Vous êtes sur la page 1sur 375

1

1
1

1
1
1
1
1

Displacement-Based Design of Reinforced


Concrete Structural Walls:
An Experimental lnvestigation of Walls with
Rectangular and T-Shaped Cross-Sections
EARTHQUAKE ENG. RES. CTR. UBRARY
Univ. of Calif. - 453 R.F.S.
1301 So. 46th St.
Richmond, CA 94804-4698 USA
(510) 2319403

by
John H. Thomsen iV
and
John W. Wallace

1
1
1

A Report on Research Sponsored by the


N ational Science Foundation
Grant No. NSF-BCS-9112962

1
1

1
1
1

Report No. CU/CEE-95-06


Department of Civil and Environmental Engineering
Clarkson U niversity

June 1995
Earthquake Eng Res Ctr Ubrary
Univ of Calif. Berkeley
1301 S. 46th St. - AFS 453

Richmond, CA 94804-4698 USA


(510) 665-3419

1
1

1
1
1
1

1
1
1
1
1

1
1
1

1
1
1

'

ABSTRACT

1
1

1
1

This report summarizes results of an experimental and analytical study of reinforced


concrete structural walls with symmetrical and unsymmetrical cross-sections, designed using a displacement-based design methodology. The primay objectives of this research were
to assess the ability of using a displacement-based approach for design of reinforced concrete
structural walls and to study the behavior of unsym.metrical "flanged" structural walls.
Four, approximately quarter-scale, structural wall specimens were constructed and

1
1
1

.....

tested under constant axial stress and reversed cyclic lateral loads. Two of the specimens
had rectangular cross-sections, whereas the other two had T-shaped cross-sections. Typical
material properties were used for all four specimens (grade 60 reinforcing steel and 4,000
psi concrete). The primary variables of the testing program were: ( 1) the shape of the wall
cross-sections, (2) the spacing and confi.guration of the transverse reinforcement, and (3)
the distribution of the vertical and horizontal web teinforcement.
Experimental results reveal that the rectangular walls exhibit stable, hysteretic behavior and possess excellent ductility and energy absorbing capabilities. Based on the results
of the rectangular walls, it is concluded that current code requirements (UBC-91, ACI
318-89) for the design of symmetrically shaped structural walls often result in overly conservative designs which should be relaxed to make the use of these walls more economical.
UBC-91 does not provide guidelines which specifi.cally address the design of unsymmetri-

cal "ftanged" structural walls which may exhibit brittle behavior, if not detailed properly.

displacement-based design procedure), these "flanged" walls can exhibit excellent behavior

1
1

accurate assessment of the effective flange width is essential.

Results of the T-shaped walls indicate, that when properly detailed (using the proposed

(ductile, hysteretic response) similar to that of the rectangular wall specimens; however, an

The analytical studies that were conducted focused on using basic principles and ex-

ii

isting monotonic modeling techniques for pred.icting the cyclic behavior of structural walls.

1
1

Analytical studies were conducted. to pred.ict: (1) flexural-strength, (2) shear-strength, (3)
stiffness, (4) ductility, and (5) buckling of longitudinal reinforcement. Excellent correlation
is observed between the analytically pred.icted. behavior and the experimentally measured.
results, indicating that monotonic modeling techniques provide an effective method for
evaluating cyclic behavior of structural walls.
Excellent correlation between analytically computed. strain distributions ( using the
proposed. displacement-based. design procedure) and experimentally measured strain distributions indicate that the recommended. displacement-based design procedure is a simple,
yet effective design technique that can be used. to design slender structural walls with any
combination of cross-section and reinforcing details. Use of this procedure results in struct ural wall designs that are directly related. to the building attributes as well as the wall
attributes, and often results in structural walls that require substantially less transverse
reinforcement than required by current codes.

1
1
1

1
1
1
1

..

iii

1
1
1
1
1
1

1
1
1

1
1
1

1
1
1
1
1
1
1
1

ACKNOWLEDGEMENTS

The work presented in this paper was funded by the National Science Foundation (NSF)
through Grant No. BCS-9112962. Opinions, conclusions and recommendations in this paper
are those of the authors, and do not necessarily represent those of the sponsor. The authors
would like to thank Sivaco New York and Florida Wire and Cable Co. for donating the
gauge wire used as transverse reinforcement, and the high~trength post-tensioning cables
used to apply the axial load. to the wall specimens, respectively. Special thanks are also due
to Potsdam Stone and Concrete ine. for their patience during concrete placement.
Special appreciation and sincere thanks are due to Dr. Gordon B. Batson, Dr. Stavroula
Pantazopoulou, Dr. John Moosbrugger and Dr. Dayakar Penumadu, for their valuable advice on this research.
The authors would like to thank graduate students, Paul Cote, Scott McConnell, Tony
Slaton-Barker, Weizhong Yan, and Chris Taylor who helped at various stages of the study,
and Cari Davey, the lab technician, who provided valuable advice and an invaluable service.
We also express extreme gratitude to undergrad.uate REU participants Mark Hugaboom,
Pam White, Matt Wells, Kari Jacot, Dan Larche, Ed Rihn, Kevin Hartnett, Derek Henessey,
and Jay Wiley who spent many hours helping construct and test the structural wall specimens.

1
1
1

-1

iv

TABLE OF CONTENTS
TITLE PAGE .................................................................. i
ABSTRACT ................................................................... ii
ACKNOWLEDGEMENTS .................................................... iv
TABLE OF CONTENTS ....................................................... V
LIST OF TABLES ........................................................... viii
LIST OF FIGURES ........................................................... ix
LIST OF SYMBOLS ......................................................... xvii
1. INTRODUCTION .............................................................. 1
1.1 General .................................................................. 1
1.2 Classification of Structural Walls ......................................... 3
1.3 Recent Trencls in Code Development ..................................... 5
1.4 Current Code Provisions ................................................. 6
1.5 Scope .................................................................... 7
1.6 Organization ............................................................ 8
2. LITERATURE REVIEW ....................................................... 9
2.1 Reinforced Concrete Structural Walls .................................... 9
2.1.~ Squat Shear Walls ............................................... 10
2.1.2 "Coupled" and "Pierced" Wall Systems .......................... 10
2.1.3 Building Frame-Shear Wall Structures ........................... 10
2.1.4 Slender Structural Walls ......................................... 12
2.1.5 Unsymmetrical, "Flanged" Structural Walls ...................... 16
2.2 Design Methodologies ................................................... 19
2.2. 1 Displacement-Ductility Based Design ............................ 19
2.2.2 Displacement-Based Design ...................................... 21
3. PROTOTYPE DESIGN ....................................................... 22
3. 1 Description of the Prototype Building ................................... 22
3.2 Simplifi.ed Design and Analysis of the Prototype Building ............... 22
3.3 Simplifi.ed Design and Analysis of the Prototype Walls .................. 24
3.3. 1 Rectangular Prototype Wall ...................................... 24
3.3.2 T-Shaped Prototype Wall (N-S Direction) ....................... 29
3.3.3 T-Shaped Prototype Wall (E-W Direction) ...................... 33
3.3.4 General Remarks ................................................ 35

1
1
1

1
1

1
1
1
1
:,
1
1
1
1
1
1
1
1

1
1
1

1
1
1
1
1

3.4 3-D Dynamic Analysis of the Prototype Building ....................... 35


4. DESCRIPTION OF MODEL WALLS ......................................... 41
4. 1 Specimen RW 1 ......................................................... 42
4.2 Specimen RW2 ......................................................... 46
4.3 Specimen TWl ......................................................... 47
4.4 Specimen TW2 ......................................................... 49
4.5 General Remarks ....................................................... 51
5. DESCRIPTION OF EXPERIMENTAL PROGRAM ........................... 53
5.1 Materials ............................................................... 53
5.2 Construction Procedure ................................................. 54
5.2. 1 Pedestal Construction ............................................ 55
5.2.2 Construction of Four Story Levels ................................ 56
5.3 Testing Apparatus ...................................................... 57
5.4 lnstrumentation and Data Acquisition .................................. 59
5.5 Testing Procedure ...................................................... 63
6. EXPERIMENTAL RESULTS ................................................. 66
6.1 Experimentally Observed Damage and Behavior ......................... 66
6.2 Lateral Load Versus Top Displacement Relations ........................ 68
6.3 Base Moment Versus Base Rotation Relations ........................... 71
6.4 Lateral Displacement Profiles ............................ : .............. 72
6.5 Shear Distortion Relations .............................................. 73

,,i

1
1
1:

6.6 Concrete Strain Profiles ................................................. 75


6. 7 Reinforcing Steel Strain Profiles ......................................... 78

6.8 Reinforcing Steel Strain Histories ....................................... 81


7. ANALYTICAL STUDIES AND COMPARISON OF RESULTS ................ 82
7. 1 Introduction ............................................................ 82
7.2 Moment-Curvature and Moment-Axial Load Relations .................. 82
7.3 Test Setup Requirements ............................................... 85
7.4 Expected Failure Modes ................................................ 86
7.4.1 Wall Flexural-Strength .......................................... 87

1
1

1
1

7.4.2 Wall Shear-Strength ............................................. 89


7.5 Displacements Due To Flexure and Shear ............................... 90
7.6 Monotonic Wall Behavior ............................................... 95
7.6. 1 Calculation of Monotonic Displacement Relations ................. 95
7.6.2 Comparisons of Monotonic Displacement Relations .............. 102

vi

1
7.7 Buckling of Longitudinal Reinforcement ................................ 105
7.8 Effectiveness of Transverse Reinforcement .............................. 109
8. VERIFICATION OF DISPLACEMENT-BASED DESIGN ................... 112
8.1 Introduction ........................................................... 112
8.2 Strain 'Distributions ................................................... 112
8.3 T-shaped Walls ........................................................ 115
8.4 Effective Flange Widths ............................................... 121
9. SUMMARY AND CONCLUSIONS ........................................... 128
9.1 Summary .............................................................. 128
9.2 Conclusions ........................................................... 129
9.2.1 General Behavior ............................................... 129
9.2.2 Displacement-Based Design of Structural Walls ................. 129
9.2.3 Structural Walls with Unsymmetrical Cross-Sections ............ 131
9.2.4 Analytical Modeling of Structural Wall Behavior ................ 133
9.3 Suggested

F\ture

Research ............................................ 134

REFERENCES .............................................................. 135


APPENDD A: DISPLACEMENT-BASED DESIGN ......................... 290
APPENDIX B: EXPERIMENTALLY OBSERVED BEHAVIOR .............. 314

1
1
1

1
1

1
1
1

'
1
1
1
1

vii

1
1
1
1

1
1
1
1
1
1
1
1
1
1
1
1

LIST OF TABLES
Table 3.1

Prototype Wall Design Parameters ................................... 142

Table 3.2

Shear-Strength Requirements For Prototype Walls ................... 143

Table 3.3

Prototype Wall Design Forces ........................................ 144

Table 3.4

Prototype Wall Design Displacements ................................ 145

Table 4.1

Design Parameters of Model Wall Specimens ......................... 146

Table 4.2

Shear-Strength Requirements For Model Walls ....................... 147

Table 5.1

Concrete Compressive and Tensile Strengths ......................... 148

Table 5.2

Concrete Mix Proportions ............................................ 149

Table 5.3

Reinforcement Properties ............................................ 150

Table 5.4

Chronology of Specimen Construction and Testing .................... 151

Table 6.1

Loads, Top Displacement and Shear Stress of Specimen RWl ......... 152

Table 6.2

Load.s, Top Displacement and Shear Stress of Specimen RW2 ......... 153

Table 6.3

Load.s, Top Displacement and Shear Stress of Specimen TW 1 ......... 154

Table 6.4

Load.s, Top Displacement and Shear Stress of Specimen TW2 ......... 155

Table 7.1

Monotonic Top Displacement Components of Rectangular Walls ...... 156

Table 7.2

Monotonic Top Displacement Components of T-shaped Walls ......... 157

Table 7.3

Allowable Spacing of Transverse Reinforcement ....................... 158

Table 7.4

Computed and Observed Drift Levels For Buckling ................... 159

Table 7.5

Confinement Effectiveness ............................................ 160

Table 8.1

Experimentally Measured Effective Flange Widths .................... 161

1
1
1
1
1
1
1

viii

1
1

LIST OF FIGURES
Fig. 3.1

Plan View of Prototype Building ..................................... 162

Fig. 3.2

Cross-Sectional Views of Rectangular Prototype Wall ................. 162

Fig. 3.3A

P-M Interaction Diagrams For Rectangular Prototype


Wall (Base level) .................................................... 163

Fig. 3.3B

P-M Interaction Diagrams For Rectangular Prototype


Wall (Fourth Story Level) ............................................ 163

Fig. 3.4

Calculated Strain Profile for Rectangular Prototype Wall ............. 164

Fig. 3.5

Moment-Curvature Relation for Rectangular Prototype Wall .......... 164

Fig. 3.6

Cross-Sectional View of T-Shaped Prototype Wall .................... 165

Fig. 3.7

Calculated Strain Profiles For T-Shaped Prototype Wall .............. 166

Fig. 3.8

P-M Interaction Diagrams For T-Shaped Prototype Wall (Base

Fig. 3.9

Moment Curvature Relations For T-Shaped Prototype Wall (Loads

level)

167

Parallel To Web) ..................................................... 167


Fig. 3.10

Moment Curvature Relations For T-Shaped Prototype Wall (Loads


Parallel To Flange) .................................................. 168

Fig. 3.11

Calculated Moment and Displacement Profiles for Rectangular


Prototype Wall ...................................................... 168

Fig. 4.1

3-Dimensional View of Rectangular Wall ............................. 169

Fig. 4.2

3-pimensional View of T-Shaped Wall ............................... 169

Fig. 4.3

RWl: Profile View Showing Reinforcement Locations ................. 170

Fig. 4.4

RWl: Cross-Sectional Views Showing Reinforcement Details .......... 171

Fig. 4.5

RWl: Photograph of First Story Reinforcement ...................... 17'2

Fig. 4.6

RWl: Photograph of Second Story Reinforcement .................... 17'2

Fig. 4.7

RWl: Calculated Strain Profile at 1.5% Lateral Drift ................. 173

Fig. 4.8

RW 1: Moment-Curvat ure Relations .................................. 173

Fig. 4.9A

RW2: Cross-Sectional View at Base of Wall .......................... 174

Fig. 4.9B

RW2: Photograph of First Story Reinforcement ...................... 174

Fig. 4.10

TWl: Cross-Sectional View at Base of Wall .......................... 175

Fig. 4.11

TW 1: Photograph of First Story Reinforcement ...................... 176

Fig. 4.12

TWl: Photograph of Web B.E. Reinforcement at Base of Wall ....... 176

Fig. 4.13

TWl: Moment-Curvature Relations .................................. 177

Fig. 4.14

TW2: Cross-Sectional View at Base of Wall .......................... 177

ix

1
1

1
1
1

1
1

1
1
1

1
1
1
1
1

Fig. 4.15

TW2: Photograph of First Story Reinforcement ...................... 178

Fig. 4.16

TW2: Photograph of Web B.E. Reinforcement at Base of Wall ....... 178

Fig. 4.17

TW2: Calculated Web Strain Profile at 1.5% Lateral Drift ............ 179

Fig. 4.18

TW2: Moment-Curvature Relations .................................. 179

Fig. 5.lA

RWl: Meuured Concrete Stress-Strain Relations .................... 180

Fig. 5.lB

RW2: Measured Concrete Stress-Strain Relations .................... 180

Fig. 5.lC

TWl: Measured Concrete Stress-Strain Relations .................... 181

Fig. 5.lD

TW2: Measured Concrete Stress-Strain Relations .................... 181

Fig. 5.2

Measured Steel Stress-Strain Relations ............................... 182

Fig. 5.3A

RWl: Pedestal Reinforcing Cage ..................................... 182

Fig. 5.3B

TW2: Pedestal Reinforcing Cage ..................................... 183

Fig. 5.4

TWl: Pedestal Formwork Prior to Casting Concrete ................. 183

Fig. 5.5

RWl: Reinforcing Steel in First Story ................................ 184

Fig. 5.6

RWl: Reinforcing Steel in Second Story .............................. 184

Fig. 5.7

RWl: 'Slip' Formwork Prior to Cuting Second Story ................. 185

Fig. 5.8

TWl: Reinforcing Steel in First Story ................................ 185

Fig. 5.9

TWl: Wall and Slab Formwork Prior to Casting First Story .......... 186

Fig. 5.10

TWl: Reinforcing Steel in Second Story .............................. 186

Fig. 5.11

TWl: Typical Slab Reinforcement Prior to Casting Second Story ..... 187

Fig. 5.12

TW 1: Reinforcing Steel at Top Story ................................. 187

Fig. 5.13

Schematic of Test Setup (For Rectangular Wall) ...................... 188

Fig. 5.14

Photograph of Test Setup (For Rectangular Wall) .................... 189

Fig. 5.15

RWl: Load Transfer Assembly (Side View) ........................... 189

Fig. 5.16

RWl: Load Transfer Assembly (End View) ........................... 190

Fig. 5.17

RWl: Hydraulic Actuator Used To Apply Cyclic Lateral Loads ....... 190

Fig. 5.18

RWl: Attachment of Lateral Support to Load Transfer Assembly ..... 191

Fig. 5.19

RWl: Tube--on-Tube Connection of Lateral Support to Reaction Wall . 191

Fig. 5.20

Steel Reference Frame U sed For Instrumentaion ...................... 192

Fig. 5.21

Instrumentation Used To Measure Pedestal Movement ................ 192

Fig. 5.22A

RWl: Wire Potentiometers Used To Measure Shear Deformations .... 193

Fig. 5.22B

TWl: Wire Potentiometers Used To Measure Shear Deformations .... 193

Fig. 5.23

RWl: Instrumentation Located on First Story Level .................. 194

Fig. 5.24

RW2: Instrumentation Located on First Story Level .................. 194

Fig. 5.25A

TW2: LVDT's Used To Measure Axial Deformations Along Web ..... 195

Fig. 5.25B

TW2: LVDT's Used To Measure Axial Deformations Along Flange ... 195

1
1

1
1
1
1

1
1

1
X

1
1

Fig. 5.26

RWl: Strain Gage Locations ......................................... 196

Fig. 5.27

RW2: Strain Gage Locations ......................................... 197

Fig. 5.28A

TWl: Strain Gage Locations in Web ................................. 198

Fig. 5.28B

TWl: Strain Gage Locations in Flange ............................... 198

Fig. 5.28C

TWl: Longitudinal Strain Gage Locations ........................... 199

Fig. 5.29A

TW2: Strain Gage Locations in Web ................................. 200

Fig. 5.29B

TW2: Strain Gage Locations in Flange ............................... 200

Fig. 5.29C

TW2: Longitudinal Strain Gage Locations ........................... 201

Fig. 5.30

RW2: Embedded Concrete Strain Gage in Boundary Element ......... 202

Fig. 5.31

TW2: Embedded Concrete Strain Gages in Web Boundary Element .. 202

Fig. 5.32

RWl: Lateral Drt Routine .......................................... 203

Fig. 5.33

RW2: Lateral Drift Routine .......................................... 203

Fig. 5.34

TW 1: Lateral Drift Routine .......................................... 204

Fig. 5.35

TW2: Lateral Drift Routine .......................................... 204

Fig. 6.1

RWl: Axial Load History ............................................ 205

Fig. 6.2

RW2: Axial Load History ............................................ 205

Fig. 6.3

TWl: Axial Load History ............................................ 206

Fig. 6.4

TW2: Axial Load History ............................................ 206

Fig. 6.5A

RWl: Lateral Load vs. Top Displacement (lnitial Cycles) ............ 207

Fig. 6.5B

RWl: Lateral Load vs. Top Displacement (All Cycles) ............... 207

Fig. 6.6A

RW2: Lateral Load vs. Top Displacement (lnitial Cycles) ............ 208

Fig. 6.6B

RW2: Lateral Load vs. Top Displacement (All Cycles) ............... 208

Fig. 6.7A

TW 1: Lateral Load vs. Top Displacement (lnitial Cycles) ............ 209

Fig. 6.7B

TWl: Lateral Load vs. Top Displacement (All Cycles) ............... 209

Fig. 6.8A

TW2: Lateral Load vs. Top Displacement (lnitial Cycles) ............ 210

Fig. 6.8B

TW2: Lateral Load vs. Top Displacement (All Cycles) ............... 210

Fig. 6.9

RWl: Measured Secant Stiffness ..................................... 211

Fig. 6.10

RW2: Measured Secant Stiffness ..................................... 211

Fig. 6.11

TWl: Measured Secant Stiffness ..................................... 212

Fig. 6.12

TW2: Measured Secant Stiffness ..................................... 212

Fig. 6.13

RWl: Base Moment vs. Base Rotation ............................... 213

Fig. 6.14

RW2: Base Moment vs. Base Rotation ............................... 213

Fig. 6.15

TWl: Base Moment vs. Base Rotation ............................... 214

Fig. 6.16

TW2: Base Moment vs. Base Rotation ............................... 214

Fig. 6.17

RWl: Lateral Displacement Profiles .................................. 215

1
1

1
1
1

1
1
1

1
1

1
1
1
1

1
xi

1
1

1
1

Fig. 6.18

RW2: Lateral Displacement Profiles .................................. 215

Fig. 6.19

TWl: Lateral Displacement Profiles .................................. 216

Fig. 6.20

TW2: Lateral Displacement Profiles .................................. 216

Fig. 6.21A

RWl: First Story Shear Distortions .................................. 217

Fig. 6.21B

RWl: Second Story Shear Distortions ................................ 217

Fig. 6.22A

RW2: First Story Shear Distortions .................................. 218

Fig. 6.22B

RW2: Second Story Shear Distortions ................................ 218

Fig. 6.23A

TWl: First Story Shear Distortions .................................. 219

Fig. 6.23B

TWl: Second Story Shear Distortions ................................ 219

Fig. 6.24A

TW2: First Story Shear Distortions .................................. 220

Fig. 6.24B

TW2: Second Story Shear Distortions ................................ 220

Fig. 6.25

Variables For Estimating Shear Distortions ........................... 221

Fig. 6.26A

RWl: Web Concrete Strain Profiles (Pos. Displacements) ............. 222

Fig. 6.26B

RWl: Web Concrete Strain Profiles (Neg. Displacements) ............ 222

Fig. 6.27A

RW2: Web Concrete Strain Profiles (Pos. Disp.) ..................... 223

Fig. 6.27B

RW2: Web Concrete Strain Profiles (Neg. Disp.) ..................... 223

Fig. 6.28A
Fig. 6.28B

TWl: Web Concrete Strain Profiles (Pos. Disp.) ..................... 224


TWl: Web Concrete Strain Profiles (Neg. Disp.) ..................... 224

Fig. 6.28C

TWl: Flange Concrete Strain Profiles (Pos. Disp.) ................... 225

Fig. 6.28D

TWl: Flange Concrete Strain Profiles (Neg. Disp.) .................. 225

Fig. 6.29A

TW2: Web Concrete Strain Profiles (Pos. Disp.) ............... ...... 226

Fig. 6.29B

TW2: Web Concrete Strain Profiles (Neg. Disp.) ..................... 226

Fig. 6.29C

TW2: Flange Concrete Strain Profiles (Pos. Disp.) ................... 227

Fig. 6.29D

TW2: Flange Concrete Strain Profiles (Neg. Disp.) .................. 227

Fig. 6.30

RW2: Concrete Strain Gage Histories ................................ 228

1
1

Fig. 6.31

TW2: Web B.E. Concrete Strain Gage Histories ...................... 228

Fig. 6.32A

RWl: Web Steel Strain Profiles

Base (Pos. Disp.) ................. 229

Fig. 6.32B

RWl: Web Steel Strain Profiles

Base (Neg. Disp.) ................. 229

Fig. 6.32C

RWl: Web Steel Strain Profiles@ First Story Level (Pos. Disp.) ..... 230

Fig. 6.32D

RWl: Web Steel Strain Profiles@ First Story Level (Neg. Disp.) ..... 230

Fig. 6.33A

RW2: Web Steel Strain Profiles @ Base (Pos. Disp.) ................. 231

Fig. 6.33B

RW2: Web Steel Strain Profiles @ Base (Neg. Disp.) ................. 231

Fig. 6.33C

RW2: Web Steel Strain Profiles

@ First

Story Level (Pos. Disp.) ..... 232

Fig. 6.33D

RW2: Web Steel Strain Profiles

@ First

Story Level (Neg. Disp.) ..... 232

Fig. 6.34A

TWl: Web Steel Strain Profiles

1
1
1
1
1
1

1
1

1
1

xii

Base (Pos. Disp.) ................. 233

1
Fig. 6.34B

TWl: Web Steel Strain Profiles@ Base (Neg. Disp.) ................. 233

Fig. 6.34C

TW 1: Web Steel Strain Profiles @ First Story Level (Pos. Disp.) ..... 234

Fig. 6.34D

TWl: Web Steel Strain Profiles@ First Story Level (Neg. Disp.) ..... 234

Fig. 6.34E

TW 1: Flange Steel Strain Profiles @ Base (Pos. Disp.) ............... 235

Fig. 6.34F

TWl: Flange Steel Strain Profiles@ Base (Neg. Disp.) ............... 235

Fig. 6.34G

TWl: Flange Steel Strain Profiles

Fig. 6.34H

TWl: Flange Steel Strain Profiles @ First Story Level (Neg. Disp.) .. 236

Fig. 6.35A

TW2: Web Steel Strain Profiles @ Base (Pos. Disp.) ................. 237

Fig. 6.35B

TW2: Web Steel Strain Profiles @ Base (Neg. Disp.) ................. 237

Fig. 6.35C

TW2: Web Steel Strain Profiles

First Story Level (Pos. Disp.) ..... 238

Fig. 6.35D

TW2: Web Steel Strain Profiles

First Story Level (Neg. Disp.) ..... 238

Fig. 6.35E

TW2: Flange Steel Strain Profiles @ Base (Pos. Disp.) ............... 239

Fig. 6.35F

TW2: Flange Steel Strain Profiles @ Base (Neg. Disp.) ............... 239

Fig. 6.35G

TW2: Flange Steel Strain Profiles

Fig. 6.35H

TW2: Flange Steel Strain Profiles @ First Story Level (Neg. Disp.) .. 240

Fig. 6.36

RWl: Reinforcing Steel Strain Gage Histories ........................ 241

Fig. 6.37

RW2: Reinforcing Steel Strain Gage Histories ........................ 246

Fig. 6.38

TWl: Reinforcing Steel Strain Gage Histories ........................ 251

Fig. 6.39
Fig. 7.1

TW2: Reinforcing Steel Strain Gage Histories ........................ 258

Fig. 7.2

RW2: Moment-Curvature Relations .................................. 265

Fig. 7.3

TWl: Moment-Curvature Relations .................................. 266

Fig. 7.4

TW2: Moment-Curvature Relations .................................. 266

Fig. 7.5

RWl: P-M Interaction Diagrams ..................................... 267

Fig. 7.6

RW2: P-M Interaction Diagrams ..................................... 267

Fig. 7.7

TWl: P-M Interaction Diagrams ..................................... 268

Fig. 7.8

TW2: P-M Interaction Diagrams ..................................... 268

Fig. 7.9A

RWl: Analytical Concrete Stress-Strain Relations .................... 269

Fig. 7.9B
Fig. 7.10A

RW2: Analytical Concrete Stress-Strain Relations .................... 269


Experimental, Nominal and Analytical Moment Capacities ........... 270

Fig. 7.10B

Experimental and Nominal Shear Capacities .......................... 270

Fig. 7.11

RWl: Base Rotation Contrihution To Measured Top Displacement ... 271

Fig. 7.12

RW2: Base Rotation Contrihution To Measured Top Displacement ... 271

Fig. 7.13

TWl: Base Rotation Contribution To Measured Top Displacement ... 272

Fig. 7.14

TW2: Base Rotation Contrihution To Measured Top Displacement ... 272

First Story Level (Pos. Disp.) ... 236

First Story Level (Pos. Disp.) ... 240

RWl: Moment-Curvature Relations .................................. 265

xiii

1
1

1
1
1

1
1

1
1
1
1
1
1

1
1
1
1
1
1
1
1
1
1
1

Fig. 7.15

Shear, Moment, Curvature and Rotation Distributions For


Cantilever Wall Specimens ........................................... 273

Fig. 7.16

RWl: Flexural Contribution To Measured First Story Displacement .. 274

Fig. 7.17

RW2: Flexural Contribution To Measured First Story Displacement .. 274

Fig. 7.18

TWl: Flexural Contribution To Measured First Story Displacement .. 275

Fig. 7.19

TW2: Flexural Contribution To Measured First Story Displacement .. 275

Fig. 7.20

Parameters For Calculating Slippage of Longitudinal


Reinforcement From The Base Pedestal .............................. 276

Fig. 7.21

Parameters For Calculating Top Displacement Caused


by Slippage of Longitudinal Reinforcement ........................... 276

Fig. 7.22

RWl: Monotonic Lateral Load-Top Displacement Envelopes .......... 2n

Fig. 7.23

RW2: Monotonic Lateral Load-Top Displacement Envelopes .......... 2n

Fig. 7.24

TW 1: Monotonic Lateral Load-Top Displacement Envelopes .......... 278

Fig. 7.25

TW2: Monotonic Lateral Load-Top Displacement Envelopes .......... 278

Fig. 7.26

RWl: Top Displacement Due To Pedestal Rotation ................... 279

Fig. 7.27

RW2: Top Displacement Due To Pedestal Rotation ................... 279

Fig. 7.28

TWl: Top Displacement Due To Pedestal Rotation .................. 280

Fig. 7.29

TW2: Top Displacement Due To Pedestal Rotation .................. 280

Fig. 7.30

Evaluation of Confinement Efiectiveness .............................. 281

Fig. 8.lA

RWl: Anal. vs. Exp. Steel Strain Distributions (Pos. Disp.) ......... 282

Fig. 8.lB

RWl: Anal. vs. Exp. Steel Strain Distributions (Neg. Disp.) ......... 282

Fig. 8.2A

RWl: Anal. vs. Exp. Concrete Strain Distributions (Pos. Disp.) ..... 283

Fig. 8.2B

RWl: Anal. vs. Exp. Concrete Strain Distributions (Neg. Disp.) ..... 283

1
1
1

Fig. 8.3A

RW2: Anal. vs. Exp. Steel Strain Distributions (Pos. Disp.) ......... 284

Fig. 8.3B

RW2: Anal. vs. Exp. Steel Strain Distributions (Neg. Disp.) ......... 284

Fig. 8.4A

RW2: Anal. vs. Exp. Concrete Strain Distributions (Pos. Disp.) ..... 285

Fig. 8.4B

RW2: Anal. vs. Exp. Concrete Strain Distributions (Neg. Disp.) ..... 285

Fig. 8.5A

TW 1: Anal. vs. Exp. Steel Strain Distributions (Pos. Disp.) ......... 286

Fig. 8.5B

TWl: Anal. vs. Exp. Steel Strain Distributions (Neg. Disp.) ......... 286

Fig. 8.6A

TWl: Anal. vs. Exp. Concrete Strain Distributions (Pos. Disp.) ..... 287

Fig. 8.6B

TWl: Anal. vs. Exp. Concrete Strain Distributions (Neg. Disp.) ..... 287

Fig. 8.7A

TW2: Anal. vs. Exp. Stee1 Strain Distributions (Pos. Disp.) ......... 288

1
1

Fig. 8.7B

TW2: Anal. vs. Exp. Steel Strain Distributions (Neg. Disp.) ......... 288

Fig. 8.8A

TW2: Anal. vs. Exp. Concrete Strain Distributions (Pos. Disp.) ..... 289

Fig. 8.8B

TW2: Anal. vs. Exp. Concrete Strain Distributions (Neg. Disp.) ..... 289

1
1

xiv

Fig. A.1

Requirements For Wall Flexural-Strength vs. Wall Height ............ 308

Fig. A.2

Elastic and lnelastic Displacement Spectra ........................... 308

Fig. A.3

Estimate of Roof Drift Ratio ......................................... 309

Fig. A.4

Relationship Between Global and Local Deformations ................ 309

Fig. A.5a

Equilibrium Requirements for Rectangular Wall Cross-Section ........ 310

Fig. A.5b

Equilibrium Requirements for Barbell-Shaped Wall Cross-Section .... 310

Fig. A.6a

Extreme Fiber Strain for Constant Wall Aspect Ratio ................ 311

Fig. A.6b

Extreme Fiber Strain for Constant Wall Axial Load .................. 311

Fig. A.7

Extreme Fiber Strain for Unsymmetrically Reinforced Walls .......... 312

Fig. A.8

Required Length of Concrete Confinement for Constant Wall


Aspect Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . . . . . . . . . . . 312

