 wN-5-f1 22.7
16 w34-V1-f2 11.4 wO-5-f1 21.3 wS-29-f2 22.8
17 wO-29-f1 11.6 wN-30-f2 22.1 wS-29-f1 22.9
18 w24-V1-f2 12.2 wU-24-f2 22.1 wV-24-f2 23.0
19 wO-29-f2 12.9 wU-34-f2 22.4 wP-35-f2 23.9
20 w28-W-f1 12.9 wU-31-f2 23.1 wN-30-f1 24.0
21 w5-O-f2 14.9 wU-27-f2 23.1 w28-W-f2 25.9
22 w5-R-f1 15.1 w5-O-f2 29.6
23 wU-31-f2 20.9 w28-W-f2 29.7
24 wU-27-f2 20.9 wP-23-f2 49.5
25 wN-30-f2 21.0 wP-35-f2 49.6
26 wU-34-f2 21.6
27 wU-24-f2 21.9
28 w28-W-f2 25.1
826

Table G.2: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Concepcion ground motion). Shading indicates that the wall also exceeded the
acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 7.4 wV-34-f1 9.0 wV-34-f1 9.1 wP-5-f1 11.0
2 wV-24-f1 0.0 wV-24-f1 7.8 wS-29-f1 9.1 wV-24-f1 9.6 wO-5-f1 11.1
3 wQ-34-f2 0.0 wV-34-f2 7.9 wV-24-f1 9.1 wN-5-f2 9.8 wO-29-f1 12.0
4 wQ-24-f2 0.0 w28-W-f1 8.7 wN-5-f2 9.3 wN-5-f1 9.8 wV-24-f1 12.2
5 wU-34-f2 0.0 wO-29-f2 8.7 w5-O-f2 9.4 wS-29-f1 10.5 w28-W-f1 12.7
6 wU-24-f2 0.0 wS-29-f1 9.0 wN-5-f1 9.4 wS-5-f1 10.6 wV-34-f1 12.8
7 wV-34-f2 0.7 wS-5-f1 9.0 wS-29-f2 9.5 wS-5-f2 10.6 wP-29-f1 12.9
8 w28-W-f1 2.5 wP-29-f1 9.1 wS-5-f2 9.6 w5-O-f2 10.9 wO-29-f2 14.5
9 wO-29-f2 2.5 wS-5-f2 9.1 wV-34-f2 9.6 wS-29-f2 11.0 wS-5-f2 20.1
10 wP-5-f1 2.5 wV-24-f2 9.1 w28-W-f1 9.6 wO-29-f2 11.0 wP-23-f2 20.8
11 w28-W-f2 2.5 wS-29-f2 9.1 wV-24-f2 9.7 wP-23-f2 11.0 wV-34-f2 21.1
12 wP-29-f1 2.7 wN-5-f2 9.3 wP-5-f1 9.7 wV-34-f2 11.0 wS-5-f1 21.1
13 wN-5-f2 3.0 wU-24-f2 9.3 wO-5-f1 9.8 wP-35-f2 11.0 w5-O-f2 21.4
14 wN-5-f1 3.0 wN-5-f1 9.3 wU-34-f2 9.8 wV-24-f2 11.1 wN-5-f2 22.6
15 wO-5-f1 3.0 w5-O-f2 9.3 wO-29-f1 9.8 w34-V1-f2 11.4 wN-5-f1 22.7
16 wS-29-f1 3.1 wP-35-f2 9.4 wP-29-f1 9.9 wU-34-f2 11.4 wS-29-f2 22.8
17 wS-5-f1 3.1 wP-23-f2 9.4 wO-29-f2 9.9 wP-5-f1 11.6 wS-29-f1 22.9
18 wV-24-f2 3.2 wP-5-f1 9.6 wS-5-f1 10.3 wO-5-f1 11.6 wV-24-f2 23.0
19 wS-29-f2 3.6 w28-W-f2 9.6 wP-35-f2 10.6 wU-24-f2 12.2 wP-35-f2 23.9
20 wS-5-f2 3.6 wU-34-f2 9.8 wP-23-f2 10.9 w24-V1-f2 12.2 wN-30-f1 24.0
21 w5-O-f2 3.7 wO-5-f1 9.8 wU-24-f2 11.2 w28-W-f1 13.0 w28-W-f2 25.9
22 wP-23-f2 4.2 wO-29-f1 9.8 w34-V1-f2 11.3 wP-29-f1 15.1
23 wP-35-f2 4.2 w5-R-f1 11.1 w24-V1-f2 12.0 wU-31-f2 21.0
24 wO-29-f1 4.9 wU-31-f2 11.1 w28-W-f2 13.0 wU-27-f2 21.0
25 wN-30-f1 5.4 wU-27-f2 11.1 wU-31-f2 14.6 wN-30-f2 21.0
26 wU-31-f1 6.2 wU-31-f1 11.1 wU-27-f2 14.6 wO-29-f1 21.2
27 wU-27-f1 6.2 wU-27-f1 11.1 wN-30-f2 14.8 w28-W-f2 22.0
28 w24-R-f1 6.2 w34-V1-f2 11.3 w5-R-f1 15.0
29 w24-T-f1 6.2 wN-30-f2 11.6 wNp-5-f1 15.1
30 wU-31-f2 6.2 w24-V1-f2 12.0 w29-R-f1 15.2
31 wU-27-f2 6.2 w34-R-f1 13.1
32 wNp-5-f1 6.5 w34-T-f1 13.1
33 wU-34-f1 6.5 wN-30-f1 13.1
34 wU-24-f1 6.5 w29-R-f1 13.5
35 w30-Np-f1 6.6 wNp-5-f1 14.9
36 wN-30-f2 6.6 w24-R-f1 15.0
37 w34-R-f1 6.7 w24-T-f1 15.0
38 w34-T-f1 6.7
39 w24-V1-f2 7.4
40 w34-V1-f2 7.8
41 w5-R-f1 7.9
42 w29-R-f1 8.4
43 w30-Np-f2 8.7
44 wQ-34-f1 8.7
45 wQ-24-f1 8.7
46 w25p-O1-f1 9.3
827

Table G.3: Sequence of exceedence of steel and concrete strain limits (Concepcion ground
motion). Shading indicates that the wall also exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 8.7 wV-34-f2 11.4 wV-34-f2 14.6 wP-5-f1 11.0
2 wV-34-f1 9.1 wV-24-f2 12.2 wV-24-f2 21.9 wO-5-f1 11.1
3 wV-24-f2 9.2 wV-34-f1 20.7 wS-29-f2 23.1 wO-29-f1 12.0
4 wNp-5-f1 9.4 wS-29-f2 20.8 wV-24-f1 12.2
5 w5-O-f2 9.4 w24-V1-f2 21.0 w28-W-f1 12.7
6 wP-5-f1 9.6 wS-5-f2 21.9 wV-34-f1 12.8
7 w5-R-f1 9.6 wU-24-f2 22.0 wP-29-f1 12.9
8 w25p-O1-f1 9.6 w28-W-f2 22.0 wO-29-f2 14.5
9 w28-W-f1 9.6 wP-35-f2 22.2 wS-5-f2 20.1
10 wV-24-f1 9.6 wP-23-f2 25.5 wP-23-f2 20.8
11 w24-V1-f2 9.6 wV-34-f2 21.1
12 wV-34-f2 9.6 wS-5-f1 21.1
13 wP-29-f1 9.7 w5-O-f2 21.4
14 wN-5-f2 9.7 wN-5-f2 22.6
15 wN-5-f1 9.8 wN-5-f1 22.7
16 w34-V1-f2 9.8 wS-29-f2 22.8
17 wO-29-f1 9.8 wS-29-f1 22.9
18 wO-29-f2 9.9 wV-24-f2 23.0
19 w29-R-f1 10.3 wP-35-f2 23.9
20 wS-29-f1 10.4 wN-30-f1 24.0
21 wS-29-f2 10.5 w28-W-f2 25.9
22 wS-5-f2 10.5
23 wS-5-f1 10.5
24 wP-35-f2 10.9
25 wP-23-f2 11.0
26 w28-W-f2 11.1
27 wU-34-f2 11.4
28 wN-30-f2 11.6
29 wU-24-f2 12.2
30 w30-Np-f1 15.1
31 wN-30-f1 16.4
32 w30-Np-f2 22.2
33 wU-27-f2 23.1
34 wU-31-f2 23.2
828

G.2 Concepcion (DBE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.5: Summary of walls exceeding acceptance criteria (Concepcion ground motion
scaled to ASCE 7 design spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.6: Walls exceeding
45
ASCE 41 rotation limits (Concepcion ground motionStrain
scaled to ASCE 7 design spectrum).
40

829
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.7: Walls exceeding
45
fragility function probabilities of 0.50 (Concepcion ground motion scaled to ASCE 7 design
spectrum). 40

830
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.8: Walls
45
exceeding strain limits (Concepcion ground motion scaled to ASCE 7 design spectrum).

831
40

35
# of Predicted walls

30

25

20

15
832

Table G.4: Sequence of exceedence of ASCE 41 acceptance criteria (Concepcion ground


motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wV-24-f1 7.9 w28-W-f1 8.9 wV-34-f1 10.6 w28-W-f1 8.5
2 w28-W-f1 8.5 wP-29-f1 9.4 wV-24-f2 10.6 wP-29-f1 8.9
3 wV-34-f1 8.5 wN-5-f2 9.5 wV-34-f2 11.2 w28-W-f2 9.3
4 w5-O-f2 8.8 wN-5-f1 10.0 wV-24-f1 11.3 wO-5-f1 9.4
5 wO-29-f2 8.8 wP-5-f1 10.1 w34-V1-f2 11.5 wO-29-f2 9.5
6 w28-W-f2 8.9 wV-34-f1 10.2 wO-29-f1 10.0
7 wP-29-f1 9.3 wV-24-f1 10.4 wV-24-f1 10.4
8 wP-5-f1 9.3 wV-24-f2 10.4 wP-5-f1 10.9
9 wN-5-f1 9.4 wV-34-f2 10.4 wV-34-f1 11.3
10 wN-5-f2 9.4 wS-29-f1 10.6 wP-23-f2 11.3
11 wP-23-f2 9.5 wS-5-f2 10.6 wN-5-f1 11.5
12 wS-29-f2 9.5 wS-5-f1 11.2 w5-O-f2 11.7
13 wS-5-f2 9.5 wS-29-f2 11.2 w5-R-f1 11.8
14 wS-5-f1 9.5 w34-V1-f2 11.3
15 wS-29-f1 9.5 w24-V1-f2 11.3
16 wP-35-f2 9.5 wO-29-f2 11.6
17 wO-5-f1 9.9 w5-O-f2 11.8
18 wO-29-f1 9.9
19 wV-34-f2 10.3
20 wV-24-f2 10.3
21 w24-V1-f2 10.4
22 w34-V1-f2 10.6
833

Table G.5: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Concepcion ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wN-5-f2 7.2 wV-34-f1 7.5 w28-W-f1 8.5 w28-W-f1 8.5
2 wV-24-f1 0.0 w28-W-f1 7.4 w28-W-f1 7.8 wO-29-f2 8.8 wP-29-f1 8.9
3 wQ-34-f2 0.0 wV-34-f1 7.4 wV-24-f1 7.9 w5-O-f2 8.9 w28-W-f2 9.3
4 wQ-24-f2 0.0 w5-O-f2 7.4 wV-34-f2 7.9 wP-29-f1 9.3 wO-5-f1 9.4
5 wU-34-f2 0.0 wV-24-f1 7.5 wN-5-f2 8.0 wP-5-f1 9.3 wO-29-f2 9.5
6 wU-24-f2 0.0 wS-29-f1 7.5 wP-29-f1 8.5 wV-34-f1 9.3 wO-29-f1 10.0
7 wP-5-f1 0.2 wN-5-f1 7.7 wP-5-f1 8.5 wV-34-f2 9.3 wV-24-f1 10.4
8 wO-5-f1 0.2 wP-5-f1 7.8 wO-29-f2 8.8 wN-5-f2 9.4 wP-5-f1 10.9
9 wS-29-f1 0.2 wP-29-f1 7.8 w5-O-f2 8.8 wN-5-f1 9.4 wV-34-f1 11.3
10 wS-5-f1 0.2 wV-34-f2 7.9 w28-W-f2 8.8 wS-29-f2 9.6 wP-23-f2 11.3
11 wS-5-f2 0.2 wV-24-f2 7.9 wU-34-f2 8.9 wS-5-f2 9.6 wN-5-f1 11.5
12 wV-24-f2 0.3 wU-24-f2 7.9 wS-29-f1 9.1 wV-24-f1 9.6 w5-O-f2 11.7
13 wS-29-f2 0.3 wO-29-f2 8.5 wV-24-f2 9.3 wO-5-f1 9.9 w5-R-f1 11.8
14 w28-W-f1 0.6 wU-34-f2 8.5 w29-R-f1 9.3 wS-29-f1 10.1
15 wV-34-f2 0.6 w28-W-f2 8.5 w5-R-f1 9.3 wS-5-f1 10.1
16 wN-5-f2 1.5 w29-R-f1 8.6 wN-5-f1 9.3 wV-24-f2 10.4
17 wN-5-f1 1.5 wO-29-f1 8.9 w34-V1-f2 9.4 w34-V1-f2 10.6
18 w5-O-f2 1.8 wS-5-f1 9.1 wN-30-f1 9.4 wU-34-f2 10.6
19 wP-29-f1 2.3 wO-5-f1 9.1 wP-23-f2 9.4 w24-V1-f2 11.2
20 wO-29-f2 2.5 wP-23-f2 9.1 wP-35-f2 9.4 wU-24-f2 11.3
21 w28-W-f2 2.5 wS-5-f2 9.1 wS-29-f2 9.5 wO-29-f1 11.8
22 wO-29-f1 2.5 wP-35-f2 9.1 wS-5-f2 9.5 w28-W-f2 11.8
23 wN-30-f1 2.6 wS-29-f2 9.1 wS-5-f1 9.5
24 wP-23-f2 3.7 w5-R-f1 9.3 wO-29-f1 9.5
25 wP-35-f2 3.7 wN-30-f1 9.3 wO-5-f1 9.5
26 w24-R-f1 4.2 w34-R-f1 9.3 wU-24-f2 10.3
27 w24-T-f1 4.2 w34-T-f1 9.3 w24-V1-f2 10.4
28 wNp-5-f1 4.2 w34-V1-f2 9.3 wU-31-f2 11.2
29 wU-31-f1 4.3 w24-R-f1 9.4 wU-27-f2 11.2
30 wU-27-f1 4.3 w24-T-f1 9.4 wN-30-f2 11.5
31 w34-R-f1 4.4 wNp-5-f1 9.5 w34-R-f1 11.7
32 w34-T-f1 4.4 wU-31-f1 10.2 w34-T-f1 11.7
33 wU-24-f1 4.7 wU-27-f1 10.2
34 w30-Np-f1 4.9 w24-V1-f2 10.3
35 wU-34-f1 5.9 wU-31-f2 10.4
36 wU-31-f2 5.9 wU-27-f2 10.4
37 wU-27-f2 5.9 wU-34-f1 11.3
38 w24-V1-f2 6.2 wN-30-f2 11.4
39 wN-30-f2 6.3 wU-24-f1 11.4
40 w5-R-f1 6.6
41 w34-V1-f2 6.7
42 w29-R-f1 6.8
43 wQ-34-f1 7.2
44 w25p-O1-f1 7.3
45 wQ-24-f1 7.3
46 w30-Np-f2 7.4
834

Table G.6: Sequence of exceedence of steel and concrete strain limits (Concepcion ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 7.4 wV-24-f2 10.4 wV-34-f2 11.4 w28-W-f1 8.5
2 w5-O-f2 7.5 wV-34-f2 10.6 wP-29-f1 8.9
3 w5-R-f1 7.8 wS-29-f2 11.3 w28-W-f2 9.3
4 wP-29-f1 7.9 wV-34-f1 11.3 wO-5-f1 9.4
5 w25p-O1-f1 7.9 w34-V1-f2 11.7 wO-29-f2 9.5
6 w28-W-f1 7.9 wP-29-f1 11.7 wO-29-f1 10.0
7 wNp-5-f1 7.9 wV-24-f1 10.4
8 wV-24-f1 7.9 wP-5-f1 10.9
9 w24-V1-f2 7.9 wV-34-f1 11.3
10 w29-R-f1 8.5 wP-23-f2 11.3
11 w34-V1-f2 8.5 wN-5-f1 11.5
12 wP-5-f1 8.5 w5-O-f2 11.7
13 wV-34-f2 8.5 w5-R-f1 11.8
14 w28-W-f2 8.5
15 wO-29-f2 8.7
16 wV-34-f1 9.3
17 wN-5-f1 9.3
18 wU-34-f2 9.3
19 wN-5-f2 9.4
20 wP-35-f2 9.4
21 wP-23-f2 9.4
22 wS-29-f1 9.5
23 wS-5-f1 9.5
24 wS-5-f2 9.5
25 wS-29-f2 9.5
26 wO-29-f1 9.9
27 wN-30-f1 10.0
28 wV-24-f2 10.2
29 wU-24-f2 10.4
30 wN-30-f2 11.4
835

G.3 Concepcion (MCE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.9: Summary of walls exceeding acceptance criteria (Concepcion ground motion
scaled to ASCE 7 MCE spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure G.10: Walls 45exceedingASCE


ASCERotation
41 rotation limitsFragility Functions
(Concepcion Strain
ground motion scaled to ASCE 7 MCE spectrum).
40

836
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.11: Walls exceeding
45
fragility function probabilities of 0.50 (Concepcion Strain
ground motion scaled to ASCE 7 MCE
spectrum). 40

837
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.12: 45
Walls exceeding strain limits (Concepcion ground motion scaled to ASCE 7 MCE spectrum).

838
40

35
# of Predicted walls

30

25

20

15
839

Table G.7: Sequence of exceedence of ASCE 41 acceptance criteria (Concepcion ground


motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wV-34-f1 7.3 w28-W-f1 7.9 w28-W-f1 8.1 wO-5-f1 7.9
2 wP-5-f1 7.4 wP-5-f1 8.0 wN-5-f1 10.3 w28-W-f2 7.9
3 wN-5-f1 7.4 wP-29-f1 8.0 wP-5-f1 10.3 wN-5-f1 8.0
4 wO-5-f1 7.4 wV-24-f1 8.0 wO-5-f1 10.3 wP-29-f1 8.6
5 wS-29-f1 7.5 w24-V1-f2 8.1 wS-29-f1 10.4 w28-W-f1 9.0
6 w28-W-f1 7.5 wV-34-f1 9.1 wS-5-f2 10.4 wO-29-f1 9.1
7 w5-O-f2 7.5 wS-5-f2 9.1 wS-5-f1 10.4 wO-29-f2 9.8
8 wS-5-f2 7.5 wS-29-f1 9.1 wV-34-f1 10.4 wP-5-f1 9.8
9 wO-29-f2 7.5 wN-5-f1 9.2 wP-29-f1 10.5 wV-34-f1 10.1
10 wS-29-f2 7.5 wS-5-f1 9.2 wV-24-f1 10.5 wS-5-f1 10.2
11 wN-5-f2 7.7 wO-5-f1 9.2 wV-24-f2 10.5 wV-24-f1 10.2
12 wP-29-f1 7.9 wN-5-f2 9.2 wS-29-f2 10.5 wP-35-f2 10.5
13 wV-24-f1 7.9 wP-35-f2 9.2 wV-34-f2 10.5 wV-34-f2 11.0
14 wS-5-f1 7.9 wS-29-f2 9.2 wO-29-f1 10.5 wS-5-f2 11.1
15 wV-34-f2 8.0 wP-23-f2 9.6 w24-V1-f2 10.6 wS-29-f2 11.2
16 w24-V1-f2 8.0 wV-34-f2 9.8 w34-V1-f2 10.7 wS-29-f1 11.3
17 w34-V1-f2 8.0 w34-V1-f2 10.1 wN-5-f2 10.9 wN-30-f1 11.4
18 w24-R-f1 8.1 wO-29-f1 10.3 wP-23-f2 11.1 w29-R-f1 11.7
19 w24-T-f1 8.1 wV-24-f2 10.4 wP-35-f2 11.1 w5-O-f2 11.7
20 w34-R-f1 8.1 wU-31-f1 10.5 wN-30-f2 11.1 wP-23-f2 11.7
21 w34-T-f1 8.1 wU-27-f1 10.5 wU-34-f2 11.5 wN-5-f2 11.8
22 w29-R-f1 8.1 wU-34-f2 11.0 wN-30-f2 11.9
23 w5-R-f1 8.3 wN-30-f2 11.0
24 wP-35-f2 9.1 wNp-5-f1 11.1
25 wV-24-f2 9.1 wO-29-f2 11.1
26 wO-29-f1 9.1 w29-R-f1 11.8
27 wP-23-f2 9.1 w5-O-f2 12.4
28 wNp-5-f1 10.3 w5-R-f1 12.6
29 wN-30-f1 10.3 w34-R-f1 12.8
30 wU-31-f1 10.5 w34-T-f1 12.8
31 wU-27-f1 10.5
32 wU-34-f1 10.5
33 wU-24-f1 10.7
34 w28-W-f2 10.8
35 wU-34-f2 10.9
36 wN-30-f2 10.9
37 w30-Np-f1 11.0
38 wU-31-f2 11.5
39 wU-27-f2 11.5
40 wU-24-f2 12.9
840