Fig. A.9

Required Length of Concrete Confinement for Unsymmetrically


Reinforcecl Walls ..................................................... 313

Fig. B.1

RWl: After 2 Cycles at 0.25% Lateral Drift .......................... 330

Fig. B.2

RWl: After 2 Cycles at 0.50% Lateral Drift .......................... 330

Fig. B.3

RWl: After 2 Cycles at 0. 75% Lateral Drift .......................... 331

Fig. B.4

RWl: After 2 Cycles at 1.0% Lateral Drift ........................... 331

Fig. B.5

RWl: North B.E. After Two Cycles at 1.0% Lateral Drift ............ 332

Fig. B.6

RWl: North B.E. After Two Cycles at 1.5% Lateral Drift ............ 332

Fig. B.7

RWl: South B.E. After Two Cycles at 2.0% Lateral Drift ............ 333

Fig. B.8

RWl: After 2 Cycles at 2.0% Lateral Drift ........................... 333

Fig. B.9

RWl: After Half Cycle at 3.0% Drift (End of Test) ................... 334

Fig. B.10

RWl: North B.E. After Half Cycle at 3.0% Drift (End of Test) ....... 334

Fig. B.11

RW2: After 2 Cycles to 0.25% Lateral Drift .......................... 335

Fig. B.12

RW2: After 2 Cycles to 0.50% Lateral Drift .......................... 335

Fig. B.13

RW2: After 2 Cycles to 0.75% Lateral Drift .......................... 336

Fig. B.14

RW2: South B.E. After 2 Cycles at 1.0% Lateral Drift ................ 336

Fig. B.15

RW2: South B.E. After 2 Cycles at 1.5% Lateral Drift ................ 337

Fig. B.16

RW2: North B.E. After 4 Cycles at 1.5% Lateral Drift ................ 337

Fig. B.17

RW2: South B.E. After 4 Cycles at 1.5% Lateral Drift ................ 338

Fig. B.18

RW2: North B.E. After 2 Cycles at 2.0% Lateral Drift ................ 338

Fig. B.19

RW2: South B.E. After 2 Cycles at 2.0% Lateral Drift ................ 339

Fig. B.20

RW2: Buckling of Longiudinal Bars in North B.E. During


Second Cycle at 2.5% Lateral Drift ................................... 339

Fig. B.21

RW2: Damage at Base of Wall After 2 Cycles at 2.5% Drift .......... 340

1
1

1
1

1
1
1

1
1
1

1
1
1

XV

1
1

1
1
1

Fig. B.22

RW2: Overall Photograph Upon Completion of Testing ............... 340

Fig. B.23

TWl: Web After 2 Cycles to 0.50% Lateral Drift ..................... 341

Fig. B.24

TWl: Flange After 2 Cycles to 0.50% Lateral Drift ................... 341

Fig. B.25

TWl: Web After 2 Cycles to 0.75% Lateral Drift ..................... 342

1
1

Fig. B.26

TWl: Flange After 2 Cycles to 0.75% Lateral Drift ................... 342

Fig. B.27

TW 1: Web After 2 Cycles to 1.0% Lateral Drift ...................... 343

Fig. B.28

TWl: Flange After 2 Cycles to 1.0% Lateral Drift .................... 343

Fig. B.29

TWl: Web B.E. After 2 Cycles at 1.0% Lateral Drift ................. 344

Fig. B.30

TWl: Web B.E. After First Cycle at 1.5% Lateral Drift .............. 344

Fig. B.31

TWl: Web B.E. After First Cycle at 1.5% Lateral Drift .............. 345

Fig. B.32

TWl: Web B.E. After Half Cycle at 2.0% Lateral Drift ............... 345

1
1

Fig. B.33

TW2: After 2 Cycles to 0.25% Lateral Drift .......................... 346

Fig. B.34

TW2: After 2 Cycles to O. 50% Later al Drift .......................... 346

Fig. B.35

TW2: Web After 2 Cycles to O. 75% Lateral Drift ..................... 347

Fig. B.36

TW2: Flange After 2 Cycles to O. 75% Lateral Drift ................... 347

Fig. B.37

TW2: After 2 Cycles to 1.0% Lateral Drift ........................... 348

Fig. B.38

TW2: Web B.E. After 2 Cycles at 1.0% Lateral Drift ................. 348

Fig. B.39

TW2: Web B.E. After First Cycle at 1.5% Lateral Drift .............. 349

1
1

Fig. B.40

TW2: Web B.E. After Second Cycle at 1.5% Lateral Drift ............ 349

Fig. B.41

TW2: Web B.E. After Third Cycle at 1.5% Lateral Drift ............. 350

Fig. B.42

TW2: Web B.E. After Fourth Cycle at 1.5% Lateral Drift ............ 350

Fig. B.43

TW2: Web B.E. After First Cycle at 2.0% Lateral Drift .............. 351

Fig. B.44

TW2: Web B.E. After Third Cycle at 2.0% Lateral Drift ............. 351

Fig. B.45

TW2: Web B.E. After Third Cycle at 2.5% Lateral Drift ............. 352

Fig. B.46

TW2: Web B.E. During First Cycle at 3.0% Lateral Drift ............. 352

Fig. B.47

TW2: Out-of-Plane Failure of Web B.E. During First Cycle


at 3. 0% Lateral Drift ................................................ 353

1
1

1
1

1
1
1

xvi

1
LIST OF SYMBOLS
a

thickness of bounday element perpendicular to the wall web

Ab

area of reinforcing bar

Acv

cross-sectional area of concrete resisting shear

Ach

cross-sectional area measured. out-to-out of transverse reinforcement

floor plan area of building

Ag

gross cross-sectional area

A.,

area of longitudinal tension steel

A~

area of longitudinal compression steel

A;

area of vertical web steel

A.sh

required. area of transverse reinforcement according to ACI 318-89

A:h

required. area of transverse reinforcement

~tr

effective area of transverse steel

Av

area of shear reinforcement

Aw

lw

length of boundary element parallel to the wall web

be//

effective flange width

length of wall cross-section in compression

c'

length of wall cross-section requiring confinement

effective depth of the wall cross-section

tw

nominal diameter of reinforcing bar


undeformed dimensions of wire pots used to measure shear distortions
deformed. dimensions of wire pots used. to measure shear distortions
center to center distance of transverse hoop legs
e

elongation of longitudinal reinforcement


modulus of elasticity for concrete
modulus of elasticity for steel
tangent modulus of steel at onset of strain hardening

1
1
1
1

1
1

1
1
1

tangent modulus of steel


stiffness

,;

concrete compressive strength

fer

critical stress

steel failure stress

xvii

1
1
1

1
1

1
1
1
1

t
1
1
1

1
1
1

fr

confinement pressure acting on concrete

fa

steel stress

boundary steel compression stress

fu
f 11

steel ultimate stress


yield stress of steel reinforcement

compression steel yield stress

web steel yield stress;

transverse steel yield stress;

distributed web steel yield stress

!rt

yield stress of transverse reinforcement

acceleration due to gravity

shear modulus

vertical height of "X" configuration used to measure shear


distortions (Section 6.5)
cross-sectional dimension of the confined core measured
center-to-center of the transverse reinforcement
total wall height
wall aspect ratio
story height

height of loading

lmportance factor according to UBC-91 (Section 3.2)

moment of inertia

jd

distance between tension reinforcement and resultant compression force

constant depending on end conditions


unsupported length of longitudinal reinforcement (Section 7. 7)
horizontal length of "X" configuration used to measure shear
distortions (Section 6. 5)

1
1

development length

yield development length

development length neecled in excess of yield


development length of bar with a standard 90 degree hook
total development length
plastic hinge length
wall length

1
1
1

xviii

1
1

moment

Mba,ae

moment at the wall base

Mcocte

moment at wall base for code prescribecl forces

Mcr

cracking moment

Mo

moment at wall base accounting for overstrength factors

yield moment

number of stories

ratio of wall area to floor plan area

Pv

area ratio of lateral reinforcement

axial load acting on wall

Pat

lateral load acting at the top of the wall

radius of gyration
system ductility factor according to UBC-91

spacing of transverse reinforcement


maximum spacing of transverse reinforcement to su ppress buckling

site soil coefficient according to ATC-03-06 or UBC-91


spectral acceleration

1
1

1
1
1

spectral displacement
wall web thickness

period of structure, seconds

U.b

average uniform bond stress

Vcode

wall shear force for code prescribecl forces

Ve:q,ected

expectecl wall shear force

Vn

nominal shear-strength

Vu

factorecl shear force

unit floor weight

weight of building

distance from critical section to point of contraflexure

seismic zone factor according to UBC-91


factor to account for material overstrength and strain hardening
of tension and distributecl reinforcement (Appendix A)
factor to account for rotation distribution over the height of the
structural wall (Section 7.5)
factor as defineci by ACI 318-89, Section 10.2. 7.3
factor to account for material overstrength and strain hardening

xix

1
1
1
1
1
1

1
1
1

1
1
1
1
1

iavg
6e

6/lex
6ah.ear
6,,

6ped,rot
6alip
6top
6top,Jlez
6.

61,1

61

61,flex
61,ah.ear
fc
fc,ma.::

EJ

1
1

1
1
1

1
1

Eah.
f.

f,

6
<P
<Pcr
<P.

<P

p
p'
p''
Ph.
Pmin
Pn
Pv
8ba.ae

e,,
8ped,rot
8alip
W,

1
1

of compression reinforcement
average shear distortion
top displacement caused by curvature over the wall height
flexural deformations
shear deformations
top displacement caused by rotation of the wall at the plastic hinge
top displacement due to pedestal rotation
top displacement caused by slippage of longitudinal reinforcement
total top displacement
top displacement due to flexural efrects
lateral roof displacement at ultimate
lateral roof displacement at yield
lateral displacement at first floor level
lateral displacement at first floor level due to flexure
lateral displacement at first floor level due to shear distortions
concrete strain
extreme fiber concrete compression strain
steel failure strain
steel strain at the onset of strain hardening
steel ultimate strain
steel yield strain
Poisson's ratio
displacement ductility ratio
curvature
cracking curvature
ultimate curvature
yield curvature
tension steel reinforcing ratio
compression steel reinforcing ratio
vertical web steel reinforcing ratio
horizontal web steel reinforcing ratio
minimum reinforcing ratio
ratio of distributed shear reinforcement per ACI 318-89, Section 21.0
vertical web steel reinforcing ratio
rotation at base of wall
plastic hinge rotation
rotation at the pedestal-floor interface
rotation caused by slippage of longitudinal reinforcement
multiplier to account for higher mode efrects on wall shear

XX

1
1
1
1

1
1

1
1
1

1
1
1
1
1
1
1

1
1
1
1

1
1
1
1
1
1

1
1
1
1
1
1
1
1
1

Chapter - 1
INTRODUCTION

1.1 General
The use of reinforced concrete shear (structural) walls to resist lateral loads (such as
those imposed by earthquakes or wind) is common in all areas of the United States. Meigs
et al. (1993) surveyed structural engineering firms in the U.S., Canada, Chile, Colombia,
Mexico, and Peru on their use of various lateral force resisting systems (LFRS ). The survey
requested the firms to report the number of buildings designed within the past five years
according to UBC 1991 definitions for LFRS and to list the factors affecting the selection of
the LFRS. The results of the survey were categorized according to the number of stories in
the buildings, the seismic zone in which they were constructed, and the type of LFRS. LFRS
were designated as: bearing/ shear walls, building frame/ shear walls, ordinary, intermediate
or special moment resisting space frames, braced frames, and dual systems. A total of 5,045
buildings were reported of which 93% were cast-in-place and 7% were precast.
The U.S. firms reported 1,135 cast-in-place buildings of which 46% were located in low
seismic regions (zones O and 1), 16% were located in moderate seismic regions (zones 2A
and 2B), and 38% were located in high seismic regions (zones 3 and 4). The firms surveyed
were divided into 5 separate geographical regions: Northwest, Southern California, Northern California, Midwest and East. The buildings in zones O and 1 were primarily designed
in the midwest and the east and a majority of the buildings incorporated shear walls. Shear
walls were used in 73% of the buildings between 1-3 stories, 56% between 4-6 stories, 41%
between 7-15 stories and 70% of the buildings over 15 stories tali. These percentages were
nearly equally divided between bearing/shear wall systems (walls designed to resist both
gravity and lateral loads) and building frame/shear wall systems (walls resist lateral loads

1
and building frame resists gravity loads). The buildings designed for moderate seismic regions (zones 2A and 2B) were alsa dominated by shear walls (6 7%) regardless of building
height. As in zones O and 1, the shear wall responses were equally distributed between
bearing/shear wall systems (31%) and building frame/shear wali systems (36%). As expected, the buildings in seismic zones 3 and 4 were designed primarily in the Northwest and
Northern and Southern Califomia. Buildings between 1-3 stories were dominated by shear
walls (85%) and once again the two types of shear wall systems were equally represented.
For buildings between 4-15 stories, shear walls still dominated the response (63%); however,
the bearing/shear wall system was rarely used. For buildings more than 15 stories tali, the
use of shear walls diminished (18%) and response was dominated by SMRSF (65%) and
dual systems (17%).
Of the 27 U.S. firms surveyed, 82% stated architectural/functional constraints as a
major factor for choosing a LFRS, 79% stated economics (including scheduling and detailing
considerations, construction costs and foundation requirements) asa major factor, and 36%
stated building performance (minimizing lateral drift levels and providing redundancy) as
a contributing factor.
The survey indicated that American, Canadian, Chilean and Peruvian structural engineers ali favored wall systems; whereas, Colombian and Mexican engineers were more likely
to design buildings without shear walls. The major factors cited for inclusion of shear walis
included: excellent performance in past eqrthquakes, ability to minimize lateral drifts, and
"simplicity of design". Poor soil conditions, tall building heights, and the absence of seismic
risk were cited as the major factors detering the use of shear walls.
Shear walis possess very large in-plane sti:.ffness and are often used to limit the amount
of lateral drift a building will sustain under lateral loadings. Reinforced concrete walis
are typicaliy designed such that during wind loadings and less intense seismic loadings,
the shear walls behave in an elastic manner thus preventing non-structural damage to the
building. However, during less frequent, more severe earthquakes, the walls are expected to

1
1
1
1
1
1
1
1
1
1
1
1
1
1

1
1
1

1
1
1

1
1
1
1
1
1
1
1
1

1
1
1
1
1
1
1

experience some amount of inelastic deformation; therefore, shear walls must be designed
to sustain inelastic deformations while maintaining their ability to carry load and dissipate
energy. During these severe earthquakes, structural and non-structural damage is expected;
however, prevention of collapse and loss of life is the main concern.
Post earthquake evaluations have indicated that reinforced concrete shear walls are
very effective at limiting damage. Observed damage is typically dependent on the building
and wall con:6.guration. These topics are discussed in the following section.

1.2 Classiflcation of Structural Walls


Structural walls are constructed in many shapes and sizes and are incorporated into
building systems in various ways, resulting in many different con.figurations and classi:6.cations. It is noted that the terms "structural wall" and "shear wall" are used interchangeably
throughout this dissertation.
Perhaps the most commonly used classi:6.cation for shear walls is their overall height
to length ratio (aspect ratio, hw/lw), Walls that have aspect ratios of one or less are
commonly referred to as short or squat walls; whereas, walls with aspect ratios of three or
greater are typically termed tall or slender. Shear walls that have aspect ratios between
one and three are commonly referred to as intermediate walls. This general classi:6.cation
alsa provides a basic distinction of expected behavioral types and failure modes. Because
of their small height to length ratio, squat walls typically behave similarly to deep beams
and their behavior is typically dominated by shear. On the other hand, behavior of slender
walls is dominated by flexure, and the effects of shear are often neglected. Intermediate
walls typically exhibit the combined effects of both shear and flexure.
Structural wall systems are often categorized by what other types of structural elements are present in the building and how neighboring walls effect one another. The most
commonly used shear wall systems in the U.S. are: {1) the bearing/shear wall system, and

(2) the building frame/shear wall system which were described above, and (3) the perimeter frame/ shear wall system with intemal gravity columns. Dual systems that incorporate
Moment Resisting Space Frames (MRSF) as well as shear walls for lateral load resistance
are also used, but less frequently.

1
1

1
1

"Coupled" wall systems are common and result when neighboring walls within a building are interconnected, either by slabs or beams. These "coupled" walls can act essentially
independently if weak coupling exists or they can act integrally if well detailed coupling
beams are provided. "Pierced" walls are similar to "coupled" walls in that they are not
solid; however, the holes in "pierced" walls are typically necessary for doors, windows, ductwork, ete. and constitute a small portion of the wall's cross-section. Differentiation between
"coupled" and "pierced" wall systems is often based on the area of the openings and the
coupling involved. If the area of the openings is sufficiently small (a conservative value of
approximately 10% or less of the cross-sectional area has been recommended, UBC-1994)
than the behavior of the "pierced" wall system is considered as a single isolated wall and
the e:ffect of the openings is neglected. "Coupled" walls are analyzed as two separate walls
with coupling elements.
The last type of classification to be discussed is based on the cross-sectional shape
(rectangular, barbell, T, C, or L-shaped, ete.) ofthe structural wall. The use of symmetrical
cross-sections (rectangular, barbell, ete.) is very common and the performance of these walls
has been studied extensively (Abrams, 1991, lists 44 references on the measured behavior of
structural walls). Due to functionality and aesthetic reasons, it is common practice to join
two or more symmetrical walls together that are arranged in orthogonal directions, resulting
in unsymmetrical (C, L, and T-shaped) "flanged" walls. The behavior of "flanged" walls is
significantly different than that of symmetrical walls. The shape of the cross-section has a
significant impact on wall behavior including: strength, sti:ffness and ductility of the wall
(Paulay, 1986; Wallace and Moehle, 1989). The influence of the flange reinforcement in
tension is of particular interest; however, there is a lack of research addressing this issue.

1
1
1
1
1
1
1
1
1

1
1
1
1
1

1
1

1
1
1
1
1
1
1
1

1
1
1
1
1
1
1
1
1

Current design practice and recent trends in code development are discussed in the following
section.

1.3 Recent Trends in Code Development


The performance of buildings in major earthquakes often leads to changes in building
codes. The March 3, 1985 earthquake in Chile is of particular interest for reinforced concrete shear walls. The earthquake of surface magnitude 7.8 occurred off the coast of central
Chile (EERI, 1986), in close proximity to the city of Va del Mar, where approximately
400 modem, reinforced concrete buildings stood {Riddell et al., 1987). A major portion of
these buildings were designed for forces similar to those used for design in areas of high
seismicity in the United States. Post-earthquake investigations (EERI, 1986; Wood et al.,
1987; Wallace and Moehle, 1989) revealed that the majority of these buildings incorporated
a considerable amount of reinforced concrete shear walls {ratios of total wall area to floor
plan area of 5-6% were common) for lateral load resistance and that these structures performed extremely well in this seismic event. in addition, it was also revealed that although
the seismic code requirements in Chile are similar to those used in the U.S., detailing requirements are less stringent and inspection is lax by U.S. standards (Wallace and Moehle,
1989; Wood, 1991).
The remarkably good performance of the structural wall dominant buildings in Va del
Mar during the 1985 earthquake inspired an increase in the amount of research on reinforced
concrete structural walls in the United States. Although many factors contributed to the
observed building performance, analytical studies indicated that light damage could be
traced to the stiffness of the structural systems, which limited the deformations imposed on
the buildings (Wood et aL, 1987; Wallace and Moehle, 1989; 1992; 1993).
Analytical studies were performed to evaluate Chilean buildings and to extrapolate
the findings to U.S. practice (Wallace and Moehle, 1992). These studies were focused on
establishing likely response characterisitics (available wall deformation capacity) of shear

wall buildings to estimate design requirements for structural walls. The analytical studies
were compared with experimental studies performed on structural walls of rectangular or
nearly rectangular shape (Carvajal and Pollner, 1983; Oesterle et aL, 1976; Shiu et al.,
1981; Oesterle, 1986; and Ali and Wight, 1991). The results indicated that the analytical
procedure used to estimate the drift capacities tends to yield conservative estimates of wall
deformation capacity, and that this procedure provides a simple, effective method for determining the need for confining steel at structural wall boundaries. Subsequent analytical
studies of reinforced concrete shear walls has lead to the development of a displacementbased design procedure (Wallace, 1994a; 1994b). The procedure uses lateral drift, as opposed to strength, and directly relates computed building response and wall attributes to
the need to provide transverse reinforcement at the wall boundaries.

1.4 Current Code Provisions

1
1
1
1

1
1

1
1

Prior to 1994, all code provisions regarding the design of reinforced concrete structural
walls were strength-based (ie. ACI 318-89, UBC-1991 ). These codes were developed to provide adequate deformability to prevent brittle behavior. Adequate deformability is achieved
through the use of heavily confined boundary elements (typically wider than the wall web,
resulting in a barbell shaped wall) whenever the extreme fiber stress due to combined axial
and lateral loads exceeds 0.2J;. The confinement within the boundary elements must extend
over the height of the wall until the extreme fiber stress falls below O. 15 J;. These boundary
elements must be capable of resisting gravity loads and overtuming moments without the
help of the wall web. These stringent detailing requirements often make it quite costly to
use a predominantly structural wall system especially in areas of high seismic risk. Studies by various researchers (Ali and Wight, 1991; Wood, 1991; Wallace and Moehle, 1992)
have shown that these code requirements are overly conservative for a majority of building
systems that utilize reinforced concrete structural walls for lateral load resistance.

in addition to the inability of these codes to differentiate between walls that require

1
1

1
1
1
1
1

1
1

1
1

specially detailed boundary elements and those that don't; the 1991 UBC does not provide

guidelines for the design of unsymmetrical "ftanged" walls. Therefore, recommendations

1
1
1
1
1
1

are designed independently. Due to the conservatism in UBC-91, the unsymmetrical walls

1
1
1
1
1
1
1
1
1
1

are needed to avoid potentially poor design practice such as the joining of two walls that

that result from this practice may yield adequate behavior; however, the provisions do not
provide a rational approach for the design of "ftanged" walls. A more effective approach
for structural wall design is the use ofa displacement-based design approach such as that
proposed by Wallace (1994a; 1995). This type of design procedure provides a direct link
between expected building response and the need for special detailing of transverse reinforcement, and eliminates the need to provide a single system ductility factor fora given
building configuration (as was the case for UBC-1991).
The most recent edition of the UBC (1994) has incorporated a displacement-based design approach to replace the strength-based approach of UBC-91. The new edition provides
guidelines for the design of unsymmetrical, "ftanged" walls; however, there are stili some
significant shortcomings in the code which must be addressed. Some areas of concem include: (1) estimates of effective ftange widths, (2) magnitudes of design displacements, and
(3) the rapid screening process.

1.5 Scope
The primary objective of this study is to evaluate the effectiveness of using a displacementbased design proced.ure for designing reinforced concrete structural walls with both symmetrical (rectangular or barbell shaped) and unsymmetrical (L-, C-, or T-shaped) cross-sections.
This is accomplished by: (1) reviewing available analytical and experimental research
on the behavior of reinforced concrete structural walls subjected to combined loadings
and identifying areas where research results are lacking, (2) developing an experimental
program to address some of the uncertainties in lateral load response of symmetrical and
unsymmetrical structural walls and, (3) conducting analytical studies to compare with the

1
experimental results to help evaluate the effectiveness of using the proposed displacementbased design procedure. Secondary objectives of the research include: (1) addressing the
problem of effective flange widths for unsymmetrical wall cross-sections, and (2) reviewing
current codes (UBC-1994) which have incorporated a displacement-based design philosophy.

1.6 Organization
This dissertation is divided into nine chapters and two appendices.

Chapter two

presents a literature review of the most recent research involving reinforced concrete structural walls and displacement-based design. Chapter three describes the design of a prototype building that incorporates rectangular and T-shaped structural walls for lateral
load resistance. Chapter four describes the four scale model structural walls tested in the
experimental phase of this study. Chapter fi.ve gives a detailed description of the experimental program and Chapter six presents the experimental results. Chapter seven describes
the analytical studies that were conducted and provides comparisons of experimental and
analytical results. Chapter eight is dedicated to verifying the reliability of the proposed
displacement-based design procedure used in this study. Chapter nine presents a summary
and lists the conclusions that were drawn from this study.
There are two appendices located at the end of this dissertation. The first appendix
describes the displacement-based design procedure used to design the walls tested in this
study. This appendix gives a general overview of the methodology and lists the necessary
equations.

1
1
1

it is not intended to be an "all inclusive" discussion of displacement-based

design; therefore, interested readers are referred to the numerous references provided. The
second appendix describes the experimentally observed behavior of each wall specimen, and
includes numerous photographs that show the damage experienced at various displacement
(drift) levels during testing.

1
1
1
1
1
1
1
1
1
1
1

1
1
1

1
1
1
1
1
1

Chapter - 2
LITERATURE REVIEW

This chapter provides an overview of previous research that has been conducted on
reinforced concrete structural walls and displacement-based design strategies.
The studies involving structural walls are broken into categories depending on their
classifications. The main emphasis of this research is on slender, solid structural walls

whose behavior is dominated by flexure; therefore, some categories of structural walls will

1
1
1
1

systems, and building frame/shear wall systems. H greater detail is desired on these partic-

1
1

1
1
1
1
1

receive limited attention, including squat shear walls, "coupled" and "pierced" shear wall

ular classifications, the reader is directed to the numerous references, particularly that by
Abrams (1991) which provides a listing of over 200 references which are listed according to
their classification (similar to those used in this chapter).
A literature review of displacement-based design procedures is presented following the
review of structural walls. For clarity, all of the studies are identified by subheadings that
list the authors of each particular study.