Table G.8: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Concepcion ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 6.2 w28-W-f1 7.2 w28-W-f1 7.5 wO-5-f1 7.9
2 wV-24-f1 0.0 wV-24-f1 6.5 wP-5-f1 7.2 w5-O-f2 7.5 w28-W-f2 7.9
3 wQ-34-f2 0.0 w28-W-f1 6.8 w5-O-f2 7.3 wO-29-f2 7.6 wN-5-f1 8.0
4 wQ-24-f2 0.0 wN-5-f2 7.0 wN-5-f2 7.3 wN-5-f2 7.7 wP-29-f1 8.6
5 wU-34-f2 0.0 w5-O-f2 7.0 wN-5-f1 7.3 wP-5-f1 7.9 w28-W-f1 9.0
6 wU-24-f2 0.0 wN-5-f1 7.0 wV-34-f1 7.3 wP-29-f1 7.9 wO-29-f1 9.1
7 wP-5-f1 0.1 wP-5-f1 7.1 wS-29-f1 7.3 wV-24-f1 7.9 wO-29-f2 9.8
8 wO-5-f1 0.1 wP-29-f1 7.1 wS-5-f1 7.3 wV-24-f2 7.9 wP-5-f1 9.8
9 wP-29-f1 0.2 wV-24-f2 7.2 wO-29-f2 7.3 wV-34-f1 8.0 wV-34-f1 10.1
10 wS-29-f1 0.2 w28-W-f2 7.2 wO-5-f1 7.4 w24-V1-f2 8.0 wS-5-f1 10.2
11 wS-5-f2 0.2 wO-29-f2 7.3 wP-35-f2 7.4 wU-24-f2 8.0 wV-24-f1 10.2
12 wS-5-f1 0.2 wP-35-f2 7.3 wS-5-f2 7.4 wV-34-f2 8.0 wP-35-f2 10.5
13 wS-29-f2 0.2 wP-23-f2 7.3 wP-29-f1 7.4 w34-V1-f2 8.1 wV-34-f2 11.0
14 wV-24-f2 0.2 wS-29-f1 7.3 wS-29-f2 7.4 wS-29-f1 9.0 wS-5-f2 11.1
15 w28-W-f1 0.2 wS-5-f1 7.3 wO-29-f1 7.4 wO-5-f1 9.0 wS-29-f2 11.2
16 wO-29-f2 0.6 wS-29-f2 7.3 wV-24-f1 7.5 wS-5-f1 9.0 wS-29-f1 11.3
17 wV-34-f2 0.6 wS-5-f2 7.3 w28-W-f2 7.5 wN-5-f1 9.0 wN-30-f1 11.4
18 w28-W-f2 0.6 wO-5-f1 7.3 wV-34-f2 7.5 wS-5-f2 9.0 w29-R-f1 11.7
19 wN-5-f2 0.6 wO-29-f1 7.3 wU-24-f2 7.5 wS-29-f2 9.1 w5-O-f2 11.7
20 wN-5-f1 0.6 wV-34-f2 7.5 wV-24-f2 7.7 wP-35-f2 9.1 wP-23-f2 11.7
21 w5-O-f2 1.8 wU-24-f2 7.5 wP-23-f2 7.8 wO-29-f1 9.2 wN-5-f2 11.8
22 wP-23-f2 2.2 wU-34-f2 7.7 w24-V1-f2 8.0 wP-23-f2 9.2 wN-30-f2 11.9
23 wN-30-f1 2.5 w24-V1-f2 7.7 wN-30-f1 8.0 wU-34-f2 10.1
24 wO-29-f1 2.5 wNp-5-f1 7.7 w29-R-f1 8.0 wN-30-f1 10.3
25 w34-R-f1 2.5 wN-30-f1 7.9 w24-R-f1 8.0 wNp-5-f1 10.3
26 w34-T-f1 2.5 w29-R-f1 7.9 w24-T-f1 8.0 wU-31-f1 10.5
27 wP-35-f2 2.5 w24-R-f1 8.0 w5-R-f1 8.0 wU-27-f1 10.5
28 w30-Np-f1 2.6 w24-T-f1 8.0 wU-34-f2 8.0 w29-R-f1 10.6
29 wU-31-f1 2.9 w5-R-f1 8.0 w34-R-f1 8.0 wU-34-f1 10.7
30 wU-27-f1 2.9 w34-R-f1 8.0 w34-T-f1 8.0 wU-24-f1 10.9
31 w24-R-f1 3.7 w34-T-f1 8.0 w34-V1-f2 8.0 wN-30-f2 11.0
32 w24-T-f1 3.7 w30-Np-f1 8.0 w30-Np-f1 8.1 w30-Np-f2 11.4
33 wNp-5-f1 3.7 w34-V1-f2 8.0 wNp-5-f1 10.3 w34-R-f1 11.8
34 wU-34-f1 4.3 w25p-O1-f1 8.1 wU-31-f1 10.4 w25p-O1-f1 12.5
35 wU-24-f1 4.4 wU-31-f1 9.1 wU-27-f1 10.4 w24-R-f1 12.5
36 wU-31-f2 4.7 wU-27-f1 9.1 wU-31-f2 10.4 w24-T-f1 12.5
37 wU-27-f2 4.7 wN-30-f2 9.2 wU-27-f2 10.4 w28-W-f2 12.5
38 w34-V1-f2 5.7 wU-31-f2 9.7 wU-24-f1 10.5 w5-R-f1 12.5
39 w24-V1-f2 6.1 wU-27-f2 9.7 wU-34-f1 10.5 w34-T-f1 12.8
40 wN-30-f2 6.3 wU-24-f1 10.4 wN-30-f2 10.9 w30-Np-f1 12.8
41 w5-R-f1 6.4 wU-34-f1 10.4 w30-Np-f2 11.2
42 wQ-24-f1 6.5 wQ-24-f1 10.5 w25p-O1-f1 12.4
43 w29-R-f1 6.5 wQ-34-f1 10.5
44 w30-Np-f2 6.8 w30-Np-f2 11.1
45 wQ-34-f1 6.8 wQ-34-f2 11.2
46 w25p-O1-f1 6.9
841

Table G.9: Sequence of exceedence of steel and concrete strain limits (Concepcion ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 6.8 w28-W-f1 8.0 wV-24-f2 10.5 wO-5-f1 7.9
2 w5-R-f1 7.2 wO-29-f2 9.6 wV-24-f1 10.5 w28-W-f2 7.9
3 wP-29-f1 7.2 wV-34-f2 9.9 wV-34-f1 10.5 wN-5-f1 8.0
4 w28-W-f1 7.2 wO-29-f1 10.3 wS-5-f1 10.7 wP-29-f1 8.6
5 w29-R-f1 7.2 wP-29-f1 10.4 wO-29-f2 11.1 w28-W-f1 9.0
6 wP-5-f1 7.2 wP-35-f2 10.4 wP-23-f2 11.2 wO-29-f1 9.1
7 wNp-5-f1 7.2 wS-5-f1 10.4 wN-5-f2 11.2 wO-29-f2 9.8
8 w25p-O1-f1 7.2 wV-24-f2 10.4 wP-35-f2 11.2 wP-5-f1 9.8
9 w5-O-f2 7.2 wS-5-f2 10.4 wS-29-f2 11.4 wV-34-f1 10.1
10 wO-29-f2 7.3 wV-24-f1 10.4 wV-34-f2 11.8 wS-5-f1 10.2
11 wS-29-f1 7.3 wV-34-f1 10.5 wS-29-f1 11.8 wV-24-f1 10.2
12 wV-34-f1 7.4 wP-5-f1 10.5 wP-35-f2 10.5
13 wN-5-f1 7.4 w24-V1-f2 10.5 wV-34-f2 11.0
14 wN-5-f2 7.4 w34-V1-f2 10.7 wS-5-f2 11.1
15 wO-29-f1 7.4 wU-34-f2 10.7 wS-29-f2 11.2
16 wS-5-f1 7.5 wN-5-f2 11.0 wS-29-f1 11.3
17 wS-29-f2 7.5 wP-23-f2 11.1 wN-30-f1 11.4
18 wV-24-f2 7.5 wO-5-f1 11.1 w29-R-f1 11.7
19 w24-V1-f2 7.5 wS-29-f2 11.2 w5-O-f2 11.7
20 w28-W-f2 7.5 wN-30-f2 11.2 wP-23-f2 11.7
21 wS-5-f2 7.5 wS-29-f1 11.5 wN-5-f2 11.8
22 wV-24-f1 7.5 w29-R-f1 11.8 wN-30-f2 11.9
23 w34-V1-f2 7.9 w5-O-f2 11.9
24 wP-35-f2 7.9 wU-24-f2 12.4
25 w30-Np-f1 8.0
26 wV-34-f2 8.0
27 wU-34-f2 8.1
28 wU-24-f2 8.1
29 wP-23-f2 9.1
30 wN-30-f1 9.2
31 wN-30-f2 9.6
32 wU-24-f1 10.4
33 wQ-24-f1 10.5
34 wU-34-f1 10.5
35 wQ-34-f1 10.5
36 wU-27-f1 10.7
37 wU-31-f1 10.7
38 w34-R-f1 10.7
39 w34-T-f1 10.7
40 w24-T-f1 11.0
41 wQ-34-f2 11.1
42 w30-Np-f2 11.2
43 w24-R-f1 12.4
842

G.4 San Pedro de la Paz (Unscaled)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.13: Summary of walls exceeding acceptance criteria (San Pedro de la Paz ground
motion).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure 45G.14: Walls exceeding ASCE 41 rotation limits (San Pedro deStrain
la Paz ground motion).
40

843
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.15: 45Walls exceeding fragility function probabilities of 0.50 (San Pedro de la Paz ground motion).
40

844
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation
Figure
45
G.16: Walls exceeding strainFragility
limits Functions Strain
(San Pedro de la Paz ground motion).

845
40

35
# of Predicted walls

30

25

20

15
846

Table G.10: Sequence of exceedence of ASCE 41 acceptance criteria (San Pedro de la Paz
ground motion). Shading indicates that the wall also exceeded the acceptance criterion for
shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 w28-W-f1 9.0 w28-W-f1 9.9 w28-W-f1 18.6 w28-W-f1 9.8
2 wN-5-f2 9.1 wN-5-f2 10.0 wV-34-f1 26.2 wO-29-f1 10.0
3 wN-5-f1 9.2 wO-29-f2 11.1 wS-29-f1 26.2 wP-29-f1 10.3
4 wP-5-f1 9.2 wP-5-f1 11.6 wN-5-f2 26.7 wO-5-f1 10.6
5 wV-34-f1 9.2 wP-29-f1 11.6 wN-5-f1 27.4 wP-5-f1 11.4
6 wS-5-f2 9.2 wV-34-f1 11.7 wP-5-f1 27.4 wO-29-f2 11.5
7 wS-29-f1 9.2 wN-5-f1 13.9 w28-W-f2 11.5
8 wS-29-f2 9.2 wS-29-f1 17.7 wN-5-f1 11.6
9 wP-29-f1 9.6 wP-35-f2 17.7 w5-O-f2 12.6
10 wV-24-f1 9.7 wS-5-f1 17.7 wV-34-f2 14.6
11 wO-5-f1 10.0 w5-O-f2 18.6 wV-24-f1 14.7
12 wP-35-f2 10.0 w24-V1-f2 18.6 wS-5-f2 14.7
13 wO-29-f1 10.1 wP-23-f2 18.6 wV-34-f1 15.3
14 w5-O-f2 10.2 wS-29-f2 20.4 wP-35-f2 17.7
15 wO-29-f2 10.2 wV-24-f1 20.6 w29-R-f1 18.1
16 w34-V1-f2 11.1 wO-5-f1 21.0 wS-5-f1 18.6
17 w28-W-f2 11.1 wO-29-f1 21.1 wN-5-f2 175.3
18 wV-24-f2 11.7 wV-34-f2 25.1
19 wS-5-f1 11.9 wS-5-f2 26.1
20 wV-34-f2 12.0 wV-24-f2 26.2
21 w24-V1-f2 12.6 w34-V1-f2 27.2
22 wP-23-f2 17.0
23 w5-R-f1 18.6
24 w24-R-f1 18.6
25 w24-T-f1 18.6
26 w29-R-f1 23.2
27 wN-30-f2 26.7
28 wN-30-f1 27.4
29 wNp-5-f1 27.5
30 w34-R-f1 50.5
31 w34-T-f1 50.5
847

Table G.11: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (San Pedro de la Paz ground motion). Shading indicates that the wall also ex-
ceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 w5-O-f2 8.2 w28-W-f1 8.7 w28-W-f1 9.0 w28-W-f1 9.8
2 wV-24-f1 0.0 wN-5-f2 8.2 wO-29-f2 8.8 w5-O-f2 9.2 wO-29-f1 10.0
3 wQ-34-f2 0.0 wN-5-f1 8.4 wP-5-f1 9.0 wN-5-f2 9.2 wP-29-f1 10.3
4 wQ-24-f2 0.0 w28-W-f1 8.6 wP-29-f1 9.0 wN-5-f1 9.5 wO-5-f1 10.6
5 wU-34-f2 0.0 wO-29-f2 8.7 wV-24-f2 9.0 wV-24-f1 9.8 wP-5-f1 11.4
6 wU-24-f2 0.0 wP-5-f1 8.7 wN-5-f2 9.0 wP-5-f1 9.8 wO-29-f2 11.5
7 wV-34-f2 2.4 wP-29-f1 8.7 w5-O-f2 9.0 wP-29-f1 9.9 w28-W-f2 11.5
8 w28-W-f1 2.4 w28-W-f2 8.7 wN-5-f1 9.1 wV-24-f2 9.9 wN-5-f1 11.6
9 wP-29-f1 2.6 wO-5-f1 8.8 wS-29-f2 9.1 wO-5-f1 10.0 w5-O-f2 12.6
10 wP-5-f1 2.9 wO-29-f1 8.8 wS-5-f2 9.1 wO-29-f2 10.3 wV-34-f2 14.6
11 wN-5-f1 3.1 wV-24-f1 8.9 wV-34-f1 9.1 wV-34-f1 10.6 wV-24-f1 14.7
12 w5-O-f2 3.1 wV-24-f2 9.0 wP-35-f2 9.1 wV-34-f2 10.6 wS-5-f2 14.7
13 wO-5-f1 3.2 wNp-5-f1 9.0 wS-29-f1 9.2 w34-V1-f2 11.1 wV-34-f1 15.3
14 wN-5-f2 3.3 wU-24-f2 9.0 wS-5-f1 9.2 wU-34-f2 11.6 wP-35-f2 17.7
15 wO-29-f2 3.5 wP-23-f2 9.0 wO-5-f1 9.2 wS-29-f1 11.9 w29-R-f1 18.1
16 wV-24-f2 3.7 wP-35-f2 9.1 wV-24-f1 9.3 w24-V1-f2 12.6 wS-5-f1 18.6
17 w28-W-f2 3.9 wV-34-f2 9.1 wU-24-f2 9.5 wU-24-f2 12.6 wN-5-f2 175.3
18 wP-23-f2 3.9 wS-29-f2 9.1 w28-W-f2 9.5 wS-5-f1 14.7
19 wO-29-f1 3.9 wS-5-f2 9.1 wO-29-f1 9.5 wS-29-f2 17.0
20 wS-29-f1 4.4 wS-29-f1 9.1 wP-23-f2 9.5 wS-5-f2 17.1
21 wS-5-f1 5.0 wS-5-f1 9.1 wV-34-f2 9.7 wP-35-f2 17.5
22 wP-35-f2 5.0 wV-34-f1 9.1 w29-R-f1 9.8 wP-23-f2 17.6
23 wS-29-f2 5.0 w29-R-f1 9.8 wU-34-f2 10.6 wO-29-f1 17.7
24 wS-5-f2 5.0 wN-30-f1 9.8 w34-V1-f2 11.1 w28-W-f2 18.6
25 wN-30-f1 5.0 wU-34-f2 9.9 wN-30-f1 11.6 wN-30-f2 26.8
26 w34-R-f1 5.5 w24-V1-f2 9.9 w5-R-f1 11.6 wN-30-f1 27.5
27 w34-T-f1 5.5 w34-V1-f2 10.6 w34-R-f1 11.6 w29-R-f1 29.3
28 wNp-5-f1 5.5 w34-R-f1 10.6 w34-T-f1 11.6 w5-R-f1 58.3
29 w30-Np-f1 5.5 w34-T-f1 10.6 w24-V1-f2 12.1
30 wU-31-f1 6.3 w5-R-f1 11.5 wNp-5-f1 12.6
31 wU-27-f1 6.3 w24-R-f1 11.6 w24-R-f1 18.6
32 wU-34-f1 6.3 w24-T-f1 11.6 w24-T-f1 18.6
33 w24-R-f1 6.3 wU-31-f1 14.7 wN-30-f2 18.6
34 w24-T-f1 6.3 wU-27-f1 14.7 wU-31-f1 26.2
35 wN-30-f2 6.6 wN-30-f2 17.5 wU-27-f1 26.2
36 wU-24-f1 6.6 w30-Np-f1 18.6
37 wU-31-f2 7.0 wU-31-f2 25.3
38 wU-27-f2 7.0 wU-27-f2 25.3
39 w24-V1-f2 7.4 wU-34-f1 26.2
40 wQ-34-f1 7.5 wU-24-f1 26.2
41 w29-R-f1 7.8 w25p-O1-f1 27.4
42 w5-R-f1 7.8
43 wQ-24-f1 7.9
44 w30-Np-f2 8.2
45 w25p-O1-f1 8.4
46 w34-V1-f2 8.5
848

Table G.12: Sequence of exceedence of steel and concrete strain limits (San Pedro de la Paz
ground motion). Shading indicates that the wall also exceeded the acceptance criterion for
shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wNp-5-f1 8.4 wV-34-f2 11.1 wV-34-f2 27.2 w28-W-f1 9.8
2 wO-5-f1 8.6 wV-24-f2 13.1 wO-29-f1 10.0
3 wO-29-f2 8.7 wP-5-f1 22.7 wP-29-f1 10.3
4 wP-29-f1 8.7 w5-R-f1 22.7 wO-5-f1 10.6
5 w28-W-f1 8.7 wV-24-f1 26.2 wP-5-f1 11.4
6 w5-O-f2 8.7 wV-34-f1 26.3 wO-29-f2 11.5
7 w5-R-f1 9.0 w24-V1-f2 26.5 w28-W-f2 11.5
8 w25p-O1-f1 9.0 wO-29-f2 26.8 wN-5-f1 11.6
9 w29-R-f1 9.0 wP-23-f2 26.8 w5-O-f2 12.6
10 wP-5-f1 9.0 wS-29-f2 26.9 wV-34-f2 14.6
11 w24-V1-f2 9.0 wV-24-f1 14.7
12 wN-5-f1 9.1 wS-5-f2 14.7
13 wN-5-f2 9.1 wV-34-f1 15.3
14 w28-W-f2 9.1 wP-35-f2 17.7
15 wS-29-f2 9.2 w29-R-f1 18.1
16 wS-29-f1 9.2 wS-5-f1 18.6
17 wV-24-f2 9.4 wN-5-f2 175.3
18 wO-29-f1 9.5
19 wP-23-f2 9.6
20 wP-35-f2 9.6
21 wV-24-f1 9.8
22 w34-V1-f2 9.9
23 wV-34-f1 10.4
24 wV-34-f2 10.6
25 wU-34-f2 10.6
26 w30-Np-f1 11.6
27 wS-5-f1 11.7
28 wS-5-f2 11.9
29 wU-24-f2 12.1
30 wN-30-f1 13.9
31 wN-30-f2 17.5
849

G.5 San Pedro de la Paz (DBE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.17: Summary of walls exceeding acceptance criteria (San Pedro de la Paz ground
motion scaled to ASCE 7 design spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation
Figure G.18: Walls exceeding
45
ASCE 41 rotation limits (SanFragility
Pedro Functions Strain
de la Paz ground motion scaled to ASCE 7 design spectrum).
40

850
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.19: Walls exceeding
45
fragility function probabilities of 0.50 (San Pedro de Strain
la Paz ground motion scaled to ASCE 7
design spectrum). 40

851
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation
Figure G.20: Walls45 exceeding strain limits (San PedroFragility
de Functions
la Paz ground motion Strain
scaled to ASCE 7 design spectrum).

852
40

35
# of Predicted walls

30

25

20

15
853

Table G.13: Sequence of exceedence of ASCE 41 acceptance criteria (San Pedro de la Paz
ground motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 w28-W-f1 8.7 w28-W-f1 9.1 w28-W-f1 17.8 w28-W-f1 9.0
2 wO-29-f2 8.7 wN-5-f2 9.2 wN-5-f1 18.2 wO-29-f1 9.3
3 wP-5-f1 9.1 wP-5-f1 10.0 wV-34-f1 18.7 wO-29-f2 9.8
4 wN-5-f2 9.1 wP-29-f1 10.0 wP-5-f1 26.2 wN-5-f1 10.0
5 wP-29-f1 9.1 wN-5-f1 11.4 wO-5-f1 26.2 wP-29-f1 10.3
6 w5-O-f2 9.1 wV-24-f1 11.4 wS-29-f1 26.2 wO-5-f1 11.4
7 wP-35-f2 9.2 wV-34-f1 11.7 wP-29-f1 26.2 w5-O-f2 11.4
8 wN-5-f1 9.2 wV-34-f2 12.0 wS-5-f1 26.2 wV-24-f1 12.0
9 wS-5-f2 9.2 wV-24-f2 12.0 wS-5-f2 26.2 wP-5-f1 12.0
10 wS-29-f2 9.2 wO-29-f2 13.1 wV-24-f2 26.3 w29-R-f1 12.1
11 wV-34-f1 9.2 wO-5-f1 13.8 wV-24-f1 26.3 wS-5-f2 12.5
12 wS-29-f1 9.2 wS-29-f1 14.0 wV-34-f2 26.3 wV-34-f2 12.8
13 wS-5-f1 9.2 wS-5-f1 14.4 wN-5-f2 26.7 wV-34-f1 12.9
14 wO-5-f1 9.2 w34-V1-f2 14.6 wP-23-f2 26.8 w28-W-f2 13.4
15 wP-23-f2 9.2 w5-O-f2 17.7 wP-35-f2 26.8 wS-5-f1 14.0
16 wO-29-f1 9.7 w24-V1-f2 17.7 w34-V1-f2 28.4 wP-35-f2 17.7
17 wV-24-f1 9.9 wO-29-f1 18.2 w29-R-f1 28.4 wS-29-f1 18.3
18 w24-V1-f2 10.0 wP-35-f2 18.7 wS-29-f2 23.2
19 w5-R-f1 10.0 wS-5-f2 18.7
20 wV-34-f2 11.2 wS-29-f2 18.7
21 wV-24-f2 11.9 wP-23-f2 20.4
22 w34-V1-f2 13.1 w29-R-f1 22.7
23 w24-R-f1 17.6 wN-30-f1 26.2
24 w24-T-f1 17.6 wN-30-f2 26.8
25 w28-W-f2 17.6 wU-34-f2 28.4
26 wN-30-f1 18.2
27 wNp-5-f1 18.2
28 w29-R-f1 18.7
29 w34-R-f1 18.7
30 w34-T-f1 18.7
31 wU-31-f1 26.2
32 wU-27-f1 26.2
33 wN-30-f2 26.8
34 wU-34-f2 28.4
35 w30-Np-f1 28.4
854

Table G.14: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (San Pedro de la Paz ground motion scaled to ASCE 7 design spectrum). Shading
indicates that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 7.4 w28-W-f1 8.5 w28-W-f1 8.8 w28-W-f1 9.0
2 wV-24-f1 0.0 w5-O-f2 8.1 w5-O-f2 8.5 wN-5-f2 9.1 wO-29-f1 9.3
3 wQ-34-f2 0.0 wN-5-f2 8.1 wO-29-f2 8.7 wP-5-f1 9.1 wO-29-f2 9.8
4 wQ-24-f2 0.0 wN-5-f1 8.4 w28-W-f2 8.7 w5-O-f2 9.1 wN-5-f1 10.0
5 wU-34-f2 0.0 w28-W-f1 8.4 wN-5-f1 8.8 wP-29-f1 9.1 wP-29-f1 10.3
6 wU-24-f2 0.0 wNp-5-f1 8.4 wN-5-f2 8.8 wV-24-f2 9.1 wO-5-f1 11.4
7 wV-34-f2 2.0 wP-5-f1 8.4 wP-5-f1 8.8 wO-29-f2 9.2 w5-O-f2 11.4
8 w28-W-f1 2.2 wP-29-f1 8.4 wS-29-f2 8.9 wN-5-f1 9.2 wV-24-f1 12.0
9 wP-29-f1 2.2 wO-29-f2 8.4 wS-5-f1 8.9 wP-35-f2 9.2 wP-5-f1 12.0
10 wP-5-f1 2.4 wS-29-f1 8.4 wS-5-f2 8.9 wS-29-f2 9.2 w29-R-f1 12.1
11 wO-29-f2 2.4 wS-5-f2 8.4 wV-24-f1 8.9 wS-5-f2 9.2 wS-5-f2 12.5
12 wV-24-f2 2.6 wS-5-f1 8.5 wP-29-f1 9.0 wS-29-f1 9.3 wV-34-f2 12.8
13 w28-W-f2 2.6 wS-29-f2 8.5 wV-24-f2 9.0 wV-34-f1 9.3 wV-34-f1 12.9
14 wN-5-f1 2.7 wP-35-f2 8.5 wP-23-f2 9.1 wS-5-f1 9.3 w28-W-f2 13.4
15 wO-5-f1 2.8 wP-23-f2 8.5 wP-35-f2 9.1 wO-5-f1 9.3 wS-5-f1 14.0
16 wO-29-f1 2.9 w28-W-f2 8.7 wU-24-f2 9.1 wV-24-f1 10.0 wP-35-f2 17.7
17 wN-30-f1 3.0 wU-24-f2 8.7 wV-34-f2 9.1 w24-V1-f2 10.0 wS-29-f1 18.3
18 w5-O-f2 3.0 wO-5-f1 8.8 wS-29-f1 9.2 wV-34-f2 11.9 wS-29-f2 23.2
19 wP-23-f2 3.0 wO-29-f1 8.8 wV-34-f1 9.2 wU-24-f2 12.0
20 wN-5-f2 3.0 wV-24-f2 8.8 wO-5-f1 9.2 w34-V1-f2 13.1
21 wS-29-f1 3.7 wU-34-f2 8.8 wO-29-f1 9.2 wO-29-f1 13.2
22 wS-5-f1 3.7 wV-24-f1 8.8 wNp-5-f1 9.6 wU-34-f2 14.6
23 wS-29-f2 3.9 wV-34-f2 9.0 w24-V1-f2 10.0 wP-23-f2 17.6
24 wS-5-f2 3.9 wN-30-f1 9.1 w5-R-f1 10.0 w28-W-f2 17.7
25 wNp-5-f1 3.9 w5-R-f1 9.1 wN-30-f1 10.0 wN-30-f1 18.2
26 w24-R-f1 3.9 w24-V1-f2 10.0 wU-34-f2 11.2 wNp-5-f1 18.2
27 w24-T-f1 3.9 w24-R-f1 10.0 w34-V1-f2 12.8 w29-R-f1 18.7
28 w34-R-f1 4.0 w24-T-f1 10.0 w29-R-f1 17.6 wU-31-f1 26.3
29 w34-T-f1 4.0 w34-R-f1 10.0 w24-R-f1 17.6 wU-27-f1 26.3
30 wP-35-f2 4.6 w34-T-f1 10.0 w24-T-f1 17.6 wN-30-f2 26.8
31 w30-Np-f1 5.0 w29-R-f1 10.2 w34-R-f1 17.8 w34-R-f1 28.4
32 wU-31-f1 5.8 w34-V1-f2 11.4 w34-T-f1 17.8 w34-T-f1 28.4
33 wU-27-f1 5.8 wN-30-f2 11.4 w25p-O1-f1 18.3 w25p-O1-f1 28.5
34 wU-34-f1 6.3 wU-31-f1 13.7 w30-Np-f1 18.7
35 wU-24-f1 6.3 wU-27-f1 13.7 wN-30-f2 20.3
36 wN-30-f2 6.3 w30-Np-f1 17.8 wU-31-f1 26.2
37 wU-31-f2 6.3 w25p-O1-f1 17.9 wU-27-f1 26.2
38 wU-27-f2 6.3 wU-24-f1 18.2 wU-24-f1 26.2
39 w34-V1-f2 6.4 wU-34-f1 18.7 wU-34-f1 26.2
40 w24-V1-f2 6.8 wU-31-f2 20.0 wU-31-f2 26.3
41 wQ-34-f1 7.4 wU-27-f2 20.0 wU-27-f2 26.3
42 wQ-24-f1 7.5 wQ-34-f1 26.2
43 w29-R-f1 7.5 wQ-24-f1 26.2
44 w5-R-f1 7.8
45 w25p-O1-f1 7.8
46 w30-Np-f2 8.0
855

Table G.15: Sequence of exceedence of steel and concrete strain limits (San Pedro de la Paz
ground motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wNp-5-f1 8.4 wV-24-f2 11.9 w29-R-f1 22.6 w28-W-f1 9.0
2 wP-5-f1 8.4 w29-R-f1 13.1 wP-5-f1 22.9 wO-29-f1 9.3
3 w5-O-f2 8.4 w5-R-f1 13.1 wV-24-f2 26.3 wO-29-f2 9.8
4 w29-R-f1 8.5 wP-5-f1 13.2 wO-29-f2 26.9 wN-5-f1 10.0
5 wO-5-f1 8.6 wV-34-f2 14.6 w28-W-f1 28.4 wP-29-f1 10.3
6 w25p-O1-f1 8.7 wO-29-f1 17.7 wP-29-f1 30.7 wO-5-f1 11.4
7 w28-W-f1 8.7 wP-29-f1 18.2 w5-O-f2 11.4
8 wO-29-f2 8.7 w5-O-f2 18.2 wV-24-f1 12.0
9 wP-29-f1 8.7 wO-5-f1 20.5 wP-5-f1 12.0
10 w28-W-f2 8.7 w28-W-f1 22.3 w29-R-f1 12.1
11 wN-5-f1 8.8 wV-34-f1 22.3 wS-5-f2 12.5
12 wN-5-f2 8.8 wS-5-f1 26.2 wV-34-f2 12.8
13 w5-R-f1 9.0 wV-24-f1 26.2 wV-34-f1 12.9
14 w24-V1-f2 9.0 wS-5-f2 26.3 w28-W-f2 13.4
15 wV-24-f2 9.1 wO-29-f2 26.8 wS-5-f1 14.0
16 w34-V1-f2 9.1 wP-23-f2 26.8 wP-35-f2 17.7
17 wV-34-f1 9.1 wP-35-f2 26.8 wS-29-f1 18.3
18 wP-23-f2 9.1 wN-5-f2 26.8 wS-29-f2 23.2
19 wP-35-f2 9.2 wS-29-f2 27.0
20 wS-29-f2 9.2 w34-V1-f2 28.4
21 wS-5-f1 9.2 wU-34-f2 28.4
22 wS-5-f2 9.2
23 wS-29-f1 9.2
24 wV-34-f2 9.3
25 wO-29-f1 9.6
26 wV-24-f1 10.0
27 wU-34-f2 11.2
28 wN-30-f1 11.4
29 wN-30-f2 11.4
30 wU-24-f2 11.9
31 w30-Np-f1 13.1
32 w34-R-f1 18.7
33 w34-T-f1 18.7
34 w24-R-f1 22.8
35 w24-T-f1 22.8
36 wU-24-f1 26.2
37 wQ-34-f2 26.9
856

G.6 San Pedro de la Paz (MCE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.21: Summary of walls exceeding acceptance criteria (San Pedro de la Paz ground
motion scaled to ASCE 7 MCE spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.22: Walls exceeding
45
ASCE 41 rotation limits (San Pedro de la Paz ground motion scaled to ASCE 7 MCE spectrum).
40

857
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.23: Walls exceeding
45
fragility function probabilities of 0.50 (San Pedro de Strain
la Paz ground motion scaled to ASCE 7
MCE spectrum). 40

858
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.24: Walls
45
exceeding strain limits (San Pedro de la Paz ground motionStrain
scaled to ASCE 7 MCE spectrum).

859
40

35
# of Predicted walls

30

25

20

15
860

Table G.16: Sequence of exceedence of ASCE 41 acceptance criteria (San Pedro de la Paz
ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 w28-W-f1 8.4 w28-W-f1 8.8 w24-V1-f2 12.1 w28-W-f1 8.3
2 wP-5-f1 8.4 wO-29-f2 8.8 w28-W-f1 12.1 w5-O-f2 9.3
3 wP-29-f1 8.5 wN-5-f2 9.2 wP-5-f1 12.3 wO-29-f1 9.4
4 wV-34-f1 8.5 wV-34-f1 9.3 w29-R-f1 12.3 wP-5-f1 9.9
5 w5-O-f2 8.7 wP-29-f1 9.3 wN-5-f1 12.3 wP-29-f1 9.9
6 wO-29-f2 8.7 wV-24-f1 11.2 wP-29-f1 12.3 w28-W-f2 9.9
7 w28-W-f2 8.7 wS-5-f1 11.2 w34-V1-f2 13.5 wO-29-f2 10.1
8 wN-5-f1 8.9 wV-34-f2 11.4 wS-29-f1 17.8 wV-34-f1 11.2
9 wN-5-f2 8.9 wN-5-f1 11.4 wV-34-f1 17.8 wO-5-f1 11.2
10 wP-35-f2 9.2 w24-V1-f2 11.5 wO-5-f1 19.7 wN-5-f1 11.5
11 wS-29-f2 9.2 wP-5-f1 11.5 wO-29-f1 19.7 wV-24-f1 11.8
12 wS-5-f2 9.2 wS-29-f1 11.8 wS-5-f1 19.8 w29-R-f1 12.1
13 wP-23-f2 9.2 w5-O-f2 12.1 wN-30-f1 19.8 wS-5-f2 12.1
14 wS-29-f1 9.2 w29-R-f1 12.1 wV-24-f1 20.1 wS-5-f1 12.7
15 wS-5-f1 9.2 w5-R-f1 12.6 wV-24-f2 20.2 wV-34-f2 13.2
16 wV-24-f1 9.3 wO-5-f1 12.6 w5-R-f1 21.0 wS-29-f1 16.1
17 w5-R-f1 9.3 wO-29-f1 12.7 wNp-5-f1 21.2 wN-30-f1 20.0
18 wV-24-f2 9.3 wN-30-f1 12.9 w24-R-f1 21.3 wU-24-f1 22.6
19 w29-R-f1 9.5 w34-V1-f2 13.3 w24-T-f1 21.3 wP-23-f2 25.6
20 wO-29-f1 9.7 wS-29-f2 13.4 wQ-24-f1 26.3 wU-34-f1 26.2
21 wO-5-f1 9.8 w30-Np-f1 13.7 wU-24-f2 26.3
22 wV-34-f2 11.2 w34-R-f1 14.2 wQ-34-f1 26.3
23 w24-V1-f2 11.4 w34-T-f1 14.2 wU-24-f1 26.3
24 w24-R-f1 11.5 wNp-5-f1 15.6 wU-34-f1 26.3
25 w24-T-f1 11.5 wS-5-f2 19.1 wU-31-f1 26.4
26 w34-R-f1 11.5 wV-24-f2 19.2 wU-27-f1 26.4
27 w34-T-f1 11.5 wU-24-f2 21.0 w30-Np-f1 26.7
28 w34-V1-f2 12.3 w24-R-f1 21.2 w34-R-f1 26.9
29 w30-Np-f1 12.4 w24-T-f1 21.2 w34-T-f1 26.9
30 wN-30-f1 12.7 wU-31-f1 25.5
31 wNp-5-f1 12.7 wU-27-f1 25.5
32 wU-24-f2 12.8 wQ-34-f1 25.6
33 wU-31-f1 19.8 wU-34-f1 25.7
34 wU-27-f1 19.8 wQ-24-f1 25.8
35 wQ-24-f1 24.1 wU-24-f1 25.8
36 wQ-34-f1 24.3
37 wU-24-f1 25.5
38 wU-34-f1 25.6
861

Table G.17: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (San Pedro de la Paz ground motion scaled to ASCE 7 MCE spectrum). Shading
indicates that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 6.8 wN-5-f1 7.9 w28-W-f1 8.4 w28-W-f1 8.3
2 wV-24-f1 0.0 wV-24-f1 7.1 w5-O-f2 8.1 w5-O-f2 8.7 w5-O-f2 9.3
3 wQ-34-f2 0.0 wN-5-f1 7.9 w28-W-f1 8.4 wO-29-f2 8.7 wO-29-f1 9.4
4 wQ-24-f2 0.0 wN-5-f2 7.9 wP-5-f1 8.4 w28-W-f2 8.8 wP-5-f1 9.9
5 wU-34-f2 0.0 wP-5-f1 7.9 wP-29-f1 8.4 wN-5-f1 8.9 wP-29-f1 9.9
6 wU-24-f2 0.0 wO-5-f1 7.9 wV-24-f2 8.4 wN-5-f2 9.2 w28-W-f2 9.9
7 wV-34-f2 1.4 wO-29-f2 7.9 wN-5-f2 8.4 wP-35-f2 9.2 wO-29-f2 10.1
8 wP-5-f1 2.0 w5-O-f2 8.1 w28-W-f2 8.5 wV-34-f1 9.3 wV-34-f1 11.2
9 w28-W-f1 2.0 wS-5-f1 8.1 wV-34-f1 8.5 wP-23-f2 9.3 wO-5-f1 11.2
10 wO-29-f2 2.0 wS-29-f2 8.1 wV-34-f2 8.5 wS-29-f2 9.3 wN-5-f1 11.5
11 wN-5-f1 2.1 wS-5-f2 8.1 wS-29-f1 8.6 wS-5-f2 9.3 wV-24-f1 11.8
12 wP-29-f1 2.1 wS-29-f1 8.1 wS-5-f1 8.6 wP-5-f1 9.3 w29-R-f1 12.1
13 w5-O-f2 2.2 wP-23-f2 8.1 wS-5-f2 8.6 wP-29-f1 9.3 wS-5-f2 12.1
14 w28-W-f2 2.2 wP-35-f2 8.1 wS-29-f2 8.6 wS-29-f1 9.3 wS-5-f1 12.7
15 wO-5-f1 2.2 wP-29-f1 8.2 wO-29-f2 8.7 wS-5-f1 9.3 wV-34-f2 13.2
16 wV-24-f2 2.2 w28-W-f1 8.3 wU-24-f2 8.8 wV-34-f2 9.3 wS-29-f1 16.1
17 wO-29-f1 2.4 w28-W-f2 8.4 wO-5-f1 8.9 wV-24-f1 10.6 wN-30-f1 20.0
18 wN-30-f1 2.4 wV-34-f2 8.4 wO-29-f1 8.9 wV-24-f2 11.3 wU-24-f1 22.6
19 wN-5-f2 2.7 wV-24-f2 8.4 wNp-5-f1 8.9 w24-V1-f2 11.4 wP-23-f2 25.6
20 w34-R-f1 2.9 wU-24-f2 8.4 wP-35-f2 8.9 wU-24-f2 11.4 wU-34-f1 26.2
21 w34-T-f1 2.9 w5-R-f1 8.4 wV-24-f1 9.0 w29-R-f1 11.5
22 wP-23-f2 3.0 wU-34-f2 8.5 wP-23-f2 9.2 wO-5-f1 11.7
23 wNp-5-f1 3.0 wO-29-f1 8.6 wU-34-f2 9.3 wO-29-f1 11.7
24 wP-35-f2 3.0 wNp-5-f1 8.8 w29-R-f1 9.3 w25p-O1-f1 12.1
25 wS-29-f2 3.1 wN-30-f1 9.3 wN-30-f1 9.3 wN-30-f1 12.3
26 wS-5-f2 3.1 w29-R-f1 9.3 w5-R-f1 9.3 w5-R-f1 12.3
27 wS-29-f1 3.1 w34-R-f1 9.3 w34-V1-f2 9.3 w34-V1-f2 12.3
28 wS-5-f1 3.1 w34-T-f1 9.3 w24-V1-f2 10.9 wNp-5-f1 12.7
29 w30-Np-f1 3.5 w34-V1-f2 9.3 w24-R-f1 11.5 w24-R-f1 12.8
30 w24-R-f1 3.7 w24-R-f1 9.3 w24-T-f1 11.5 w24-T-f1 12.8
31 w24-T-f1 3.7 w24-T-f1 9.3 w34-R-f1 11.5 wU-34-f2 12.8
32 wU-31-f1 5.2 wN-30-f2 9.3 w34-T-f1 11.5 w34-R-f1 12.8
33 wU-27-f1 5.2 w24-V1-f2 10.9 w25p-O1-f1 11.5 w34-T-f1 12.8
34 wU-34-f1 5.5 wU-31-f1 11.2 wU-24-f1 12.3 w30-Np-f1 13.2
35 w5-R-f1 5.5 wU-27-f1 11.2 w30-Np-f1 12.3 wU-31-f1 19.9
36 wN-30-f2 5.7 w25p-O1-f1 11.5 wU-34-f1 13.5 wU-27-f1 19.9
37 wU-24-f1 5.8 wU-24-f1 11.5 wQ-34-f1 13.7 wQ-24-f1 24.0
38 wQ-24-f1 6.3 w30-Np-f1 12.1 wU-31-f1 16.6 wQ-34-f1 25.0
39 wU-31-f2 6.3 wU-34-f1 12.3 wU-27-f1 16.6 wU-24-f1 25.5
40 wU-27-f2 6.3 wQ-34-f1 12.3 wQ-24-f1 21.2 wU-34-f1 25.6
41 wQ-34-f1 6.3 wQ-24-f1 12.3
42 w34-V1-f2 6.3 wU-31-f2 19.1
43 w24-V1-f2 6.6 wU-27-f2 19.1
44 w29-R-f1 6.7
45 w25p-O1-f1 6.8
46 w30-Np-f2 7.5
862

Table G.18: Sequence of exceedence of steel and concrete strain limits (San Pedro de la
Paz ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wNp-5-f1 7.9 wV-34-f2 11.2 w29-R-f1 12.1 w28-W-f1 8.3
2 w5-O-f2 8.2 w28-W-f1 11.5 wV-24-f2 12.1 w5-O-f2 9.3
3 w25p-O1-f1 8.3 wP-5-f1 11.9 wP-5-f1 12.2 wO-29-f1 9.4
4 wP-5-f1 8.3 wV-24-f2 12.0 wP-29-f1 12.3 wP-5-f1 9.9
5 wP-29-f1 8.3 w29-R-f1 12.1 w28-W-f1 12.8 wP-29-f1 9.9
6 w5-R-f1 8.3 w5-R-f1 12.1 wO-5-f1 12.9 w28-W-f2 9.9
7 w28-W-f1 8.3 wP-29-f1 12.1 wV-34-f2 13.5 wO-29-f2 10.1
8 w29-R-f1 8.4 w24-V1-f2 12.1 wO-29-f1 16.7 wV-34-f1 11.2
9 wN-5-f1 8.4 wV-24-f1 12.1 wN-5-f1 21.2 wO-5-f1 11.2
10 wO-29-f2 8.4 wO-29-f1 12.7 wV-34-f1 25.5 wN-5-f1 11.5
11 w24-V1-f2 8.4 wN-5-f1 12.8 wS-29-f1 25.5 wV-24-f1 11.8
12 wN-5-f2 8.4 wO-5-f1 12.9 wV-24-f1 25.5 w29-R-f1 12.1
13 w28-W-f2 8.4 wV-34-f1 13.2 wS-5-f1 25.6 wS-5-f2 12.1
14 w34-V1-f2 8.5 wS-5-f1 13.5 wQ-24-f1 26.3 wS-5-f1 12.7
15 wV-34-f1 8.5 w34-V1-f2 13.5 wQ-34-f1 26.3 wV-34-f2 13.2
16 wO-5-f1 8.8 wU-24-f2 21.0 wU-34-f1 26.4 wS-29-f1 16.1
17 wO-29-f1 8.8 wS-29-f1 21.4 wU-24-f1 26.4 wN-30-f1 20.0
18 wV-34-f2 8.9 wQ-34-f1 24.1 wNp-5-f1 26.5 wU-24-f1 22.6
19 wP-23-f2 9.2 wQ-24-f1 24.1 wU-27-f1 26.5 wP-23-f2 25.6
20 wP-35-f2 9.2 wU-24-f1 25.5 wU-31-f1 26.5 wU-34-f1 26.2
21 wS-29-f2 9.2 wU-34-f1 25.5
22 wS-5-f2 9.2 wN-30-f1 25.6
23 wS-29-f1 9.2 wU-31-f1 25.6
24 wS-5-f1 9.2 wU-27-f1 25.6
25 w30-Np-f1 9.3 wNp-5-f1 26.2
26 wN-30-f2 9.3
27 wV-24-f1 10.0
28 wV-24-f2 10.0
29 wU-24-f2 10.9
30 wU-34-f2 11.2
31 w24-R-f1 11.5
32 w24-T-f1 11.5
33 wN-30-f1 11.7
34 w34-R-f1 12.3
35 w34-T-f1 12.3
36 wQ-24-f1 17.7
37 wU-24-f1 17.7
38 wQ-34-f1 19.7
39 wU-34-f1 19.8
40 wU-31-f1 25.5
41 wU-27-f1 25.5
863

G.7 Llolleo (DBE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.25: Summary of walls exceeding acceptance criteria (Llolleo ground motion scaled
to ASCE 7 design spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.26: Walls
45
exceeding ASCE 41 rotation limits (Llolleo ground motion scaled to ASCE 7 design spectrum).
40

864
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.27: Walls exceeding
45
fragility function probabilities of 0.50 (Llolleo ground motion scaled to ASCE 7 design spectrum).
40

865
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.28:45 Walls exceeding strain limits (Llolleo ground motion scaled Strain
to ASCE 7 design spectrum).