2.1

Reinforced Concrete Structural Walls


Previous research related to reinforced concrete structural walls has focused primarily

on isolated walls and wall systems with symmetrical cross-sections (rectangular and barbell
shaped). One of the primary objectives of this research is to extend the knowledge and
understanding of these symmetrical walls to unsymmetrical, "flanged" walls. To accomplish
thls task, a review of previous studies of symmetrical walls was conducted to establish what
has been sufficiently documented and where research results are lacking. The next step
involved reviewing previous st udies of unsymmetrical, ''flanged" walls; however, very little
research has been conducted in this area.
Eathquake Eng Res Ctr Llbrary

Unlv of C811f. Berkeley


1301 S. 46th St. - RFS 463
Rlchmond, CA 94804-4898 USA
(510) 685-3419

1
2.1.1

Squat Shear Walls

As mentioned previously, squat shear walls typically have aspect ratios of one or less and
their behavior is often dominated by shear. The effects of shear typically lead to premature
stiffness and strength degradation anda reduced ability to dissipate energy (Paulay, 1986);
however, if walls have properly detailed web reinforcement then it is possible to ensure
inelastic flexural response (Paulay, et al., 1982). Because of the brittle nature of shear
failures, research has focused on trying to inhibit shear failures so that fiexural-strength will
govern their behavior. A detailed review of research on squat shear walls was not conducted;
however, those interested can find numerous references in the reports by Abrams (1991) and
Ali and Wight (1990).

2.1.2

"Coupled" and "Pierced" Wall Systems

Other topics that are not reviewed in great detail include, "coupled" and "pierced" wall
systems. Although the effects of openings in struct ural walls is not a primary concem in
this particular study, a majority of the research performed in these areas also falls into the
category of slender walls; therefore, pertinent studies are discussed in the following sections.
For a more detailed review of "coupled" and "pierced" wall systems the reader is referred
to the report by Abrams (1991) which provides 16 references specifically focussing on the
measured behavior of "coupled" wall systems.

2.1.3 Building Frame/Shear Wall Structures


Structures utilizing a building frame/shear wall combination are very popular in the
U.S., especially in zones of high

seisnic

risk where this type of construction accounts for

approximately 50% of the reinforced concrete buildings under 15 stories tall (Meigs, et al.,
1993).

10

1
1
1
1

1
1

1
1
1

1
1
1

1
1

1
1

1
1
1
1
1
1
1
1
1
1

Goodsir:
Research by Goodsir et al. (1985} was aimed at developing a design methodology for
interconnected frame-wall buildings. The method uses a capacity-based design philosophy
and incorporates procedures used for structural walls and ductile moment-resisting frames.
The method starts by using a routine static elastic lateral load analysis to determine "code"
level earthquake forces. The beams were than designed for the combination of lateral and
gravity loads. Columns were designed for moments in excess of the code values which
accounted for maximum input from the flexural overstrength of the beams and moment
magnification due to the participation of higher mode response. The structural walls were
than designed based on a linear flexural-strength envelope. Ali elements were designed
to sustain relatively high levels of shear to avoid the possibility of shear failure.

The

research involved extensive analytical studies of simplified buildings designed using the
proposed methodology, as well as experimental studies of four approximately 1/3 scale
structural walls. The analytical studies involved dynamic analyses of the simplified buildings
subjected to the El Centro and Pacoima Dam accelerograms. These analytical studies lead
to the conclusion that "the proposed methodology is logical and straightforward and should
provide buildings so designed, and carefully detailed, with excellent seismic resistance."
The main objective of the experimental portion of the study was to investigate the
adequacy of the New Zealand code provisions concerning hoop reinforcement and section
instability. Of the four walls that were tested, three had rectangular cross-sections and
one hada T-shaped cross-section. The rectangular walls had 1500 x 100 mm (59 x 4 in.)

1
1
1
1
1

cross-sections and the T-wall was comprised ofa 1300 x 100 mm (51 x 4 in.) web anda 700
x 100 mm (27.5 x 4 in.} flange (aspect ratios, hw/lw, were approximately 2.2). Material
strengths varied for the four wall specimens. Concrete compressive strengths ranged from
23.2 to 40.3 MPa (3,360 to 5,850 psi) and steel yield strengths ranged between 345 and 450
MPa (50 and 65 ksi). The main variables were the amount of hoop reinforcement in the
boundary zone, the level of axial load and the cross-sectional shape. Hoop reinforcement was

11

1
typically recommended over the outer half of the expected depth of the compression zone,
regardless of the expected maximum concrete compressive strain. The walls were tested
with eccentrically acting axial loads varying between 0.03 and 0.2A 9 J;, and subjected to
increasing levels of fully reversing displacement controlled lateral load.
The failure of ali four specimens initiated in the unconfined concrete immediately adjacent to the well-confined core. The first rectangular wall was subjected to the highest level
of axial stress and experienced a crushing failure at a displacement ductility (6) of 4. The
other three specimens had lower axial stress levels and each failed at a displacement ductility of 6. Failure was due to the combined effects of high ductility demands and axial force
eccentricity arising from out-of-plane displacements. The amount of hoop reinforcement
used in the boundary elements was sufficient for both confinement and buckling purposes;
however, it was concluded that the hoop reinforcement should be extended further into
the section when large compressive zones are expected. The increase in depth should be a
function of the expected ductility demand on the wall. Goodsir recommended that future
studies be conducted that: (1) address the necessary depth which the hoop reinforcement
extends into the cross-section, and (2) assess the amount of necessary hoop reinforcement

The report by Abrams (1991) lists 85 references pertaining to the behavior of frames
and frame-wall systems.

2.1.4 Slender Structural Walls


A large number of studies have been conducted that focus on isolated slender structural
Many

additional studies can be found in the list of references.

Portland Cement Association (PCA):


Perhaps the most extensive study of slender structural walls to date was conducted at
the Construction Technology Laboratories of the Portland Cement Association in Skokie,

12

1
1
1
1
1
1
1

to help indicate how conservative the current New Zealand code provisions were.

walls; however, only the most relevant studies will be reviewed in this section.

1
1
1
il
il
1

1
1

1
1
1
1

1
1
1
1
1

1
1
1
1
1

1
1
1
1

lliinois (Oesterle et al., 1976; Oesterle et al., 1979; Shiu et al., 1981). The first two stages
of this study (Oesterle et al., 1976; 1979) involved the construction and testing of sixteen
solid structural wall specimens of approximately 1/3 scale. The primary test variables
were the shape of the wall cross-section (rectangular, barbell, and fianged), the amount of
fiexural reinforcement at the wall boundaries, the amount of transverse reinforcement in the
boundary zones, the amount of horizontal shear reinforcement, the level of axial stress, the
concrete strength, and the load history. The main purpose of the study was to determine
the ductility, energy dissipation capacity, and strength of structural walls and to develop
design criteria for walls in earthquake resistant structures.
The sixteen test specimens were all designed based on the 1971 ACI Building Code
and were tested under monotonic or reversed cyclic lateral loads applied to the top of the
specimens. Axial loads acting on the wall specimens ranged from zero to approximately

0.09A 9 J;. All specimens were 15 feet (4.57 m) tall, 75 inches (1.9 m) long and 4 in. (102
mm) wide. The aspect ratio (hw/lw) for all specimens was 2.4. The barbell shaped walls
had 12 in. (305 mm) square boundary elements and the flanged walls had 36 in. (914 mm)
long and 4 in. (102 mm) thick fianges at each end of the wall web. Concrete compressive
strengths ranged from 3,165 to 7,775 psi (21.8 to 53.6 MPa) and steel yield strengths ranged
from 59.5 to 74.2 ksi (410 to 512 MPa).
The results of the study showed that two types of wall behavior were observed depending on the shear stress level. Behavior of walls subjected to shear stresses less than
3v'"Pc psi (0.25

v'"Pc MPa) was govemed by buckling of the flexural reinforcement and loss

of concrete confinement within the core; whereas, the capacity of walls subjected to shear
stresses greater than 7

v'"Pc psi (0.64 v'"Pc MPa) were governed by web crushing.

The ductil-

ity of structural walls decreased as nominal shear stress levels increased. The nominal shear
stresses on the wall specimens ranged from 1.4 to 13.8v'"Pc psi (0.1 to 1.1

v'"Pc MPa) and

the wall ductilities, as determined from measured rotations, ranged from approximately 3
to 8. It was also determined that shear deformations within the hinge region contribute

13

1
significantly to overall specimen deformation; however, they are coupled to the flexural rotations. The effect of this coupling is that shear distortions increase with increasing levels
of flexural hinging. Results also showed that structural walls could possess good inelastic
deformation capacity, especially if sufficient transverse reinforcement was provided in the
boundary elements. This confinement served several purposes including: increasing allowable compressive strain in the concrete, preventing main longitudinal reinforcement from
buckling, containing concrete after cracking, and improving shear capacity and stiffness.
It was also concluded that the confinement reinforcement was required only in the plastic
hinge region of the wall (usually over a height approximately equal to the length of the
wall), and that proper anchorage of horiwntal web reinforcement into the boundary ele-ments was essential. Additional horizontal web steel, beyond that required by ACI-1971,
did not improve strength or ductility in walls govemed by web crushing. Strengths and
ductilities of walls subjected to reversed lateral loads were lower than those calculated for
walls tested under monotonic loads, and behavior was more dependent upon the previous
maximum displacement level than it was on the entire previous load history.
The third study performed at the PCA was conducted by Shiu et al.

(1981) and

involved the testing of wall specimens with and without openings. The main objectives
of this study were: to determine effects of openings on strength and deformation capacity
of structural walls under simulated earthquake loadings, and to verify design criteria and

Both walls

were 18 ft (5.49 m) tall, 75 in (1.91 m) long and 4 in (10.2 cm) thick, resulting in an
aspect ratio (hw/lw) of 2.88.

1
1
1
1
1
1
1
1

u
u

reinforcement details for earthquake resistant structural walls with openings.


Two, six story, approximately 1/3 scale rectangular walls were tested.

The two walls were identical except for the inclusion of

openings in the second wall. The openings were 12.5 x 18 in. (317 x 457 mm) and were
located at the center of the wall at each story level. The openings simulated typical window
openings and accounted for approximately 8.3% of the wall area. The walls were designed

u
1

based on the 1971 ACI Building Code and the 1976 Uniform Building Code. Concrete

14

1
1

1
1

1
1
1
1

compressive strengths were approximately 3,200 psi (22 MPa) and steel yield strengths
were approximately 65 ksi (448 MPa) for both specimens.
The study concluded that load capacity and deformation characteristics of the solid
wall were very similar to those of the "pierced" wall. Shear stresses were 4.6

J'I'c MPa)

and 4.8

J'I'c psi (0.40 J'I'c MPa),

J'I'c psi (0.38

and displacement ductilities (6) were 4.3

and 3.5 for the solid and "pierced" wall, respectively. Both specimens exhibited similar
energy dissipation capacities. The design practice of placing discontinuous reinforcement
to each side of the openings functioned well. No special diagonal reinforcement was used
in the lintels and stress concentrations around the openings were not observed to affect the
behavior of the "pierced" wall. Different failure modes were observed for each specimen.

1
1
1
1
1

1
1
1
1
1
1

The solid wall lost its load carrying capacity through "sliding shear" along a horizontal crack
in the first story level; whereas, the "pierced" wall lost capacity through shear-compression
failure at the boundary elements. It was also concluded that premature failure of lintels
between openings can be avoided by designing lintels to remain elastic until the flexural
capacity of the structural wall is reached.
Ali and Wight:
Ali and Wight (1990, 1991) tested four, approximately 1/5 scale, walls which were
fi.ve stories high and had barbell shaped cross-sections. All four specimens were 140 in.
(3.56 m) tall and 48 in. (1.22 m) long resulting in an overall aspect ratio of 2.92. One of
the specimens was solid whereas the other three incorporated staggered door openings. A
longitudinal boundary element reinforcing ratio of 0.03 was provided and hoops (3/16 inch
diameter wire) were spaced at 2.5 in. (64 mm) over the lower two stories. This amount of
transverse reinforcement was approximately 50% of that required by current codes at the
time of the study (UBC-1985). Web reiif'orcement was a single curtain of #2 (6.4 mm)
deformed hars both vertically and horizontally which resulted in a uniformly distributed
reinforcing ratio of 0.003 (slightly larger than the code minimum of 0.0025).

All four

specimens were constructed of concrete with compressive strengths of approximately 5,000

15

-
psi (34.5 MPa). The specimens were tested under constant axial stress of approximately 400
psi (2.8 MPa; 0.08/;) and reversed cyclic lateral loads were applied at the top of the wall
by a hydraulic actuator. The tests were displacement controlled and cycled to increasing
levels of d.rift ratio ( top displacement divided by wall height, 6/ hw).
The results of the study revealed that all four specimens exhibited ductile flexural
behavior up to an average drift ratio of 1%. At drift ratios of 1.25 and 1.5% the three
specimens with staggered openings experienced shear-compression failures in the portion of
the wall between the opening and the compression edge; however, the solid wall experienced

1
1

1
:

il

drift ratios up to 3% without any significant loss of strength and a relatively minor loss of
stiffness. Based on the results of the four lightly reinforced walls tested, the following
conclusions were drawn: (1) slender structural walls subjected to moderate levels of axial
stress (O.IOJ;) and shear stress (3 to 4

vfc psi) can experience 1% d.rift without major

damage, and higher d.rift levels if moderately confined boundary elements are provided,
(2) the UBC-1985 requirements for wall boundary element confinement can be significantly
relaxed for these types of walls, and (3) a staggered door opening configuration is a viable
alternative to the in-line openings created by a coupled wall system; however, openings
located too close to the edge of the wall can trigger a premature shear-compression failure.

2.1.5

Unsymmetrical, "Flanged" Structural Walls

Relatively few studies have been conducted that have focussed on "flanged" walls;
therefore, behavior of these walls is not as well understood as that of symmetrical walls.
The need to evaluate the effects of flanges is instrumental in evaluating many aspects of
wall behavior, including strength, stiffness and ductility.
Priestley and Limin:
Priestley and Limin (1990) conducted a study of T-shaped reinforced masonry shear
walls. The study was aimed at developing a Structural Component Model (SCM) capable

16

~t
1
1
1
il

1
1

1
1
1
1

1
1

of predicting the non-linear response histories of flanged wall elements suhjected to seismic
The fi.rst phase of their study involved experimentally testing four flanged wall

loads.

1
1
1

in the experimental investigation included: the amount of longitudinal reinforcement, the

models subjected to pseudo-static cyclic lateral loading. The main variahles considered

flange width, and the presence or ahsence of confi.nement in the web at the base of the wall.
The walls were 12 ft. (3.66 m) ta11, 46 in. {1.17 m) long (aspect ratio, hw/lw, of
3.13), and had flange lengths of either 104 in. (2.64 m) or 200 in. (5.08 m). They were
constructed with 6 in. (152 mm) nominal concrete masonry units (actual dimension of
5.625 in., 143 mm), and fully grouted; thus providing effectively monolithic action of weh

I
1
1
1
1

1
1
1
1
1
1
1

and flange. Vertical reinforcement consisted of either #6 (19 mm) or #4 (13 mm), grade
60 hars at 16 in. (406 mm) on center. Transverse weh reinforcement was essentially the
same for all four walls and consisted of #4 (13 mm) grade 60 hars at 8 in. (203 mm) on
center over the full height of the walls. Transverse flange reinforcement consisted of #4 (13
mm) hars at 16 in. (406 mm) vertical centers. One of the walls included 1/8 in. (3 mm)
thick steel confi.ning plates at the end of the web over the lower 7 mortar bed courses. The
walls were tested under axial and reversed cyclic lateral loads. A uniform axial stress at the
wall hase (including self weight) of 100 psi (0.69 MPa; approximately 0.05/:,.) was provided
and lateral loads were applied by a douhle-acting hydraulic actuator at the top of the wall.
Standard prism tests resulted in compressive strengths ranging from 1.8 to 2.6 ksi (12.4 to
17. 9 MP a); however, prism strengths predicted using mechanics of masoncy and relating
block strength and grout strength, resulted in predicted strengths that were 10-28% higher.
Steel reinforcement was all grade 60; however, yield strengths ranged hetween 68 and 75 ksi
(469 and 517 MPa).
Results of the experimental tests showed that hysteretic behavior of the walls was
characterized hy thin loops (poor energy dissipation capahility), particularly on repeated
cycles. Wall failure was sudden and brittle, and initiated by a compression failure in the toe
of the web; however, behavior could be drastica1ly improved by including confi.nement of the

17

.
web. It was also concluded that shear d.isplacements were significant when the web was in
compression, accounting for up to 30% of the total displacement, despite the slender nature
of the walls. Finally, deflection calculations based on a modi:6.ed elastoplastic approach
which acknowledges the spread of elastic strains caused by diagonal-flexure shear cracking,
agreed well with experimental values.

Sittipunt and Wood:


Sittipunt and Wood (1993) conducted an experimental and analytical study of Cshaped "flanged" structural walls at the University of Illinois at Urbana-Champaign. Two
isolated, C-shaped wall specimens were tested under reversed cyclic lateral loads (applied

1
'.I

1
1

.,,,,

parallel to the webs of the specimens) and constant axial stress of approximately 0.065J;.
The walls consisted of two 36 in. (0.91 m) long parallel webs and a 60 in. (1.52 m) long
connecting fiange. Both walls were 9 ft. (2. 74 m) tall (aspect ratios, hw/lw, of 3) and 3
in. (76 mm) thick. The main longitud.inal reinforcement in each web consisted of 10 #3
(9.5 mm) hars, four of which were placed at the web flange intersection while the other
six were placed at the boundary at the free end of the web. Transverse reinforcement
consisted of hoops and cross-ties made from #10 gage (approximately 1/8 inch diameter)
wire. The transverse reinforcement was spaced at 2 in. (51 mm) and was approximately
66% of that required by ACI 318-89. The d.ifference between the two specimens was the
amount of uniformly distributed web steel. Web reinforcement consisted ofa single layer of
#2 (6.4 mm) deformed hars in each d.irection. One specimen had a web reinforcing ratio
of 0.0025 (minimum allowed by ACI 318-89) while the other specimen had twice as much
web reinforcement. Failure of both specimens was caused by crushing of the concrete in
the boundary region at the free end of the web and buckling of longitudinal reinforcement.
After failure, the region of crushed concrete extended nearly the entire length of the web.
It was concluded that all of the vertical reinforcement in the fiange was effective as
tension reinforcement ind.icating that the effective fiange width could be as high as ten
times the thickness of the fiange. The use of a too small effective fiange width can lead

18

1
1

"ti

il

1
1
1
1

1
1
1
1

to an underestimation of the wall strength which could result in inadequate transverse


reinforcement in the boundary elements opposite the intersecting flange. When the fiange
was in compression, very low compressive stresses were observed. Conclusions indicate
that behavior of flanged walls are governed by the ability of the web to resist the large
compressive strains (concrete confinement and buckling of longitudinal reinforcement) that
develop when the flange is in tension. Finite element analyses were performed to compare
with the experimentally observed behavior. The models successfully simulated the cyclic
behavior of the C-shaped walls and thus it was concluded that the models could be used
to investigate the behavior of C-shaped walls with different configurations. The authors

1
1

1
1

1
1
1

suggested that some improvements were necessary in the material models to increase the
capacity of the analytical modeling procedure. These improvements include, among other
things, a linkage element which accurately models slip between concrete and reinforcing
hars, anda better shear stress function for the concrete material model

2.2 Design Methodologies


As discussed previously, current codes (UBC-1994) have recently incorporated a displacement based design approach to evaluate detailing requirements at wall boundaries, to
replace the strength-based approach that has been used for many years. The studies that
are discussed herein are a few of the studies that helped direct designers and code officials away from the inhibitive strength-based design methodologies toward more "rational"
displacement-based design methodologies.

2.2.1

Displacement-Ductility Based Design

Paulay:
Early research by Paulay (1986) focused on comparing ductility demands as a design

tool as opposed to strength (such as incorporated into the Uniform Building Codes prior to

1
1

19

1994). Paulay suggested relating global displacement ductility to required wall details. The

first step is calculating the global displacement ductility demand using inelastic response
spectra (ie. ATC-03-06, 1978). Based on the global demand, a local curvature ductility
demand can be calculated for each wall. The local curvature ductility demand is then
compared with the available wall cross-sectional ductility, for a given set of details, to

1
1
1

1
.

determine whether the details are adequate.


The main drawback to this method is that it must be assumed that each individual
wall will undergo the same displacement ductility as the entire building. This is generally

not the case for buildings having a variety of wall cross-sections. Because different walls

....

yield at different displacements, and each wall is required to displace an amount equal to

"'

the building displacement, different displacement ductilities will be required for different
walls (Wallace and Moehle, 1989 ).

Moehle:
Moehle (1992) suggested that a simpler and more direct way of determining the necessary details is to work explicity with building displacements as opposed to displacement
ductility. This approach is useful in the planning stages when decisions are made regarding the control of displacements and in the final design stages when details for structural
and non-structural elements are established. A ductility-based approach may be adequate
for certain cases such as when inelastic response is distributed uniformly throughout the
structure; however, when inelastic response is not uniformly distributed, the local demand
and capacity are not the same for all elements and a displacement-based design may be
preferred. Regardless of the design hasis (strength, ductility or displacement) a comprehensive analysis should consider displacements, forces and ductilities. Moehle concluded
that the displacement-based approach can be used to establish proportions and layout that
will control drift demand, and to determine structural and non-structural details that will
ensure adequate performance. Compared to force or ductility-based design procedures, a
displacement-based approach is both simple and effective.

20

JI
ti

il
1
1
1

1
1
1
1

1
1

1
1

1
1

1
1
1
1

1
1

2.2.2

Displacement-Based Design

Wallace:
Other studies (Moehle and Wallace, 1989; Wallace and Moehle, 1992; Wallace, 1994a;
1995) have focused on developing a displacement-based design procedure for reinforced concrete structural walls. Wallace (1994a) developed a procedure that compares the expected
displacement capacity with the displacement demands on the building thus eliminating the
need to provide a single system ductility factor for a given building configuration. The
proposed procedure relates the need for special confinement with the expected building
response, thus reducing the likelihood of overdesigned walls.
The primary variables affecting the wall details are: the ratio of wall area to floor
plan area, the wall aspect ratio and configuration, the axial load, and the reinforcement
ratios. This approach results in wall designs that are directly dependent upon the building
configuration; therefore, any change in building configuration will have a direct effect on
the wall details ( transverse reinforcement). The displacement-based design approach also
provides guidelines for the design of unsymmetrical "flanged" walls. Applications of this
design procedure (Wallace and Thomsen, 1995) have shown it to be flexible and effective
for evaluating structural wall behavior.
The displacement-based design procedure developed by Wallace is used extensively
throughout this study, and evaluation of its effectiveness is a primary objective of this
research. For this reason, a detailed description of the procedure seems warranted, and is
provided in Appendix A at the end of this dissertation.

1
1
1

1
1

21

1
Chapter - 3

1
PROTOTYPE DESIGN

3.1

Description of the Prototype Building


A prototype building, representing a typical multi-story office building located in an

area of high seismic risk, was used to assist in determining the wall geometry and reinforcing
details for the testing program. The prototype building is six stories tali and incorporates
rectangular and T-shaped reinforced concrete structural walls as well as concrete momentresisting frames to resist lateral and gravity loads. This type of construction is quite common
in areas of moderate to high seismic risk (Meigs et al., 1993). The prototype building has
plan dimensions of 180 x 100 feet (55 x 30.5 m) and has a story height of 12 feet (3.66 m).
Figure 3. 1 shows a plan view of the prototype building.
This chapter describes both the simpli:6.ed and the detailed proc~ures that were used
to arrive at the final design of the prototype building. in general, simpli:6.ed procedures were
used wherever possible and results were veri:6.ed using more detailed procedures. Design and
analysis of the building and the individual structural walls are discussed in sections 3.2 and
3.3, respectively.

3.2

Simplifi.ed Design and Analysis of the Prototype Building


The simpli:6.ed analysis of the prototype building was divided into two main parts. The

:6.rst part involved analyzing the building's strength requirements; whereas, the second part
focused on estimating the expected building displacements. Detailing requirements for the
individual walls are addressed in section 3.3.

1
1
1
1
1
1

1
1
1

Building Strength Reguirements:

22

1
1

1
1
1
1
1

The forces imposecl on the prototype building were calculatecl basecl on the 1991 Uniform Building Code (UBC-91). Building dead loads were estimatecl as 175 psf (8.38 kPa);
whereas, live loads of 40 and 20 psf (1.92 and 0.96 kPa) were usecl for floors and the roof,
respectively. Wind loads were calculatecl assuming a basic wind speecl of 80 mph (129 kph).
Earthquake loads were calculatecl using the static lateral force proceclure outlined in UBC91, section 2334. A design base shear (Vcode) was calculatecl using Equation 34-1 and the
following values: Z=0.4, 8=1.2,

1.0 (importance factor) and Rw=12 (dual system). A

design base shear of 1,506 kips (6,699 kN) was calculatecl for the prototype building. Basecl
on a story height of 12 feet (3.66 m) and a triangular shapecl load distribution over the

'

height of the building, a base overturning moment of 'rl,288 kip-feet (98,023 kN-m) was
calculatecl for the building. Due to the relatively large stiffnesses of the structural walls

comparecl to the concrete frames, it was assumecl that the structural walls would resist all

1
1

of the lateral loads imposecl on the building. Discussion of the individual structural walls

An estimate of the expectecl lateral roof displacement of the building was calculatecl

1
1
1

1
1
1
1

is presentecl in the next section.


Building Displacement Estimates:

using the displacement-basecl design procedure recommendecl by Wallace (1994a, 1995).


Appendix A presents a detailecl discussion of the proposecl displacement-based design proceclure that will be referrecl to many times throughout this dissertation; therefore, the reader
may find it beneficial to familiarize themselves with the A ppendix and its references before
proceeding any further.
Basecl on the dimensions of the prototype building and the structural walls (Fig. 3.1),
a ratio of wall area to total floor plan area of 0.00711 (p = O. 711 % ) was calculatecl. The
height and length dimensions of the prototype walls result in an overall aspect ratio (hw/lw)
of 4.5. Basecl on these values, use of Eq. A6 results in a roof drift estimate of 1.07% which
corresponds to a roof displacement of 9.24 in. (234 mm).
Once the global building response has been estimated, the evaluation of individual

23

1
structural components can be conducted., as described. in the following sections.

:
3.3 Simplifled Design and Analysis of the Prototype Walls
A rectangular wall and a T-shaped. wall oriented. in the N-S direction (Fig. 3.1) were
chosen as prototype walls to aid in the design of the wall specimens tested. as part of the
experimental program. The prototype walis are 72 ft. (21.9 m) tali, 16 ft. (4.88 m) long and
16 in. (406 mm) thick. The flange and web of the T-shaped wali have the same dimensions
as the rectangular wall The design of the rectangular and the T-shaped prototype walis

1
J.I

are described in the following two subsections, respectively.

3.3.1

Rectangular Prototype Wall

This section describes the simplifi.ed procedures used to design the rectangular prototype wall The procedures include simple hand calculations (UBC-91 equations and Appendix A equations) as well as sectional analyses using the BIAX (Wallace, 1992) computer program.

The aspects of the wall design that are considered include: flexural-

strength, transverse reinforcement in the boundary zones (confi.nement, buckling), and


shear-strength. The computer analyses were used to verify certain aspects of the hand
calculations including strength, ductility, compressive strain and depth of compression zone.

Flexural-Strength Reguirements:
It was conservatively assumed that the structural walis in each direction would resist

ali of the lateral loads imposed. on the building because of their relatively large stiffnesses
compared with the frames. The factored. building base shear (1.4 x Vcoc:1e) of 2,108 kips
(9,376 kN) was evenly divided among the six structural walls oriented in each principal
direction resulting in a factored base shear of approximately 351 kips (1,561 kN) per wall.
It is noted that the T-shaped walis have larger moments of inertia and therefore will attract
larger lateral forces than the rectangular walis; however, simplified design practice might

24

~1
1
il

1
il
il

1
1
1

involve even distribution of lateral force to ali walls having similar web dimensions, thus

this approach was used. Assuming a triangular loading over the height of the wall, the base

the wall. An axial load of O. lOAwf~ was estimated for the rectangular prototype wall based

1
(

1
1
1
1

1
1

1
1
1
1
1
1

shear force results in an overturning moment of 16,848 k-ft (22,846 kN-m) at the base of

on its tributary area.


Longitudinal boundary reinforcement was designed to resist the combined effects of
the lateral and axial loads and consists of 10-#11 (36 mm) hars at each boundary (p

p' = A 8 /twlw = 0.00508). Uniformly distributed web steel was based on code prescribed
minimums (p" ~ 0.0025) and consists of two curtains of #5 (16 mm) hars at 14 in. (356
mm) in both the horizontal and vertical directions (Ph = Pv = 0.00277). Figure 3.2a shows
the cross-section and reinforcing details at the base of the rectangular prototype wall, and
Table 3.1 summarizes the wall design parameters for both prototype walls.
Figure 3.3a plots two P-M interaction diagrams (4>=1.0, and <t>=O. 7) for the base of
the rectangular prototype wall. The curves were constructed using BIAX and assuming
unconfined concrete (!~=4 ksi, 27.6 MPa) and an elastic, perfectly plastic stress-strain
relation for the reinforcing steel {/11 =60 ksi, 414 MPa). Three load cases are plotted on Fig.
3.3a to verify that the prototype wall has adequate strength. The first load case consists of
the factored gravity loads only; whereas, the second and third load cases consist of factored
gravity and lateral (earthquake) loads (ACI 318-89, Section 9.2). It can be seen that all
three load cases lie within the design interaction curve, indicating adequate strength of the
section.

Transverse Reinforcement Reguirements:


Due to the generally over-conservative detailing requirements of UBC-91 for confining steel at boundaries of symmetrical wall cross-sections (Ali and Wight, 1990; Wallace
and Moehle, 1992), considerable liberties were taken in the design of the transverse reinforcement. Transverse reinforcement was designed using the displacement-based approach
recommended by Wallace (1995) as opposed to current ACI provisions.