866
40

35
# of Predicted walls

30

25

20

15
867

Table G.19: Sequence of exceedence of ASCE 41 acceptance criteria (Llolleo ground motion
scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded the
acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 w28-W-f1 29.3 w28-W-f1 30.5 wV-34-f1 38.9 w28-W-f2 30.4
2 w5-O-f2 30.0 wP-5-f1 34.4 w28-W-f1 31.3
3 wO-29-f2 30.0 wN-5-f1 34.5 wO-5-f1 33.0
4 wN-5-f2 30.4 wN-5-f2 34.5 wP-29-f1 33.3
5 wP-29-f1 30.4 wV-34-f1 34.6 wP-5-f1 33.3
6 wP-5-f1 30.4 wV-24-f1 36.5 wO-29-f2 33.8
7 wV-24-f1 31.4 wV-34-f2 36.8 wO-29-f1 34.4
8 wN-5-f1 31.6 w34-V1-f2 36.8 wV-34-f2 36.5
9 wO-5-f1 31.7 wP-29-f1 38.9 wV-34-f1 36.7
10 wV-34-f1 32.0 w24-V1-f2 45.4 wV-24-f1 37.1
11 w28-W-f2 33.2 wS-29-f1 45.7 wS-5-f2 37.4
12 wS-29-f1 33.4 wS-5-f1 46.8 wS-29-f1 38.3
13 wS-5-f1 33.4 wO-5-f1 50.5 wS-29-f2 38.3
14 wP-35-f2 33.5 wS-5-f1 38.9
15 wS-5-f2 33.5 w5-O-f2 42.2
16 wS-29-f2 33.9 wN-5-f1 45.0
17 wO-29-f1 34.3 wP-23-f2 50.0
18 wP-23-f2 34.8 wN-5-f2 50.0
19 wV-34-f2 36.0 wP-35-f2 52.6
20 w24-V1-f2 36.5
21 wV-24-f2 36.7
22 w34-V1-f2 36.7
23 w5-R-f1 51.7
868

Table G.20: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Llolleo ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 w28-W-f1 20.3 w28-W-f1 21.2 w28-W-f1 29.6 w28-W-f2 30.4
2 wV-24-f1 0.0 wP-29-f1 20.6 wN-5-f2 23.1 w5-O-f2 30.1 w28-W-f1 31.3
3 wQ-34-f2 0.0 wP-5-f1 20.6 wO-29-f2 29.2 wP-29-f1 30.5 wO-5-f1 33.0
4 wQ-24-f2 0.0 w5-O-f2 20.8 w5-O-f2 29.2 wP-5-f1 30.5 wP-29-f1 33.3
5 wU-34-f2 0.0 wO-29-f2 20.8 w28-W-f2 29.3 wN-5-f2 32.1 wP-5-f1 33.3
6 wU-24-f2 0.0 wV-34-f1 22.3 wP-29-f1 29.6 wO-29-f2 33.2 wO-29-f2 33.8
7 wV-34-f2 0.2 wN-5-f2 22.4 wP-5-f1 29.6 wN-5-f1 33.4 wO-29-f1 34.4
8 w5-O-f2 0.2 wN-5-f1 22.4 wV-34-f2 29.7 wO-5-f1 33.4 wV-34-f2 36.5
9 w28-W-f1 0.4 wV-24-f1 22.7 wV-24-f2 29.7 wV-24-f2 33.7 wV-34-f1 36.7
10 wN-5-f1 0.4 w28-W-f2 29.2 wN-5-f1 30.0 wS-29-f1 34.4 wV-24-f1 37.1
11 wP-5-f1 0.4 wV-34-f2 29.6 wV-24-f1 30.5 wV-34-f1 34.4 wS-5-f2 37.4
12 wN-5-f2 0.4 wV-24-f2 29.6 wV-34-f1 31.0 wS-5-f1 34.4 wS-29-f1 38.3
13 wO-5-f1 0.4 wU-34-f2 29.6 wO-5-f1 31.6 wS-5-f2 34.4 wS-29-f2 38.3
14 wP-29-f1 0.4 wU-24-f2 29.7 wO-29-f1 31.7 wS-29-f2 34.4 wS-5-f1 38.9
15 wS-29-f1 0.4 wO-5-f1 30.0 wU-34-f2 32.6 wV-34-f2 34.7 w5-O-f2 42.2
16 wS-5-f2 0.4 wS-29-f1 30.0 wP-23-f2 33.0 wV-24-f1 35.9 wN-5-f1 45.0
17 wS-5-f1 0.5 wO-29-f1 30.0 wP-35-f2 33.0 w34-V1-f2 36.7 wP-23-f2 50.0
18 wS-29-f2 0.5 wS-5-f1 30.0 wS-29-f2 33.0 wU-34-f2 36.7 wN-5-f2 50.0
19 wO-29-f1 0.5 wS-5-f2 30.0 wU-24-f2 33.2 w24-V1-f2 37.6 wP-35-f2 52.6
20 wV-24-f2 0.5 wS-29-f2 30.0 wS-5-f2 33.4 wO-29-f1 45.2
21 wO-29-f2 0.6 wP-35-f2 30.3 wS-29-f1 33.4 wU-24-f2 45.3
22 wP-23-f2 0.7 wP-23-f2 30.3 wS-5-f1 33.4 wP-23-f2 50.0
23 wP-35-f2 1.1 wN-30-f1 30.4 w24-V1-f2 33.7 w5-R-f1 51.8
24 w28-W-f2 1.8 w29-R-f1 32.9 w34-V1-f2 36.7 wP-35-f2 52.6
25 wN-30-f1 2.7 wNp-5-f1 33.1 wN-30-f1 38.9
26 w34-R-f1 3.3 w24-V1-f2 33.3 w34-R-f1 38.9
27 w34-T-f1 3.3 w24-R-f1 33.7 w34-T-f1 38.9
28 wNp-5-f1 3.5 w24-T-f1 33.7 w29-R-f1 44.9
29 w30-Np-f1 3.5 w34-R-f1 33.8 w5-R-f1 51.3
30 w24-R-f1 3.5 w34-T-f1 33.8
31 w24-T-f1 3.5 w34-V1-f2 34.7
32 wU-24-f1 5.9 wN-30-f2 34.9
33 wU-34-f1 6.3 wU-31-f1 36.7
34 w24-V1-f2 7.2 wU-27-f1 36.7
35 w5-R-f1 7.4 w5-R-f1 38.9
36 w29-R-f1 7.4
37 w34-V1-f2 8.5
38 wN-30-f2 8.7
39 wU-31-f1 10.3
40 wU-27-f1 10.3
41 w30-Np-f2 11.9
42 wU-31-f2 13.0
43 wU-27-f2 13.0
44 wQ-34-f1 14.3
45 wQ-24-f1 14.6
46 w25p-O1-f1 21.1
869

Table G.21: Sequence of exceedence of steel and concrete strain limits (Llolleo ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 20.9 wV-34-f2 36.7 w5-R-f1 51.2 w28-W-f2 30.4
2 w5-R-f1 21.2 wV-24-f2 45.4 w28-W-f1 31.3
3 wNp-5-f1 23.2 wV-34-f1 50.0 wO-5-f1 33.0
4 w28-W-f1 29.0 w5-R-f1 51.2 wP-29-f1 33.3
5 w25p-O1-f1 29.0 wP-5-f1 33.3
6 w5-O-f2 29.2 wO-29-f2 33.8
7 wO-29-f2 29.2 wO-29-f1 34.4
8 wP-29-f1 29.6 wV-34-f2 36.5
9 w29-R-f1 29.6 wV-34-f1 36.7
10 wP-5-f1 29.6 wV-24-f1 37.1
11 w34-V1-f2 29.6 wS-5-f2 37.4
12 w28-W-f2 29.6 wS-29-f1 38.3
13 w24-V1-f2 29.7 wS-29-f2 38.3
14 wV-34-f1 30.4 wS-5-f1 38.9
15 wV-24-f2 30.5 w5-O-f2 42.2
16 wN-5-f1 31.6 wN-5-f1 45.0
17 wN-5-f2 31.6 wP-23-f2 50.0
18 wO-29-f1 31.7 wN-5-f2 50.0
19 wV-34-f2 32.0 wP-35-f2 52.6
20 wP-23-f2 33.0
21 wP-35-f2 33.0
22 wS-29-f1 33.4
23 wS-5-f1 33.4
24 wS-29-f2 33.4
25 wS-5-f2 33.4
26 wV-24-f1 33.7
27 wU-34-f2 35.2
28 wN-30-f1 35.4
29 wU-24-f2 36.0
30 w30-Np-f1 38.9
31 wN-30-f2 44.7
870

G.8 Llolleo (MCE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.29: Summary of walls exceeding acceptance criteria (Llolleo ground motion scaled
to ASCE 7 MCE spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.30: Walls
45
exceeding ASCE 41 rotation limits (Llolleo ground motion Strain
scaled to ASCE 7 MCE spectrum).
40

871
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.31: Walls exceeding
45
fragility function probabilities of 0.50 (Llolleo ground motion scaled to ASCE 7 MCE spectrum).
40

872
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.32:
45
Walls exceeding strain limits (Llolleo ground motion scaledStrain
to ASCE 7 MCE spectrum).

873
40

35
# of Predicted walls

30

25

20

15
874

Table G.22: Sequence of exceedence of ASCE 41 acceptance criteria (Llolleo ground motion
scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also exceeded the
acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wV-34-f1 27.5 w28-W-f1 33.7 wN-5-f1 44.8 w28-W-f1 29.7
2 wN-5-f2 27.6 wN-5-f2 33.8 wP-5-f1 44.8 w28-W-f2 29.7
3 wV-24-f1 28.1 wN-5-f1 34.3 wV-34-f1 49.6 wP-29-f1 33.2
4 w28-W-f1 29.0 wV-24-f1 36.5 wO-29-f2 33.8
5 wO-29-f2 29.3 wV-34-f1 36.5 wO-29-f1 34.3
6 wP-5-f1 29.7 w34-V1-f2 36.7 wO-5-f1 34.8
7 wP-29-f1 29.7 wS-5-f1 36.8 wP-5-f1 36.1
8 w5-O-f2 30.1 wS-29-f1 36.8 wS-29-f2 36.5
9 w28-W-f2 30.2 wV-34-f2 38.1 wV-34-f2 36.5
10 wN-5-f1 31.4 wP-5-f1 38.9 wV-34-f1 36.5
11 wO-5-f1 33.4 wP-29-f1 38.9 w5-O-f2 36.9
12 wO-29-f1 34.3 wO-5-f1 39.3 wS-5-f1 36.9
13 wP-35-f2 34.3 wO-29-f1 39.5 wV-24-f1 37.1
14 wS-29-f1 34.4 w24-V1-f2 44.8 wS-5-f2 37.4
15 wP-23-f2 34.7 wN-30-f1 44.8 wP-23-f2 38.8
16 wS-5-f1 34.7 wN-5-f1 38.9
17 wS-29-f2 34.7 wS-29-f1 44.3
18 wS-5-f2 34.7 wP-35-f2 44.8
19 wV-34-f2 36.5
20 w24-V1-f2 36.5
21 w34-V1-f2 36.7
22 wV-24-f2 36.7
23 w29-R-f1 39.0
24 wN-30-f1 39.5
25 wNp-5-f1 44.8
26 w24-R-f1 44.8
27 w24-T-f1 44.8
28 wU-24-f2 44.9
29 wU-31-f1 49.6
30 wU-27-f1 49.6
31 w5-R-f1 60.4
875

Table G.23: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Llolleo ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that
the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 w28-W-f1 8.5 w28-W-f1 23.8 wV-34-f1 27.5 w28-W-f1 29.7
2 wV-24-f1 0.0 w5-O-f2 10.3 wV-34-f1 24.6 w28-W-f1 29.0 w28-W-f2 29.7
3 wQ-34-f2 0.0 wP-5-f1 10.6 wV-24-f1 25.0 wO-29-f2 29.3 wP-29-f1 33.2
4 wQ-24-f2 0.0 wO-29-f2 10.8 wN-5-f2 25.6 wP-5-f1 29.8 wO-29-f2 33.8
5 wU-34-f2 0.0 wP-29-f1 11.1 w5-O-f2 26.8 wP-29-f1 29.8 wO-29-f1 34.3
6 wU-24-f2 0.0 wN-5-f2 11.5 wS-29-f2 26.9 wV-24-f1 30.0 wO-5-f1 34.8
7 wV-34-f2 0.2 wN-5-f1 13.3 wS-5-f2 26.9 w5-O-f2 30.1 wP-5-f1 36.1
8 w5-O-f2 0.2 wV-34-f1 14.3 wS-5-f1 26.9 wN-5-f2 31.4 wS-29-f2 36.5
9 wS-29-f2 0.2 wV-24-f1 24.0 wN-5-f1 27.2 wV-34-f2 31.5 wV-34-f2 36.5
10 wS-5-f1 0.2 wV-34-f2 24.9 wV-34-f2 28.3 wN-5-f1 33.4 wV-34-f1 36.5
11 wS-5-f2 0.2 wS-29-f1 26.5 wS-29-f1 28.7 wO-5-f1 34.3 w5-O-f2 36.9
12 wS-29-f1 0.2 wS-5-f1 26.5 wP-29-f1 29.0 wS-29-f2 34.8 wS-5-f1 36.9
13 wO-29-f2 0.2 wS-5-f2 26.5 wP-5-f1 29.0 wP-23-f2 34.8 wV-24-f1 37.1
14 wP-23-f2 0.2 wS-29-f2 26.5 wO-29-f2 29.3 wS-5-f2 34.8 wS-5-f2 37.4
15 w28-W-f1 0.4 wO-5-f1 26.6 w28-W-f2 29.3 wP-35-f2 34.8 wP-23-f2 38.8
16 wP-29-f1 0.4 wP-35-f2 26.6 wU-34-f2 29.3 wV-24-f2 36.5 wN-5-f1 38.9
17 wN-5-f1 0.4 wP-23-f2 26.8 wV-24-f2 29.7 wS-5-f1 36.5 wS-29-f1 44.3
18 wN-5-f2 0.4 wV-24-f2 27.5 wU-24-f2 30.0 wS-29-f1 36.5 wP-35-f2 44.8
19 wO-5-f1 0.4 wU-34-f2 28.1 wO-5-f1 31.7 w34-V1-f2 36.7
20 wP-5-f1 0.4 wO-29-f1 28.1 wO-29-f1 33.4 wO-29-f1 37.0
21 wO-29-f1 0.4 wU-24-f2 28.6 wP-35-f2 33.4 w24-V1-f2 37.4
22 wP-35-f2 0.4 w28-W-f2 28.7 wP-23-f2 33.8 wU-24-f2 37.4
23 wV-24-f2 0.5 wN-30-f1 29.7 w24-V1-f2 36.5 wN-30-f1 44.8
24 w28-W-f2 1.6 w24-V1-f2 30.6 w5-R-f1 36.7 wNp-5-f1 44.8
25 w24-R-f1 2.0 w34-V1-f2 31.5 w34-V1-f2 36.7 wU-34-f2 49.3
26 w24-T-f1 2.0 w34-R-f1 31.5 wN-30-f1 37.0
27 wN-30-f1 2.2 w34-T-f1 31.5 wNp-5-f1 38.4
28 wN-30-f2 2.3 w29-R-f1 31.5 w29-R-f1 38.9
29 w34-R-f1 2.5 w5-R-f1 33.7 w34-R-f1 38.9
30 w34-T-f1 2.5 wNp-5-f1 33.9 w34-T-f1 38.9
31 wNp-5-f1 2.7 wN-30-f2 34.8 wU-31-f1 44.5
32 w30-Np-f1 2.7 w24-R-f1 36.5 wU-27-f1 44.5
33 wU-34-f1 3.7 w24-T-f1 36.5 w24-R-f1 44.8
34 wU-24-f1 4.6 wU-31-f1 36.5 w24-T-f1 44.8
35 w24-V1-f2 5.0 wU-27-f1 36.5
36 w5-R-f1 5.3 wU-24-f1 36.8
37 w29-R-f1 5.3 wU-34-f1 44.5
38 w34-V1-f2 6.4 w25p-O1-f1 45.1
39 w30-Np-f2 7.3
40 wU-31-f1 7.5
41 wU-27-f1 7.5
42 wQ-34-f1 8.3
43 wU-31-f2 10.0
44 wU-27-f2 10.0
45 w25p-O1-f1 10.6
46 wQ-24-f1 10.7
876

Table G.24: Sequence of exceedence of steel and concrete strain limits (Llolleo ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 8.8 wV-34-f2 36.7 wO-5-f1 45.0 w28-W-f1 29.7
2 w5-R-f1 23.8 wV-24-f2 37.4 w28-W-f2 29.7
3 w5-O-f2 26.9 wV-34-f1 44.5 wP-29-f1 33.2
4 wV-34-f1 27.5 wO-29-f1 44.8 wO-29-f2 33.8
5 wV-24-f2 27.5 wO-5-f1 44.8 wO-29-f1 34.3
6 wNp-5-f1 27.6 wV-24-f1 49.6 wO-5-f1 34.8
7 wN-5-f1 28.1 w28-W-f1 50.1 wP-5-f1 36.1
8 w34-V1-f2 28.1 wS-29-f2 36.5
9 wV-34-f2 28.1 wV-34-f2 36.5
10 wN-5-f2 28.1 wV-34-f1 36.5
11 wP-5-f1 28.4 w5-O-f2 36.9
12 wP-29-f1 28.4 wS-5-f1 36.9
13 w28-W-f1 28.4 wV-24-f1 37.1
14 w25p-O1-f1 28.6 wS-5-f2 37.4
15 w24-V1-f2 28.7 wP-23-f2 38.8
16 w29-R-f1 29.0 wN-5-f1 38.9
17 w28-W-f2 29.0 wS-29-f1 44.3
18 wO-29-f2 29.2 wP-35-f2 44.8
19 wV-24-f1 30.0
20 wO-29-f1 31.8
21 wP-35-f2 33.4
22 wU-34-f2 33.4
23 wP-23-f2 33.8
24 wS-29-f1 34.3
25 wN-30-f1 34.4
26 wS-5-f1 34.4
27 wS-29-f2 34.4
28 wS-5-f2 34.7
29 wN-30-f2 34.8
30 w30-Np-f1 36.5
31 wU-24-f2 37.4
32 wU-24-f1 49.6
877

G.9 Matanzas (DBE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.33: Summary of walls exceeding acceptance criteria (Matanzas ground motion
scaled to ASCE 7 design spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure G.34: Walls45exceedingASCE Rotation


ASCE 41 rotation limitsFragility Functions
(Matanzas ground motionStrain
scaled to ASCE 7 design spectrum).
40

878
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.35: Walls exceeding
45
fragility function probabilities of 0.50 (Matanzas ground motion scaled to ASCE 7 design
spectrum). 40

879
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.36: 45Walls exceeding strain limits (Matanzas ground motion scaledStrain
to ASCE 7 design spectrum).

880
40

35
# of Predicted walls

30

25

20

15
881

Table G.25: Sequence of exceedence of ASCE 41 acceptance criteria (Matanzas ground


motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 w28-W-f1 24.9 wN-5-f2 26.0 wP-5-f1 26.6 w28-W-f1 24.9
2 wN-5-f2 25.0 wN-5-f1 26.4 wN-5-f1 26.6 wO-29-f1 26.4
3 wN-5-f1 25.1 wP-5-f1 26.5 wO-5-f1 26.6 wO-29-f2 26.9
4 wS-29-f2 25.1 wO-5-f1 26.5 wN-5-f2 27.1 wO-5-f1 27.0
5 wP-5-f1 25.4 wS-29-f1 26.5 wP-35-f2 27.2 wP-5-f1 27.1
6 wO-5-f1 25.4 wP-35-f2 26.5 wP-23-f2 27.2 wP-29-f1 27.2
7 wS-29-f1 25.5 wP-29-f1 26.5 wV-24-f2 31.3 wV-34-f2 27.7
8 wV-34-f1 25.6 wS-5-f1 26.6 wV-34-f1 41.3 w5-O-f2 28.4
9 wP-23-f2 26.0 wS-5-f2 26.6 wV-34-f2 42.2 w28-W-f2 32.7
10 wP-35-f2 26.0 wO-29-f1 26.6 wS-29-f2 44.1 wP-35-f2 33.7
11 wS-5-f2 26.0 wS-29-f2 26.6 wS-5-f2 44.2 wN-5-f1 33.9
12 wS-5-f1 26.0 wV-34-f1 26.7 wS-29-f1 44.2 wV-24-f1 37.5
13 wP-29-f1 26.0 wV-24-f1 26.8 wS-5-f1 44.2 wV-34-f1 38.4
14 wV-24-f1 26.2 wV-24-f2 26.8 w28-W-f1 52.0 wS-5-f2 41.3
15 wO-29-f1 26.4 wP-23-f2 27.1 w24-V1-f2 52.1 wS-5-f1 41.3
16 wV-34-f2 26.7 wN-30-f2 27.2 wU-24-f2 44.4
17 wV-24-f2 26.7 w28-W-f1 30.7 wN-5-f2 44.4
18 wO-29-f2 27.1 wV-34-f2 31.2
19 wN-30-f2 27.1 w5-O-f2 31.3
20 w5-O-f2 27.7 w24-V1-f2 31.3
21 w28-W-f2 30.3 w34-V1-f2 31.8
22 w24-V1-f2 30.7 wO-29-f2 32.2
23 w34-V1-f2 31.6
24 w5-R-f1 38.1
25 w29-R-f1 39.1
26 w34-R-f1 39.3
27 w34-T-f1 39.3
28 wU-34-f2 43.1
29 wU-31-f2 44.3
30 wU-27-f2 44.3
31 w24-R-f1 52.1
32 w24-T-f1 52.1
882

Table G.26: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Matanzas ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 20.8 w28-W-f1 24.6 w28-W-f1 24.9 w28-W-f1 24.9
2 wV-24-f1 0.0 w28-W-f1 21.5 wN-5-f1 24.7 w5-O-f2 25.0 wO-29-f1 26.4
3 wQ-34-f2 0.0 wO-29-f2 22.0 wO-29-f2 24.7 wN-5-f2 25.0 wO-29-f2 26.9
4 wQ-24-f2 0.0 wV-24-f1 22.9 wN-5-f2 24.7 wN-5-f1 25.5 wO-5-f1 27.0
5 wU-34-f2 0.0 w5-O-f2 23.4 wP-5-f1 24.7 wP-5-f1 25.5 wP-5-f1 27.1
6 wU-24-f2 0.0 wP-5-f1 23.7 wO-5-f1 24.7 wV-34-f1 25.7 wP-29-f1 27.2
7 w5-O-f2 0.5 wN-5-f1 24.0 wP-29-f1 24.9 wS-29-f2 26.0 wV-34-f2 27.7
8 w28-W-f1 0.5 wN-5-f2 24.1 wV-34-f2 24.9 wP-23-f2 26.0 w5-O-f2 28.4
9 wV-34-f2 0.5 wP-29-f1 24.3 wV-24-f2 25.0 wO-29-f2 26.0 w28-W-f2 32.7
10 wN-5-f2 0.6 w28-W-f2 24.6 w5-O-f2 25.0 wS-5-f2 26.1 wP-35-f2 33.7
11 wN-5-f1 0.7 wU-34-f2 24.7 w28-W-f2 25.0 wV-24-f1 26.2 wN-5-f1 33.9
12 wP-5-f1 0.7 wO-29-f1 24.7 wP-23-f2 25.0 wO-5-f1 26.4 wV-24-f1 37.5
13 wO-5-f1 0.7 wO-5-f1 24.7 wS-29-f2 25.0 wP-29-f1 26.5 wV-34-f1 38.4
14 wV-24-f2 0.9 wS-29-f1 24.7 wP-35-f2 25.0 wP-35-f2 26.5 wS-5-f2 41.3
15 wP-29-f1 1.0 wP-35-f2 24.7 wS-5-f2 25.0 wO-29-f1 26.5 wS-5-f1 41.3
16 wS-29-f1 1.1 wS-5-f1 24.7 wS-5-f1 25.1 wS-29-f1 26.5 wU-24-f2 44.4
17 wO-29-f2 1.1 wS-5-f2 24.7 wS-29-f1 25.1 wS-5-f1 26.5 wN-5-f2 44.4
18 w28-W-f2 1.1 wS-29-f2 24.8 wV-24-f1 25.1 wV-24-f2 26.8
19 wN-30-f1 1.4 wV-34-f2 24.9 wO-29-f1 25.4 wV-34-f2 26.8
20 wO-29-f1 1.4 wV-24-f2 24.9 wV-34-f1 25.5 wN-30-f2 27.1
21 w34-R-f1 1.6 wU-24-f2 24.9 wN-30-f2 26.6 w24-V1-f2 30.8
22 w34-T-f1 1.6 wP-23-f2 25.0 wU-34-f2 26.9 wU-24-f2 31.2
23 w24-R-f1 1.8 wNp-5-f1 25.2 w34-V1-f2 27.0 w34-V1-f2 31.6
24 w24-T-f1 1.8 wN-30-f2 26.5 wU-24-f2 27.2 wU-34-f2 31.6
25 wNp-5-f1 2.0 wN-30-f1 26.5 w24-V1-f2 30.7 w28-W-f2 32.3
26 w30-Np-f1 2.0 w25p-O1-f1 26.6 wN-30-f1 30.8 w29-R-f1 39.1
27 wP-35-f2 2.2 wU-31-f1 26.6 w5-R-f1 38.1 w5-R-f1 39.1
28 wS-5-f1 2.9 wU-27-f1 26.6 w29-R-f1 38.2 wU-31-f2 44.3
29 wS-5-f2 2.9 wU-31-f2 26.6 w24-R-f1 38.2 wU-27-f2 44.3
30 wS-29-f2 2.9 wU-27-f2 26.6 w24-T-f1 38.2
31 wP-23-f2 3.2 w34-V1-f2 26.9 w34-R-f1 39.1
32 wU-31-f1 5.2 w24-V1-f2 27.7 w34-T-f1 39.1
33 wU-27-f1 5.2 w5-R-f1 28.9 wU-31-f2 41.1
34 wU-24-f1 6.8 w24-R-f1 30.7 wU-27-f2 41.1
35 wU-34-f1 6.8 w24-T-f1 30.7 wU-31-f1 41.3
36 w24-V1-f2 6.8 w29-R-f1 30.7 wU-27-f1 41.3
37 w34-V1-f2 7.2 w34-R-f1 30.8
38 wN-30-f2 7.2 w34-T-f1 30.8
39 w29-R-f1 7.5 w30-Np-f1 39.3
40 w5-R-f1 7.5 wU-24-f1 41.3
41 wU-31-f2 8.3 wU-34-f1 41.3
42 wU-27-f2 8.3 wQ-34-f2 44.3
43 w30-Np-f2 8.6 w30-Np-f2 44.3
44 wQ-24-f1 15.9
45 wQ-34-f1 16.2
46 w25p-O1-f1 21.8
883

Table G.27: Sequence of exceedence of steel and concrete strain limits (Matanzas ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 23.4 wP-35-f2 26.6 wO-29-f2 27.2 w28-W-f1 24.9
2 w5-R-f1 23.8 wO-29-f2 27.1 wV-24-f2 31.3 wO-29-f1 26.4
3 w5-O-f2 24.0 wP-23-f2 27.2 wS-29-f2 44.3 wO-29-f2 26.9
4 wO-29-f2 24.7 wN-5-f2 27.2 wP-23-f2 44.3 wO-5-f1 27.0
5 wN-5-f1 24.8 wS-29-f2 27.3 wV-34-f2 44.5 wP-5-f1 27.1
6 w29-R-f1 24.9 wV-34-f2 30.5 wP-29-f1 27.2
7 w28-W-f1 24.9 wV-24-f2 31.2 wV-34-f2 27.7
8 wP-29-f1 24.9 wV-24-f1 41.3 w5-O-f2 28.4
9 w25p-O1-f1 24.9 w28-W-f2 44.4 w28-W-f2 32.7
10 wP-5-f1 24.9 w5-O-f2 51.5 wP-35-f2 33.7
11 w34-V1-f2 24.9 w24-V1-f2 52.1 wN-5-f1 33.9
12 w28-W-f2 24.9 wV-24-f1 37.5
13 w24-V1-f2 25.0 wV-34-f1 38.4
14 wNp-5-f1 25.0 wS-5-f2 41.3
15 wN-5-f2 25.0 wS-5-f1 41.3
16 wP-35-f2 25.1 wU-24-f2 44.4
17 wP-23-f2 25.1 wN-5-f2 44.4
18 wO-29-f1 25.4
19 wS-29-f1 25.5
20 wS-29-f2 25.5
21 wV-34-f1 25.7
22 wS-5-f1 26.0
23 wS-5-f2 26.0
24 wV-34-f2 26.2
25 wV-24-f1 26.2
26 wN-30-f1 26.5
27 wN-30-f2 26.5
28 wV-24-f2 26.6
29 wU-34-f2 27.0
30 wQ-34-f2 27.2
31 wU-24-f2 31.2
32 w30-Np-f1 33.8
884

G.10 Matanzas (MCE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.37: Summary of walls exceeding acceptance criteria (Matanzas ground motion
scaled to ASCE 7 MCE spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation
Figure G.38: Walls45exceeding ASCE 41 rotation limitsFragility Functions
(Matanzas ground motionStrain
scaled to ASCE 7 MCE spectrum).
40

885
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.39: Walls exceeding
45
fragility function probabilities of 0.50 (Matanzas ground motion scaled to ASCE 7 MCE spec-
trum). 40

886
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.40:45Walls exceeding strain limits (Matanzas ground motion scaled to ASCE 7 MCE spectrum).

887
40

35
# of Predicted walls

30

25

20

15
888

Table G.28: Sequence of exceedence of ASCE 41 acceptance criteria (Matanzas ground


motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 w28-W-f1 23.8 wN-5-f2 25.9 wP-5-f1 26.5 w28-W-f1 23.2
2 wV-24-f1 25.0 wN-5-f1 26.4 wN-5-f1 26.5 wO-5-f1 26.0
3 wS-29-f2 25.0 wP-5-f1 26.4 wO-5-f1 26.5 wO-29-f1 26.5
4 wS-5-f2 25.1 wO-5-f1 26.5 wN-5-f2 26.6 wO-29-f2 26.6
5 wN-5-f2 25.1 wP-29-f1 26.5 wP-35-f2 26.6 w28-W-f2 26.7
6 wS-5-f1 25.1 wO-29-f1 26.5 wS-29-f1 26.6 w5-O-f2 26.7
7 wV-34-f1 25.1 wP-35-f2 26.5 wS-5-f2 26.6 wP-5-f1 26.8
8 wP-23-f2 25.1 wS-29-f1 26.5 wS-5-f1 26.8 wP-29-f1 26.9
9 wS-29-f1 25.1 wS-5-f1 26.5 wV-34-f1 26.9 wS-29-f1 27.2
10 wN-5-f1 25.4 wS-5-f2 26.6 w28-W-f1 27.0 wV-34-f2 27.7
11 wP-5-f1 25.4 wS-29-f2 26.6 w34-V1-f2 30.1 wV-24-f1 28.6
12 wO-5-f1 25.5 wV-34-f1 26.8 w29-R-f1 37.3 wN-5-f1 29.0
13 wP-35-f2 26.0 wV-24-f1 26.8 wV-24-f2 37.6 wS-5-f2 29.4
14 wP-29-f1 26.0 wV-24-f2 26.8 w24-V1-f2 37.6 wV-34-f1 29.5
15 wO-29-f1 26.4 w28-W-f1 26.9 wV-24-f1 37.6 wS-5-f1 32.9
16 wN-30-f1 26.5 w34-V1-f2 27.0 wV-34-f2 37.7 wS-29-f2 33.1
17 wNp-5-f1 26.5 wP-23-f2 27.3 wP-35-f2 33.7
18 wN-30-f2 26.6 w24-V1-f2 28.9
19 wV-34-f2 26.6 w28-W-f2 30.1
20 wV-24-f2 26.7 wV-34-f2 31.1
21 w34-V1-f2 26.9 wO-29-f2 35.6
22 w28-W-f2 27.0 w29-R-f1 37.3
23 wO-29-f2 27.0 w5-O-f2 37.7
24 w34-R-f1 27.0
25 w34-T-f1 27.0
26 w24-V1-f2 28.6
27 w24-R-f1 28.9
28 w24-T-f1 28.9
29 w5-O-f2 29.4
30 wU-34-f2 30.1
31 w29-R-f1 37.3
889

Table G.29: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Matanzas ground motion scaled to ASCE 7 MCE spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 w28-W-f1 11.9 wV-34-f1 17.6 w28-W-f1 24.5 w28-W-f1 23.2
2 wV-24-f1 0.0 wV-34-f1 12.5 wV-24-f1 18.1 w5-O-f2 25.1 wO-5-f1 26.0
3 wQ-34-f2 0.0 wV-24-f1 13.0 w28-W-f1 23.2 wV-24-f1 25.1 wO-29-f1 26.5
4 wQ-24-f2 0.0 wN-5-f2 13.3 w5-O-f2 23.4 wS-29-f2 25.1 wO-29-f2 26.6
5 wU-34-f2 0.0 wV-34-f2 18.0 wP-29-f1 23.8 wN-5-f2 25.5 w28-W-f2 26.7
6 wU-24-f2 0.0 wN-5-f1 18.1 wP-5-f1 23.8 wN-5-f1 25.5 w5-O-f2 26.7
7 wP-5-f1 0.1 wP-5-f1 20.0 wO-29-f2 24.1 wV-34-f1 25.6 wP-5-f1 26.8
8 wO-5-f1 0.2 w5-O-f2 21.4 wV-34-f2 24.5 wO-29-f2 26.0 wP-29-f1 26.9
9 wS-29-f1 0.2 wO-29-f2 22.9 wN-5-f1 24.8 wP-23-f2 26.0 wS-29-f1 27.2
10 wS-5-f1 0.2 wP-29-f1 23.1 wO-5-f1 24.8 wP-35-f2 26.0 wV-34-f2 27.7
11 w28-W-f1 0.2 wU-24-f2 23.4 wN-5-f2 24.8 wS-5-f2 26.1 wV-24-f1 28.6
12 wV-24-f2 0.2 wS-29-f1 23.4 wS-29-f2 25.0 wP-5-f1 26.4 wN-5-f1 29.0
13 wS-5-f2 0.2 w28-W-f2 23.4 wS-5-f2 25.0 wO-5-f1 26.4 wS-5-f2 29.4
14 w5-O-f2 0.4 wS-5-f2 23.4 wS-5-f1 25.0 wP-29-f1 26.4 wV-34-f1 29.5
15 wN-5-f2 0.4 wS-29-f2 23.4 wS-29-f1 25.0 wO-29-f1 26.4 wS-5-f1 32.9
16 wP-29-f1 0.5 wS-5-f1 23.5 wP-23-f2 25.0 wS-29-f1 26.5 wS-29-f2 33.1
17 wV-34-f2 0.5 wV-24-f2 23.8 wP-35-f2 25.0 wS-5-f1 26.5 wP-35-f2 33.7
18 wS-29-f2 0.5 wU-34-f2 24.1 wO-29-f1 25.5 wNp-5-f1 26.6
19 wN-5-f1 0.6 wO-5-f1 24.8 wV-24-f2 25.6 wN-30-f2 26.6
20 wO-29-f2 0.7 wO-29-f1 24.8 wU-34-f2 26.4 wV-24-f2 26.8
21 w28-W-f2 0.7 wP-35-f2 24.8 w34-V1-f2 26.5 wV-34-f2 26.8
22 wO-29-f1 1.2 wP-23-f2 25.0 wN-30-f1 26.5 wU-34-f2 26.9
23 wN-30-f1 1.2 wNp-5-f1 25.9 wNp-5-f1 26.5 w34-V1-f2 26.9
24 w34-R-f1 1.4 wN-30-f2 26.0 wN-30-f2 26.6 w24-V1-f2 28.9
25 w34-T-f1 1.4 w34-V1-f2 26.4 w25p-O1-f1 26.6 wU-24-f2 28.9
26 wNp-5-f1 1.4 wN-30-f1 26.4 wU-24-f2 26.6 w28-W-f2 28.9
27 w30-Np-f1 1.4 w25p-O1-f1 26.5 w28-W-f2 26.9 w29-R-f1 37.3
28 w24-R-f1 1.6 wU-31-f1 26.6 wU-31-f2 26.9
29 w24-T-f1 1.6 wU-27-f1 26.6 wU-27-f2 26.9
30 wP-35-f2 1.6 w34-R-f1 26.6 w5-R-f1 26.9
31 wP-23-f2 2.0 w34-T-f1 26.6 w29-R-f1 26.9
32 wU-24-f1 2.3 wU-31-f2 26.6 w34-R-f1 26.9
33 wU-34-f1 2.7 wU-27-f2 26.6 w34-T-f1 26.9
34 wU-31-f1 3.6 w29-R-f1 26.8 w24-V1-f2 28.6
35 wU-27-f1 3.6 wU-34-f1 26.8 w24-R-f1 28.9
36 wU-31-f2 5.6 w5-R-f1 26.9 w24-T-f1 28.9
37 wU-27-f2 5.6 w24-R-f1 26.9 wU-31-f1 37.4
38 w34-V1-f2 5.7 w24-T-f1 26.9 wU-27-f1 37.4
39 wN-30-f2 5.8 w30-Np-f1 27.0
40 w24-V1-f2 6.0 w24-V1-f2 27.0
41 w29-R-f1 7.1 wU-24-f1 37.4
42 w5-R-f1 7.1
43 w30-Np-f2 7.3
44 wQ-34-f1 8.2
45 wQ-24-f1 11.7
46 w25p-O1-f1 13.0
890

Table G.30: Sequence of exceedence of steel and concrete strain limits (Matanzas ground
motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 7.8 wP-35-f2 26.6 wO-29-f2 27.3 w28-W-f1 23.2
2 w5-R-f1 23.1 wS-5-f2 26.6 wV-24-f2 37.6 wO-5-f1 26.0
3 wP-29-f1 23.2 wO-29-f1 26.6 wV-34-f2 38.7 wO-29-f1 26.5
4 w28-W-f1 23.2 wO-29-f2 27.2 wO-29-f2 26.6
5 w25p-O1-f1 23.2 w34-V1-f2 30.1 w28-W-f2 26.7
6 w29-R-f1 23.2 wU-34-f2 30.1 w5-O-f2 26.7
7 w5-O-f2 23.4 wV-34-f2 30.5 wP-5-f1 26.8
8 wO-29-f2 23.4 wV-24-f2 31.3 wP-29-f1 26.9
9 wP-5-f1 23.8 wV-24-f1 37.6 wS-29-f1 27.2
10 w34-V1-f2 24.5 w24-V1-f2 37.6 wV-34-f2 27.7
11 wNp-5-f1 24.8 wV-34-f1 37.6 wV-24-f1 28.6
12 wV-24-f1 25.1 wN-5-f1 29.0
13 wS-29-f2 25.1 wS-5-f2 29.4
14 wV-24-f2 25.1 wV-34-f1 29.5
15 wP-35-f2 25.1 wS-5-f1 32.9
16 wS-5-f2 25.1 wS-29-f2 33.1
17 wS-5-f1 25.1 wP-35-f2 33.7
18 wP-23-f2 25.1
19 wV-34-f2 25.1
20 wN-5-f1 25.4
21 wN-5-f2 25.4
22 wS-29-f1 25.5
23 wV-34-f1 25.6
24 wN-30-f2 26.0
25 wO-29-f1 26.0
26 wU-34-f2 26.4
27 wN-30-f1 26.4
28 w28-W-f2 26.5
29 wU-24-f2 26.6
30 wQ-24-f2 26.6
31 w30-Np-f1 26.9
32 w24-V1-f2 26.9
33 wQ-34-f2 27.3
34 w34-R-f1 30.1
891

G.11 Valparaiso Almendral (DBE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.41: Summary of walls exceeding acceptance criteria (Valparaiso ground motion
scaled to ASCE 7 design spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure G.42: Walls 45exceedingASCE


ASCERotation
41 rotation limitsFragility Functions
(Valparaiso ground motionStrain
scaled to ASCE 7 design spectrum).
40

892
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.43: Walls exceeding
45
fragility function probabilities of 0.50 (Valparaiso ground motion scaled to ASCE 7 design
spectrum). 40

893
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.44: 45
Walls exceeding strain limits (Valparaiso ground motion scaled to ASCE 7 design spectrum).