25

1
Basecl on the maximum roof drift eBtimate of 1.07%, the ultimate curvature demand
on hoth the rectangular and T-shapecl walls is 0.01854/lw (Eq. A9). Recalling that lw =
192 in. (4.88 m), an ultimate curvature (<l>u) of 0.0000966 rads/inch (0.000038 rads/cm) is
expectecl at the hase of the walls. Using the results of Eq. A9, the extreme fiber concrete
compressive strain can he calculatecl using Eq. A11, where p, p 1 , p 11 , lw, and tw were de:finecl
ahove, "Y

= 1.25, o= 1.5, J.,, = 60 ksi (414 MPa), J! = 4 ksi (27.6 MPa), P=O.lOAwf:=l,229

kips (5,467 kN), and

/3

0.85. Suhstitution of terms into Eq. A11 results in an extreme

concrete compressive strain

(Ec,maz)

of 0.0040 which is recommendecl as the limiting strain

for which special transverse reinforcement is requirecl for concrete confinement. Based on the
estimates of ultimate curvature and extreme compressive strain, the depth of compression
zone was estimatecl at 41.1 inches (104 cm). Figure 3.4 plots the expected strain distrihution
for the rectangular wall. Basecl on the strain distrihution, special transverse reinforcement
for concrete confinement was conservatively provided over a minimum length of lw / 1O =
19.2 in. (48.8 cm) at each end of the wall.
Required transverse reinforcement was designecl hased on the guidelines set forth in
Appendix A for

Ec,maz ~

0.004. Details of the transverse reinforcement locatecl in the

houndary zones were govemecl hy huckling restrictions (as opposed to concrete confinement)
and consist of #4 (13 mm) hoops and cross-ties spacecl at 8 in. (Fig. 3.2a). The special
transverse reinforcement is providecl over the hottom 1-1/3 stories (lw) which corresponds
to an upper hound estimate of the plastic hinge length (Paulay, 1986).
Based on ACI 318-89 requirements (equations 21-3 and 21-4) the transverse reinforcement should he placecl at a spacing of approximately 3.2 inches (81 mm). The transverse
reinforcement requirecl according to the proposecl displacement-hasecl design approach corresponds to appro:ximately 40% of that specifiecl hy current ACI requirements.
A moment-curvature analysis was performecl on the rectangular prototype wall to verify
previous estimates of curvature capacity, extreme fiber compression strain, and depth of
the compression zone. The moment-curvature plot is shown in Fig. 3.5 and was calculatecl

26

1
1
:

1
1

il
1
11

il
1

il

u
1
il
1
il
1

1
1

1
1

1
1

1
1

'
1
1

1
1
1

assuming unconfined concrete (1;=4 ksi, 27.6 MPa), anda probable steel stress-strain curve
to account for material over-strength and strain hardening (/11=65 ksi, 448 MPa; fu=90 ksi,
621 MPa). A line showing the expected ultimate curvature demand on the wall is plotted
in Fig. 3.5 and indicates good performance. Figure 3.5 also indicates where buckling of the
longitudinal steel is expected to control design based upon the 8 in. (203 mm) spacing of
transverse reinforcement in the boundary zones. This point was determined using Eq. A 13
and a probable stress-strain relationship for the reinforcing steel. Based on the computed
analysis, the expected ultimate curvature corresponds to a compressive strain at the extreme
fiber of 0.0041, anda 42.9 in. (109 mm) depth of the compression zone. These values are
listed in Table 3. 1, and comparisons with previously calculated values indicate excellent
correlation between the two approaches.

Shear-Strength Reguirements:
An estimate of maximum shear stress expected to develop in the wall is required to
ensure that adequate shear-strength is provided. Based on current design provisions (ACI
318-89, Eq. 21-6) the wall shear-strength can be computed using Eq. A16, where Ac is the
concrete area bounded by the web thickness (16 in., 406 mm) and the length of the section
in the direction of shear force considered (192 in., 259 cm), ; = 4,000 psi (27.6 MPa), ! 11
= 60,000 psi (414 MPa), and Pn is the ratio of distributed shear reinforcement on a plane
perpendicular to

Ac

(0.00277). Substitution of these values into Eq. A16 results in a

nominal shear strength of 899 kips (4,000 kN). Using a capacity reduction factor of 4>=0.85
(for shear capacity when flexural-strength is obtained prior to shear-strength) results in <t>Vn
= 764 kips (3,398 kN).
The maximum expected wall shear stress was calculated using two different approaches.
The first was a simplified approach based on Eq. Al7; whereas the second was a more

detailed approach based on the moment-curvature analysis. The unfactored base shear

force expected to develop in the wall can be estimated using Eq. Al 7, where w,, is taken as

1
1

27

(Vcoae) calculated using UBC, Equation 34-1, is 251 kips (1,116 kN). The maximum shear

1
4/3 and M 0 / Mcode is taken as 2. Based on Eq. Al 7 the maximum shear force expected to
act on the wall is 669 kips (2,976 kN) which is less than the factored nominal shear capacity
of the wall. The maximum expected shear force results in a shear stress of 218 psi (1,503
kPa) which corresponds to approximately 3.44...J'H. This stress is well below the limit of

6...J'H recommended by Aktan and Bertero (1985).


Based on the moment-curvature plot shown in Fig. 3.5, the moment capacity of this
wall at the expected ultimate curvature is approximately 300,000 in-kips (33,900 kN-m).
Assuming a uniformly distributed load pattern over the height of the building (to account
for higher mode effects), this moment corresponds to a shear force of approximately 694
kips (3,087 kN) and a shear stress of 226 psi (1,558 kPa). Shear-strengths and expected
shear stresses for both the rectangular and the T-shaped prototype walls are listed in Table
3.2.

;I
1
1

1
1

General Reguirements:
The required :flexural-strength ofa cantilever structural wall decreases over the height
of the wall thus allowing designers to terminate :flexural reinforcement at specific locations;
however, adequate flexural-strength must be provided over the entire wall height to ensure
that inelastic behavior occurs primarily in the well-detailed region at the base of the wall
(plastic hinge region). Based on an assumed linear moment envelope with an offset at
the base of the wall of lw (Paulay, 1986), it was determined that 40% of the longitudinal
boundary steel could be terminated at the fourth story. Figure 3.2b shows the cross-section
and reinforcing details of the rectangular prototype wall for the upper three stories of the
prototype building. The longitudinal reinforcement has been cut by 40% and the transverse
reinforcment is based on code prescribed minimums, and consists of #5 (16 mm) "U-shaped"
stirrups at the same spacing as the uniformly distributed web steel (14 inches; 35.6 cm).
Figure 3.3b plots two P-M interaction diagrams for the wall cross-section shown in
Fig. 3.2b (upper three stories of the prototype wall). The plot was calculated in the same
manner as Fig. 3.3a and the same three load cases are plotted in this figure except the axial

28

1
1

1
1
1

1
il
1
il

loads are from the upper three floors only and the moment is extrapolated from the linear
moment envelope with an offset at the hase of the wall of lw. it can he seen that the three
load cases lie within the limits of the plot, indicating adequate strength.

-1
,..
1
I'

'1
1
1
1

1
1
1

3.3.2 T-Shaped Prototype Wall (N-S Direction)


This section descrihes the simplified procedures used to design the T-shaped prototype wall and addresses the major differences hetween symmetrical (ie. rectangular) and
unsymmetrical (ie. T-shaped) cross-sections.

Flexural-Strength Reguirements:
The design hase shear was evenly divided among the six structural walls (four rectangular and two T-shaped), therefore the T-shaped prototype wall was designed to resist the
same lateral loads as the rectangular wall; however, the hehavior of the T-shaped wall will
differ significantly from that of the rectangular sections that form the T-shaped wall The
application of the proposed code format (Wallace, 1995; Appendix A) to T-shaped walls is
presented in the following suhsections.

Transverse Reinforcement Reguirements:


The hehavior ofa T-shaped structural wall is greatly influenced hy the direction of
the applied lateral loads. When the lateral load causes the flange of the wall to he in
compression, the tensile forces that develop in the weh reinforcing steel must he halanced hy
compressive forces in the fl.ange. Because the compression zone is relatively wide (effective
flange width), the depth of the compression zone is relatively short and compressive strains
are rarely large enough to warrant special transverse reinforcement for concrete confinement
(tc,maz :::;

0.004).

The critical condition occurs when the lateral loads are reversed, and the flange goes
into tension. The large tensile forces that develop in the flange reinforcement (houndary
and uniformly distrihuted) must he halanced hy compressive forces in the weh; however, the

29

web area is relatively narrow, thus the concrete in this area must be capable of withstanding
large compressive strains. Detailing of the special transverse reinforcement needed in the
web boundary zone requires an estimate of the effective flange width. The effective flange
width depends on the deformation (roof drift) imposed on the wall (Pantazopoulou and
Moehle, 1990); however, for design, simplified approaches for estimating the effective flange
width can be used. For effective tension flange widths, Paulay (1986) recommends using a
conservatively large estimate of tw plus hw/2 on each side of the web; whereas, ACI 318-89
recommends a width of approximately tw plus hw/4 on each side of the web if requireme:11ts
for T-beams are extrapolated to cantilever walls. Using either estimation results in an
effective width that significantly exceeds the length of the flange; therefore, the entire 16 ft.
(4.88 m) flange was assumed to be effective.

The extreme fiber compressive strain was calculated using Eq. A11 and the following
values: P=0.05AwJ;=l,178 kips (5,240 kN), and p=0.0168 (assuming all flange steel is effective as tension reinforcement). All other values were identical to those used for calculating
the compressive strain in the rectangular wall Equation A 11 resulted in a compressive
strain estimate of 0.00873 and the depth of the compression zone was estimated to be 90.4
inches (230 cm). Figure 3. 7b shows a plot of the strain profile for the T-shaped prototype
wall when the flange is in tension (critical condition). The strain profile indicates that the
portion of the wall where compressive strains are larger than 0.004 is 49 inches (124.5 cm),
which exceeds the minimum required length of lw/10=19.2 inches (49 cm). To accomodate this large depth, the spacing of the 10 - #11 (36 mm) hars in the web boundary was
increased from 6 to 12 inches (152 to 305 mm). Figure 3.6 shows the cross-section and
reinforcing details at the base of the T-shaped prototype wall. Based on the maximum
compressive strain estimate of 0.00873, Eq. A14 requires that the transverse reinforcement
in the web boundary correspond to 100% of that recommended by ACI 318-89. Transverse
reinforcement consists of 2-#4 (13 mm) hoops and one cross-tie spaced at 3 inches (Fig.
3.6); alternatively, #5 (16 mm) hoops and cross-ties at the ACI recommendecl maximum

30

......"
1

1
1
1
il

1
il
1
1

1
1
1
1

spacing of 4 inches (102 mm) could have been used.


When the flange of the T-shaped wall is in compression Paulay suggests an e:ffective
flange width of approximately tw plus hw / 4 on each side of the web. This estimate is
approximately 50% smaller than the estimate for the e:ffective tension flange; however,
the estimate still indicates that the entire flange is e:ffective when in compression. Use of
equation All (with reinforcing ratios listed in Table 3.1) results in a compressive strain

estimate of 0.00089 anda 9.23 in. (23.4 cm) depth of compression zone. Figure 3. 7c plots

special transverse reinforcement for concrete confinement is not neecled in the flange. Due

1
1
1
1

1
1
1

1
1
1
1

the strain profile for the T-shaped wall when the flange is in compression and indicates that

to cyclic loading of the wall, normal strains in the flange range from 0.00981 in tension (Fig.
3.7b) to 0.00089 in compression (Fig. 3.7c). The large tensile strains in the flange cause the
concrete to crack and the reinforcement to yield When the loads are reversed, the cracks
must be closed before the concrete can resist any of the compressive load; therefore, the
steel must resist the entire compressive force, which causes stability (buckling) concems.
Minimum amounts of transverse reinforcement are provided (for cyclic loads) in the :flangeweb intersection and consist of #4 (13 mm) hoops spaced at 12 inches (305 mm; Fig. 3.6).
The validity of this stability concem (in the compression flange) is questionable and will be
addressed in subsequent chapters.
Figure 3.8 plots two P-M interaction diagrams (<t>=l.0, and <t>=O. 7) for the base of
the T-shaped prototype wall. The curves were calculated using the same assumptions as
for the rectangular wall (uncon:6.ned concrete, elastic-perfectly plastic steel stress-strain
relationship, 4 ksi concrete, and 60 ksi reinforcing steel). Three ACI 318-89 load cases are
plotted on Fig. 3.8 verifying the strength of the T-shaped prototype wall.
Several moment-curvature analyses (Fig. 3.9) were performed on the T-shaped prototype wall to verify previous estimates of compressive strain and depth of compression zone.
All of the analyses were performed using a probable steel stress-strain curve (/11 =65 ksi, 448
MPa; f u=90 ksi, 621 MPa) and the analyses that include confined concrete incorporated the

31

stress-strain relationship recommended by Saatcioglu and Razvi (1992). The three specific
cases which are plotted include: (1) flange in compression with unconfined concrete, (2)
flange in tension with unconfined concrete, and (3) flange in tensi on with confined concrete.
It can be seen in Fig. 3. 9 that when the flange is in compression the wall is extremely

ductile and the moment capacity is similar to that of the rectangular wall (Fig. 3.5). At
the expected ultimate curvature of 0.0000966 rads/in. (0.000038 rads/cm), the computed
results indicate a maximum compressive strain of 0.0011, anda 11.4 in. (29 cm) depth of
compression zone (Table 3.1), verifying that special transverse reinforcement is not needed
for concrete confinement at the web-flange intersection.
The two cMes plotted in Fig. 3.9, for the flange in tension, verify that the web must have
well confined concrete in order to reach its ultimate curvature demand without experiencing
a loss of strength. At the expected curvature demand (t/>=0.0000966 rads/in., for a drift
of 1.07%), the computed results (with confined concrete) indicate a compressive strain of
approximately 0.0075 and a 74.1 inch (188 cm) depth of compression zone (Table 3.1).
These results indicate that the simplified hand calculations provide conservative estimates
of the maximum expected compressive strain and the depth of the compression zone. The
configuration of the special transverse reinforcement in the web boundary zone is shown in
Fig. 3.6.
Shear-Strength Requirements:
Special attention should be placed on evaluating shear-strength requirements for Tshaped walls due to the influence of the cross-section on behavior. The required shearstrength should be based on the expected wall flexural capacity (and an assumed force
distribution over the height of the wall), as the wall shear stresses will be significantly
higher than those caused by code prescribed forces or the code prescribed forces as modified by Eq. Al 7. Flexural-strength can be estimated from the moment-curvature analysis
(using probable material properties) plotted in Fig. 3.9. The analysis indicates a probable
flexural-strength of approximately 600,000 kip-inches (67,800 kN-m) at the expected ulti-

32

1
11
il

1
1
1
il

1
1
1
il

1
1

1
1
il
1
1

1
1
1

mate curvature. A maximum expected shear force of 1,389 kips (6,178 kN) was calculated
assuming a uniform load distribution over the wall height. Table 3.2 lists the shear-strengths
and expected shear stresses for both prototype walls, and indicates that the expected shear
demand on the T-shaped wall is double that of the rectangular wall. Based on this estimate
of expected shear, the required shear reinforcement (based on ACI 318-89, Eq. 21-6) con-

sists of #5 {16 mm) hars at 7 inches {178 mm). This shear corresponds to a shear stress of

limit {Aktan and Bertero, 1985), but less than the 8J"JI recommended by UBC-91 (for walls

approximately 7.lJ"JI which is slightly greater than the

6J"JI recommended as an upper

sharing a common lateral force ). The shear stress in the wall could be reduced by increasing

1
1
1
1
1
1

1
1
1

the web thickness or lengthening the wall, or by reducing the wall's flexural-strength by
removing some of the longitudinal reinforcement at the web-flange intersection.

3.3.3 T-Shaped Prototype Wall (E-W Direction)


The amount of transverse reinforcement required at the boundaries of the flange is
governed by lateral forces acting in the E-W direction since flange compressive strains are
relatively small for loads acting in the N-S direction. To evaluate the required details at
the boundary of the fl.anges, the fl.ange can be considered as a rectangular wall; however,
the effects of the reinforcement in the wall web (perpendicular to the fl.ange) must be considered. If the reinforcement in the web (perpendicular to the fl.ange) is considered to be
effective tension reinforcement, the tension reinforcing ratio is increased from 0.00508 to
0.01846 and Eq. All results in a compressive strain of 0.00946 and a 97.9 in. (249 cm)
depth of compression zone. This compression zone depth indicates that the neutral axis
is located at the web; thus, the strains in the web reinforcement are small (approximately
zero) indicating that the web steel does not act as effective tension reinforcement. Altema-

tively, if the reinforcement in the web (perpendicular to the fl.ange) is considered to act as

1
1
1

from 0.00277 to 0.015 and Eq. All results in a compressive strain of 0.00539 and a 55.8

uniformly distributed web steel, the uniformly distributed steel reinforcing ratio is increased

33

1
in. (142 cm) depth of compression zone. These two approaches yield confl.icting results
indicating that the true solution lies somewhere between these two estimates; therefore, a
moment-curvature analysis was conducted to determine which approach provided the most
reasonable results. Table 3.1 lists the results from the moment-curvature analysis and one
of the analyses using Eq. All (web reinforcement acting as uniformly distributed steel).
Based on the moment-curvature analysis (Fig. 3.10) using probable material properties
with confined concrete (Saatcioglu and Razvi, 1992), the maximum compressive strain and
depth of compression zone were determined to be 0.007 and 72.9 in. (185 cm), respectively
(at <!>u=0.0000966 rads/inch; 0.000038 rads/cm). The computed results indicate that Eq.
A 11 may yield unconservative or overly conservative results depending on how the web
steel (perpendicular to the ftange) is incorporated into the equation; therefore, for these
cases, conducting a moment-curvature analysis is recommended. The computed estimate
of compressive strain indicated the need for special transverse reinforcement at the flange
boundaries over a depth of approximately 31.2 in. (79.2 cm). To accomodate this depth,
spacing of the longitudinal reinforcement in the boundaries was increased from 6 to 8 inches

1
1

1
1
:

1
1

1
1

(152 to 203 mm) and transverse reinforcement consisted of 2-#4 (13 mm) hoops and one
cross-tie at 5 inch (127 mm) spacings (Fig. 3.6).
Shear-strength was verified by comparing the nominal shear capacity of the section
against two different estimates of the maximum expected shear demand. A nominal shear
capacity of 899 kips (3,999 kN) was calculated for the T-shaped prototype, in the E-W
direction, using Eq. A16. The first estimate of the maximum expected shear was calculated using Eq. Al 7 and resulted in an expected shear force of 669 kips (2,976 kN) which
corresponds to a shear stress of 218 psi (3.44Jf:). The second estimate of the maximum
expected shear was based on the moment-curvature analysis (Fig. 3.10). At the expected
ultimate curvature (<l>u) of 0.0000966 rads/in., the plot indicates a flexural-strength of approximately 370,000 in-kips (41,810 kN-m). Assuming a uniformly distributed load pat tem
over -the height of the wall, this moment corresponds to a shear force of approximately 856

34

1
1
1

1
1
1
1

il

1
1
1

1
1
1
1
1

1
1
1
1
1
1

1
1

1
1
1

kips (3,807 kN) anda shear stress of 279 psi (4.41J'Fc). The nominal shear-strength of the
wall exceeds both estimates of maximum expected shear force and the expected shear stress
levels are well below the recommended limit of 6

J'Fc, indicating that the wall has adequate

shear-strength when loads act parallel to the fiange. The shear-strength requirements are
listed in Table 3.2.

3.3.4

General Remarks

The simplified design procedures used in the preceeding sections provide an effective
yet fiexible method for designing structural walls subjected to lateral and axial loads. The
results of the simplified hand calculations can be easily verified using sectional analysis
programs which are readily available. These simplified procedures yield conservative results
and provide an excellent method for the iterative task of designing structural walls.
These simplified procedures can be used as "stand alone" guidelines for conservative
design of structural walls or they can be used in conjunction with more detailed computer
analyses to yield more economical designs. Detailed computer analyses of the entire building
system are necessary to evaluate the contribution of all components of the lateral force
resisting system (structural walls, moment resisting frames, ete.), and they also provide
verification of member design forces and estimates of building displacements. The following
section describes a detailed 3-dimensional computer analysis of the prototype building and
compares the results with those of the simplified procedures described in the preceeding
sections.

3.4 3-D Dynamic Analysis of the Prototype Building


The SAP90 (CSI, 1990) computer program was used to conduct 3-dimensional static
and dynamic elastic analyses of the prototype building. The objectives of the analyses were:
( 1) to verify that. the simplified procedures discussed in previous sections yield conservative

35

1
results, (2) to evaluate forces acting on T-shapecl structural walls versus rectangular walls
(due to different stiffnesses), (3) to verify moment distribution over the height of the walls,
(4) to verify assumecl displacement profiles over the building height, (5) to study the effects
of cracking on building performance, and (6) to compare the results of static and dynamic
analyses. The UBC-91 equivalent lateral force procedure was usecl to calculate the lateral
loads usecl in the static analysis; whereas, the UBC-91 equivalent code spectrum for soil
type 2 was usecl for the dynamic response spectrum analyses (Rw=12 and gross-sections;

Rw=l and crackecl sections).


The prototype building was modelecl as a 3-D frame building with rigid diaphragms
at each story level. Preliminary dimensions of the beams and columns of the prototype
building (Fig 3.1) were based on gravity loads (175 psf dead loads; 40 and 20 psf live loads
for floors and roof, respectively) and tributary areas. Interior columns were 28 x 28 in. (686
x 686 mm); whereas, exterior columns were 24 x 24 in. (610 x 610 mm). Beam dimensions
variecl from 18 x 30 in. (457 x 762 mm) to 12 x 24 in. (305 x 610 mm) depending on the
beam's span and tributary area. The structural walls were modelecl as columns at the wall
centroid with section properties basecl on the prototype cross-sections (moments of inertia
were calculatecl for the rectangular and T-shapecl sections, respectively). Beams framing
into the structural walls were modelecl with rigid end offsets from the centroid of the wall
to the eclge of the wall. The results of the detailecl 3-dimensional analyses are describecl in
the remainder of this section which is subdividecl into two categories: (1) strength concerns,
and (2) displacement concerns.

Strength Concerns:
Design forces for structural members were calculatecl using three different analyses.
The first analysis involvecl hand calculations and the simplifiecl static lateral force proceclure
outlinecl in UBC-91, section 2334. The entire design base shear (assuming walls resistecl
100% of the lateral load) was then evenly distributecl among the six structural walls orientecl
in each principle direction (disregarding variations in wall stiffnesses) of the building. The

36

1
J
1
1
1
1
1
1
1
1

1
1
1
1
1
1
1

1
1

second analysis also incorporated the static lateral force procedure; however, the lateral

forces were used as design loads in a 3-D static computer analysis. The third analysis was

demands, the analysis was performed using a force reduction factor (Rw) of 12 (dual system),

1
1
1
1

1
1
1
1

1
1
1

a dynamic response spectrum computer analysis. For the purposes of evaluating strength

and ali structural members were modeled using gross-section properties. A comparison of
member forces calculated using the three approaches is presented in Table 3.3.
The results listed in Table 3.3 indicate that the UBC simplified analysis (assuming
equal distribution of force to ali walls) over-predicts the forces acting on the rectangular
walls and slightly under-predicts the forces acting on the T-shaped walls. Results from the
static computer analysis indicate that the six structural walls resist approximately 85% of
the total lateral force; whereas, the frames resist approximately 15%. Results also indicate
that each rectangular wall resists approximately 66% of the loads that each T-shaped wall
resists (due to variations in wall stiffnesses). Eventhough the predicted loads acting on the
T-shaped walls are slightly higher than the loads for which they were originally designed, the
critical combinations of axial load and moment stili plot within the design P-M interaction
diagram (Fig. 3.8), indicating that the walls possess adequate strength. The computer
analyses predict lower member forces acting on the rectangular walls; therefore, these load
cases also fall within the limits of the design P-M curve (Fig. 3.3).
The dynamic response spectrum analysis was performed using a force reduction factor
of 12 which indicates that the building is expected to experience inelastic behavior. The
member forces resulting from this analysis can be compared directly with the results of the
other analyses (UBC simplified and static computer analyses ). When a dynamic response
spectrum analysis is conducted, UBC-91 requires that a sufficient number of modes are
considered such that 90% of the total mass participates in the calculation of response for
each principal horizontal direction. Requirements also state that the base shear determined
using this procedure must be at least 90% of that calculated using the static lateral force

1
1
1

procedure. Adherence to these requirements was verified. The member design forces cal-

37

culated from the design spectrum analysis show good correlation with the forces calculated
using the static analyses.
The moment distribution over the height of the building was verified using results
from the static and dynamic (Rw=12) computer analyses.

Two moment profiles for a

rectangular wall are plotted in Fig. 3.11. The first profile corresponds to the static analysis
and indicates that the wall experiences reversed curvature at upper levels; whereas, the
profile based on the dynamic response spectrum analysis indicates single curvat ure over the
entire wall height. This discrepancy is caused by the dynamic response spectrum analysis
using absolute values; however, it is not of major concem because the strength provided
by code prescribed minimum amounts of steel, is considerably higher than the required
strength at this level. The plotted moment profiles indicate good correspondence with each
other and verify that the rectangular wall posesses adequate flexural-strength over its entire

1
1
1

1
1
1
1
1

height.
Ductility of structural members is also a major concern facing design engineers. An
estimate of the maximum expected displacement (or curvature) is needed so that adequate
ductility can be provided. Discussion of displacement estimates are presented in the following subsection.

Displacement Concerns:
Common design practice dictates that during moderate and severe earthquakes, buildings are expected to experience inelastic behavior, resulting in the need for adequate ductility which comes in the form of special detailing in critical regions. A conservative estimate of
maximum expected displacement is required to ensure that adequate detailing is provided.
The results of various analyses are discussed in the remainder of this section.

1
1

1
1
1
1

Simplified hand calculations, using the displacement-based design procedure of Appendix A, resulted in a fundamental building period estimate of 1.02 seconds anda maximum expected roof displacement of 9.23 inches (234 mm; 1.07% lateral drift ). The period

38

1
1
1

1
1
1
1
1

estimate was calculated assuming cracked section stiffness and the roof displacement estimate was approximated as 1.5 times the spectral displacement (to account for the difference
between the displacement ofa single degree of freedom oscillator and the building system the
oscillator represents). The displacement and period estimates of this analysis and various
dynamic response spectrum analyses are listed in Table 3.4.
Two dynamic response spectrum analyses were conducted to compare displacement

estimates with those calculated using Appendix A equations. Both dynamic computer

1
1

different combination of member stiffnesses to account for load induced cracking of concrete.

1
1
1
1
1

'1
1
1
1

analyses incorporated a force reduction factor (Rw) of 1; however, each analysis used a

The first dynamic analysis used gross section properties of all structural members (similar
to the analysis used for evaluating building strength). The results of this analysis represent
the expected building displacements if the building is expected to remain elastic at all
times (no cracking of concrete). During moderate to severe ground motions, the building is
expected to experience cracking of concrete which results in a decrease in building stiffness
and increased estimates of the fundamental period and the building displacements. To
account for these increases caused by cracking of concrete, the results of this analysis were
multiplied by

\/'2.

The second analysis accounted for cracking of concrete by using 50%

of the gross-section inertias ( 50% I 9 ) for all of the structural members. The results of all
dynamic analyses are listed in Table 3.4 to allow for comparisons.
Based on gross-section properties the maximum displacement estimate is only 6. 71
in. (17 cm); however, multiplying these results by

v'2 provides excellent

correlation with

the analysis based on 50% 19 , and the simple analysis using Appendix A. The procedures
outlined in A ppendix A provide an extremely simple yet accurate method for estimating
maximum expected building displacements. Due to the assumption that the building is
constructed with rigid diaphragms at each story level, the displacement profile of each
individual structural wall is equivalent to the displacment profile of the building. The
displacement profile ofa rectangular wall, based on the analysis using 50% 19 , is plotted in

39

1
1

Fig. 3.11.

General Remarks:
The prototype builcling and the prototype walls were designed and analyzed using
simplified hand calculations (Appendix A) and verified using a sectional analysis computer
program (BIAX). Detailed 3-D computer analyses of the entire building system indicate that
the simplified procedures provide designers with a simple yet effective method of estimating
required strength and expected building displacements, and the procedures tend to yield
conservative results. The four scale model specimens tested in this study were designed
using the simplified procedures and are described in detail in the following chapter.

1
1
1
1
1
1
1

1
1

1
1
1

1
1
1
40

1
1

1
1

Chapter - 4

1
1

1
1
1

1
1
1

1
1
1

DESCRIPTION OF MODEL WALLS

The rectangular and T-shaped prototype walls described in the preceed.ing chapter
aided in the design of the four, one-quarter scale, wall specimens tested in this study. The
same design procedures were employed for the model walls; however, some modifications
were made to demonstrate the flexibility of the proposed displacement-based design procedure.
Although the 3-D elastic computer analyses performed in the preceed.ing chapter indicated that the prototype walls could be designed for 1.07% drift, all four of the model
wall specimens were designed for a conservative upper-bound lateral roof drift estimate of
1.5%. This level was selected because, at lower drift levels, relatively small amounts of
transverse steel are need.ed such that once the walls were deformed beyond the design drift
level, buckling of longitudinal reinforcement would occur for all specimens. By selecting a
design drift level of 1.5%, requirements for concrete confi.nement and rebar buckling could
be investigated jointly.
Of the four wall specimens tested in this study, specimens #1 (RWl) and #3 (RW2)
had rectangular cross-sections whereas specimens #2 (TWl) and #4 (TW2) were T-shaped.
The one-quarter scale was chosen based on the capacity of available hydraulic actuators
and height limitations in the structures lab at Clarkson University. Each rectangular wall

1
1
1
1

specimen was constructed asa 12 foot (3.66 m) tall cantilever with a cross-section of 4 x 48
in. (102 mm x 1.22 m; Fig. 4.1). T-shaped walls had identical flange and web dimensions
(Fig. 4.2). The model walls included only the bottom four stories (3 foot story heights)
because lateral load was applied by a single hydraulic actuator acting at two-thirds of the
height of the scaled down walls (the resultant of the triangular load acting over the wall

41

height occurs at the top of the fourth story). The height and length of the scale model walls
result in an aspect ratio (hw/lw) anda moment to shear span ratio (Mu/Vulw) of 3.
Figures 4. 1 and 4.2 show 3-dimensional views of the one-quarter scale rectangular and
T-shaped walls, respectively. For clarity, Fig. 4.2 does not include the slabs that were
placed at each fl.oor level of the T-shaped walls (the design and details of the typical fl.oor
slab are presented in the following chapter). The remainder of this chapter is dedicated to
the design and analysis of the four model wall specimens.

4.1

Specimen

RWl

Specimen RWl closely resembles a one-quarter scale model of the rectangular prototype
wall described in the preceeding chapter (concrete dimensions are exactly one-quarter scale
and reinforcing ratios are similar).

Strength Reguirements:
Longitudinal reinforcement was comprised of 8-#3 (9.5 mm) deformed hars (Grade
60, 414 MPa) placed in each boundary zone resulting in tension and compression reinforcing ratios (p=A 8 /twlw and p' = A~/twlw) of 0.00458. All longitudinal reinforcement was
continuous over the wall height and had 90-degree hooks that were embedded into a large
pedestal which provided fixity at the base of the wall (Fig. 4.3). It was realized that in
actual construction there would be lap splices over the wall height; however, spacing limitations within the cross-section made splices difficult to construct and their exclusion was
not critical to the overall objectives of the research.
Uniformly distributed web steel was provided based upon code prescribed minimums
and consisted of two curtains of #2 (6.4 mm) deformed hars spaced at 7.5 in. (191 mm) in
each direction (Figs. 4.4 through 4.6). This resulted in a reinforcing ratio (p" = A:ftwlw)
of 0.00327. It is noted that horizontal web steel was terminated in the boundary zone with

1
1
1
1
1
1
1
1

1
1
1
1
1
1
1

a 90 degree hook as opposed to UBC-91 practice of developing the hooked bar within the

42

1
1

1
1
1
1
1
f
1
1

boundary zone. Table 4. 1 lists the reinforcing details and design parameters for all four
model specimens.

Detailing at Wall Boundaries:


The transverse reinforcement for the scale model walls was comprised of 3/16 inch (4. 75
mm) diameter smooth wire (A=0.0276 in2 , 17.8 mm 2 ) that was annealed to yield material
properties similar to those of Grade 60 (414 MPa) reinforcing steel (material properties are
discussed in detail in Chapter 5 ).
Although transverse reinforcement at the boundaries of the model walls could have
been extrapolated from the prototype wall designs, the design process was repeated for
the model walls due to slight differences between the prototype and model specimens. A
modified version of Eq. A9 (Wallace and Moehle, 1992) was usep. to calculate the ultimate

1
1
1
1
1
1
1
1
1

1
1

curvature demand on the model walls. The modified equation replaces the 1/2 multiplier on
the wall aspect ratio with 2/3 to account for the different elastic curvature distribution over
the wall height caused by a single lateral load at the top of the wall (such as used during the
experimental portion of this study) as compared with an inverse triangular lateral loading.
Based on a maximum roof drift estimate of 1.5%, andan aspect ratio (hw/lw) of 3.0, the
modified version of Eq. A9 results in an ultimate curvature demand on the wall of <i>u =

0.0275/lw. Recalling that lw = 48 in. (122 mm), the ultimate curvature demand is

<>. =

0.00057 radians/inch (0.000022 rads/mm).


Based on the roof drift and ultimate curvature estimates, the maximum expected concrete compressive strain at the extreme fiber was computed using Eq. All; where, P

0.10A 9 J;,

=4

ksi (27.6 MPa), f,11

= 60

ksi (414 MPa), and the reinforcing ratios are

listed in Table 4. 1. Substitution of these values into Eq. All resulted in a maximum
compressive strain of 0.00603. Since the expected compressive strain at the extreme fiber
exceeded 0.004, special transverse reinforcement was needed for concrete confinement and
to suppress buckling of the longitudinal reinforcement. The transverse reinforcement was
evaluated using the strain profile plotted in Fig. 4. 7. The strain profile indicates that the

43

portion of the wall where strains exceed 0.004 is 3.56 inches (90 mm); however, a minimum
depth of confinement of lw/10 = 4.8 inches (122 mm) is recommended (Wallace, 1995).
Transverse reinforcement for specimen RWl was governed by limitations set forth to
control buckling of the longitudinal reinforcement and consisted of one hoop and two crossties ata spacing of 3 inches (76 mm; 8db). The special transverse reinforcement was provided
over the bottom 4 ft. ( 1.22 m; lw) of the wall which represented an upper bound estimate
of the plastic hinge length (Paulay, 1986). Above the potential plastic hinge zone (l,,

= 4ft;

1
1
1
1
1

1.22 m), minimum transverse reinforcement was used and consisted of "U-shaped" stirrups
spliced with the horizontally distributed web reinforcement (Figs 4.4 through 4.6).
Figure 4.3 shows a profile view of specimen RWl, and Fig. 4.4 shows the cross-sections
of specimen RWl at the three distinct levels indicated in Fig. 4.3. The first level (section
A-A) is the well detailed section at the base of the wall where inelastic deformations are
expected to concentrate (Fig. 4.5 shows a photograph of the reinforcement at this level).
The second level (section B-B) is just above the well detailed region where the special
transverse reinforcement is no longer required (photograph in Fig. 4.6), and the third level
(section C-C) is located at the top of the wall specimen where the top assembly is attached.

1
1
1

1
1

The amount of uniformly distributed web steel (and U-shaped stirrups) was doubled at the
fourth story level (section C-C) to help prevent any local distress due to load transfer from
the act uator to the wall specimens. Details for the other specimens at this level are similar
and therefore are not discussed in subsequent sections.

Verification Using Sectional 'Analyses:


The sectional analysis program BIAX (Wallace, 1992) was used to verify the results
of the displacement-based design procedure and to predict expected wall behavior. Figure
4.8 plots the results of three separate moment-curvature analyses of specimen RWL The
first analysis was performed assuming that all concrete was unconfined; whereas, the other
two analyses incorporate two different stress-strain relationships for confined concrete (the
Modified Kent-Park relationship recommended. by Park, Priestley and Gill, 1982; and the

44

1
1
1
1

1
1
1

1
1
1
1
1
1

relationship recommended by Saatcioglu and Razvi, 1992}. All of the analyses were performed using a pro bable steel stress-strain curve to account for material over-strength and
strain hardening (/y = 65 ksi, 448 MPa; f u

= 90 ksi, 621

MPa).

Figure 4.8 includes aline indicating the maximum expected curvature demand on the
wall (tl>u=0.00057 rads/in, 0.000022 rads/mm). The plot clearly indicates that special transverse reinforcement is required at the boundaries of the rectangular wall and that the confined concrete stress-strain relationship proposed by Saatcioglu and Razvi predicts slightly

higher deformation capacity than does the Modified Kent-Park relationship. It is noted that

1
1
1

of longitudinal steel and hoop spacing, as well as the different confinement effects in the two

1
1

the relationship recommended by Saatcioglu and Razvi (1992) accounts for the distribution

directions of the confined rectangular boundary zone. The relationship has been verified
using experimental data for circular, square and rectangular columns, confined with spirals and/or rectilinear transverse reinforcement, subjected to eccentric and concentric axial
loads. Given these attributes, the results using this relationship were considered to be more
accurate and thus used for design purposes. Based on the plot for confined concrete (using
relation recommended by Saatcioglu and Razvi}, the ultimate curvature corresponds to a
maximum compressive strain of 0.005, and a 8.8 in. (224 mm) depth of compression zone.
These values are listed in Table 4.1 to compare with the results obtained using Appendix
A equations and indicate the simplified design approach tends to yield conservative results.
Figure 4.8 alsa indicates that buckling of longitudinal reinforcement is expected to occur at

a curvature of approximately 0.0008 rads/in (based on a hoop and tie spacing of 3 inches,

for the reinforcing steel. Analytical predictions of longitudinal reinforcement buckling are

1
1

1
1

8db} This estimate was calculated using Eq. A13 and a probable stress-strain relationship

described in detail in Chapter 7.


The moment-curvature analyses presented in Fig. 4.8 were alsa used to check shear
demands on the rectangular wall.

At the ultimate expected curvature, the wall has a

flexural-strength of approximately 4,500 in-kips (509 kN-m). Based on a cantilever wall

45

1
height of 144 in. (366 cm), this flexural capacity corresponds to an expected shear force of
approximately 32.6 kips (145 kN). Use of Eq. A16 results in a nominal shear capacity of
62 kips (276 kN) indicating adequate shear-strength. The maximum expected shear force
corresponds to an expected shear stress of approximately 2.69v'JI psi (0.22v'JI MPa) which
is less than the recommended limit of

6v'JI psi (0.5v'JI MPa; Aktan and Bertero, 1985).

Table 4.2 summarizes shear-strength requirements for the model wall specimens.

Deformation Dependent Shear-Strength:


Previous studies of RC columns (Aschheim and Moehle, 1992; Priestley et al., 1994)
and high-strength RC shear walls (Kabeyasawa et al., 1994) have shown that shear-strength
of these structural members decreases as the displacement ductility demand increases. Currently, a relationship of this type, for normal strength RC shear walls, does not exist; however, it could be valuable during design to ensure that adequate shear-strength is provided
hased on an estimate of the expected displacement ductility demand. For the specimens
tested in this study, it is anticipated that the shear force corresponding to flexural-strength
will be less than the residual shear-strength (Priestley et al., 1994), thus ensuring a ductile
flexural response. The maximum expected curvature demand of 0.00057 rads/in (1.5% lateral drift) corresponds to a curvature ductility (tf,) of approximately 8, and results in a displacement ductility (c5) of approximately 4.5. At the design drift level, the maximum shear
stress is only about 3v'JI psi (0.25v'JI MPa) ; therefore, degradation of shear-strength is

1
1

1
1
1
1

1
1
1
1

not anticipated, as might be the case for walls with lower aspect ratios.

4.2

Specimen

RW2:

Specimen RW2 has identical longitudinal and uniformly distributed web reinforcement
as RWl (Table 4.1); however, the transverse reinforcement consists of single hoops (no
cross-ties) at a spacing of 2 inches (51 mm). According to ACI 318-89, equations 21-3 and
21-4, this configuration results in approximately the same value of A.,h as the configuration
used for specimen RWl (spacing of steel has decreased but so has the quantity of steel).

46

1
1
1

1
1

1
1
1
1
1

1
1
1
1
1

The closer spacing of the hoops was selected to suppress buckling of the longitudinal steel;
therefore, concrete confinement (crushing of confined core) was expected to control wall
behavior. Figures 4.9 A&B show the cross-section anda photograph of specimen RW2 at
the base level where the special transverse reinforcement is located (compared to Figs. 4.4A
and 4.5 for specimen RWl). It is noted that removal of the. cross-ties from the boundary
zones violates ACI 318-89 requirements (Section 21.4.4.3) for spacing between legs of hoops
or ties. Reinforcing details at sections B-B and C-C are identical to those of specimen RWl
(Fig. 4.4b,c ).

Sectional Analyses:
The cross-section of specimen RW2 is identical to that of specimen RW 1 except for a
slight change in the configuration of the transverse reinforcement in the boundary elements.
The change in configuration of the transverse reinforcement has a negligible effect on ali
three moment-curvature analyses performed for specimen RWl (Fig.

4.8); except that

buckling of longitudinal boundary hars in specimen RW2 is expected to occur at an ultimate


curvature in excess of 0.002 rads/in. The estimates of compressive strain and depth of

1
1
1
1
1
1
1
1
1

compression zone for specimen RW2 are identical to those of RWl (Table 4.1), as well as
the shear-strength requirements (Table 4.2).

4.3

Specimen TWl:
As discussed previously, in Chapter 1, current codes (UBC-91) do not give recommen-

dations for the design of "fianged" walls. One possible design philosophy involves designing
the web and fiange of T-shaped walls independently and simply joining them together.
Based on this scenario, the T-shaped wall would simply be two rectangular walls joined
together. This philosophy was utilized for the design of specimen TWl, eventhough poor
performance is likely as discussed in the following paragraphs.
The flange and web of specimen TW 1 had reinforcing details that were identical to
those of specimen RWL Figures 4. 10 and 4. 11 show a schematic and a photograph of the

47

1
reinforcing details at the base of specimen TWl (section A-A), respectively. Figure 4. 12
shows a close-up photograph of the web boundary element of specimen TW 1.
Figure 4. 13 plots the results of three moment-curvature analyses performed on specimen
TWl (material properties identical to those used in analysis of specimen RWl). The first
analysis was performed assuming that the flange was in compression and that all concrete
was unconfined; whereas, the remaining two analyses were performed assuming that the
flange was in tension and that the web boundary element was confined as shown in Figure
4.12.
It can be seen from the first analysis that the wall possesses adequate strength and

ductility when the flange is in compression even without the presence of special transverse
reinforcement in the flange. The computed results indicate a maximum compressive strain
estimate of 0.0014 and a 2.5 in. (64 mm) depth of compression zone. These results indicate that the transverse reinforcement provided in the flange-web intersection and the
flange boundaries could be relaxed (designed based on recommended minimums); however,
it should be noted that the transverse reinforcement located in the flange boundaries is
typically governed by loads acting parallel to the flange, whereas only loads acting parallel
to the web are considered in this study.
Two analyses were performed assuming that the flange was in tension and that the
web boundary element was confined as shown in Figs. 4. 10 through 4. 12. Each analysis incorporated a different stress-strain relationship for confined concrete; however, Fig.
4.13 indicates the likelihood of poor wall behavior at the expected maximum curvature,
regardless of which stress-strain relationship was used. These analyses indicate the major
behavioral differences between the flange being in tension (critical case) and compression,

1
1
1
1
1
1
1
1
1
1
1

1
1
1

and also indicate that special detailing is more likely for the web of T-shaped cross-sections
(TW 1) compared with similarly reinforced rectangular walls (RW 1).
Based on Fig. 4.13, the wall hasa maximum flexural-strength of approximately 9,600
in-kips (1,085 kN-m). This flexural capacity corresponds to an expected shear force of

48

1
1
1
1

'I
1
1

1
1
1
1
1
1
1

1
1

1
1
1
1
1
1
1

approximately 68 kips (302 kN; assuming a 12 faot loading height), which exceeds the
nominal shear-strength of 62 kips (276kN; Table 4.2). This indicates that shear may have
a significant influence on specimen behavior.
Eventhough all of the preceeding analyses indicate the likelihood of poor wall behavior
(due to the inadequacies of designing the web and flange independently), specimen TWl
was constructed in this manner to emphasize the behavioral differences between symmetrical (rectangular) and unsymmetrical (T-shaped) wall cross-sections and to illustrate the
consequences of a poor conceptual design. The fallowing section describes the design of
specimen TW2 which accounts far the anticipated behavior of the T-shaped wall which
were purposely ignored during the design of specimen TWl.

4.4 Specimen TW2:


Specimen TW2 has the same cross-sectional dimensions as TW 1; however, the reinfarcement was modified based upon preliminary analyses of the previous specimens and
the displacement-based design procedure described in Appendix A. Figures 4.14 and 4.15
show a schematic and a photograph of the reinfarcing steel at the base of specimen TW2,
respectively. Figure 4.16 shows a close-up photo of the web boundary element far TW2
(compared to Fig. 4. 12 for specimen TWI ).
When the flange of the T-wall is in compression, extremely small concrete compressive
strains are expected; therefore, no special transverse reinforcement is needed for concrete
confinement in the flange. Minimal amounts of transverse reinforcement were provided at
the web-flange intersection and at the flange boundaries and consisted of hoops without
cross-ties at a spacing of 4 inches (102 mm; 10. 7db; width of wall). It is noted that confinement at the flange boundary zones is likely ta be controlled by lateral forces acting parallel
to the flange; however, minimal confinement was provided based on the loads acting parallel
to the web.
49

1
The critical condition occurs when the flange is in tension and the web must develop a
very large compressive force to equilibrate the large tensile forces that develop in the flange
reinforcement (see discussion in Section 3.3.2). Equation All was used to estimate the
maximum expected compressive strain at the extreme fiber and the depth of the compression
zone, assuming that all steel in the flange is effective as tension reinforcement. Based on
an axial load of O. lOAgJ; and using the reinforcing ratios listed in Table 4. 1, a compressive
strain of 0.0137 was expected at the boundary of the web. The depth of the compression
zone was calculated to be 24.04 in. (611 mm). Based on the strain profile plotted in Fig.
4.17, the length of compression zone where strains exceed 0.004 is 17.02 in. (432 mm) which
is greater than the minimum required length of lw/10 = 4.8 in. (122 mm). To facilitate
confinement over this depth, the spacing of the 8-#3 (9.5 mm) longitudinal hars at the web
boundary was increased from 2 in. to 4 in. (51 to 102 mm) and 2 additional #2 (6.4 mm)
longitudinal hars were added to the boundary zone. Transverse reinforcement consisted of
two hoops and one cross-tie (Fig. 4.14). The hoops nearest the free end of the web enclosed
6-#3 (9.5 mm) longitudinal hars and were spaced at 1.25 in. (32 mm). The second hoop
and cross-tie enclosed the 6-#3 (9.5 mm) and 2-#2 (6.4 mm) longitudinal hars that were set
back from the end of the wall and were spaced at 1.5 inches (38 mm). The larger spacing of
these hoops was based on the lower values of compressive strain expected at this location.
[The transverse reinforcement in specimen TWl hada 3 in. (76 mm; 8db) vertical spacing
and was only provided over a 6. 75 in. ( 171 mm) depth at the free end of the web indicating
insufficient depth and spacing of confinement.]
The amount of uniformly distributed web steel in specimen TW2 was increased to
account for the higher shear demands associated with an increase in flexural-strength (due
to flange steel acting as tensile reinforcement). The spacing of the #2 (6.4 mm) web hars
was decreased from 7.5 to 5.5 in. (191 to 140 mm) in both the horizontal and vertical
directions, yielding a reinforcing ratio (p"

A~/twlw) of 0.00445 (Table 4.1).

Figure 4. 18 plots the results of three moment-curvature analyses performed on specimen

50

1
1
1
1
1
1

1
1

1
1
1
1
1
1
1

1
1
1

TW2. The :6.rst analysis was performed assuming that the flange was in compression and the
results are similar to those obtained for specimen TWl (Fig. 4.13). The other two analyses

were performed assuming that the flange was in tension and that the web boundary zone

1
1
1

confined concrete predicts good behavior at the expected maximum curvature; whereas, the

was confined. The analysis using the Saatcioglu and Razvi (1992) stress-strain relation for

analysis using the Modified Kent-Park relation predicts a complete loss of strength prior
to the expected maximum curvature. The results using the Saatcioglu and Razvi relation
are considered to be more reliable than those using the Modified Kent-Park relation, as
discussed in the preceeding section. Additional analyses described in Chapter 7 will verify
this conclusion.

1
1
1
1
1

Based on Fig. 4. 18, the wall has a maximum flexural-strength of approximately 9,500
in-kips (1,074 kN-m) at the expected ultimate curvature. This flexural capacity corresponds
to an expected shear force of approximately 69.4 kips (309 kN). Due to the decreased
spacing of uniformly distributed web steel, specimen TW2 hasa nominal shear capacity of
75.6 kips (336 kN) which exceeds the maximum expected shear force (Table 4.2), indicating
adequate shear-strength of the section. The maximum expected shear force corresponds to
an expected shear stress of approximately 5. 72,J"'H psi (0.4 7

J"H MPa) which is less than

the recommended limit of 6,J"'H psi (0.5,J"'H MPa) recommended by Aktan and Bertero
(1985).

1
1
1
1
1
1
1

4.5

General Remarks:
The design of four quarter-scale model wall specimens was described in the preceeding

sections. One of these specimens (TWl) was purposely designed poorly to illustrate differences between the behavior of symmetrical and unsymmetrical cross-sections and to show
the results of poor conceptual design practice (designing web and flange independently and
simply joining them together ). The remaining specimens were designed using a combination
of the displacement-based design procedure recommended by Wallace (Appendix A) and the

51

1
1
computer sectional analysis program BIAX. Behavior of these three walls is expected to be
dominated by fl.exure and indicate stable, hysteretic response. The following two chapters
describe the experimental program and present the experimental results, respectively.

1
1
1

1
1

1
1

1
1
1

1
1
1

1
52

1
1

1
1
Chapter - 5

1
1
1

1
1
1
1
1

1
1

1
1

1
1
1
1
1

DESCRIPTION OF EXPERIMENTAL PROGRAM

This chapter gives detailed descriptions of: 1) construction materials, 2) specimen


construction, 3) testing apparatus, 4) instrumentation and data acquisition, and 5) testing
proced ures.

5.1

Materials
A total of four wall specimens were constructed and tested. Two specimens had rectan-

gular cross-sections (RWl and RW2) and two had T-shaped cross-sections (TWl and TW2).
The design concrete compressive strength was 4,000 psi (27.6 MPa); however, strengths at
the time of testing ranged from 4,156 to 8,462 psi (28. 7 to 58.4 MPa), with a mean compressive strength at the first story height of 5,471 psi (37. 7 MPa). Tahle 5. 1 lists the concrete
compressive strengths at the fi.ve distinct levels of each specimen at 28 days, and at the
testing date. Figure 5. 1 plots the measured concrete stress-strain relationships for the four
wall specimens at their respective testing dates. The mix design proportions for one cuhic
yard of concrete are listed in Tahle 5.2.
Three different types of reinforcing steel were used in this study: (1) typical Grade 60
(414 MPa), deformed #3 (9.5 mm) hars as longitudinal reinforcement, (2) deformed #2
(6.4 mm) hars as uniformly distrihuted weh steel, and (3) 3/16 inch (4.75 mm) diameter
smooth wire as transverse reinforcement in the houndary regions. Figure 5.2 plots the
measured stress-strain relations for each type of reinforcement. The plotted stress-strain
curves terminate at the point when the strain gages hroke. Tahle 5.3 lists the mechanical
properties of the three different types of reinforcement used in this study. Ultimate strains

53

were calculated by dividing the elongation of the steel by known gage lengths. The #3 (9.5
mm) hars used as longitudinal boundary steel had typical Grade 60 properties; however,
the other two types of steel exhibited some atypical behavior.

The deformed #2 (6.4

mm) reinforcing hars used as uniformly distributed web reinforcement hada yield stress of
approximately 65 ksi (448 MPa); however, they did not exhibit a well defined yield plateau.
The 3/ 16 inch (4. 75 mm) diameter smooth wire had to be annealed to lower the yield stress
from approximately 80 ksi (552 MPa) to approximately 65 ksi (448 MPa); however, the
strain hardening region of the annealed wire is relatively ftat (Fig. 5.2).
All of the #3 hars and #2 hars used in this study were taken from the same heat;
therefore, variations of material properties were eliminated. All of the 3/16 in. (4. 75 mm)
diameter smooth wire was taken from a single roll; however, it was annealed in two batches.
The annealing process consisted of heating the steel from room temperature (70F; 21 C) to
l,165F (629C) over a period of approximately 2.5 hours and then allowing the steel to cool
slowly back to room temperature over a period of several hours. This annealing procedure
was used for each batch; therefore, variations in stress-strain characteristics between batches
were found to be negligible.

5.2

Construction Procedure
The test specimens were constructed in fi.ve different lifts; one for the pedestal (base-

block) and one each for the four different story levels. In general, the four model walls
were constructed in pairs (concrete at each story level placed on same day). The first pair
that was constructed consisted of specimens RW 1 and TW 1. Construction of the second
pair, consisting of specimens RW2 and TW2, started after completion of the first pair.
A chronology of specimen construction and testing is listed in Table 5.4. The following
sections give a detailed description of the typical procedures used during construction of
the model wall specimens.

1
1
1

1
1
1

1
1

1
1

1
1
1

54

1
1

1
1
1

5.2.1

Pedestal Construction

The fi.rst step of the construction process was to construct the pedestal (base-block)

1
1

for the specimen. This pedestal provided a stiff foundation for the specimen that could
be rigidly anchored to the strong floor in the laboratory. The pedestal also provided a
means of anchoring the post-tensioning cables that were used to apply axial stress to the

specimens during testing. The pedestals for the two rectangular walls measured 76 x 16

1
1

walls were also 27 in. (0.69 m) deep; however, the web was 24 x 66 in. (0.61 x 1.68 m)

1
1

in. (1.93 x 0.41 m) in plan and were 27 in. (0.69 m) deep. The pedestals for the T-shaped

and the flange was 16 x 60 in. (0.41 x 1.52 m) in plan. All longitudinal reinforcement was
firmly anchored into the pedestal between 10 and 16 in. (254 to 406 mm) and terminated
with a 90 degree hook. The reinforcing cages for the pedestals were designed to resist the
shear forces and bending moments induced by the ultimate loads developed at the base of
the wall specimens. A conservative approach was used to design the pedestals to prevent
excessive cracking of the concrete and to prevent possible pullout failures of the longitudinal
wall reinforcement. Photographs of the pedestal reinforcing cages for specimens RWl and

1
1
1

1
1
1
1
1

TW2 are shown in Figures 5.3 A&B, respectively. Pedestal reinforcing cages for specimens
RW2 and TW 1 were identical to those shown for RW 1 and TW2, respectively.
The formwork for the pedestals was constructed of 2x4's and 1/2 in. (13 mm) plywood.
The formwork was oiled prior to placing the concrete to allow for easy formwork removal.
Plywood templates were used to hold all reinforcement in place during the casting of the
concrete. Strain gages located near the base of the wall were mounted on the reinforcing
steel and waterproofed prior to casting the pedestal. Lead wires were soldered to the gages
and led out of the formwork through small holes drilled in the plywood. Eight pieces of
1-1/2 in. (38 mm) diameter PVC pipe were cast in the pedestal to allow high-strength
steel tie-down rods to pass through the pedestal into the strong floor. Two pieces of 3/ 4 in.
(19 mm) diameter PVC were used to allow the 0.6 in. (15 mm) diameter post-tensioning
cables to pass through the pedestal and be anchored (four pieces were used for the T-shaped

55

1
specimens). The post-tensioning cables were anchored in the bottom of the pedestal with a
reusable set of chucks anda 3/8 in. (9.5 mm) thick steel plate to help transfer the bearing
load. Figure 5.4 shows a photograph of the pedestal formwork and templates for specimen
TWl, just prior to concrete placement.
Concrete with a design compressive strength of 4,000 psi (27.6 MPa) was ordered from
a ready-mix plant. The wall specimens were cast in an upright position and construction
joints were cleaned and roughed up to allow good bond between story levels. Concrete
was finished and covered with wet burlap and plastic for approximately one week after
placement. The formwork was typically stripped between four days and seven days after
concrete placement.

5.2.2 Construction of Four Story Levels


nce the pedestal was finished, construction of the wall began.. The first step was
to tie all uniformly distributed horizontal web steel and transverse boundary steel to the
longitudinal steel that was already in place.

Strain gages were installed at all critical

locations and lead wires were led out of the formwork through small holes in the plywood.
The four story levels were constructed using "slip" formwork that could easily be stripped
and reused on the next story. The rectangular specimens were constructed withou t fl.oor
slabs; whereas, the T-shaped wall specimens were constructed with a slab at every fl.oor
level. The formwork was always checked for plumbness and braced prior to casting. The
concrete was allowed to cure for approximately five days before the forms were stripped and
the next level was started. The same procedure was followed for all four story levels.
Figures 5. 5 through 5. 7 show photographs of specimen RW 1 during various stages of
construction. Figures 5.8 through 5.12 show photographs of specimen TWl during various
stages of construction. The construction procedures used for specimens RW2 and TW2 were
identical to those used for specimens RWl and TWl, respectively. The testing apparatus,
instrumentation and testing procedure are described in the following sections.