894
40

35
# of Predicted walls

30

25

20

15
895

Table G.31: Sequence of exceedence of ASCE 41 acceptance criteria (Valparaiso ground


motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wV-34-f1 22.1 w28-W-f1 28.2 wN-5-f2 33.0 w28-W-f2 27.4
2 wN-5-f2 22.5 wN-5-f2 29.2 wN-5-f1 35.6 wP-29-f1 27.8
3 wN-5-f1 22.9 wN-5-f1 29.8 wP-5-f1 35.6 wO-5-f1 29.2
4 w28-W-f1 25.3 wP-5-f1 30.7 wO-5-f1 35.7 wO-29-f1 29.8
5 wP-29-f1 25.7 wO-5-f1 31.8 w28-W-f1 37.7 wO-29-f2 31.2
6 wP-5-f1 25.7 wS-5-f2 31.9 w5-O-f2 37.7 w28-W-f1 35.0
7 wO-29-f2 26.1 wP-35-f2 32.9 wO-29-f2 37.7 wN-5-f1 36.1
8 wV-24-f1 27.3 wO-29-f2 33.4 w34-V1-f2 49.1 wP-5-f1 37.7
9 wS-29-f2 27.5 w5-O-f2 33.4 wV-34-f1 38.9
10 w5-O-f2 27.7 w24-V1-f2 33.5 wV-24-f1 46.3
11 wO-5-f1 27.9 wV-34-f1 34.8
12 wS-5-f2 27.9 wV-24-f1 35.2
13 wS-29-f1 28.0 wV-34-f2 35.4
14 wP-23-f2 29.2 w34-V1-f2 35.6
15 wO-29-f1 29.7 w28-W-f2 35.6
16 wP-35-f2 29.8 wP-29-f1 35.6
17 wS-5-f1 30.7 wO-29-f1 35.6
18 wV-34-f2 33.4 wS-29-f1 35.7
19 w24-V1-f2 33.4 wS-5-f1 35.7
20 w28-W-f2 33.4 wP-23-f2 36.2
21 wV-24-f2 34.8 wS-29-f2 39.4
22 w5-R-f1 35.2 w5-R-f1 43.9
23 w34-V1-f2 35.5 wV-24-f2 48.1
24 wN-30-f1 35.7
25 wNp-5-f1 35.7
26 w29-R-f1 38.2
27 w34-R-f1 45.8
28 w34-T-f1 45.8
29 w24-R-f1 45.8
30 w24-T-f1 45.8
31 wU-34-f2 49.6
896

Table G.32: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Valparaiso ground motion scaled to ASCE 7 design spectrum). Shading indicates
that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 6.2 wV-34-f1 13.5 w28-W-f1 25.7 w28-W-f2 27.4
2 wV-24-f1 0.0 wV-24-f1 12.2 wN-5-f2 14.4 wV-34-f1 26.6 wP-29-f1 27.8
3 wQ-34-f2 0.0 wV-34-f2 12.2 w5-O-f2 17.5 wV-24-f1 27.4 wO-5-f1 29.2
4 wQ-24-f2 0.0 wN-5-f2 12.2 wN-5-f1 17.9 wO-29-f2 27.7 wO-29-f1 29.8
5 wU-34-f2 0.0 wN-5-f1 14.0 wS-29-f2 22.1 w5-O-f2 27.8 wO-29-f2 31.2
6 wU-24-f2 0.0 w5-O-f2 14.3 wS-5-f2 22.1 wP-29-f1 28.2 w28-W-f1 35.0
7 wP-5-f1 0.2 wS-29-f1 17.2 wP-5-f1 22.1 wV-34-f2 28.2 wN-5-f1 36.1
8 wO-5-f1 0.2 wS-29-f2 17.2 wS-29-f1 22.1 wP-5-f1 28.2 wP-5-f1 37.7
9 wP-29-f1 0.2 wS-5-f2 17.2 wS-5-f1 22.1 wN-5-f2 28.3 wV-34-f1 38.9
10 wS-29-f1 0.2 wO-29-f2 17.2 wO-5-f1 22.1 wN-5-f1 28.7 wV-24-f1 46.3
11 wS-5-f1 0.2 wP-5-f1 17.2 wO-29-f2 22.2 wO-5-f1 28.8
12 wS-5-f2 0.2 wS-5-f1 17.3 wP-35-f2 22.5 wP-35-f2 30.7
13 wS-29-f2 0.2 wP-23-f2 17.5 wV-24-f1 22.6 wS-29-f1 30.7
14 wV-24-f2 0.2 wP-35-f2 17.5 w28-W-f1 23.0 wS-5-f2 30.7
15 wN-5-f1 0.4 wO-5-f1 17.8 wV-34-f2 23.8 wS-29-f2 30.7
16 wN-5-f2 0.4 wP-29-f1 17.9 wP-29-f1 25.6 wS-5-f1 30.7
17 w5-O-f2 0.4 wV-24-f2 21.1 w28-W-f2 25.7 wP-23-f2 31.2
18 wO-29-f2 0.4 wO-29-f1 22.1 wU-34-f2 25.7 wO-29-f1 31.8
19 wP-23-f2 0.5 w28-W-f1 22.4 wV-24-f2 26.6 wV-24-f2 33.4
20 wP-35-f2 0.5 w28-W-f2 23.1 wU-24-f2 27.3 wU-24-f2 33.4
21 wV-34-f2 0.5 wU-24-f2 23.3 wO-29-f1 27.9 w24-V1-f2 33.4
22 w28-W-f1 0.6 wNp-5-f1 23.3 w34-V1-f2 28.3 w28-W-f2 33.4
23 wO-29-f1 0.7 wU-34-f2 24.4 wP-23-f2 28.3 wU-34-f2 35.5
24 wU-31-f1 0.9 wN-30-f1 25.7 wN-30-f2 32.9 w34-V1-f2 35.5
25 wU-27-f1 0.9 w34-V1-f2 28.2 w24-V1-f2 33.4 w5-R-f1 43.8
26 w24-R-f1 1.0 w34-R-f1 28.2 w5-R-f1 35.1
27 w24-T-f1 1.0 w34-T-f1 28.2 wN-30-f1 35.6
28 wN-30-f2 1.0 w5-R-f1 29.1 wNp-5-f1 35.7
29 wN-30-f1 1.0 w29-R-f1 29.4 w29-R-f1 38.2
30 w28-W-f2 1.1 wN-30-f2 30.7 w24-R-f1 38.2
31 w34-R-f1 1.3 w24-V1-f2 32.9 w24-T-f1 38.2
32 w34-T-f1 1.3 w24-R-f1 33.4 w34-R-f1 45.8
33 wU-31-f2 1.6 w24-T-f1 33.4 w34-T-f1 45.8
34 wU-27-f2 1.6 wU-31-f1 35.7 w30-Np-f1 45.8
35 wU-24-f1 1.6 wU-27-f1 35.7
36 wU-34-f1 1.6 w30-Np-f1 45.8
37 wNp-5-f1 3.9 wU-31-f2 47.2
38 w30-Np-f1 4.1 wU-27-f2 47.2
39 w24-V1-f2 4.7 wU-34-f1 49.1
40 w34-V1-f2 5.1
41 wQ-34-f1 10.2
42 w5-R-f1 13.1
43 wQ-24-f1 13.1
44 w29-R-f1 13.5
45 w30-Np-f2 14.0
46 w25p-O1-f1 14.7
897

Table G.33: Sequence of exceedence of steel and concrete strain limits (Valparaiso ground
motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also exceeded
the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 w5-O-f2 18.9 wV-24-f2 33.4 wV-34-f2 49.1 w28-W-f2 27.4
2 wNp-5-f1 21.5 wV-34-f2 35.5 wP-29-f1 27.8
3 wS-29-f1 22.2 wP-29-f1 35.7 wO-5-f1 29.2
4 wP-5-f1 22.2 wP-35-f2 35.7 wO-29-f1 29.8
5 wN-5-f1 22.8 w34-V1-f2 49.1 wO-29-f2 31.2
6 wN-5-f2 22.8 w28-W-f1 35.0
7 wO-29-f2 22.9 wN-5-f1 36.1
8 w29-R-f1 23.1 wP-5-f1 37.7
9 w25p-O1-f1 23.3 wV-34-f1 38.9
10 w24-V1-f2 23.3 wV-24-f1 46.3
11 w5-R-f1 23.7
12 w28-W-f1 23.7
13 wP-29-f1 25.0
14 w34-V1-f2 25.6
15 wV-34-f2 25.7
16 w28-W-f2 25.7
17 wV-34-f1 26.6
18 wV-24-f2 26.8
19 wV-24-f1 27.3
20 wO-5-f1 27.8
21 wO-29-f1 27.8
22 wS-29-f2 27.9
23 wP-35-f2 27.9
24 wU-34-f2 28.2
25 wP-23-f2 29.2
26 wS-5-f2 30.7
27 wS-5-f1 30.7
28 wN-30-f2 30.7
29 wN-30-f1 31.8
30 wU-24-f2 32.9
31 w30-Np-f1 38.2
32 w34-R-f1 49.6
33 w34-T-f1 49.6
898

G.12 Valparaiso Almendral (MCE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.45: Summary of walls exceeding acceptance criteria (Valparaiso Almendral ground
motion scaled to ASCE 7 MCE spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.46: Walls exceeding
45
ASCE 41 rotation limits (Valparaiso Almendral ground motion scaled to ASCE 7 MCE spectrum).
40

899
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.47: Walls exceeding
45
fragility function probabilities of 0.50 (Valparaiso Almendral ground motion scaled to ASCE 7
MCE spectrum). 40

900
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation
Figure G.48: Walls45exceeding strain limits (ValparaisoFragility Functions
Almendral ground motionStrain
scaled to ASCE 7 MCE spectrum).

901
40

35
# of Predicted walls

30

25

20

15
902

Table G.34: Sequence of exceedence of ASCE 41 acceptance criteria (Valparaiso Almendral


ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wN-5-f2 13.1 w28-W-f1 14.6 w28-W-f1 33.1 w28-W-f2 14.6
2 w28-W-f1 13.7 wP-5-f1 15.4 wV-24-f1 35.2 wP-29-f1 15.0
3 wO-29-f2 14.1 wP-29-f1 15.4 w24-V1-f2 35.2 w28-W-f1 16.3
4 wP-29-f1 14.5 wO-29-f2 27.9 wP-5-f1 35.2 wO-5-f1 25.8
5 wP-5-f1 14.5 wV-34-f1 28.5 wV-34-f2 35.2 wN-5-f1 28.4
6 w5-O-f2 14.9 w5-O-f2 28.8 wP-29-f1 35.2 wO-29-f1 31.6
7 w28-W-f2 15.0 wN-5-f2 31.1 wS-29-f2 35.3 wO-29-f2 32.3
8 wN-5-f1 17.2 wN-5-f1 31.6 wN-5-f1 35.6 wP-5-f1 34.6
9 wO-5-f1 18.1 wO-5-f1 31.7 wO-5-f1 35.7 w5-O-f2 35.2
10 wV-24-f1 21.9 wP-23-f2 32.2 wO-29-f1 35.8 wS-5-f1 36.0
11 wV-34-f1 22.4 wS-29-f1 33.0 wO-29-f2 37.7 wV-24-f1 36.7
12 wS-29-f1 27.2 w34-V1-f2 33.1 w5-O-f2 37.7 wV-34-f1 37.5
13 wS-5-f1 27.2 w28-W-f2 33.5 w29-R-f1 40.1 w29-R-f1 37.6
14 wS-29-f2 27.6 wO-29-f1 33.6 w5-R-f1 44.0 wS-29-f1 49.9
15 wS-5-f2 27.6 wS-5-f1 33.6 wN-5-f2 49.2 wV-34-f2 53.5
16 w34-V1-f2 28.3 w24-V1-f2 34.8 wV-34-f1 54.4
17 wV-34-f2 29.2 wV-24-f1 35.1
18 w24-V1-f2 29.3 wS-29-f2 35.1
19 wP-23-f2 31.2 wV-34-f2 35.2
20 wP-35-f2 31.2 wS-5-f2 35.2
21 wO-29-f1 31.6 wP-35-f2 35.2
22 wV-24-f2 33.0 wV-24-f2 35.2
23 w29-R-f1 33.1 wN-30-f1 35.8
24 w34-R-f1 33.1 wNp-5-f1 35.8
25 w34-T-f1 33.1 w5-R-f1 38.1
26 w24-R-f1 33.1 w29-R-f1 40.0
27 w24-T-f1 33.1 w24-R-f1 48.0
28 w5-R-f1 35.1 w24-T-f1 48.0
29 w30-Np-f1 35.2
30 wU-34-f2 35.6
31 wN-30-f1 35.7
32 wNp-5-f1 35.7
33 wU-24-f2 48.0
903

Table G.35: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Valparaiso Almendral ground motion scaled to ASCE 7 MCE spectrum). Shading
indicates that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-24-f1 4.5 wN-5-f2 12.0 w28-W-f1 13.7 w28-W-f2 14.6
2 wV-24-f1 0.0 wV-34-f1 4.8 w5-O-f2 12.0 wP-5-f1 14.6 wP-29-f1 15.0
3 wQ-34-f2 0.0 wN-5-f2 11.3 wN-5-f1 12.4 wP-29-f1 14.6 w28-W-f1 16.3
4 wQ-24-f2 0.0 w5-O-f2 11.3 w28-W-f1 13.0 wV-24-f2 14.6 wO-5-f1 25.8
5 wU-34-f2 0.0 wN-5-f1 11.3 wO-29-f2 13.3 wO-29-f2 14.9 wN-5-f1 28.4
6 wU-24-f2 0.0 w28-W-f1 11.4 wP-29-f1 13.7 w5-O-f2 15.0 wO-29-f1 31.6
7 wP-5-f1 0.1 wP-5-f1 11.7 wP-5-f1 13.7 wV-34-f2 15.4 wO-29-f2 32.3
8 wO-5-f1 0.1 wO-29-f2 11.7 w28-W-f2 13.7 wN-5-f2 17.2 wP-5-f1 34.6
9 wS-29-f1 0.1 wP-29-f1 12.0 wV-24-f2 13.8 wN-5-f1 18.1 w5-O-f2 35.2
10 wP-29-f1 0.1 wO-5-f1 12.3 wV-34-f2 14.5 wV-34-f1 25.9 wS-5-f1 36.0
11 wS-5-f1 0.2 wP-35-f2 12.4 wU-24-f2 14.6 wO-5-f1 26.3 wV-24-f1 36.7
12 wS-5-f2 0.2 w28-W-f2 13.0 wU-34-f2 15.4 wV-24-f1 27.7 wV-34-f1 37.5
13 wS-29-f2 0.2 wV-24-f2 13.1 w24-V1-f2 15.4 w28-W-f2 27.9 w29-R-f1 37.6
14 wN-5-f2 0.2 wP-23-f2 13.1 wN-30-f1 15.4 w34-V1-f2 28.4 wS-29-f1 49.9
15 wN-5-f1 0.2 wO-29-f1 13.2 wO-5-f1 15.7 wU-34-f2 28.5 wV-34-f2 53.5
16 w5-O-f2 0.2 wU-24-f2 13.4 wP-35-f2 17.2 w24-V1-f2 29.3
17 wV-24-f2 0.2 wV-34-f2 13.7 wS-29-f2 17.2 wU-24-f2 29.4
18 wP-35-f2 0.2 wU-34-f2 13.7 wS-5-f2 17.2 wO-29-f1 31.7
19 wO-29-f1 0.3 w29-R-f1 14.5 wS-29-f1 18.1 wS-29-f1 31.8
20 w28-W-f1 0.3 wN-30-f1 14.6 wS-5-f1 18.1 wS-5-f1 31.8
21 wO-29-f2 0.4 w5-R-f1 14.6 wO-29-f1 18.1 wP-23-f2 32.2
22 wP-23-f2 0.4 w34-V1-f2 15.4 wV-34-f1 18.1 wP-35-f2 32.2
23 wV-34-f2 0.5 w34-R-f1 15.4 wV-24-f1 20.6 wS-5-f2 32.9
24 wU-31-f1 0.6 w34-T-f1 15.4 w5-R-f1 25.7 wS-29-f2 32.9
25 wU-27-f1 0.6 w24-R-f1 15.4 w34-V1-f2 25.7 w25p-O1-f1 35.2
26 wU-31-f2 0.6 w24-T-f1 15.4 wP-23-f2 26.8 w24-R-f1 35.2
27 wU-27-f2 0.6 w24-V1-f2 15.4 w29-R-f1 28.4 w24-T-f1 35.2
28 wN-30-f1 0.7 wS-29-f1 15.7 w34-R-f1 28.4 wN-30-f1 35.3
29 wN-30-f2 0.7 wNp-5-f1 16.0 w34-T-f1 28.4 w34-R-f1 35.3
30 w28-W-f2 0.8 wS-5-f2 16.4 w24-R-f1 33.1 w34-T-f1 35.3
31 wU-24-f1 0.9 wS-5-f1 16.8 w24-T-f1 33.1 wNp-5-f1 35.7
32 w24-R-f1 0.9 wS-29-f2 16.8 w30-Np-f1 33.1 w5-R-f1 37.7
33 w24-T-f1 0.9 w30-Np-f1 28.4 wNp-5-f1 33.3 w29-R-f1 37.8
34 w34-R-f1 1.0 wN-30-f2 31.8 w25p-O1-f1 35.2 w30-Np-f1 47.8
35 w34-T-f1 1.0 w25p-O1-f1 33.4 wU-31-f2 35.2
36 w34-V1-f2 1.3 wU-31-f1 35.1 wU-27-f2 35.2
37 wU-34-f1 1.3 wU-27-f1 35.1 wN-30-f2 35.7
38 w24-V1-f2 1.6 wU-31-f2 35.2 wU-24-f1 47.6
39 wNp-5-f1 2.0 wU-27-f2 35.2 wU-34-f1 47.6
40 w30-Np-f1 2.1 wU-34-f1 35.2
41 w29-R-f1 5.0 wU-24-f1 35.2
42 w30-Np-f2 6.0 wQ-34-f1 47.6
43 wQ-34-f1 7.2 wQ-24-f1 47.6
44 wQ-24-f1 8.6
45 w5-R-f1 11.1
46 w25p-O1-f1 11.4
904

Table G.36: Sequence of exceedence of steel and concrete strain limits (Valparaiso Almendral
ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wNp-5-f1 12.4 wP-29-f1 33.6 w28-W-f1 35.3 w28-W-f2 14.6
2 w5-R-f1 13.0 wV-24-f2 34.6 w5-R-f1 37.7 wP-29-f1 15.0
3 w25p-O1-f1 13.0 w5-R-f1 34.8 wP-29-f1 40.1 w28-W-f1 16.3
4 w29-R-f1 13.0 w24-V1-f2 35.2 wP-5-f1 47.0 wO-5-f1 25.8
5 wP-5-f1 13.0 wV-34-f2 35.2 wV-24-f2 47.6 wN-5-f1 28.4
6 w5-O-f2 13.0 wV-34-f1 35.2 wO-29-f1 31.6
7 wN-5-f2 13.1 wS-29-f2 35.2 wO-29-f2 32.3
8 wN-5-f1 13.2 w28-W-f1 35.2 wP-5-f1 34.6
9 wO-29-f2 13.3 wU-24-f2 35.2 w5-O-f2 35.2
10 w28-W-f1 13.4 wS-5-f1 35.3 wS-5-f1 36.0
11 wP-29-f1 13.7 wN-5-f1 35.4 wV-24-f1 36.7
12 w34-V1-f2 13.7 wO-29-f1 35.7 wV-34-f1 37.5
13 w28-W-f2 13.7 w29-R-f1 37.7 w29-R-f1 37.6
14 w24-V1-f2 13.8 wO-29-f2 37.8 wS-29-f1 49.9
15 wV-34-f1 14.5 wP-5-f1 38.9 wV-34-f2 53.5
16 wV-24-f2 14.5 wV-24-f1 47.5
17 wO-5-f1 15.1 w34-V1-f2 54.4
18 wV-34-f2 15.4
19 wO-29-f1 16.8
20 wS-29-f1 18.1
21 wU-34-f2 25.3
22 wP-35-f2 26.3
23 wS-5-f1 26.3
24 wS-29-f2 27.2
25 wS-5-f2 27.2
26 wV-24-f1 27.5
27 w30-Np-f1 28.3
28 wU-24-f2 28.9
29 wP-23-f2 31.1
30 wN-30-f1 31.6
31 wN-30-f2 31.8
32 w24-R-f1 35.2
33 w24-T-f1 35.2
34 w34-R-f1 35.2
35 w34-T-f1 35.2
905

G.13 Vina del Mar - Centro (DBE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.49: Summary of walls exceeding acceptance criteria (Vina del Mar - Centro ground
motion scaled to ASCE 7 design spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation
Figure G.50: Walls exceeding
45
ASCE 41 rotation limits Fragility
(Vina Functions Strain
del Mar - Centro ground motion scaled to ASCE 7 design
spectrum). 40

906
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.51: Walls exceeding
45
fragility function probabilities of 0.50 (Vina del Mar - Strain
Centro ground motion scaled to ASCE 7
design spectrum). 40

907
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure G.52: Walls 45exceedingASCE Rotation


strain Fragility Functions
limits (Vina del Mar - Centro ground motionStrain
scaled to ASCE 7 design spectrum).