56

1
1

1
1

1
1

1
1

1
1
1
1
1

1
1

1
1
1

5.3

search Laboratory (SERL) at Clarkson University. Figures 5.13 and 5.14 show a schematic

1
1
1
1
1
1
1

Testing Apparatus
The wall specimens were tested in an upright position in the Structural Engineering Re-

and a photograph of the overall test setup. The walls were rigidly mounted to the strong
floor through the use of eight 1-1/4 in. (32 mm) diameter high-strength steel tie-down rods.
The rods passed through the specimen's pedestal (PVC pipes were used to provide holes in
the pedestals) and were threaded into 2H nuts (the nuts were welded to a Cl5x33.9 steel
section) that are permanently embedded approximately 40 in. (1.02 m) into the laboratory's strong floor. Nuts and 1/2 in. (13mm) thick bearing plates were used at the top of
the tie-down rods to anchor the pedestal firmly to the floor. A layer of grout was placed
between the pedestal and the strong floor (and between the bearing plates and the top of
the pedestal) to ensure uniform contact and even load distribution.
A specially fabricated steel load transfer assembly was used to transfer both axial and
lateral loads to the wall specimen. (The assembly is similar to that used by Ali and Wight,
1990). The steel assembly was made of back-to-back channel sections (C-12x68's) welded
together with a O. 75 in. (19 mm) gap between them to allow the post-tensioning strands
to pass through. The load transfer assembly was attached to the wall specimen using 1
in. (25 mm) diameter threaded rods that were cast integrally with the fourth story of each

specimen. Six threaded rods were used to attach the assembly to the rectangular walls

1
1
1

transfer assembly and the top of the wall to evenly distribute the axial load applied by

whereas nine were used on the T-walls. A thin layer of grout was placed between the load

the hydraulic jacks. Figures 5. 15 and 5. 16 show photographs of the load transfer assembly
mounted on the top of specimen RWl, from aside view andan end view, respectively.
Axial stress was applied by hydraulic jacks mounted on top of the load transfer assembly
(Fig. 5. 15). Post-tensioning strands were anchored in the specimen's pedestal and run up
through the back-to-back channels and through the hollow jacks. A set of reusable chucks

was placed above the hydraulic jack to grip the cable and the jacks could then tension the

57

1
1
cables placing a compressive stress on the wall. Two jacks were used to apply axial stress
to the rectangular walls whereas four jacks were used on the T-shaped walls. A constant
axial stress of approximately O. 10!~ was maintained throughout the duration of each test.
The wall axial stress ranged from 429 to 470 psi (2.96 to 3.24 MPa), depending upon the
specimen's compressive strength at the time of testing.
Cyclic lateral loads were applied to the walls by a 125 kip {556 kN) hydraulic actuator
that was mounted horizontally to a reaction wall approximately 15 feet (4.6 m) above the
strong floor. A 1-1/2 inch (38 mm) thick steel plate was welded on one end of the load
transfer assembly {Fig. 5.15), to allow for attachment of the actuator to the specimen. A
large steel device was bolted to the plate and the actuator was than attached to the device
using a frictionless steel pin. Both ends of the actuator used a pin connection that allowed
the actuator to rotate freely (in the vertical plane) throughout the tests. Figure 5.17 shows
a photograph of the hydraulic actuator mounted at the top of specimen RWL Hydraulic
fluid was supplied to the actuator by an electric pump; whereas, the hydraulic jacks used
for axial load were adjusted through the use of hand pumps. The electronic equipment used
to control the tests is described in detail in the following section.
Out-of-plane support was provided during each test to provide stability to the wall
specimen. A steel truss was placed between the top of each wall specimen and the reaction
wall parallel to the specimen's web. The truss consisted of two channel sections lying flat
with small angles acting as diagonal members to keep the channels parallel. The webs
of the channels were bolted directly to the load transfer assembly {Figure 5.18).

The

other ends of the channels were attached to the reaction wall using a tube-on-tube sliding
connection {Figure 5.19) that allowed free movement parallel to the load direction but
prevented perpendicular motion. In this manner, twisting of the wall was eliminated and
out-of-plane restraint was provided; however, the wall was free to move in the direction of
the applied lateral load, and very little vertical restraint was provided {since the 10 foot
long channel sections were laid flat ).

58

1
1
1
1

1
1
1

1
1
1

1
1
1

1
1

5.4

tions in each wall specimen. Wire potentiometers (wire pots), linear potentiometers (linear

Instrumentation and Data Acquisition


lnstrumentation was used. to measure displacements, loads, and strains at critical loca-

pots) and linear variable differential transducers (LVDT's) were used. to measure displacement response. Load cells were used. to measure axial and lateral loads, and strain gages
were used. to measure strains in the concrete and the reinforcing steel. Figures 5.20 through

1
1

5.31 show photographs of the various types of instrumentation that are discussed. in the
following paragraphs.
A rigid steel reference frame was mounted. to the strong floor at the end of the wall
specimen to provide a location for instrumentation to be attached.. Figure 5.20 shows a
photograph of the four wire potentiometers that were mounted. to the reference frame to

measure lateral displacements at each floor level over the height of the specimen. A linear

not sliding as a rigid body along the strong floor (Fig. 5.21). Two additional linear poten-

potentiometer was also mounted. horizontally on the ped.estal to ensure that the wall was

tiometers were mounted. vertically at each end of the ped.estal to measure ped.estal rotation

1
1

1
1
1
1

1
1

caused. by uplift of the ped.estal from the floor (Fig. 5.21). The two wire potentiometers
mounted. at the top two stories of the wall specimens hada stroke of ten inches ( 5 in.,
127 mm); whereas, all other wire potentiometers had strokes of five inches ( 2.5 in., 64
mm). All of the linear potentiometers hada stroke of 4 inches ( 2 in., 51 mm).
Shear deformations were measured. through the use of wire potentiometers mounted on
the bottom two stories of each specimen. Two wire pots were mounted. on each story in an
'X' configuration. The wire pots were bolted to small steel angles that were epoxied. to the
walls. Small hooks were epoxied on the wall at the opposite corner of the story and attached.
to the wire pots with lead wires. Figures 5.22 A&B show photographs of the wire potentiometers used. to measure shear deformations on specimens RWl and TWl, respectively.
ldentical configurations were used. for specimens RW2 and TW2. The relationships used. to
compute the experimental shear deformations (from the displacement measurements) are

59

discussed in detail in Chapter 6.


Axial displacements at the wall boundaries were measured using two wire potentiometers mounted directly to the wall ends. The potentiometers were bolted to small steel angles
that were epoxied to the wall boundaries 30 inches (76 cm) above the pedestal. Wall base
rotations were calculated by dividing the difference in relative axial displacements, measured by the wire potentiometers, by the distance between the wire potentiometers ( 52 in.,
1.32 m). The measured base rotations include the effects of slippage of the longitudinal
reinforcement as well as the flexural deformations over the bottom 30 inches (76 cm) of the
wall. Figure 5.23 shows a photograph of the instrumentation located on the first story level
of specimen RWL The two wire potentiometers used to measure axial displacements at the
wall boundaries can be seen in the :figure.
Axial displacements were also measured along the length of each wall specimen using
linear variable differential transducers (LVDT's). The LVDT's were mounted vertically,
over a known gage length, at various positions along the length of each wall so that axial
strains could be determined. The LVDT's are extremely sensitive, accurate instruments;
however, they had relatively short strokes of 0.2 inches which were often exceeded when
axial strains got large. The LVDT's were mounted inside hollow steel tubes with aluminum
rods acting as extensions for the LVDT's. The LVDT's were mounted over a 9 in. {229
mm) gage length at the base of the wall and axial strains were calculated by simply dividing
the axial displacement by the gage length. The rectangular walls were instrumented with
LVDT's spaced along the length of the wall; whereas, the T-shaped walls had LVDT's
mounted along the web and the outside face of the flange. Figures 5.23 and 5.24 show the
locations of the LVDT's mounted on rectangular specimens RWl and RW2, respectively.
Figures 5.25A&B show the locations of the LVDT's mounted on the web and flange of
specimen TW2 {identical locations were used for TWl).
Axial loads were measured using hollow pancake load cells and the lateral cyclic loads
were measured using a 150 kip (667 kN) load cell mounted in series with the hydraulic

60

1
1
1

1
1

1
1
1
1
1
1
1

1
1
1
1
1

1
1
1

1
1
1
1
1
1
1
1

1
1

1
1
1
1
1
1

actuator. The load cell used to measure lateral loads can be seen in Figure 5. 17.
The strains in the reinforcing steel were measured through the use of strain gages. Ali
three types of reinforcement were monitored (longitudinal steel, uniformly distributed web
steel, and transverse boundary steel). Three different strain gage sizes were used depending
on the type of reinforcement being monitored. Strain gages with gage lengths of 0.25, 0.125,
and 0.062 inches (6.4, 3.2, and 1.6 mm) were used for the longitudinal, web and transverse
boundary steel, respectively. Ali of the strain gages were ordered from the Measurements
Group and had stock numbers of EP-08-250BG-120 (used on #3 longitudinal hars), EA06-125BT-120 (used on #2 web steel), and EA-06-062AK-120 (used on 3/16 inch diameter
transverse boundary reinforcement).
The longitudinal reinforcement (#3 hars, 9.5 mm) were instrumented at two distinct
levels; just above the wali-pedestal interface (approximately 1 in.

(25 mm) above the

pedestal), and at the first floor level (approximately 37 in. (940 mm) above the pedestal).
The vertical web steel ( #2 hars, 6.4 mm) were also instrumented at these two levels. The
horizontal web steel was instrumented at various heights of the wall as well as various
positions along the length of the wali. The transverse boundary reinforcement (3/16 inch
(4.8 mm) diameter wire) was heavily instrumented at the critical locations of the crosssection. Both boundary elements of the rectangular walls were evenly instrumented because
of symmetry. Gages were installed on the transverse boundary reinforcement located just
above the pedestal, and approximately 9 and 18 inches above the pedestal. The flanges ofthe
T-walls were not expected to experience significant damage; therefore, only a few gages were
installed at these locations. A majority of the damage was expected at the web boundary
region; therefore, this boundary region was heavily instrumented. Figures 5.26 and 5.27
show the location of strain gages mounted on the reinforcing steel of specimens RWl and
RW2, respectively. Figures 5.28 and 5.29 show the location of strain gages mounted on the
reinforcing steel of specimens TWI and TW2, respectively. These figures show the location
of strain gages mounted. on the reinforcing steel in the web as well as the flange. All of the

61

gages were systematically labeled so that data could be easily analyzed after completion of
the tests. The labeling scheme is discussed in detail in Chapter 6 where strain gage data is
presented.
Concrete strains within the boundary regions of specimens RW2 and TW2 were measured using embedded concrete strain gages. These gages were located approximately 2
in. ( 51mm) above the pedestal. They were oriented vertically and the measured strains
corresponded well with the axial strains measured using the LVDT' s, as will be discussed in
Chapter 6. Figures 5.30 and 5.31 show photographs of the embedded concrete strain gages
located in the boundary regions of specimens RW2 and TW2, respectively (RW2 had one
gage in each boundary element; whereas, TW2 had two gages in the web boundary element
and none in the web-flange intersection).
Two strain gage conditioner boxes, each having ten independent channels, were used
to supply power to the linear and wire potentiometers, and the pancake load cells used to
measure axial load. Each independent channel supplied a user specified amount of excitation
(de volts) and received an output signal that was then amplified to 10 volts. A Keithley
data acquisition system (DAS) was used to scan all channels at user specified intervals
throughout the duration of each test. The Keithley DAS performed all analog to digital
conversions so that the data could be read and stored ona personal computer. The Keithley
DAS was equipped with an AMM2 board, two AIM3 boards, and four AIM9 boards. The
AIM3 boards could each accomodate 16 channels of differential inputs. Displacement data
from the wire and linear potentiometers were channeled through these boards. The signals
from the AIM3 boards were then fed to the AMM2 board to perform all analog to digital
conversions.

The AMM2 board also had capacity for eight differential inputs and the

pancake load cells were wired directly to this board. Each AIM9 board was specifically set
up for signal conditioning of two LVDT's.
Additional instrumentaion was needed to accomodate the large amount of strain gage
data that was to be collected. Eight EXP-GP conditioning boards were used to provide the

62

1
1
1
1
1
1
1
1
1

1
1
1
1
1
1
1
1
1

1
1
1

1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1

necessary excitation and signal conditioning fora maximum of 64 strain gages (8 gages per
EXP-GP board). The eight EXP-GP boards were all connected in series (daisy-chained)
and then connected to a DAS-8 board located in the personal computer.

The DAS-8

board performed all analog to digital conversions for all incoming strain <lata. A special
software package called "VIEWDAC" was used to write a program that controlled the data
acquisition for all tests. VIEWDAC could access the Keithley DAS as well as the DAS-8
board, collect data at user specified intervals, and give real-time plots of test results that
were monitored throughout each test.
The hydraulic actuator used to apply the cyclic lateral loads was controlled by an MTS
458.10 Micro Console anda Wavetek Function Generator (Model 185). The 458.10 Console
has conditioners for the lateral load celi as well as a Temposonic displacement transducer
that was mounted directly to the hydraulic actuator. The Console allowed the versatility
of running the tests under load control or displacement control; however, all of the tests
performed in this study were run under displacement control.

5.5

Testing Proced ure


nce the specimens were entirely constructed, they were painted white so that cracks

in the concrete could be more easily detected. The specimen was than hoisted into place
using the fi.ve ton overhead crane and mounted to the strong floor. Eight high-strength
tie-down rods ( two tie-down pads) were used to attach the specimen to the floor. A thin
layer of grout was placed between the pedestal and the floor to level the specimen and
assure uniform contact.
The next step was to attach the load transfer assembly to the top of the wall (Fig.
5.15). A layer of grout was placed between the steel top assembly and the concrete to help
distribute axial load evenly over the entire wall cross-section. The steel top assembly was
attached by 1 inch (25 mm) diameter threaded rods that were anchored in the top story

63

of the wall. Six rods were used for the rectangular specimen and nine were used for the
T-shaped specimens.
nce the top assembly was in place, the post-tensioning cables were run up through
the steel channels and through the load cells and jacks. A set of reusable chucks were placed
on each cable and the axial load was ready to be applied. The hydraulic actuator was then
mounted between the reaction wall and the steel top assembly (Fig. 5.17). Both ends of the
actuator were attached to clevices with frictionless pins to allow rotation of the actuator in
the plane of the wall specimen. The out-of-plane support system was then mounted between
the top of the wall specimen anda second reaction wall (Figs. 5.18 and 5.19), parallel to
the wall web.
The next step of the testing procedure was to mount all of the necessary instrumentation. nce the instrumentation had been properly mounted, all lead wires were attached to
the data acquisition system (or directly to the personal computer) and the lead wires were
taped to the fl.oor (to prevent them from being tripped over). The instrumentation and the
software program were then tested and all channels were zeroed prior to testing.
Testing was initiated by taking a reading from all instruments prior to load application.
The axial load was than applied through the use of hand pumps. Readings were taken
periodically during application of axial load. nce the axial load reached its specified level,
valves on the hand pumps were shut off to ensure that the load would remain constant
throughout the test. The axial load could be monitored continuously by either the load
cells, whose values were printed on the computer screen throughout testing, or by the
pressure gages attached directly to the hand pumps.
The reversed cyclic lateral drifts were then applied under displacement control. The
first drift level to be applied was approximately O. 1%, followed by 0.25%. The drift level

1
1
1
1
1
1
1
1
1
1

1
1
1
1

to 1.0% and then each test varied somewhat depending on its behavior up to that point.

1
1

64

was then increased in 0.25% increments up to 1.0%. The drift was then increased in 0.5%
increments up to 3.0%. At least two complete cycles were performed at each drift level up

1
1

1
1

Figures 5.32 through 5.35 plot the actual lateral drift routines that were followed during
testing for specimens RWl, RW2, TWl and TW2, respectively. it can be seen from the
figures that specimens RWl and TWl were subjected to a fewer number of cycles than
were RW2 and TW2. Cracks were marked at peak displacements in each direction for every

cycle. The cracks were marked with two different color markers. One color was used to

the second color was used to mark cracks caused by pushing on the specimen (southward

1
1
1
1

mark all cracks caused by pulling on the specimen (northward displacements); whereas,

displacements). Every crack was labeled with a number that corresponded to the <lata point
number that was recorded by the DAS computer program. Cracks were no longer marked
when the specimen began to experience significant splitting and spalling at the boundary
regions.
Chapter 6 presents the experimental results from all of the tests and A ppendix B
provides a complete description of the experimental observations made during each test.
N umerous photographs are included in the appendix which show each specimen at specified
drift levels, providing a visual "history" of each test. The photographs allow comparisons

to be made between specimen behavior which help reveal how variations of the test pa-

1
1
1

experimentally observed behavior; whereas, comparisons of experimentally measured and

rameters affected the behavior of the specimens. Chapter 6 discusses comparisons of the

analytically predicted behavior are discussed in Chapter 7.

1
1
1
1
1

65

Chapter - 6
EXPERIMENTAL RESULTS

An overview of the experimental results of the four wall specimens is presented in this
chapter. Experimentally obtained. data that are presented include: 1) observed damage and
behavior, 2) lateral load-top displacement relations, 3) base moment-base rotation relations,
4) lateral displacement pro:files, 5) shear distortion relationships, 6) concrete (normal) strain
pro:files, 7) reinforcing steel strain pro:files, and 8) reinforcing steel strain histories. Sections
6. 1 through 6.8 present and discuss the experimental data listed above, respectively. Comparisons of experimental behavior of each of the four wall specimens is emphasized..

6.1

Experimentally Observed Damage and Behavior


As described in Chapter 5, a constant axial load of approximately O. 10A 9 J~ was applied

to each specimen prior to subjecting the specimen to the reversed cyclic lateral drift routines shown in Figs. 5.32 through 5.35. Due to variations in concrete compressive strengths
(specimens RW2 and TW2 had relatively large compressive strengths compared to RW 1
and TWl), it was difficult to achieve the desired axial load leveJ of 0.10A 9

J: on ali speci-

mens; therefore, the axial loads measured during testing ranged from 0.07A 9 J~ to O. 10A 9 J~.
Figures 6.1 through 6.4 plot the axial load histories for specimens RWl, RW2, TWl, and
TW2, respectively.

In general, two cycles were performed at each drift level; therefore, odd numbered cycles represent the first cycle performed at a given drift level and the even numbered cycles
represent the repeat cycles. Each cycle began by subjecting the specimens to northward
displacements (actuator was pulling the specimen; Fig. 5.13) which are considered positive

66

1
1
1
1
1
1

1
1
1
1
1
1
1
1
1
1
1
1

1
1

1
1

1
1
1
1
1

displacements for the remainder of this discussion.

The second half of each cycle con-

sisted. of southward displacements (pushing the specimen) which are considered. negative
displacements. The lateral loads causing positive displacements are considered. positive,
and likewise, lateral loads causing negative displacements are considered. negative.
Tables 6. 1 through 6.4 present the peak lateral displacements, loads, and shear stresses
associated. with each cycle for specimens RWl, RW2, TWl and TW2, respectively. Shear
stresses are presented in terms of

,JH,

where /~ is the measured. concrete compressive

strength (psi) at the first story level at the time of testing.

The concrete compressive

strength of the specimen at this level is listed. in the footnotes of each table, as well as in
Table 5. 1. The shear stresses in Tables 6. 1 and 6.2 are calculated. based. on gross crosssectional areas of the rectangular wall specimens (192 in 2 ; 1,239 cm 2 ); whereas, shear
stresses in Tables 6.3 and 6.4 are based. on effective web cross-sectional area (alsa 192 in 2 ;
l,239cm 2 ) of the T-shaped. specimens.
A detailed. description of the experimentally observed. damage and behavior of each test
specimen is presented. in Appendix B. The appendix summarizes the observations that were
made while the tests were in progress and is meant to act asa supplement for the interested.
reader. The observations are based. on visible damage (cracking, splitting, crushing, buckling, ete.), as well as observations that were ded.uced. from instrumentation results during

the tests (yielding of reinforcement, displacement/drift measurements, normal strains along

1
1
1

photographs taken at various displacement (drift) levels during testing. These photographs

1
1
1

the length of the wall, ete.). The observed. damage for each wall is shown in a series of

demonstrate the type and extent of damage that each specimen experienced and allows for
easy comparisons with other specimens.
Discussion in Appendix B refers to figures that are presented. in this chapter; therefore,
it is recommended. that this chapter be read in its entirety before reading the appendix.
Readers that are interested solely in the experimental results, and have little or no interest in
what occurred. during the actual tests, may skip Appendix B without any loss of continuity;

67

1
1
however, the photographs presented in the appendix provide an excellent "history" of each
test.

6.2 Lateral Load Versus Top Displacement Relations

The measured lateral load versus top displacement relations are plotted in Figs. 6.5
through 6.8 for specimen RWl, RW2, TWl, and TW2, respectively. The lateral load versus
top displacement relations for each specimen are presented in two separate plots to show
the changes in stiffness experienced by each specimen. The fi.rst fi.gure (designated with an
'A') plots the initial cycles of the test up to a lateral drift level of O. 75%. The second fi.gure
(designated with a 'B') plots all drift cycles for the entire test (see Figs. 5.32 through 5.35
for lateral drift routine of each specimen). These fi.gures also indicate the axial load level
that each specimen was subjected to during testing, as well as the maximum shear stress.
The axial load and shear stress levels are based on measured concrete strengths at the base
of the wall at the time of testing (Table 5.1).
The lateral loads plotted in Figs. 6.5 through 6.8 were recorded directly from the 150
kip load cell attached to the lateral hydraulic actuator. The horizontal component of the
applied axial load (due to rotation at the top of the wall) is neglected in these plots, which
results in an error of no more than approximately 5% at a lateral drift level of 2%. Top
displacements were recorded by a wire potentiometer at the fourth story level (12 feet above
the wall-pedestal interface), mounted to the rigid steel reference frame that was bolted to
the strong floor (Fig. 5.20). A lead wire was attached from the top of the wall specimen to
the wire potentiometer.
Figures 6. 5 and 6.6 indicate that the rectangular specimens exhibit stable hysteretic behavior and excellent ductility and energy absorbing capabilities. Specimen RWl (Fig. 6.5A)
experienced a decrease in stiffness during the fi.rst eyde at O. 75% lateral drift, due to yielding
of the longitudinal boundary reinforcement (see Appendix B for a detailed description of

1
1

1
1
1
1
1
1
1
1
1
1

the observed behavior and its influence on the measured results), and completed two cycles

68

1
1
1

1
1
1
1
1
1
1
1
1
1

at 2.0% lateral drift (designed for 1.5% drift) before experiencing a loss of strength (caused
by buckling of longitudinal boundary reinforcement) during the :6.rst cycle attempted at
3.0% lateral drift (Figure 6.5B). Specimen RW2 (Fig. 6.6) has similar strength and stiffness
characteristics as RWl; however, RW2 was subjected to many additional cycles compared
to RWl (Fig. 5.33), and did not experience a significant loss of strength even after two
complete cycles at 2.5% lateral drift. The improved behavior is attributed to the closer
spacing of the transverse boundary reinforcement which delayed the onset of buckling of
the longitudinal reinforcement.

Figures 6. 7 and 6.8 indicate that behavior of the T-shaped walls is dependent upon the
direction of the applied lateral loads. When the flange is in compression, the strength is
only slightly larger than that of the rectangular walls; however, the ductility is significantly
enhanced. When the flange is in tension a significant increase in strength and stiffness is
noticed; however, ductility may be adversely affected as shown by specimen TWl (Figure
6. 7). The sudden loss of strength during the first cycle to 1.5% lateral drift was caused
by a brittle buckling failure in the web.

The failure is attributed to the lack of well-

confined concrete (transverse reinforcement spread over adequate depth of compression


zone and spaced close enough to suppress buckling of longitudinal reinforcement) in the
web boundary region (Fig. 4.12) where high compressive strains were expected (Figure
4. 17). Transverse reinforcement in the web boundary region of specimen TW2 was designed
using the displacement-based design procedure outlined in Appendix A, resulting in a closer

spacing of transverse reinforcement over an increased depth of the cross-section (Figure

of which direction the lateral load is acting. It can be seen that specimen TW2 did not

4.16). Figure 6.8 indicates that specimen TW2 exhibits stable, ductile behavior regardless

experience a loss of strength until the second and third cycles at a lateral drift level of

of-plane (stability) failure. The maximum shear stress levels indicated on Figs. 6.7 and 6.8

1
1

approximately 2.5%, at which time the web boundary element began to experience an out-

were calculated for the critical case of the flange in tension, and indicate that shear demands

69

1
on the T-shaped walls are approximately twice as large as they are on the rectangular walls.

Secant Stiffness:
It can be seen in Figs. 6. 5 through 6. 8 that each specimen experienced significant
changes (decreases) in their stiffness as drift levels increased. It is also evident however,
that repeat cycles at each drift level caused little or no additional stiffness degradation of
the specimen even at high drift levels approaching the end of each test. The repeat cycles
at each drift level followed the initial cydes very dosely, giving a dear indication of the
specimen's stability.
It is of interest to determine the magnitude of stiffness degradation experienced by
each specimen; therefore, measured values of lateral load at the peak displacement for each
initial cycle were divided by the corresponding displacement to provide a measure _of the
specimen's effective secant stiffness. Figures 6.9 through 6.12 plot the calculated values
of secant stiffness at each odd numbered (initial) cycle for specimens RWl, RW2, TWl,
and TW2, respectively. it is noted that these values represent average values (stiffnesses
are greater at the beginning of each eyde) and are used as a hasis for comparisons with
other specimens. The stiffness calculated for eyde

# 1 gives a reasonable approximation

of the uncracked stiffness for each specimen; whereas, the stiffness calculated for cycle #7
(O. 75% drift) gives an estimate of the specimen's stiffness at the onset of first yielding of the
longitudinal boundary hars. In general, the secant stiffness of each specimen at the onset
of first yielding was approximately 50% of its uncracked (gross) stiffness.
The stiffness is plotted for all initial cydes for each direction of applied lateral loads
(north and south). Figures 6.9 and 6.10 indicate that the stiffness of the rectangular walls
are nearly equal in both directions indicating symmetry of these specimens (slight variations
in stiffness during the initial cydes are attributed to minor offsets at the start of each test).
Figures 6.11 and 6. 12 reveal that the T-shaped walls are considerably stiffer under nega tive
displacements (flange in tension), indicating that the unsymmetrical shape of the walls and
the direction of loading has a significant impact on wall stiffness. During initial cycles, the

70

1
1
1
1
1
1
1
1
1
1

1
1

1
1

1
1

1
1

1
1

1
1
t
1
1

1
1
1
1
1
1
1
1
1

stiffness of the T-shaped walls under positive displacements are approximately 80% of the
stiffness when subjected to negative displacements; however, as drift levels increase beyond
yield, the variation in stiffness between the two directions increases. At the design drift
level of 1.5%, the stiffness of specimen TW2 under positive displacements is approximately
20% of its initial gross-section stiffness and approximately 50% of the stiffness measured
for loads acting in the opposite direction (negative displacements).

Energy Absorption:
The la ter al load versus top displacement plots shown in Figs. 6. 5 through 6.8 provide
a convenient means for evaluating and comparing the energy absorbing capacity of each
specimen. The amount of energy absorbed by a specimen during any cycle is defined as the
area enclosed within the hysteretic loop for that eyde. Structures capable of absorbing more
energy are less susceptible to damage (and collapse) during severe ground motions; therefore,
structural members that have load-deflection curves with wide, full loops are desirable over
members that reveal narrow, closed loops.

The post-yield behavior of these walls was

dominated by flexure and energy was absorbed primarily by widening of existing cracks
and yielding of longitudinal reinforcement. The gradual widening of the hysteretic loops
in Figs. 6.5 and 6.6 indicate that the rectangular walls posess excellent energy absorbing
capabilities. Figures 6. 7 and 6.8 indicate the presence of some minor "pinching" which
is attributed to the increased role of shear in the T-shaped specimens.

This pinching

results in hysteretic loops of the T-shaped walls being narrower and not as full as those
of the rectangular walls; however, the T-shaped walls still exhibit good energy absorbing
capabilities (except for specimen TWl when the flange is in tension).

6.3 Base Moment Versus Base Rotation Relations


Measured base moment versus base rotation relations are plotted in Figs. 6. 13 through
6. 16. Moment-rotation relationships provide an indication of the contribution of flexural
response to overall specimen behavior. it is seen that the shape of the plots in Figs. 6. 13

71

1
through 6.16 closely resemble the shapes in Figs. 6.5B through 6.8B, indicating that specimen behavior is dominated by flexural response. A detailed discussion of the contributions
of flexure and shear to overall specimen behavior, presented in Chapter 7, verifies this observation. Other observations drawn from these plots are similar to those discussed in Section
6.2, for lateral load versus top displacement relations.
The moment at the base of the wall was calculated as the sum of primary and secondary
moments. The primary moment was calculated as the product of the applied lateral load
and the loading height (150 in., 3.8 m). The P-Ll moment was calculated as the product
of the applied axial load, measured using the pancake load cells, and the measured lateral
deflection at the top of the wall. The moment caused by the horizontal component of the
axial load was determined to be negligible; therefore, it was not included in the calculation.
The base rotations were measured by two wire potentiometers mounted 30 inches (762
mm) above the pedestal at each end of the wall (Figure 5.23). Rotations were calculated
by dividing the difference in relative displacements by the distance between the wire potentiometers (52 in., 1.32 m); therefore, the measurements include the effects of flexural
curvature over the lower 30 inches (762 mm) of the wall and slippage of longitudinal reinforcement from the pedestal; however, the measurements are not influenced by the presence

1
1
1
1

1
1
1
1

of shear distortions.

6.4

Lateral Displacement Profiles


Measured lateral displacement profiles are plotted in Figs. 6. 17 through 6.20. The

displacements plotted in the figures were recorded at each "floor level" using wire potentiometers mounted to the steel reference frame which was bolted to the strong floor at the
end of the wall specimen (Fig. 5.20). The plotted data points are the peak displacement
readings taken during the first eyde at each drift level. The symbols used represent a
particular drift level as shown in the legend, and each plot indicates the north and south
directions which represent positive and negative displacements, respectively.

72

1
1
1

1
1

1
1

1
1
1

1
1
1
1

1
1
1
1
,

1
1
1
1

1
1
1

Figures 6. 17, 6. 18 and 6.20 show that displacement readings taken from the lower two
story levels tend to plot non-linearly; whereas, the readings from the third and fourth story
levels tend to plot linearly with the second story leveL This indicates that a majority of the
wall deformation is occuring over the first story level, as expected ( moment is largest at the
base of the cantilevered wall specimens causing more curvature at this level). This trend
becomes more pronounced as drift levels increase, indicating that inelastic deformations are
concentrated in the plastic hinge region at the base of the wall. This trend is not as distinct
in Fig. 6. 19 due to the.lack of inelastic deformation experienced by specimen TW 1 prior to
failure.

6.5 Shear Distortion Relations


Measured shear distortions over the bottom two stories of each wall specimen are
plotted in Figs. 6.21 through 6.24. The shear distortions for each specimen are presented
in two separate plots. The first figure (designated with an 'A') plots the measured shear
distortion over the first story height; whereas, the second figure (designated with a 'B')
plots the measured shear distortion over the second story level. The shear distortions were
calculated using the displacement data collected by the wire potentiometers mounted in an

"X" configuration on the lowest two levels of each wall specimen (Figure 5.22). The average
shear distortion bavg) was calculated as:
(d~ - d)d - (d2 - d2)d2

2hl

iavg =

(6.1)

w here, d 1 and d2 are the original (undeformed) dimensions of the two wire potentiometers,

d~ and d are the deformed dimensions, h is the vertical height and

1 is the horizontal length

of the "X" configuration. The variables are shown in Fig. 6.25.


Based on comparisons of Figures 6.21 and 6.22, A & B, it is concluded that the shear
distortions at the second story level of the rectangular walls are significantly less than
the shear distortions measured over the first story level, as expected. When the wall is

73

1
in the elastic range, the flexural deformations increase in proportion to the height above
the pedestal cubed (6ex is proportional to l 3 ); therefore, flexural deformations rapidly
outgrow the shear deformations as the height above the pedestal increases.

When the

wall enters the inelastic range, the deformation is concentrated in the plastic hinge region
located at the base of the wall. The deformations occurring above the hinging region are
still predominantly elastic, thus the measured shear deformations are substantially lower at
the second story level.
The flexural-strength of the T-shaped specimens is approximately twice as large when
the flange is in tension (compared to the flange in compression), indicating that the shear
force should also be approximately twice as large. Based on Figures 6.23A and 6.24A, it
can be seen that the measured shear distortions at the first story level of the T-shaped walls
are larger when the flange is in compression than when the flange is in tension, eventhough
the shear force is substantially larger when the flange is in tension. lnitially, this may appear to be contradictory; however, based on the experimentally observed damage (diagonal
crack patterns), the following explanation is offered. When the flange is in compression,
the depth of the compression zone is extremely small ("'-' 3 in.; 76 mm), and large inelastic
tensile strains are developed in the web, resulting in substantial fiexural and shear cracking
(diagonal shear cracks extend the entire length of the web). The inelastic shear distortions
measured under this loading condition are relatively high, eventhough the measured shear
force is comparitively low (half as large as the shear expected under reversed loading condition). When the flange is in tension, the wall stiffness increases and the depth of the
compression zone is approximately half the wall length. Under this loading condition, less
damage (diagonal cracking) was witnessed; therefore, relatively small shear distortions were
measured, eventhough the shear force was approximately twice as large.
Figures 6.23B and 6.24B indicate that shear distortions in the second story of the
T-shaped walls are smaller than they are in the first story and appear to have similar
magnitudes under each direction of loading (eventhough shear force is substantially larger

74

1
1

1
1
1

1
1
1
1

1
1
1
1
1
1

1
1
1
1
1

1
1

when the flange is in tension). Similar reasoning to that used in the previous paragraph for
the first story shear distortions can be used to explain these observations (inelastic shear
distortions caused by relatively low shear force are approximately equal to the elastic shear
distortions caused by a larger shear force in the opposite direction).
The amount of lateral displacement at the first floor level attributed to the shear
distortions can be calculated as:
l,shear

= 36 bavg)

(6.2)

where, the height of each story is 36 inches (914 mm). The lateral displacements caused by
the second story shear distortions can be evaluated in asim.Har manner using the calculated
shear distortion at that level. The contribution of shear distortion to overall specimen

1
1
1

1
1
1
1
1

response is discussed in detail in chapter 7.

6.6

Concrete Strain Proflles


Measured normal strain profiles at the base of the walls are plotted in Figs. 6.26

through 6.29. The concrete strain profiles for each rectangular specimen are presented in
two separate plots; whereas, the concrete strain profiles for each T-shaped wall are presented
in four plots. The figures designated as 'A' and 'B' plot the measured web strain profiles for
the lateral load acting in the north (positive) and south (negative) directions, respectively.
The figures designated as 'C' and 'D' plot the measured flange strain profiles (T-shaped
walls only) for the case of the flange in compression and the flange in tension, respectively.
The measured concrete normal strains were calculated using the displacement data collected
from the Linear Variable Differential Transducers (LVDT's) mounted along the length of
the wall specimens (Figs. 5.23-5.25 ). The strains were calculated as the displacements
divided by the known gage length of 9 in. (229 mm). The plotted data points were taken
at the peak lateral displacement during the first cycle at each drift level. The symbols used
in the plots represent a particular drift level as shown in the legends.

75

Earthquake Eng Res Ctr Ubrary


Unlv of Callf. Berkeley
1301 S, 46th St. - RFS 453
Rlchmond, CA 94804-4698 USA

(S10) 885-3419

1
Figures 6.26 A&B and 6.27 A&B indicate that the strain profiles for positive and negative displacements are mirror images of each other, indicating symmetry of the rectangular
wall specimens. Figures 6.28 A&B and 6.29 A&B indicate that the strain profiles for the
T-shaped specimens are significantly different under positive and negative displacements.
Figures 6.28 and 6.29 indicate that when the flange of a T-shaped wall is in compression
the depth of the compression zone is relatively small and the strain gradient is relatively
large (low compressive strains in the flange and relatively large tensile strains in the web ).
When loads are reversed and the flange ofa T-shaped wall is in tension, the depth of the
compression zone is relatively large and the strain gradient is relatively low (compressive
strains in the web are approximately equal to the tensile strains in the flange). The strain

1
1

1
1

profiles plotted in Fig. 6.28A show slightly different results than those shown in Fig. 6.29A.

It is believed that the displacement data recorded by the LVDT mounted 6 in. (152 mm)
from the outside edge of the flange of specimen TWl is slightly erroneous. The LVDT was
bearing against a thin washer located at the end of one of the high-strength tie-down rods
used to secure the specimen to the strong floor. The washer "cupped" upward when the
wall was subjected to negative displacements, causing the LVDT to indicate slightly larger
compressive strains than were actually being experienced. A linear interpolation between
the LVDT's located on each side of the erroneous LVDT may provide a more accurate
estimate of the strain distribution at this location.

The flange strain distributions plotted in Figs. 6.28 C&D and 6.29 C&D were recorded
by LVDT's mounted on one side of the flange. The strains are assumed to be symmetrical
about the centerline of the flange; therefore, the figures plot the measured strains across
the entire flange width. Figure 6.29C indicates that compressive strains are nearly uniform
across the flange until drift levels of approximately 2.0%. At these relatively high drift
levels, the longitudinal reinforcement located at the web-flange intersection experiences
significantly higher tensile strains (when the flange is in tension) than the flange steel
located further from the web. The relatively large tensile strains cause cracks to open

76

1
1

il
1

1
1

1
1
1

1
1
1

1
1
1
1
1

1
1
1
1

1
1

wider at the web-flange intersection. When loads are reversed, these cracks must close
before the concrete can effectively develop compressive strains. Figure 6.29C indicates that
the concrete located further from the web-flange intersection begins to develop compressive
strains prior to the cracks at the intersection being fully closed. Specimen TW 1 (Fig. 6.28C)
does not indicate similar behavior because the specimen failed prior to the flange ever being
subjected to these relatively high levels of tension (specimen TWl failed at approximately
1.25% drift under negative disp.lacements).
The concrete strain was also measured using embedded concrete strain gages. These
gages were only used in the boundary regions of specimens RW2 and TW2. One gage was
placed in each boundary region of specimen RW2; w hereas, two gages were placed in the
web boundary region of specimen TW2, and none were placed in the web-flange intersej:tion.
The data collected using the embedded concrete strain gages are plotted in Figs. 6.30 and
6.31 for specimens RW2 and TW2, respectively. Figure 6.31 also plots the strain results
measured by an LVDT, for comparison. The LVDT was located 6 in. ( 152 mm) from the
end of the web of specimen TW2, which lies between the locations of the two embedded
strain gages. The strain in these figures is plotted versus the data point number. Tables
6.1 through 6.4 {column 2) list the data point numbers which correspond to each lateral
drift eyde; thus, cross-referencing the appropriate table with the figure provides a method
for determining the measured strains at various drift levels. Figure 6.31 indicates that the
strains based on the LVDT readings show good correlation with the embedded gages when
in compression; however, the LVDT's predict substantially larger tensile strains than the
embedded gages. This is attributed to the LVDT being mounted to the cover concrete

which experiences significantly more cracking compared to the confined core. The results

1
1
1
1

221). Similar results would be expected for the other specimens.

of the LVDT start to vary from the embedded gages at a drift level of O. 75% {data point

77

6. 7

Reinforcing Steel Strain Profiles


Reinforcing steel strain gage data were also used to plot normal strain profi.les along

the length of the specimens. The results are plotted in Figs. 6.32 through 6.35. The steel
strain profi.les are plotted for two distinct levels. The first level is at the base of the wall;
therefore, these profi.les can be compared directly to those shown in Figs. 6.26 through
6.29, which were calculated using the LVDT displacement data. The second level for which
the steel strain profi.les are plotted, is at the top of the first story level (approximately 36
inches above the wall-pedestal interface). The strain profi.les at each level are plotted for
the lateral load acting in the north (positive) and south (negative) directions, respectively.
The profi.les for each rectangular specimen are presented in four separate plots; whereas, the
strain profiles for each T-shaped wall are presented in eight plots. The figures designated as
'A' and 'B' plot the measured steel strain profi.les at the base of the wall web (approximately
2 inches above the pedestal); whereas, the figures designated as 'C' and 'D' plot the steel
strain profi.les along the web at the top of the first story level ( approximately 36 inches above
the wall-pedestal interface). The figures designated as 'E' through 'H' plot the measured
flange steel strain profiles (T-shaped walls only) at the two distinct levels for the cases of

1
1
1

1
1
1
1

the flange in compression and in tension, respectively. The plotted data points were taken
at the peak lateral displacement during the first cycle at each drift level. The symbols used
in the plots represent a particular drift level as shown in the legend.

it is evident from Figs. 6.32A&B and 6.33A&B that the steel strains do not plot linearly
along the length of the rectangular walls. At moderately low drift levels (~ 0.5% lateral
drift), the strains tend to plot linearly; however, once the reinforcing steel yields (in tension
or compression) it possesses a permanent set in that direction. The figures indicate that the
longitudinal boundary steel nearest the outermost edge of the rectangular walls experience

1
1

develop the same level of tensile strain that they could prior to yielding in compresison. The

1
1

78

"strain relaxation", which can be attributed to several effects. As drift levels increase, the
longitudinal boundary reinforcement yields in compression, and thereafter these hars can not

1
longitudinal reinforcement located nearer to the neutral axis never experiences high levels

1
1
1
1

of compressive strain; therefore, these hars develop larger tensile strains than the boundary
reinforcement. The "relaxation" may alsa be attributed to bond deterioration between the
steel and the concrete, caused by increasing levels of damage as drift levels increase. The
bond deterioration allows the reinforcing steel to "slip" thus preventing large tensile strains
from developing. The effects of "strain relaxation" are not as evident in specimen RW2
(Fig. 6.33 A&B) as they are in specimen RWl (Fig. 6.32 A&B). The compressive strength
of concrete is known to effect the bond strength between steel and concrete (ACI 318-89);
therefore, it is plausible that the higher compressive strength of RW2 (Table 5. 1) provided
better bond strength which suppressed slippage of the longitudinal boundary reinforcement
in this specimen ( transverse reinforcement in specimen RW2 was spaced closer than in RW 1
which also allows for better bond).
Based on these plots, the wall exhibits approximately a 6 inch ( 152 mm) depth of

compression zone when drift levels exceed 0.5%. Figures 6.32 C&D and 6.33 C&D indicate

steel experiences less strain relaxation at this level. These plots also indicate that the depth

1
1
1

1
1
1

1
1

that strain profiles above the first story height tend to plot more linearly and the reinforcing

of the compression zone above the first story height has increased to approximately half
of the wall length. The moment at this location is approximately 75% of the moment
at the base of the wall; therefore, the measured strain levels at this location are smaller.
The boundary steel yields in tension (at approximately 1.0% lateral drift); however, the
measured compressive strains in the opposite boundary are relatively low (sligthly less than
yield).
Reinforcing steel strain profiles for the T-shaped walls indicate substantially different
behavior than the rectangular walls. Figures 6.34A and 6.35A indicate that when the fiange
is in compression the depth of the compression zone is approximately 3 inches (76 mm) which
is less than the fiange thickness. The compressive strains within the fiange are extremely
small compared to the tensile strains at the web boundary. When loads are reversed and the

79

flange is placed in tension, the depth of the compression zone is increased to approximately
20 inches (508 mm), and the compressive strains at the web boundary begin to exceed yield
strains at drift levels as low as O. 75%. Tensile strains in the flange approach yield at a
drift level of approximately 1.5%; however, the largest tensile strains are experienced by the
longitudinal reinforcement located between the flange and the neutral axis (these hars are
in tension regardless of the loading direction and never experience significant compressive
strains).
Figures 6.34C and 6.35C indicate that strains at the top of the first story level plot
more linearly and that the depth of the compression zone ranges dramatically depending
on the drift level. The flange reinforcement does not yield in compression at this level;
however, the web boundary reinforcement yields in tension at a drift level of approximately
1.0%. Figures 6.34D and 6.35D also reveal more linear strain behavior, and indicate that
the depth of the compression zone is greater than 30 inches (762 mm), regardless of drift
level. In general, the tension and compression steel at this level remain elastic until drift
levels of approximately 1.5%.
Figures 6.34E and 6.35E indicate that compressive strains in the flange reinforcement
are extremely uniform and below yield until drift levels of approximately 2.5%. When in
tension, the flange reinforcement exhibits uniform tensile strains until drift levels approach
2.0%, at which time the reinforcing steel at the web-flange intersection experienced significantly higher tensile strains than the steel located further from the web. Figures plotting
the tensile strains of the flange reinforcement also plot a horizontal line indicating the yield
strain of the reinforcement to allow for easy determination of the drift level at which yielding
occurs.
Figures 6.34G and 6.35G indicate that compressive strains in the flange reinforcement
above the first story height are well below yield, regardless of drift level Figures 6.34H and
6.35H indicate that tensile strains at this level are linear and remain less than the yield
strain until drift levels of approximately 1.5%.

80

1
1

1
1
1
'I

1
1

1
1

1
1
1

1
1
1
1
1

1
6.8

Reinforcing Steel Strain Histories


Reinforcing steel strain histories for all four specimens are plotted in Figs. 6.36 through

6.39, respectively. The figures plot results from all three types of reinforcement (longitudinal, uniform web, and transverse boundary) at various locations along the cross-section as

1
1
1

well as various heights above the wall-pedestal interface. Each plot is labeled with letters
and numbers which indicate the type of reinforcement, the distance above the pedestal, and
which bar the gage is mounted on. Strain gages mounted on longitudinal reinforcement
are labeled with an "L". The number preceeding the ''L" indicates which bar the gage is
mounted on (see Figs. 5.26 through 5.29). Longitudinal gages with the number 37 following the "L" indicate that the gages were mounted approximately 37 inches above the
wall-pedestal interface (just above the first story level). Gages mounted on the uniformly

1
1
1

east, and west. Gages mounted on transverse boundary reinforcement are labeled with

point number and in general, each graph plots the results from two strain gages ( usually

1
1

distributed web steel were labeled with a "W", and the number following the "W" indicates
the height above the pedestal. The letters following the number represent north, south,

either an "H" if mounted on a hoop, or a "T" if mounted on a cross-tie. The numbers


represent the height above the wall-pedestal interface and the letters following the numbers
indicate location (north, south, east, and west). All of the graphs plot strain versus data

mounted in similar locations on the same type of reinforcement). Tables 6.1 through 6.4 list
the data point numbers that correspond to each drift level allowing for correlation between
the measured strains and the drift levels at which they occurred.
The yield strains for the various types of reinforcement are listed in Table 5.3 and

their stress-strain relationships are plotted in Figure 5.2. Based on these values, it can

Results from the strain histories plotted for the transverse boundary reinforcement can

1
1
1

be determined from the strain histories when yielding of the reinforcement first occurred.

be used to evaluate the confinement effectiveness. A detailed discussion of confinement


effectiveness is presented in Chapter 7.

81

1
Chapter - 7
ANALYTICAL STUDIES AND COMPARISON
WITH EXPERIMENTAL RESULTS

1
7 .1

Introd uction
Comparisons of analytically obtained and experimentally measured wall behavior are

presented in this chapter. Section 7.2 presents moment-curvature and moment-axial load
relations that are based on the measured material properties of each specimen (compared
with the assumed material properties used for design purposes in Chapter 4). Section 7.3
discusses the test setup requirements based on preliminary analytical work. Section 7.4 discusses the expected failure mode of each wall specimen and emphasis is placed on discussing
the wall specimens' flexural and shear strengths. Section 7.5 discusses the contributions of
flexure and shear to specimen displacements measured at the first story level and at the
top of each specimen. Section 7.6 describes analytically calculated monotonic wall behavior including: (1) calculation of monotonic lateral load-top displacement relations, and (2)
comparisons of the monotonic lateral load-top displacement relations with experimentally
o btained results for each specimen. Section 7. 7 describes the analytical method used to
predict the displacement (drift) levels at which local buckling of longitudinal reinforcement
is expected, and Section 7.8 evaluates the effectiveness of the transverse reinforcement in
providing confinement within the boundary regions of each specimen.

1
1

1
7.2 Moment-Curvature and Moment-Axial Load Relations
Uniaxial moment-curvature relations were computed for monotonic loading using the
computer program BIAX (Wallace, 1992). For the analyses, plane sections were assumed to

82

1
1

1
1

1
1
1

1
1
1
1

1
1
1

1
1
1
1
1

1
1
1
1

remain plane and nonlinear material stress-strain relations were used. Two different stress
strain relations for unconfined and confined concrete were used. The first stress-strain
relation is the Modi:fied Kent-Park relation (Park, Priestley and Gill; 1982); whereas, the
second relation is that proposed by Saatcioglu and Razvi (1992). The measured concrete
compressive strengths of each specimen were incorporated into the respective relationships
to yield analytical results that can be compared directly to the experimentally obtained
results. The steel stress-strain relationship used in the analytical models was identical to
the measured relationship (for the #3 reinforcing hars) shown in Figure 5.2. It is noted that
the moment-curvature relations do not include the effects of tension stiffening (as described
by Vecchio and Collins, 1986).
The computed moment-curvature relations for specimens RWl, RW2, TWl, and TW2
are plotted in Figs. 7.1 through 7.4, respectively. The moment-curvature relations presented
in these :figures incorporate the measured material properties (presented in Chapter 5) of
each particular wall specimen. Figures 7.1 and 7.2 present two separate moment-curvature
plots for each rectangular wall, representing the two different concrete stress-strain relations.
The moment-curvature relations for the T-shaped specimens depend on the direction of the
applied load; therefore, Figs. 7.3 and 7.4 show four plots ( two with the flange in tension
and two with the flange in compression). it is noted that these relations assume that the
entire fl.ange is effective (in tension and compression). It can be seen in the figures for the
T-shaped walls that when the flange is in compression, use of either concrete stress-strain
relation results in identical behavior. However, when the flange is in tension the predicted
behavior is dependent upon which stress-strain relation is used.
A limiting compressive strain of 0.015 was used in the analyses for the rectangular

specimens and the T-shaped walls when the web was in compression. lt is important to note
that these plots do not take into account when buckling of the longitudinal reinforcement
is expected to occur; therefore, the ductility levels shown may be misleading if buckling is
not adequately prevented. Buckling of longitudinal reinforcement is discussed in detail in

83

Section 7. 7.
A much lower limiting compressive strain of 0.003 was used for the T-shaped walls
when the flanges were in compression due to the significant curvature ductility that exists
at this relatively low strain level when loads act in this direction. Buckling of longitudinal
reinforcement in the flange is not a concern when loads act in this direction because the
depth of the compression zone is very small (less than the flange thickness) and the expected
compressive strains are quite low (:5 0.003). it is also noted that the concrete cover would
still be intact at this strain level.
The plotted relations (Figs. 7. 1 through 7.4) clearly indicate that strength, stiffness
and ductility are all directly affected by the shape of the cross-section and the direction of
the applied lateral load. The plots for the rectangular specimens give a reference against
which the strength, stiffness and ductility of the T-shaped walls can be compared. The
moment-curvature relations for the T-shaped walls with the flange in compression indicate
a slight increase in strength and stiffness but a dramatic increase in ductility. The relations
for the case of the flange in tension indicate a substantial increase in strength (approximately
double the strength of the rectangular wall); however, a substantial decrease in ductility
is also evident. This dramatic decrease in ductility is of particular interest because of the
possibility of brittle behavior.
it can also be seen from the moment-curvature relations that the concrete stress-strain
relation proposed by Saatcioglu and Razvi (1992) predicts approxim'ately equal strengths
but significantly more ductility than the Modified Kent-Park relation (Park, Priestley and
Gill, 1982) when compressive strains exceed 0.004. For the case of the T-shaped walls when
the flange is in compression, the strain levels are so small that the two models predict almost
identical behavior; however, the two relations predict different behavior for the rectangular
walls and the T-shaped walls when the flange is in tension.
Moment-axial load interaction diagrams, for specimens RWl, RW2, TWl and TW2 are
plotted in Figs. 7.5 through 7.8, respectively. These interaction diagrams were also based

84

1
1
1
1
1

1
1

1
1
1

1
1
1
1
1
1

1
1

1
1
1

1
1

on the two different stress-strain relationships (Modified Kent-Park and Saatcioglu and
Razvi), and were calculated fora maximum compressive strain of 0.003. All of the plotted
relations were based on measured material properties (presented in Chapter 5), and confined
concrete relations were used for the concrete located in the boundary regions (although
this is not expected to have a significant impact since the relations were computed for a
maximum compressive strain of 0.003 and the unconfined concrete stress-strain relation
is identical for the two confined stress-strain relations). Figures 7.5 and 7.6 show that
the predicted moment and axial load capacities are almost identical regardless of which

confined stress-strain relation is used. Figures 7. 7 and 7.8 reveal that the predicted P-M

1
1
1
1
1

stress-strain relation that is incorporated, when a limiting compressive strain of 0.003 is

interaction diagrams for the T-shaped specimens are essentially independent of the confined

used.
Figures 7. 9 A&B plot the analytical concrete stress-strain relations used for specimens
RW 1 and RW2, respectively. The figures plot the unconfined and confined relations based on
the Modified Kent-Park relation as well as the Saatcioglu and Razvi model. The measured
concrete compressive strengths (Table 5.1) of the specimens were used for the peak stress
levels. The unconfined models are identical for the two relations; therefore, they plot on
top of each other. It is evident that the confined relation proposed by Saatcioglu and Razvi
predicts substantially higher strain capabilities than the Modified Kent-Park relation. The
concrete stress-strain relations for the T-shaped specimens would be similar to those shown
for the rectangular specimens; therefore, they are not presented.

1
1
1
1

7.3 Test Setup Requirements


The analytical work described in section 7.2 is very similar to that described in Chapter
4 except that measured material properties were used in section 7.2 and assumed material
properties were used for design in Chapter 4. The work presented in Chapter 4 was performed prior to the experimental testing in order to determine the adequacy of available

85

1
testing equipment and to assist in evaluating instrumentation requirements.
Based on the moment-curvature and moment-axial load relations it was determined
that the maximum lateral load needed to fail any of the wall specimens was approximately
70 kips (specimen TW2 when flange is in tension). The 125 kip (556 kN) hydraulic actuator
and the 150 kip (667 kN) load cell used to apply and measure the lateral load were adequate
for this purpose. Based on the analytical study, maximum displacements at the top of wall
were estimated to determine travel requirements for displacement transducers and actuators.
Based on an estimated maximum lateral drift of 3.0%, a maximum wall top displacement
of approximately 4.5 inches (114 mm) was determined. The wire potentiometers used on
the upper two stories to measure lateral displacements had 5 inch ( 127 mm) strokes
and the remaining wire potentiometers ali had strokes of 2.5 inches ( 64 mm). All of
the linear potentiometers had 2 inch ( 51 mm) strokes and the 125 kip (556 kN) lateral
actuator had a 6 inch ( 152 mm) stroke.

7 .4

Expected Failure Modes


Analytical studies performed on the rectangular wall specimens indicated that the

specimens were anticipated to fail in flexure after developing large inelastic deformations.
The ultimate failure mode of the rectangular walls was expected to be buckling of the
longitudinal boundary reinforcement (specimen RWl) or loss of confinement in the confined
core of the boundary regions (specimen RW2). Shear was expected toplaya minor role in
the behavior of the rectangular specimens.
The two T-shaped specimens were expected to behave differently from the rectangular
walls as well as each other. Failure of specimen TWl was expected to occur at the web
boundary element, when the flange was in tension. Due to the presence of large compressive
strains and the lack of adequate confinement {large spacing of hoops and inadequate depth
of confined core), a buckling failure was expected in this boundary region. The failure
of specimen TW2 was also expected to occur in the web boundary zone but after many

86

1
1

1
1
1
1
1
1

1
1
1
1

1
1
1
1
1

1
1
1

1
1
1

1
1
1
1
1
1
1

cycles of inelastic deformation. The ultimate failure of TW2 was expected to be a stability
failure (buckling of the web boundary zone over a large number of hoop spacings). Shear
was expected to contribute more to the behavior of the T-shaped walls (opposed to the
rectangular walls) due to the increase in flexural-strength; however, behavior was still
expected to be flexurally dominant.
Discussions of structural wall flexural-strengths and shear-strengths are presented in
the following two subsections, respectively.

7.4.1

Wall Flexural-Strength

The analytically predicted flexural-strength of the wall specimens were determined


from the moment-curvature analyses described above (Figures 7. 1 through 7.4), for a maximum extreme fiber compressive strain of 0.003. The expected moment capacities of the
two rectangular walls are approximately 4,600 in-kips (520 kN-m); whereas, the expected
moment capacities of the T-walls are approximately 7,400 in-kips (836 kN-m) when the
flange is in compression and approximately 10,000 in-kips (1,130 kN-m) when the flange is
in tension. It is noted that the analytical moment capacities of the T-shaped walls were
determined assuming that the entire flange width was effective in tension and compression
(as discussed in Chapter 4).
The analytically predicted moment capacities are compared with the nominal moment
capacities and the experimentally measured moment capacities in Fig. 7.lOA for all four
specimens. The nominal moment capacities were calculated assuming that the entire flange
was effective; however, an elastic perfectly plastic stress-strain curve was used for the reinforcing steel (Grade 60 steel and neglecting strain hardening), and the compressive strength
of the concrete was assumed to be 4 ksi (27.6 MPa). The experimentally measured moment
capacities that are presented in the figure were taken directly from the base moment versus
base rotation plots (Figs. 6.13 through 6.16) for each specimen.

87

The rectangular walls possess approximately equal strengths regardless of the loading
direction; therefore, the moment capacities of each rectangular wall are plotted for only one
direction of loading. The T-shaped walls have different strengths depending on the direction
of loading; therefore, moment capacities are presented for loads acting in the positive ( +)
and negative (-) directions.

It is evident from Fig.

1
1

1
1

1
7. lOA that the experimentally measured flexural-strengths

of the rectangular walls are slightly larger than the nominal and the analytically predicted
strengths, indicating that each of these walls experienced a flexural failure. Figure 7. lOA also
indicates that the experimentally measured strengths of the T-shaped walls under negative
loading (flange in tension) are slightly larger than the nominal and analytically predicted
strengths; however, the experimentally measured strengths under positive loading (flange in
compression) are approximately equal to the nominal moment capacities and significantly
less than the analytically computed strengths. Several possible causes of this discrepancy
are discussed; however, it is uncertain which (if any) of these effects is the main cause. If

1
1
1
1

the measured moment capacity does not reach or exceed the predicted capacity, then it
may be assumed that the specimen is experiencing a shear failure opposed to a flexural
failure; however, based on the experimentally observed behavior and the discussion of shear
capacity in the following section, this is not believed to be the case. It is plausible; however,
that the shear demand on the T-shaped walls prevented the boundary reinforcement from
acting solely as flexural reinforcement, thus lowering the measured capacity. Another effect
contributing to the low measured moment capacity is the eccentrically applied axial load
which causes an initial offset in the lateral load. This offset causes the experimentally
plotted results to be shifted downward, yielding a low estimate of the measured strength ( the
effects of the eccentric axial load are discussed in detail in Section 7.6.2). The analytically
computed moment capacity was based on monotonically increasing lateral loads and the
assumption that the strain distribution is linear over the entire wall length. The analytical
model also assumes that the boundary reinforcement will strain harden, resulting in a

88

1
1
1

1
1

1
high estimate of predicted strength. The experimentally measured strain data indicate the

1
1
1

1
1
1
1

1
1
1

presence of "strain relaxation" in the boundary reinforcement (due to the cyclic lateral
loading) which prevents the steel from developing the tensile strains (and stresses) that are
predicted analytically.