908
40

35
# of Predicted walls

30

25

20

15
909

Table G.37: Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - Centro
ground motion scaled to ASCE 7 design spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wV-34-f1 38.1 w28-W-f1 43.2 wN-5-f2 49.2 w28-W-f1 43.1
2 wV-24-f1 39.9 wP-5-f1 43.2 w28-W-f1 55.9 wP-29-f1 43.7
3 w28-W-f1 40.3 wP-29-f1 43.3 wO-29-f2 56.4 wO-29-f1 44.4
4 wN-5-f1 42.2 wO-29-f2 43.7 wO-5-f1 47.9
5 wN-5-f2 42.2 wN-5-f1 43.8 wP-5-f1 55.4
6 w5-O-f2 42.8 wN-5-f2 43.9 wN-5-f1 55.9
7 wO-29-f2 42.8 w5-O-f2 44.7 wV-24-f1 65.0
8 wP-5-f1 43.1 wV-34-f1 47.8 wV-34-f1 66.1
9 wP-29-f1 43.2 wO-5-f1 48.6 wV-34-f2 66.1
10 w24-V1-f2 43.3 wO-29-f1 48.6
11 wO-5-f1 43.4 wS-29-f1 48.7
12 w28-W-f2 43.7 wP-23-f2 49.2
13 wNp-5-f1 43.8 w28-W-f2 55.4
14 wO-29-f1 43.8 w24-V1-f2 55.4
15 wP-35-f2 43.9 wP-35-f2 56.9
16 wS-29-f2 45.3 wS-5-f2 57.0
17 wS-5-f2 45.3 wS-29-f2 57.0
18 wS-29-f1 45.3 wS-5-f1 57.1
19 wS-5-f1 45.3 wV-24-f1 65.0
20 wP-23-f2 45.3
21 wV-24-f2 45.7
22 w34-V1-f2 47.8
23 w5-R-f1 49.0
24 wV-34-f2 54.9
25 w34-R-f1 55.9
26 w34-T-f1 55.9
27 w29-R-f1 62.3
28 w24-R-f1 64.6
29 w24-T-f1 64.6
910

Table G.38: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - Centro ground motion scaled to ASCE 7 design spectrum). Shad-
ing indicates that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 w5-O-f2 22.9 wN-5-f2 32.8 wV-34-f1 41.7 w28-W-f1 43.1
2 wV-24-f1 0.0 wN-5-f2 22.9 w5-O-f2 33.1 wN-5-f2 42.3 wP-29-f1 43.7
3 wQ-34-f2 0.0 wN-5-f1 23.7 w28-W-f1 34.9 w5-O-f2 42.4 wO-29-f1 44.4
4 wQ-24-f2 0.0 wV-34-f1 30.6 wN-5-f1 37.0 w28-W-f1 42.8 wO-5-f1 47.9
5 wU-34-f2 0.0 wV-24-f1 31.9 wP-5-f1 37.0 wP-5-f1 43.2 wP-5-f1 55.4
6 wU-24-f2 0.0 wP-5-f1 32.8 wO-5-f1 37.0 wV-24-f2 43.2 wN-5-f1 55.9
7 wP-5-f1 0.7 wO-5-f1 32.9 wV-34-f1 37.0 wP-29-f1 43.2 wV-24-f1 65.0
8 wO-5-f1 0.7 wO-29-f2 33.1 wS-29-f1 37.0 wV-24-f1 43.3 wV-34-f1 66.1
9 wN-5-f2 0.7 w28-W-f1 34.3 wS-29-f2 37.0 w24-V1-f2 43.3 wV-34-f2 66.1
10 wN-5-f1 0.8 wP-29-f1 34.3 wS-5-f2 37.0 wU-24-f2 43.3
11 wP-29-f1 0.8 wU-24-f2 35.2 wV-24-f1 38.6 wN-5-f1 43.4
12 w28-W-f1 0.8 w28-W-f2 35.2 wS-5-f1 39.2 wO-5-f1 43.5
13 w5-O-f2 1.0 wP-35-f2 35.6 wP-29-f1 40.3 wO-29-f2 43.7
14 wV-34-f2 1.3 wS-29-f2 36.6 wV-34-f2 40.3 w28-W-f2 44.7
15 wS-29-f1 2.2 wS-5-f2 36.6 wO-29-f2 41.3 wV-34-f2 45.2
16 wS-5-f1 2.2 wS-5-f1 36.6 wV-24-f2 41.7 wS-5-f2 45.3
17 wS-29-f2 2.2 wP-23-f2 36.6 wP-35-f2 42.4 wP-35-f2 45.3
18 wS-5-f2 2.2 wS-29-f1 36.6 wU-24-f2 42.8 wS-29-f2 45.3
19 wV-24-f2 2.3 wO-29-f1 37.0 w28-W-f2 42.8 wS-29-f1 45.3
20 wO-29-f2 2.4 wV-24-f2 37.0 w24-V1-f2 43.2 wS-5-f1 45.4
21 wP-23-f2 2.4 wV-34-f2 37.5 wN-30-f1 43.3 wO-29-f1 45.4
22 wP-35-f2 2.4 wU-34-f2 39.9 wO-29-f1 43.3 w34-V1-f2 47.8
23 wO-29-f1 2.6 wNp-5-f1 42.8 wNp-5-f1 43.7 wU-34-f2 47.9
24 w28-W-f2 4.9 w29-R-f1 43.2 wP-23-f2 43.9 wP-23-f2 48.1
25 wN-30-f1 5.9 w24-V1-f2 43.2 wU-34-f2 44.2 w5-R-f1 49.0
26 wN-30-f2 5.9 wN-30-f1 43.2 w24-R-f1 44.2 w29-R-f1 62.4
27 w34-R-f1 9.9 w24-R-f1 43.2 w24-T-f1 44.2
28 w34-T-f1 9.9 w24-T-f1 43.2 w34-V1-f2 47.5
29 w30-Np-f1 10.1 w5-R-f1 43.2 w29-R-f1 47.9
30 wNp-5-f1 10.3 w34-R-f1 43.2 w34-R-f1 48.0
31 wU-31-f1 10.9 w34-T-f1 43.2 w34-T-f1 48.0
32 wU-27-f1 10.9 w34-V1-f2 44.2 w5-R-f1 48.9
33 w24-R-f1 11.3 wN-30-f2 45.4 wN-30-f2 49.2
34 w24-T-f1 11.3 w30-Np-f1 55.9
35 wQ-34-f1 22.6 wU-31-f1 57.1
36 wU-24-f1 22.6 wU-27-f1 57.1
37 wU-31-f2 22.6
38 wU-27-f2 22.6
39 wU-34-f1 22.6
40 w34-V1-f2 22.7
41 wQ-24-f1 22.9
42 w29-R-f1 23.0
43 w5-R-f1 23.0
44 w30-Np-f2 23.4
45 w25p-O1-f1 23.7
46 w24-V1-f2 24.1
911

Table G.39: Sequence of exceedence of steel and concrete strain limits (Vina del Mar -
Centro ground motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wNp-5-f1 34.8 wP-29-f1 48.6 w28-W-f1 43.1
2 wP-5-f1 34.9 wO-29-f1 48.8 wP-29-f1 43.7
3 wO-29-f2 35.2 wV-24-f2 55.4 wO-29-f1 44.4
4 w25p-O1-f1 35.5 wV-34-f2 56.4 wO-5-f1 47.9
5 w5-O-f2 35.6 wP-5-f1 55.4
6 wV-34-f2 39.9 wN-5-f1 55.9
7 w34-V1-f2 39.9 wV-24-f1 65.0
8 w5-R-f1 40.3 wV-34-f1 66.1
9 wP-29-f1 40.3 wV-34-f2 66.1
10 w29-R-f1 40.3
11 w28-W-f1 40.3
12 wV-34-f1 40.5
13 wV-24-f2 40.5
14 w24-V1-f2 41.1
15 wN-5-f1 42.2
16 wN-5-f2 42.2
17 wO-5-f1 42.8
18 w28-W-f2 42.8
19 wO-29-f1 42.8
20 wV-24-f1 43.3
21 w30-Np-f1 43.3
22 wP-35-f2 43.4
23 wS-5-f1 43.4
24 wS-29-f1 43.5
25 wU-24-f2 43.7
26 wP-23-f2 43.8
27 wS-29-f2 45.3
28 wS-5-f2 45.3
29 wN-30-f2 45.4
30 wN-30-f1 45.5
31 wU-34-f2 47.4
912

G.14 Vina del Mar - Centro (MCE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.53: Summary of walls exceeding acceptance criteria (Vina del Mar - Centro ground
motion scaled to ASCE 7 MCE spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.54: Walls exceeding
45
ASCE 41 rotation limits (Vina del Mar - Centro ground motion scaled to ASCE 7 MCE spectrum).
40

913
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions


Figure G.55: Walls exceeding
45
fragility function probabilities of 0.50 (Vina del Mar - Strain
Centro ground motion scaled to ASCE 7
MCE spectrum). 40

914
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure G.56: Walls45exceedingASCE Rotation


strain Fragility Functions
limits (Vina del Mar Strain
- Centro ground motion scaled to ASCE 7 MCE spectrum).

915
40

35
# of Predicted walls

30

25

20

15
916

Table G.40: Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - Centro
ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall also
exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 w28-W-f1 36.5 wV-34-f1 42.0 w28-W-f1 44.2 w28-W-f1 37.6
2 wO-29-f2 38.8 wV-24-f1 43.1 wV-24-f2 45.5 wO-29-f1 43.4
3 wN-5-f2 39.1 w28-W-f1 43.3 wN-5-f1 48.6 wP-29-f1 43.6
4 wV-34-f1 39.2 wN-5-f2 43.4 wP-5-f1 48.6 wV-24-f1 43.8
5 wV-24-f1 39.5 wN-5-f1 43.4 wO-5-f1 48.6 wO-5-f1 44.1
6 w5-O-f2 40.8 wP-5-f1 43.4 wN-5-f2 48.6 wN-5-f1 44.3
7 wV-34-f2 41.3 wO-5-f1 43.4 wP-35-f2 48.8 w28-W-f2 44.3
8 wS-5-f1 41.3 wS-29-f1 43.5 wO-29-f2 55.5 wV-34-f1 44.7
9 wS-29-f2 41.3 wS-5-f2 43.5 w5-O-f2 55.5 wO-29-f2 48.0
10 wS-5-f2 41.3 wS-5-f1 43.5 w5-R-f1 63.8 w29-R-f1 57.5
11 wS-29-f1 41.3 wP-35-f2 43.5
12 wN-5-f1 41.7 wS-29-f2 43.5
13 wP-5-f1 41.8 wP-29-f1 43.5
14 wV-24-f2 42.2 w5-O-f2 43.7
15 wP-29-f1 43.3 wV-24-f2 43.7
16 wO-5-f1 43.4 wV-34-f2 43.7
17 wP-35-f2 43.4 w24-V1-f2 43.8
18 wO-29-f1 43.4 w34-V1-f2 44.2
19 w28-W-f2 43.7 wO-29-f2 44.7
20 w24-V1-f2 43.7 w28-W-f2 44.8
21 w34-V1-f2 44.1 wP-23-f2 48.0
22 wP-23-f2 44.1 wO-29-f1 48.6
23 w5-R-f1 44.1 w5-R-f1 55.1
24 w34-R-f1 44.2
25 w34-T-f1 44.2
26 w29-R-f1 44.3
27 w24-R-f1 44.3
28 w24-T-f1 44.3
29 wU-34-f2 48.0
30 wNp-5-f1 48.7
31 wN-30-f2 48.7
917

Table G.41: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - Centro ground motion scaled to ASCE 7 MCE spectrum). Shading
indicates that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wN-5-f2 22.2 w28-W-f1 25.6 w28-W-f1 38.8 w28-W-f1 37.6
2 wV-24-f1 0.0 w5-O-f2 22.4 wP-29-f1 36.5 wV-34-f1 39.3 wO-29-f1 43.4
3 wQ-34-f2 0.0 w28-W-f1 23.8 wO-29-f2 36.7 wV-24-f1 40.0 wP-29-f1 43.6
4 wQ-24-f2 0.0 wN-5-f1 24.5 w5-O-f2 36.7 wN-5-f2 42.1 wV-24-f1 43.8
5 wU-34-f2 0.0 wP-5-f1 24.6 wV-34-f1 37.0 wS-5-f1 43.0 wO-5-f1 44.1
6 wU-24-f2 0.0 wO-29-f2 25.3 wN-5-f2 37.2 wS-29-f2 43.0 wN-5-f1 44.3
7 wP-5-f1 0.3 wP-29-f1 25.6 wN-5-f1 37.5 wS-5-f2 43.1 w28-W-f2 44.3
8 wO-5-f1 0.3 wV-34-f1 25.7 wP-5-f1 37.5 wV-34-f2 43.1 wV-34-f1 44.7
9 w28-W-f1 0.4 w28-W-f2 25.7 w28-W-f2 38.7 wP-5-f1 43.3 wO-29-f2 48.0
10 w5-O-f2 0.5 wV-24-f1 26.2 wV-24-f1 38.8 wV-24-f2 43.3 w29-R-f1 57.5
11 wN-5-f2 0.5 wV-34-f2 36.4 wS-29-f2 39.2 wP-29-f1 43.3
12 wP-29-f1 0.6 wO-5-f1 36.9 wS-5-f2 39.2 w5-O-f2 43.4
13 wN-5-f1 0.7 wP-35-f2 37.2 wS-29-f1 39.2 wN-5-f1 43.4
14 wV-34-f2 0.8 wO-29-f1 37.5 wS-5-f1 39.2 wO-29-f2 43.4
15 wO-29-f2 1.0 wV-24-f2 37.6 wV-34-f2 39.4 wO-5-f1 43.4
16 wV-24-f2 1.0 wU-24-f2 37.6 wU-24-f2 39.4 wP-35-f2 43.4
17 w28-W-f2 1.1 wS-29-f2 37.8 wV-24-f2 40.1 wS-29-f1 43.4
18 wP-23-f2 1.9 wS-5-f2 37.8 wU-34-f2 40.1 wO-29-f1 43.5
19 wP-35-f2 1.9 wS-5-f1 37.9 wO-5-f1 41.7 wU-24-f2 43.7
20 wS-29-f2 1.9 wP-23-f2 37.9 wO-29-f1 41.8 w24-V1-f2 43.7
21 wS-5-f2 2.0 wS-29-f1 38.2 wP-23-f2 43.0 w28-W-f2 43.7
22 wS-5-f1 2.0 wU-34-f2 38.7 wP-35-f2 43.0 w34-V1-f2 44.1
23 wS-29-f1 2.0 w5-R-f1 41.3 w29-R-f1 43.3 wU-34-f2 44.1
24 wO-29-f1 2.1 w34-V1-f2 43.0 w24-V1-f2 43.7 wP-23-f2 46.6
25 w24-R-f1 4.9 wN-30-f1 43.3 w34-V1-f2 44.0 wNp-5-f1 48.7
26 w24-T-f1 4.9 w29-R-f1 43.3 w5-R-f1 44.1 wN-30-f2 48.7
27 wN-30-f1 4.9 w24-V1-f2 43.3 wN-30-f1 44.1 w5-R-f1 55.0
28 wN-30-f2 4.9 wNp-5-f1 43.4 w34-R-f1 44.2 w29-R-f1 55.1
29 w34-R-f1 5.9 wN-30-f2 43.4 w34-T-f1 44.2
30 w34-T-f1 5.9 wU-31-f1 43.5 w24-R-f1 44.2
31 wNp-5-f1 10.1 wU-27-f1 43.5 w24-T-f1 44.2
32 w30-Np-f1 10.1 wU-31-f2 43.5 wNp-5-f1 44.3
33 wU-31-f1 10.5 wU-27-f2 43.5 w30-Np-f1 44.3
34 wU-27-f1 10.5 w25p-O1-f1 43.5 wN-30-f2 45.8
35 wU-34-f1 11.0 w24-R-f1 43.8
36 w34-V1-f2 11.0 w24-T-f1 43.8
37 wU-24-f1 14.9 w34-R-f1 44.1
38 wU-31-f2 20.0 w34-T-f1 44.1
39 wU-27-f2 20.0 w30-Np-f1 44.2
40 wQ-34-f1 21.6 wU-34-f1 48.0
41 wQ-24-f1 22.2
42 w5-R-f1 23.0
43 w30-Np-f2 23.0
44 w29-R-f1 23.0
45 w24-V1-f2 23.2
46 w25p-O1-f1 23.9
918

Table G.42: Sequence of exceedence of steel and concrete strain limits (Vina del Mar -
Centro ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 w5-R-f1 25.6 wV-24-f2 43.7 wV-24-f2 43.8 w28-W-f1 37.6
2 w29-R-f1 25.7 w24-V1-f2 43.8 wV-34-f2 44.7 wO-29-f1 43.4
3 w28-W-f1 25.7 w28-W-f1 44.3 wP-29-f1 43.6
4 wP-29-f1 25.7 wV-34-f2 44.6 wV-24-f1 43.8
5 w5-O-f2 25.9 w34-V1-f2 48.0 wO-5-f1 44.1
6 wO-29-f2 25.9 wP-29-f1 48.6 wN-5-f1 44.3
7 wO-5-f1 36.8 wP-35-f2 48.7 w28-W-f2 44.3
8 wNp-5-f1 37.2 wO-29-f2 48.7 wV-34-f1 44.7
9 wN-5-f2 37.6 wO-29-f1 48.9 wO-29-f2 48.0
10 w25p-O1-f1 37.6 w29-R-f1 57.5
11 wP-5-f1 37.6
12 w28-W-f2 38.8
13 w34-V1-f2 38.8
14 wN-5-f1 39.1
15 wV-34-f1 39.3
16 wV-24-f2 39.4
17 w24-V1-f2 39.4
18 wV-24-f1 39.4
19 wV-34-f2 39.9
20 wS-5-f1 41.3
21 wS-29-f1 41.3
22 wS-5-f2 41.3
23 wS-29-f2 41.3
24 wO-29-f1 41.7
25 wP-35-f2 43.4
26 wP-23-f2 43.4
27 wN-30-f2 43.4
28 wN-30-f1 43.5
29 wU-24-f2 43.7
30 wU-34-f2 44.0
31 w30-Np-f1 44.2
32 w34-R-f1 44.3
33 w34-T-f1 44.3
919

G.15 Vina del Mar - El Salto (DBE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.57: Summary of walls exceeding acceptance criteria (Vina del Mar - El Salto
ground motion scaled to ASCE 7 design spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.58: Walls exceeding
45
ASCE 41 rotation limits (Vina del Mar - El Salto ground motion scaled to ASCE 7 design
spectrum). 40

920
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.59: Walls exceeding
45
fragility function probabilities of 0.50 (Vina del Mar - El Salto ground motion scaled to ASCE 7
design spectrum). 40

921
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.60: Walls exceeding
45
strain limits (Vina del Mar - El Salto ground motion scaled to ASCE 7 design spectrum).

922
40

35
# of Predicted walls

30

25

20

15
923

Table G.43: Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - El
Salto ground motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wV-34-f1 40.4 wN-5-f2 55.1 w28-W-f1 60.7 w28-W-f1 53.4
2 w28-W-f1 53.3 wN-5-f1 55.6 wV-34-f1 60.8 wO-5-f1 55.1
3 wN-5-f2 53.4 wP-5-f1 55.7 wN-5-f2 61.8 wO-29-f1 55.6
4 wP-5-f1 53.8 wO-5-f1 55.7 wP-5-f1 68.3 wO-29-f2 56.2
5 wS-29-f2 54.2 wP-23-f2 56.3 wN-5-f1 68.3 wP-29-f1 57.1
6 wP-23-f2 54.2 wP-29-f1 56.8 wO-5-f1 68.3 wV-34-f2 60.2
7 wS-5-f2 54.2 wP-35-f2 56.8 w5-O-f2 69.3 wN-5-f1 60.7
8 wN-5-f1 54.2 wS-29-f2 57.3 wO-29-f2 69.4 wV-24-f1 60.8
9 wV-24-f1 54.3 w28-W-f1 57.4 w28-W-f2 69.4 wP-35-f2 60.8
10 wO-5-f1 54.6 wO-29-f2 59.5 wP-29-f1 71.4 wP-5-f1 61.1
11 wP-29-f1 54.6 wV-34-f1 60.5 w34-V1-f2 71.4 w29-R-f1 61.2
12 wS-29-f1 54.7 wS-5-f2 60.7 w24-V1-f2 71.5 wS-5-f1 67.3
13 wP-35-f2 54.7 wS-29-f1 60.7 wV-24-f2 72.9 wV-34-f1 72.2
14 wS-5-f1 54.7 w34-V1-f2 60.7 wS-5-f2 72.9
15 wO-29-f1 55.6 wS-5-f1 60.8 wV-34-f2 72.9
16 wO-29-f2 56.3 wV-24-f1 60.8 wS-29-f1 72.9
17 wV-34-f2 57.4 w24-V1-f2 61.0
18 w5-O-f2 57.8 w5-O-f2 61.2
19 w28-W-f2 59.0 wO-29-f1 68.2
20 w24-V1-f2 59.0 wV-34-f2 68.9
21 w34-V1-f2 60.2 w28-W-f2 69.3
22 wV-24-f2 60.2 wV-24-f2 69.3
23 w34-R-f1 60.8 w29-R-f1 71.3
24 w34-T-f1 60.8 w5-R-f1 71.3
25 w24-R-f1 61.2
26 w24-T-f1 61.2
27 wU-24-f2 61.9
28 wN-30-f1 68.2
29 wNp-5-f1 68.3
30 wN-30-f2 68.8
31 w5-R-f1 68.8
32 w29-R-f1 71.3
33 w30-Np-f1 71.4
924

Table G.44: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - El Salto ground motion scaled to ASCE 7 design spectrum).
Shading indicates that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 27.5 wV-34-f1 39.4 w5-O-f2 54.2 w28-W-f1 53.4
2 wV-24-f1 0.0 wV-24-f1 35.8 wV-24-f1 40.0 wN-5-f2 54.2 wO-5-f1 55.1
3 wQ-34-f2 0.0 wV-34-f2 39.9 wN-5-f2 50.4 wN-5-f1 54.6 wO-29-f1 55.6
4 wQ-24-f2 0.0 wV-24-f2 40.3 w5-O-f2 50.4 wP-5-f1 54.6 wO-29-f2 56.2
5 wU-34-f2 0.0 wN-5-f2 40.4 wN-5-f1 50.4 wO-5-f1 54.6 wP-29-f1 57.1
6 wU-24-f2 0.0 wN-5-f1 40.4 w28-W-f1 52.6 wS-5-f2 54.8 wV-34-f2 60.2
7 wV-34-f2 2.3 wS-29-f1 42.3 wO-29-f2 53.0 wS-29-f2 54.8 wN-5-f1 60.7
8 wP-5-f1 4.2 wS-5-f1 42.3 wU-34-f2 53.0 wS-29-f1 54.8 wV-24-f1 60.8
9 wS-29-f1 4.2 w28-W-f1 45.4 wP-29-f1 53.3 wV-34-f1 54.8 wP-35-f2 60.8
10 wN-5-f1 4.7 wP-29-f1 45.4 wV-34-f2 53.4 wS-5-f1 54.8 wP-5-f1 61.1
11 wO-5-f1 4.7 wO-29-f2 45.7 wP-5-f1 53.8 wO-29-f2 55.1 w29-R-f1 61.2
12 wN-5-f2 6.1 wP-5-f1 46.0 wO-5-f1 53.8 wP-23-f2 55.1 wS-5-f1 67.3
13 wP-29-f1 6.6 wS-5-f2 46.3 wS-29-f1 53.8 wV-24-f1 55.5 wV-34-f1 72.2
14 w5-O-f2 7.3 w5-O-f2 46.4 wP-35-f2 53.9 wP-35-f2 55.6
15 wS-29-f2 9.2 wS-29-f2 46.4 wS-29-f2 54.1 wP-29-f1 55.7
16 wS-5-f2 9.2 wO-5-f1 50.8 wS-5-f1 54.1 wO-29-f1 55.7
17 wS-5-f1 9.2 wP-35-f2 50.9 wS-5-f2 54.1 w28-W-f1 56.5
18 wV-24-f2 9.6 wU-24-f2 52.4 wP-23-f2 54.2 wV-24-f2 59.0
19 wO-29-f2 9.7 wU-34-f2 52.9 wO-29-f1 54.6 wV-34-f2 59.0
20 wP-23-f2 10.6 w28-W-f2 53.0 wV-24-f2 54.8 w24-V1-f2 59.1
21 wO-29-f1 10.9 wO-29-f1 53.0 w28-W-f2 57.3 w28-W-f2 60.2
22 wP-35-f2 11.8 wP-23-f2 53.4 wN-30-f2 57.3 w34-V1-f2 60.4
23 w28-W-f1 13.2 wNp-5-f1 55.1 w29-R-f1 57.5 wU-34-f2 60.7
24 w28-W-f2 15.2 wN-30-f2 55.2 wU-24-f2 59.0 wU-24-f2 61.0
25 wN-30-f1 15.2 wN-30-f1 55.8 w24-V1-f2 59.0 wNp-5-f1 62.1
26 wU-31-f1 15.5 w25p-O1-f1 56.8 wN-30-f1 59.1 wN-30-f1 68.3
27 wU-27-f1 15.5 w29-R-f1 57.4 w34-V1-f2 59.1 wN-30-f2 69.4
28 w34-R-f1 16.0 w34-V1-f2 57.5 wNp-5-f1 60.6 w29-R-f1 71.3
29 w34-T-f1 16.0 w24-V1-f2 59.0 w34-R-f1 60.7 w5-R-f1 71.3
30 wU-24-f1 16.8 w24-R-f1 59.0 w34-T-f1 60.7
31 wU-31-f2 16.9 w24-T-f1 59.0 wU-31-f1 60.8
32 wU-27-f2 16.9 w34-R-f1 59.0 wU-27-f1 60.8
33 wU-34-f1 17.3 w34-T-f1 59.0 w24-R-f1 61.2
34 wN-30-f2 19.1 w5-R-f1 60.6 w24-T-f1 61.2
35 w24-R-f1 20.3 w30-Np-f1 60.7 w25p-O1-f1 61.9
36 w24-T-f1 20.3 wU-31-f1 60.7 w5-R-f1 68.8
37 w34-V1-f2 21.7 wU-27-f1 60.7 w30-Np-f1 71.3
38 w24-V1-f2 22.1 wU-31-f2 60.8 wU-31-f2 72.8
39 w30-Np-f1 26.3 wU-27-f2 60.8 wU-27-f2 72.8
40 wNp-5-f1 28.6 wU-24-f1 60.8
41 wQ-34-f1 37.0 wU-34-f1 60.8
42 w5-R-f1 37.0 w30-Np-f2 69.4
43 w29-R-f1 37.0
44 wQ-24-f1 38.6
45 w30-Np-f2 39.4
46 w25p-O1-f1 46.0
925

Table G.45: Sequence of exceedence of steel and concrete strain limits (Vina del Mar - El
Salto ground motion scaled to ASCE 7 design spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 45.7 wV-34-f2 60.2 wV-24-f2 69.3 w28-W-f1 53.4
2 w5-O-f2 46.4 wV-24-f2 61.0 wV-34-f2 71.9 wO-5-f1 55.1
3 wNp-5-f1 50.9 wP-5-f1 61.2 wO-29-f1 55.6
4 wN-5-f2 50.9 wP-23-f2 68.8 wO-29-f2 56.2
5 wV-24-f2 52.3 wS-29-f2 68.9 wP-29-f1 57.1
6 w24-V1-f2 52.4 w5-O-f2 69.4 wV-34-f2 60.2
7 w25p-O1-f1 52.4 w24-V1-f2 69.4 wN-5-f1 60.7
8 wV-34-f1 52.5 wO-29-f2 69.4 wV-24-f1 60.8
9 wP-29-f1 52.6 wP-29-f1 69.4 wP-35-f2 60.8
10 w5-R-f1 52.6 wP-35-f2 69.4 wP-5-f1 61.1
11 w28-W-f1 52.6 w28-W-f1 71.4 w29-R-f1 61.2
12 wV-34-f2 52.9 w34-V1-f2 71.4 wS-5-f1 67.3
13 w34-V1-f2 52.9 wS-5-f2 72.9 wV-34-f1 72.2
14 wO-29-f2 53.0 wV-24-f1 72.9
15 wP-5-f1 53.3
16 w29-R-f1 53.3
17 wN-5-f1 53.8
18 wP-35-f2 54.2
19 wP-23-f2 54.2
20 wS-5-f2 54.3
21 wO-29-f1 54.6
22 wS-29-f1 54.6
23 wS-29-f2 54.7
24 wS-5-f1 54.8
25 wV-24-f1 54.9
26 wN-30-f2 55.2
27 wN-30-f1 55.7
28 wU-34-f2 56.8
29 w28-W-f2 56.9
30 w30-Np-f1 60.7
31 wU-24-f2 61.0
32 w24-T-f1 61.8
33 w24-R-f1 61.9
34 w30-Np-f2 69.5
35 w34-R-f1 71.3
36 w34-T-f1 71.3
926

G.16 Vina del Mar - El Salto (MCE)

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
1

ld
IO

LS

CP
ne

ne

ne

all

h
DS

DS

DS

DS

us
Yie

Sp
No

No

No

Cr
(a) Shear limit not considered.

Observed > Predicted


Observed = Predicted
Observed < Predicted

45 ASCE Rotation Fragility Functions Strain

40

35
# of Predicted walls

30

25

20

15

10

0
r

r
1

ld

r
IO

LS

CP
ne

ne

ne

all

h
ea

ea

ea
DS

DS

DS

DS

us
Yie

Sp
No

No

No
Sh

Sh

Sh
Cr

(b) Shear limit considered.

Figure G.61: Summary of walls exceeding acceptance criteria (Vina del Mar - El Salto
ground motion scaled to ASCE 7 MCE spectrum).
(a) f1: None (b) f1: IO (c) f1: LS (d) f1: CP (e) f1: Shear

(f) f2: None (g) f2: IO (h) f2: LS (i) f2: CP (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation
Figure G.62: Walls exceeding
45
ASCE 41 rotation limits Fragility
(VinaFunctions
del Mar - El Salto Strain
ground motion scaled to ASCE 7 MCE
spectrum). 40

927
35
# of Predicted walls

30

25

20

15

10
(a) f1: DS1 (b) f1: DS2 (c) f1: DS3 (d) f1: DS4 (e) f1: Shear

(f) f2: DS1 (g) f2: DS2 (h) f2: DS3 (i) f2: DS4 (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

ASCE Rotation Fragility Functions Strain


Figure G.63: Walls exceeding
45
fragility function probabilities of 0.50 (Vina del Mar - El Salto ground motion scaled to ASCE 7
MCE spectrum). 40

928
35
# of Predicted walls

30

25

20

15

10
(a) f1: None (b) f1: Yield (c) f1: Spall (d) f1: Crush (e) f1: Shear

(f) f2: None (g) f2: Yield (h) f2: Spall (i) f2: Crush (j) f2: Shear

Observed > Predicted


Observed = Predicted
Observed < Predicted

Figure G.64: Walls 45exceedingASCE Rotation


strain Fragility Functions
limits (Vina del Mar Strain
- El Salto ground motion scaled to ASCE 7 MCE spectrum).

929
40

35
# of Predicted walls

30

25

20

15
930

Table G.46: Sequence of exceedence of ASCE 41 acceptance criteria (Vina del Mar - El
Salto ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear.

ASCE 41
IO LS CP Shear
Name Time Name Time Name Time Name Time
1 wN-5-f2 40.2 w28-W-f1 41.9 wN-5-f1 56.6 wP-29-f1 42.6
2 wN-5-f1 40.2 wN-5-f2 45.5 wN-5-f2 56.7 w28-W-f1 44.4
3 w28-W-f1 40.3 wN-5-f1 46.5 wP-5-f1 56.8 wO-29-f1 45.5
4 wP-5-f1 41.1 wV-34-f1 53.9 wS-29-f1 56.8 wO-5-f1 46.1
5 wP-29-f1 41.1 wP-5-f1 55.6 wO-5-f1 56.8 w28-W-f2 47.1
6 wO-29-f2 41.4 wS-29-f1 55.8 wV-34-f1 58.6 wO-29-f2 51.7
7 w5-O-f2 41.5 wV-24-f1 56.1 w28-W-f1 60.7 wV-34-f2 53.4
8 w28-W-f2 42.3 wS-5-f1 56.2 wP-29-f1 60.7 wP-5-f1 53.7
9 wV-34-f1 43.6 wP-23-f2 56.2 wS-5-f2 60.8 wV-24-f1 55.5
10 wO-5-f1 43.8 wP-35-f2 56.6 w34-V1-f2 60.8 wS-5-f2 55.8
11 wS-29-f1 43.8 wO-5-f1 56.6 wP-35-f2 60.8 wV-34-f1 56.1
12 wS-5-f2 44.7 wP-29-f1 56.6 wO-29-f2 61.3 w5-O-f2 60.7
13 wS-5-f1 44.7 wS-5-f2 56.7 wS-29-f2 68.9 wN-5-f1 60.7
14 wO-29-f1 46.5 wS-29-f2 56.7 wP-23-f2 68.9 w29-R-f1 61.3
15 wP-35-f2 46.5 wO-29-f1 56.8 wV-24-f1 68.9
16 wP-23-f2 47.0 w24-V1-f2 57.1 w5-O-f2 69.3
17 wV-24-f1 49.7 wV-34-f2 58.6 w28-W-f2 69.4
18 wS-29-f2 55.7 wV-24-f2 58.6 w24-V1-f2 71.5
19 wV-24-f2 56.7 w34-V1-f2 58.9 wV-24-f2 72.9
20 wV-34-f2 56.8 wO-29-f2 59.3 wV-34-f2 72.9
21 w24-V1-f2 57.0 w5-O-f2 59.4
22 w5-R-f1 57.4 w28-W-f2 60.3
23 w34-V1-f2 57.9 w5-R-f1 60.7
24 w34-R-f1 59.1 w29-R-f1 61.8
25 w34-T-f1 59.1
26 w24-R-f1 60.7
27 w24-T-f1 60.7
28 wN-30-f1 61.3
29 wNp-5-f1 61.3
30 w29-R-f1 61.7
31 wN-30-f2 68.8
32 wU-34-f2 71.7
931

Table G.47: Sequence of exceedence of 0.50 probability of damage occurrence using fragility
functions (Vina del Mar - El Salto ground motion scaled to ASCE 7 MCE spectrum).
Shading indicates that the wall also exceeded the acceptance criterion for shear.

Fragility Functions
DS1 DS2 DS3 DS4 Shear
Name Time Name Time Name Time Name Time Name Time
1 wV-34-f1 0.0 wV-34-f1 20.7 wV-34-f1 26.6 wN-5-f2 41.0 wP-29-f1 42.6
2 wV-24-f1 0.0 wV-24-f1 21.8 w28-W-f1 39.6 w5-O-f2 41.0 w28-W-f1 44.4
3 wQ-34-f2 0.0 wV-34-f2 26.2 wN-5-f2 39.8 w28-W-f1 41.0 wO-29-f1 45.5
4 wQ-24-f2 0.0 wN-5-f2 26.7 w5-O-f2 39.8 wO-29-f2 41.5 wO-5-f1 46.1
5 wU-34-f2 0.0 wV-24-f2 26.7 wN-5-f1 40.1 wP-5-f1 41.9 w28-W-f2 47.1
6 wU-24-f2 0.0 w28-W-f1 36.9 wP-5-f1 40.1 wP-29-f1 41.9 wO-29-f2 51.7
7 wV-34-f2 1.9 wP-5-f1 39.0 wO-5-f1 40.1 wV-24-f2 42.8 wV-34-f2 53.4
8 wP-5-f1 2.4 wP-29-f1 39.1 wO-29-f2 40.2 wN-5-f1 44.6 wP-5-f1 53.7
9 wO-5-f1 2.4 w5-O-f2 39.2 wP-29-f1 40.3 wO-5-f1 44.7 wV-24-f1 55.5
10 wN-5-f2 3.3 wN-5-f1 39.4 wV-24-f2 41.0 wV-34-f2 47.1 wS-5-f2 55.8
11 wN-5-f1 3.3 wO-5-f1 39.4 wV-24-f1 41.0 wV-24-f1 49.8 wV-34-f1 56.1
12 wS-29-f1 3.4 wS-29-f1 39.5 wP-35-f2 41.0 wV-34-f1 50.3 w5-O-f2 60.7
13 wS-5-f1 3.4 wS-5-f1 39.5 w28-W-f2 41.0 wS-29-f1 54.1 wN-5-f1 60.7
14 wP-29-f1 4.2 wO-29-f2 39.5 wU-24-f2 41.1 wO-29-f1 55.6 w29-R-f1 61.3
15 wS-29-f2 4.2 wS-29-f2 39.5 wV-34-f2 41.9 wS-5-f1 55.7
16 wS-5-f2 4.2 wS-5-f2 39.5 w5-R-f1 41.9 wS-5-f2 55.8
17 wV-24-f2 4.2 wP-35-f2 39.5 wP-23-f2 42.8 wS-29-f2 55.8
18 w5-O-f2 4.5 wP-23-f2 39.8 wO-29-f1 43.2 wP-23-f2 56.1
19 w28-W-f1 5.8 w28-W-f2 39.9 wS-5-f1 43.2 wP-35-f2 56.2
20 wO-29-f2 7.4 wU-34-f2 40.0 wS-29-f2 43.6 w24-V1-f2 57.1
21 wP-23-f2 7.4 wO-29-f1 40.1 wS-5-f2 43.6 wU-24-f2 57.1
22 wP-35-f2 8.3 wU-24-f2 40.6 wS-29-f1 43.6 w5-R-f1 57.5
23 wO-29-f1 8.5 wNp-5-f1 40.6 wU-34-f2 46.2 w34-V1-f2 57.9
24 wU-31-f1 10.9 w29-R-f1 41.9 w29-R-f1 49.9 wU-34-f2 58.9
25 wU-27-f1 10.9 w5-R-f1 41.9 w24-V1-f2 52.2 w28-W-f2 59.4
26 wN-30-f1 13.2 wN-30-f1 41.9 wN-30-f2 56.6 wNp-5-f1 61.1
27 w28-W-f2 14.0 w24-R-f1 41.9 wNp-5-f1 56.7 w29-R-f1 61.8
28 w34-R-f1 14.9 w24-T-f1 41.9 wU-31-f1 56.8 wN-30-f1 62.2
29 w34-T-f1 14.9 w34-V1-f2 46.3 wU-27-f1 56.8 wN-30-f2 68.9
30 wN-30-f2 15.1 w34-R-f1 47.1 w34-V1-f2 57.9 w25p-O1-f1 71.5
31 wU-31-f2 15.5 w34-T-f1 47.1 wN-30-f1 58.9
32 wU-27-f2 15.5 w24-V1-f2 49.8 w34-R-f1 59.1
33 wU-24-f1 15.5 wU-31-f1 55.8 w34-T-f1 59.1
34 wU-34-f1 15.5 wU-27-f1 55.8 w24-R-f1 60.7
35 w24-R-f1 15.8 wN-30-f2 56.2 w24-T-f1 60.7
36 w24-T-f1 15.8 wU-31-f2 56.7 w30-Np-f1 60.7
37 w30-Np-f1 16.0 wU-27-f2 56.7 wU-31-f2 60.9
38 w34-V1-f2 18.0 wU-24-f1 56.8 wU-27-f2 60.9
39 w24-V1-f2 20.7 wU-34-f1 56.8 w25p-O1-f1 61.1
40 wQ-34-f1 22.3 w30-Np-f1 59.1 w30-Np-f2 69.4
41 wNp-5-f1 22.5 w25p-O1-f1 60.7
42 wQ-24-f1 23.6 w30-Np-f2 69.4
43 w5-R-f1 23.6
44 w29-R-f1 24.9
45 w30-Np-f2 26.4
46 w25p-O1-f1 35.0
932

Table G.48: Sequence of exceedence of steel and concrete strain limits (Vina del Mar - El
Salto ground motion scaled to ASCE 7 MCE spectrum). Shading indicates that the wall
also exceeded the acceptance criterion for shear.

Strain
Yield Spall Crush Shear
Name Time Name Time Name Time Name Time
1 wO-5-f1 37.2 wP-35-f2 56.6 wV-24-f2 69.4 wP-29-f1 42.6
2 w5-R-f1 39.1 wO-29-f1 56.6 wV-34-f2 72.0 w28-W-f1 44.4
3 w29-R-f1 39.6 w5-O-f2 56.7 wO-29-f1 45.5
4 w28-W-f1 39.6 wV-24-f2 57.0 wO-5-f1 46.1
5 wP-29-f1 39.6 wV-34-f2 57.9 w28-W-f2 47.1
6 w5-O-f2 39.8 wS-5-f2 60.8 wO-29-f2 51.7
7 wNp-5-f1 39.8 wV-24-f1 60.8 wV-34-f2 53.4
8 wO-29-f2 39.9 wP-5-f1 61.3 wP-5-f1 53.7
9 wN-5-f1 40.1 w29-R-f1 61.3 wV-24-f1 55.5
10 wN-5-f2 40.1 w28-W-f1 61.8 wS-5-f2 55.8
11 wP-5-f1 40.2 wP-29-f1 65.2 wV-34-f1 56.1
12 w25p-O1-f1 40.2 wO-29-f2 68.8 w5-O-f2 60.7
13 w24-V1-f2 40.6 wP-23-f2 68.8 wN-5-f1 60.7
14 wO-29-f1 40.6 wS-29-f2 68.9 w29-R-f1 61.3
15 wV-24-f2 41.0 w24-V1-f2 69.4
16 w28-W-f2 41.0 w34-V1-f2 71.4
17 w34-V1-f2 41.9 wU-24-f2 71.5
18 wV-34-f1 41.9
19 wS-29-f1 43.6
20 wS-5-f1 43.6
21 wP-23-f2 43.6
22 wP-35-f2 43.8
23 wS-29-f2 43.8
24 wV-34-f2 46.2
25 wN-30-f1 46.6
26 wU-34-f2 47.1
27 wV-24-f1 49.7
28 wS-5-f2 52.3
29 wU-24-f2 52.7
30 wN-30-f2 56.2
31 w30-Np-f1 58.9
32 w34-R-f1 60.7
33 w34-T-f1 60.7
34 w24-R-f1 60.7
35 w24-T-f1 60.7
36 wQ-34-f2 68.9
37 w30-Np-f2 69.4

You might also like