7.4.2 Wall Shear-Strength


The wall shear-strengths were calculated using ACI 318-89 requirements to ensure
adequate shear-strength at flexural failure. According to ACI 318-89, equation 21-6, the
nominal shear-strength of structural walls with an aspect ratio (hw / lw) greater than two,
is given by:
(7.1)
where Acv is the cross-sectional area bounded by the web thickness and the length of the
section in the direction of shear force considered,

is the concrete compressive strength,

Pn is the ratio of distributed shear reinforcement ona plane perpendicular to plane of Acv,

and f y is the yield strength of the reinforcement.


The nominal shear capacities (calculated using Eq. 7.1) are compared with the maximum experimentally measured shear forces in Fig. 7. lOB, for all four specimens. The
measured shear forces presented in the figure correspond with the maximum applied lateral
load (using data collected by the lateral load cell), or they can be calculated as the max-

imum measured moment capacity of each specimen divided by the height to the applied

1
1
1

it is evident from Fig. 7.10B that the maximum measured shear force acting on the

lateral load (150 in.).

rectangular wall specimens was substantially less (approximately 50%) than the nominal
shear capacity derived from Equation 7.1; eventhough, the amount of uniformly distributed
web reinforcement (horizontally and vertically) in each rectangular wall was only slightly
greater than the minimum amount required (Pn = 0.00327 > Pmin = 0.0025).

Earthquake Eng Res Ctr Ubrary


Univ of Calif. Berkeley
1301 S. .:.blr St. RFS 453
P:. '~md, CA 94804-4698 USA
(51 O) 665-3419

89

1
The shear demand on the T-shaped walls was substantially larger than the demand on
the rectangular walls due to the dramatic increase in the flexural-strength of the T-shaped
walls when the flanges were in tension. This increase in shear demand was unaccounted for
in the design of specimen TWl; however, specimen TW2 was designed to resist the increased
shear demand by decreasing the spacing of the uniformly distributed web reinforcement (p"
= 0.00445 in the horizontal and vertical directions).

The analytical shear capacities of

each T-shaped specimen are the same in both loading directions; however, the measured
shear force is substantially larger under negative loading (flange in tension). It is shown
in Fig. 7. lOB that under positive loading the measured shear is approximately 50-70% of
the nominal capacity; however, under negative loading the measured shear is approximately
equal to the nominal capacity.
The maximum measured shear stress on specimens TWl and TW2 were 4.85~ (psi)
and 5.46y'H (psi), measured at drift levels of approximately 1.25% and 2.5%, respectively.
These drift levels correspond to displacement ductility ratios of approximately 2.9 and 5.0,
respectively; however, it is noted that the effects of pedestal rotation will cause these ratios
to decrease slightly. Based on the experimentally observed behavior of the wall specimens,
it appears that degradation of shear-strength is not a problem when shear stresses are kept
below 6y'H, psi (0.5y'H MPa).
The relative contributions of flexure and shear to the overall response of each specimen
is discussed in the following section.

7 .5

Displacements Due To Flexure and Shear


As discussed in the previous section, it was anticipated that shear would contribute

to the overall response of the wall specimens; however, flexure was expected to dominate
the response. It was also anticipated that shear would contribute more to the response of
the T-shaped specimens than it would to the rectangular walls, due to the larger flexural
strengths of the T-shaped walls. It is of interest to determine the relative contributions of

90

1
1
1

1
1
1
1
1
1
1

1
1
1
1
1
1

1
1
1

flexure and shear to overall specimen response by comparing the displacements caused by
the effects of shear and flexure for each particular specimen.
Top Displacements:

As discussed in Chapter 6, moment-rotation relationships provide a good indication of


the contribution of flexural behavior to overall specimen response. Based on the simil~rities
of the moment-rotation and lateral load-top displacement plots for each specimen, it is
evident that the behavior of ali four specimens was predominantly flexural.

the contribution of flexural rotations at the base of the wall to the overall response, the
measured top displacements were compared with top displacements calculated using first
story deformations. The top displacements caused by the first story deformations (also
referred to as base rotations) were calculated as:
Otop,flex = (hw - h)Baae

+ 61

1
1
1
1

(7.2)

where 6top,flex is the top displacement due to the first story deformations {dominated by
flexural rotations), hw is the total height of the wall {144 in.),

1
1

To assess

is the height of the first

story (36 in.), 6ooae is the flexural base rotation, and 61 is the measured lateral displacement
at the first story level. Equation 7.2 assumes that all flexural rotation occurs in the first
story level and that above the first story the wall displaces as a rigid body. The calculated
top displacements (using Eq. 7.2) are not entirely flexural because the first story lateral
displacement {61 ) also includes the effects of shear and slippage; however, their contribution
to overall top displacement is relatively small. The contribution of shear to first story
displacement is discussed in the following subsection.
Figures 7.11 through 7. 14 plot the measured top displacements and the top displacements calculated using Eq. 7.2 for specimens RWl, RW2, TWl, and TW2, respectively.
The measured top displacements are plotted as a 45 degree line against which the base
rotation ( first story deformation) components can be compared. The experimentally measured top displacements plotted in these figures have been adjusted to compensate for the

91

1
1
effects of pedestal rotation (detailed discussion of pedestal rotation is presented in Section
7.6.2); therefore, the magnitudes of the top displacements may be significantly less than
the actual measured values (especially for the T-shaped specimens under negative loading,
where pedestal rotation accounted for as much as 20% of the top displacement ).
It can be seen in all four fi.gures that the contribution of fi.rst story deformations (base
rotations) to top displacement increased slightly as displacement levels increased. In general,
Figs. 7. 11 through 7. 14 indicate that base rotations accounted for approximately 85% of
the top displacement regardless of specimen shape (rectangular or T-shaped) or loading
direction. It is noted that the above results would indicate even higher estimates of fl.exural
contribution if the fl.exural deformations over the second, third and fourth stories were
included in the calculations; however, the percentages would decrease slightly if the top
displacements due to first story shear distortions were subtracted. The following subsection
discusses the infl.uence of shear and fl.exure on the measured first story lateral displacements.

First Story Displacements:


(a) Analytical Procedure
It is possible to determine the first story lateral displacements caused by fl.exural rotations if some assumptions are made regarding the curvature and rotation distributions over
the height of the wall. Figure 7. 15 shows the expected shear and moment diagrams for a
typical cantilever wall specimen as well as the assumed curvature and rotation distributions.
According to Hiraishi (1983), the bending deformation component in a fl.exurally dominant
structural wall can be accurately estimated as:
61,Jle:: =

where

6,/le::

a6h

(7.3)

is the fl.exural component of the first story lateral displacement, 6 is the

rotation at height h, and o is a factor that accounts for the rotation distribution over the
height of the wall. The term a is effectively the ratio of the shaded rotation region in Fig.
7.15 to the rectangular region surrounding the rotation plot, and can be determined by

92

1
1

1
1
1
1
1

,,

1
1
1

1
1
1
1
1
1
1
1
1
1
1

1
1
1
1
1
1
1

integrating the expression for rotation, or it can be assumed based on the shape (quadratic)
of the rotation distribution plot. A value of 0.67 was used for o in the following calculations.
Since the lateral displacements at the first story level are being considered, h was taken as
36 in. (914 mm) and 8 values were taken as the flexural base rotations (same values as used
in the previous subsection).
Figures 7.16 through 7.19 plot the measured first story lateral displacements and the
calculated flexural component of the displacement for specimens RWl, RW2, TWl, and
TW2, respectively. The experimentally measured first story displacements were modified
to account for the effects of pedestal rotation; therefore, the gap between the flexural
component and the experimentally measured values is attributed solely to shear distortions
in the first story. It is apparent from these plots that shear plays a larger role in first
story displacements than it did for top displacements; however, displacements are stili
dominated by flexural effects. Figures 7. 16 through 7. 19 indicate that flexure accounted
for approximately 55% to 75% of the first story displacements for the rectangular and the
T-shaped wall specimens, regardless of lateral drift level.
(b) Experimental
The analytically predicted first story displacements due to shear (gap between measured displacements and analytically predicted flexural displacements in Figs. 7. 16 through
7.19) were compared with the experimentally measured shear displacements to verify the
accuracy of the analytical procedure presented in the preceeding paragraphs. The experimentally measured first story shear distortions for the rectangular and T-shaped walls
(Figs. 6.21 through 6.24) were used in conjunction with Eq. 6.2 to calculate the experimental shear displacements at the first story height. Comparisons of the analytically and
experimentally calculated first story shear displacements of all four wall specimens yielded
excellent correlation, verifying the accuracy of the analytical technique.
(c) Elastic Theory

During the early stages of testing (prior to yielding of any reinforcement), the wall

93

1
1
specimens remain in the elastic range (drift levels

O. 5%), and elastic theory can be used

to calculate the percentage of displacements caused by shear and flexure.

Based on a

cantilever wall with a concentrated tip load, the top displacement due to flexure is given as:
6 flex

where

Pat

_ Pat h!

3EI

(7.4)

is the applied lateral load, hw is the height of the wall, E is the elastic modulus

and I is the moment of inertia of the cross-section. For the same loading condition, the
displacement due to shear is given as:

shear

Pat hw

5AG

(7.5)

w here A is the cross-sectional area and G is the shear modulus defined as:
G=

where

2(l+)

(7.6}

is Poisson's ratio (typically taken as ,...., 0.2 for concrete).

Based on elastic theory for the rectangular walls, it was determined that flexure accounted for 93% of the top displacement in the uncracked region (drift level of approximately
0.1 % ) and 96% of the top displacement after cracking occurred (drift level of approximately
0.5%}. It was also determined that flexure accounted for 44% of the first story displacement
in the uncracked region and 61 % of the first story displacement after cracking occurred.
For elastic theory of the T-shaped walls, it was determined that flexure accounted for
85% of the top displacement in the uncracked region ('"'-' 0.1% drift) and 92% of the top
displacement after cracking occurred ("' 0.5% drift). It was also determined that flexure
accounted for 26% of the first story displacement in the uncracked region and 41 % 9f the
first story displacement after cracking occurred.
In general, elastic theory predicted that flexure would account for slightly larger percentages of top displacement than the analytical procedures used in the previous section.

94

1
1

1
1
1
1
1

1
1
1
1
1

1
1
1
1

1
1
1
1
1
1

On the other hand, elastic theory predicted smaller percentages of first story displacements
due to flexural effects than the analytical procedures used in (a).
Based on the results of this section, it is concluded that the behavior of each specimen
was dominated by flexure and that shear played a relatively minor role, as expected for
specimens with aspect ratios (hw/lw) of 3. If adequate shear-strength is provided for the
structural wall (Eq. 7.1), and shear stresses are kept below

6Jl'c (psi),

it is anticipated

that shear will not have a signi:ficant impact on wall behavior. Neglecting the contribution of shear deformation on displacement response results in a conservative estimate of
flexural deformations for a displacement-based design methodology, and thus errs on the
conservative side.

1
1
1
1
1
1

1
1
1

1
1

7.6

Monotonic Wall Behavior


Wall responses for monotonically increasing lateral loads, are computed for each indi-

vidual specimen and compared with the specimen's measured hysteretic response. Section
7.6.1 discusses the calculation of the monotonic lateral load-top displacement relations. The
influence of elastic curvature, inelastic curvature, reinforcement slip, shear deformation, and
pedestal rotation on the lateral load-top displacement relations are discussed. Section 7.6.2
compares the monotonic lateral load-top displacement relations with the experimentally
obtained results for each specimen.

7 .6.1

Calculation of Monotonic Lateral Load-Top Displacement

Relations
Analytically derived monotonic relations for lateral load versus top displacements (Pa.t
vs. top) are compared with the experimentally measured results. The analytical relations
are computed by numerically integrating calculated curvatures over the height of the wall
specimens and adding displacements caused by the slippage of the reinforcement from the

95

1
base pedestal. The analytical procedure is based on the assumption that the computed
flexural response will yield good correlation with the experimentally measured response;
therefore, displacements due to shear are not included in the analytical model. Based on
the results of the previous section it is anticipated that errors resulting from the exclusion
of shear will not exceed approximately 10-15% for any of the wall specimens.
The calculated P lat vs.

top

relations include points corresponding to: cracking of the

concrete, yielding of the longitudinal tension reinforcement, and approximately four points
in the post-yielding range. The moment-eurvature relations shown in Figs. 7.1 through 7.4
are used to determine Mcr, </>er, My and </>y, as well as additional sets of moment-eurvature
data points to give a full range of expected behavior.

where

Mbaae

Mbaae

= -H-

(7.7)

is the moment taken from the moment-eurvature plots, and H is the loading

height between the point of application of the lateral load and the base of the wall (150
inches; 3.81 m). The top displacements include the effects of three distinct components and
are calculated as:
top

where

6e

the wall,

= el + pl + alip

(7.8)

is the elastic top displacement caused by the elastic curvature over the height of
pl

is the inelastic (plastic) top displacement caused by inelastic rotation in the

plastic hinge region at the base of the wall, and

Calip

is the wall top displacement caused by

slippage of the longitudinal reinforcement from the pedestal.


The displacement corresponding to cracking of the concrete includes only the contribution from the elastic curvature over the height of the wall. At this displacement level,
contributions from plastic rotation and slippage of longitudinal reinforcement from the
pedestal were determined to be negligible. The top displacement corresponding to yielding
of the longitudinal reinf~rcement includes the effects of elastic displacement and slippage

96

1
1

The lateral load corresponding to the given moment is calculated using:


Plat

1
1

1
1
1
1
1

1
1
1
1
1

1
1
1

of reinforcement from the pedestal; however, plastic rotation is stili negligible at this level.

displacement.

1
1
1
1
1
1
1
1
1

1
1
1
1

1
1

After the longitudinal reinforcement has yielded, all three effects contribute to the top

The contribution of each component included in Eq.

7. 8 is summarized in Tables

7.1 and 7.2 for the rectangular and T-shaped specimens, respectively.

The monotonic

displacements of the T-shaped walls are dependent upon which direction the lateral load is
acting; therefore, Table 7.2 lists the displacements for the lateral load acting in each direction
(flange in compression and flange in tension). The tables list the total top displacements
at three different levels (cracking of concrete, yielding of longitudinal reinforcement, and
one post-yield point near the peak displacement); as well as, the displacement attributed
to each component and the percentage of top caused by that component.

Elastic Curvature:
The elastic top displacement of a simple cantilever with a tip load is given by:

Pat h!
3EI

(7.9)

where hw is the height of the wall (144 in.; 3.66 m), and El is the stiffness of the wall and
is given by the simple expression:
M
El= -

(7.10)

</>

which represents the slope of the moment-curvature diagram.

It is shown in Tables 7. 1 and 7.2 that the elastic displacement accounts for 100% of
the top displacement at the lateral drift level that causes cracking of the concrete. The
percentage of top displacement caused by elastic curvature drops to approximately 95%
at the drift level that corresponds to yielding of the longitudinal reinforcement for the
rectangular and T-shaped specimens. When displacements approach their ultimate values,
elastic curvature accounts for approximately 10-15% of the total top displacement in the
rectangular walls and in the T-shaped walls when the flange is in compression; however,

97

they account for 20-60% of the total top displacement when the flange of the T-shaped walls
are in tension (depending on the ductility level of the T-shaped wall). It is evident from
these results that as the total top displacement increases, the percentage of the displacement
caused by elastic curvature decreases.

1
1
1

Inelastic Curvat ure:


The inelastic (plastic) top displacement is calculated using:

pl

= (<p -

'Py)

lp ( hw

- ; )

(7.11)

where <p is the curvature at the prescribed level, 'Py is the yield curvature, hw is the height
of the structural wall (144 in.; 3.66 m), and

lp

is the plastic hinge length. The plastic hinge

length was calculated using the following expression (Corley, 1966):

lp

= 0.5d + 0.2v'd (~)

(7.12)

where d is the effective depth of the member, and z is the distance from the critical section to
the point of contraflexure. Substituting values of d

= 44 inches (1.12 m) and z = 150 inches

(3.81 m; point of lateral load application), yields a plastic hinge length of 26.5 inches (673
mm). This plastic hinge length falls within the range recommended by other researchers
(Paulay, 1986) who suggest plastic hinge lengths for structural walls should lie between
0.5lw and lw, This lower-bound estimate of the plastic hinge length (26.5 in.; 0.55lw) yields

slightly larger top displacements than would be estimated ifa larger plastic hinge length
were used; however, based on experimentally observed damage at the base of the wall, this
length is believed to be reasonable, and thus it was used in the analyses. It is noted that the
plastic hinge length will increase as deformation levels increase; however, the length used
in this study was assumed to be constant for simplicity, and it is anticipated that use ofa
varying plastic hinge length would not result in significantly different results.

It is shown in Tables 7. 1 and 7.2 that the inelastic (plastic) displacement does not
account for any of the top displacement at the lateral drift levels that cause cracking of the

98

1
1

1
1
1
1

1
1

1
1
1
1
1
1

1
1
1
1

1
1
1
1
1
1

1
1

1
1
1

concrete or yielding of the longitudinal reinforcement. When the drift levels begin to exceed
those that cause yielding of the longitudinal reinforcement, the amount of top displacement
attributed to inelastic curvature (rotation in the plastic hinge region) increases dramatically.
Table 7. 1 indicates that as the displacements of the rectangular walls approach their peak
values (approximately 4 in.; 102 mm), the plastic displacements account for approximately
80% of the total top displacement. Table 7.2 indicates that the percentage (""' 80%) is similar
for the T-shaped walls when their flanges are in compression; however, the percentage ranges
from approximately 40% to 80% when the flanges are in tension, depending on the ductility
of the structural wall under this loading condition. It is evident from these results that
as the total top displacement gets larger, the percentage of the displacement caused by
inelastic curvature increases, verifying conclusions that were made in Section 7.5.

Reinforcement Slip:
Studies performed by previous researchers (Oesterle et al., 1976; 1979) indicate that
slippage of longitudinal reinforcement contributes very little to overall response (compared
to flexural contributions) for this type of specimen; however, the presence of cracking at
the wall-pedestal interface, suggested that slippage of longitudinal reinforcement from the
pedestal may be occurring. The following procedure was used to estimate the effects of
slippage of the longitudinal reinforcement from the base pedestal.
The length of reinforcement needed within the pedestal to develop the stress within
the bar was calculated by:
ld

= _!_ = f sAb
Ubdb1r

(7.13)

Ubdb1r

where T is the tensile force in the bar, calculated as the product of the steel stress (!8 )
and the cross-sectional area of the bar (Ab), Ub is the average uniform bond stress between
the concrete and the steel, and

db

is the nominal diameter of the longitudinal bar. Values

of average uniform bond stress ( ub) typically range between 700 and 1,100 psi (Devries,
Moehle and Hester; 1991, or Filippou, Popov and Bertero; 1983). A representative bond

stress of 1,000 psi (6.9 MPa) was arbitrarily selected for use in this study. Substituting

99

1
values of Is=

111 =

63 ksi (434 MPa), Ab

0.11

in 2

(71

mm 2 ),

and db = 3/8 inch (9.5 mm)

yields a development length of 5.88 inches (149 mm).


According to the provisions of section 12.5 of ACI 318-89, a development length (ldh)
of no less than 6 in. (152 mm) is required for a #3 (9.5 mm) deformed bar in tension
terminating in a standard 90 degree hook. A minimum hook extension of 12db (4.5 inches;

1
1

1
1

114 mm) is alsa required. It is noted that the longitudinal boundary steel used in all four test

specimens met (and exceeded) all ACI 318-89 requirements for anchorage of longitudinal
reinforcement. Assuming a linear steel stress distribution (valid prior to rebar yielding)
over the length ld results in a longitudinal bar elongation of:

where

Is

Is

T
21rubdbAbEs

-ld = - - - - -

2E8

(7. 14)

divided by 2E8 is the average strain over the length ld and Es is Young's modulus

for the reinforcing steel The rotation at the base of the wall caused by the elongation of
the longitudinal reinforcement is given by:

(7.15)

where jd is the distance between the longitudinal reinforcement and the resultant of the
compressive stress block (estimated as approximately 36 in.; 914 mm). Figure 7.20 is a
schematic showing the base ofa wall specimen and all the parameters necessary to determine
the base rotation caused by slippage of the longitudinal reinforcement from the pedestal.
nce the rotation at the base of the wall has been determined, the top displacement, caused
by slippage, is calculated as:
6slip

where

hw

= (} sliphw

(7. 16)

is the height of the wall (144 inches; 3.66 m). Figure 7.21 shows a schematic ofa

wall specimen and the parameters needed to calculate the top displacement caused by the
slippage of the longitudinal reinforcement from the pedestal.

100

1
1
1
1
1
1
1
1
1
1

1
1
1
1

1
1

Equations 7.14 and 7.15 are valid only prior to yielding of the reinforcement (E 11 is
constant); therefore, a different method of determining the elongation is need.ed. for postyield behavior.
Equation 7.13 is used to calculate the development length after yielding of the longitu-

1
1
1

1
1
1

dinal reinforcement. The average strain over the length required. to develop the bar stress
in excess of the yield stress is interpolated. from stress-strain diagrams (Chapter 5). Multiplying this average strain by the length required. to develop the bar stress in excess of yield
[ld = ldt - ldy, where ldt is the total development length and ldy is the yield development

length (ldt and ldy were calculated. using Eq. 7-13)] estimates the inelastic elongation in
the longitudinal reinforcing bar. The base rotation due to inelastic elongation can then be
added. to the rotation due to elastic elongation to compute the total base rotation caused.
by slippage (9alip).
The effect of reinforcement slip on the total top displacement is summarized. in Tables
7.1 and 7.2. At drift levels corresponding to cracking of concrete, it was determined. that
slippage of the longitudinal reinforcement from the ped.estal hasa negligible effect on top
At drift levels corresponding to yield of longitudinal reinforcement, Tables 7. 1 and 7.2
indicate that slippage accounts for approximately 5% of the total top displacement for the
rectangular and T-shaped. wall specimens. When top displacements approach their peak

values, Table 7.1 indicates that reinforcement slip accounts for between 1% and 8% of the

Kent-Park relation pred.icting much lower stress levels in the steel which result in lower

total top displacement for rectangular walls. The variation is the result of the Modified.

elongation estimates.

1
1

1
1

Table 7.2 indicates that slippage accounts for approximately 9%

of the total top displacement when the flange of the T-shaped. walls are in compression;
however, it accounts for only about 2% when the flange is in tension because the steel
stress is approximately at yield (strain hardening is not as prevalent when loads act in this
direction). These results verify other researchers conclusions that slippage of longitudinal
reinforcement contributes very little to overall behavior for this type of specimen.

101

1
7.6.2

Comparisons of Monotonic Lateral Load-Top Displacement

Relations
Monotonic lateral load-top displacement relations are plotted with the measured results
in Figures 7.22 through 7.25 for specimens RWl, RW2, TWl and TW2, respectively.
Due to the lack ofa ''perfect" testing arrangement, some extemally induced effects were
introduced into the measured responses (load and displacement ), that must be accounted
for prior to comparing the experimentally measured lateral load-top displacement relations
with the analytically computed monotonic relations. The experimentally induced effects
needing consideration include; pedestal rotation and initial moment caused by eccentric
axial loads. The following two subsections discuss these two effects, respectively.

Effect of Pedestal Rotation on Monotonic Behavior:


The experimentally measured top displacements were recorded from a reference frame
that was mounted firmly to the strong floor in the structural engineering research laboratory
(SERL); therefore, the measured displacements are relative to the reference frame and
not relative to the pedestal. The influence of pedestal rotation and pedestal sliding are
not accounted for in the analytical models; therefore, they must be accounted for in the
presentation of the experimental results. A linear potentiometer mounted at the base of each
wall specimen measured the lateral sliding of the pedestal during testing. The data collected
from these linear potentiometers indicated that rigid body pedestal sliding displacements
were extremely small ( < 1% of total top displacements) and thus were neglected.
Linear potentiometers mounted vertically at both ends of the pedestal were used to
measure pedestal rotations. The rotation at the pedestal-floor interface ( Bped,rot) was calculated as the difference in displacements measured by the linear potentiometers mounted
on the two ends of the pedestal, divided by the distance between the two potentiome-

1
1
1
1

1
1

1
1

1
1
1

1
1

ters (82 inches for the rectangular walls and 72 inches for the T-shaped walls). The top

102

1
1

1
1
1

1
1
1
1
1

displacements resulting from these pedestal rotations were calculated as:

ped,rot

= 171

Op,e.d,rot

(7. 17)

where 171 is the sum of the wall height (144 in.) and the pedestal depth (27 inches). Figures
7.26 through 7.29 plot the top displacements caused by pedestal rotations for specimens
RWl, RW2, TWl, and TW2, respectively. The plots shown in Figs. 7.22 through 7.25, that
compare the experimentally measured response versus the analytically computed monotonic
response have been modified to account for the effects of pedestal rotation by subtracting the
displacements caused by the pedestal rotations ( ped,rot) from the total top displacements

(top). The top displacements caused by pedestal rotation account far approximately 9% of
the total top displacement for the rectangular specimens. When the flange of the T-shaped

walls were in compression, the pedestal rotation accounted for approximately 7% of the total

1
1

rotations accounted for approximately 18% of the total top displacements (T-shaped walls

1
1
1
1

1
1
1
1

top displacement; however, when the flange of the T-shaped walls were in tension, pedestal

were much stronger under this loading condition and thus pedestal rotations were larger
since the same number of tie-downs were used in all four tests).

Initial Moment Caused by Eccentric Axial Load:


The axial load acting on the rectangular walls was applied using two hydraulic jacks
that were symmetrically placed about the specimen's centroid, resulting in a concentric
axial load. The axial load acting on the T-shaped walls was applied using four jacks which
were not placed symmetrically about the wall's centroid. The resultant axial load from the
four jacks had a small amount of eccentricity that resulted in an initial moment about the
specimen's centroidal axis. This initial moment caused curvature over the height of the wall
which resulted in an initial displacement offset at the top of the wall. The magnitude of
lateral load necessary to cause an equal displacement was calculated and used as the initial
load offset. lnitial offsets of O. 75 kips (3.3 kN) and 1.6 kips (7.2 kN) were used for specimens
TWl aid TW2 which correspond to approximately 1.2% and 2.0% of the ultimate lateral

103

1
1
load for each specimen, respectively. A larger offset was used for TW2 due to the higher
axial load level.

General:
The experimental results plotted in Figs. 7.22 through 7.25 have been modified to account for both external effects. It is evident from Figs. 7.22 and 7.23 that the analytically
predicted monotonic relations show excellent correlation with the experimental results, eventhough the effects of shear were neglected. The analytical relations based on the concrete
stress-strain relation proposed by Saatcioglu and Razvi predicts slightly higher strengths
and slightly more ductility than the predicted behavior based on the Modified Kent-Park
relation. It is noted that these analytical relations are based on the monotonic momentcurvature relations plotted in Figs. 7. 1 and 7.2 which do not account for buckling of the
longitudinal reinforcement; thus, these analytically predicted relations do not account for
buckling. Buckling of the longitudinal reinforcement is discussed in detail in Section 7. 7.

In general, the analytical relations plotted in Figs. 7.24 and 7.25 also reveal good
correlation with the experimental results; however, some discrepancies exist between experimentally observed and analytically predicted behavior. When the fl.ange of the T-shaped
wall is in compression, both concrete stress-strain relations predict identical behavior because extreme fiber compressive strains remain quite low (ec < 0.003). It is evident in Figs.
7.24 and 7.25 that the analytical relations slightly overpredict the strength and stiffness of
the T-shaped walls when the fl.ange is in compression (as discussed in Section 7.4. 1). It is
noted that, if the lateral displacements caused by shear were added to the computed results, the correlation between experimental and analytical relations would improve slightly
( analytically predicted stiffness would decrease slightly).
When the flange of the T-shaped wall is in tension the analytically predicted behavior
is dependent on the confined concrete stress-strain relation used. Prior to yielding of the
longitudinal reinforcement, both relations accurately predict strength and stiffness values
for the T-shaped walls; however, the two relations predict different post-yield response. The

104

1
1
1

1
1

1
1
1

1
1
1

1
1

1
1

1
1
1
1

relation proposed. by Saatcioglu and Razvi pred.icts more ductility than that pred.icted. using
the Modified. Kent-Park relation. nce again, it is noted. that buckling of the longitudinal
reinforcement must be accounted. for to accurately pred.ict specimen behavior.

7. 7

B uckling of Longit udinal Reinforcement


This section descri bes the analytical st udies that were performed. to predict w hen local

1
1

1
1
1
1

buckling of longitudinal reinforcement would occur. The analytically pred.icted. behavior


is compared. with the experimentally observed. behavior. The effects of buckling were witnessed in some manner in all four wall specimens. Many analytical approaches have been
recommended for determining displacement (drift) levels at which buckling of longitudinal
reinforcement is expected.; however, based. on previous studies (Thomsen and Wallace, 1994)
it has been shown that the approach proposed. by Mau (1990) is a relatively simple and
reliable proced.ure. Mau suggested. that a critical tie spacing to bar diameter ratio

(s/d,)

exists for which tangent modulus theory can be used. to pred.ict buckling loads if the actual
ratio is less than the critical ratio. Based. on measured. material properties for the #3 (9.5
mm) reinforcing hars used., a critical s/ d, ratio of 5.8 was determined. The actual spacing
to bar diameter ratios used. in the four wall specimens ranged. between 3.33 and 8. Specimens RWl and TWl each hada s/db ratio of 8 which exceed.s the critical ratio; therefore,
according to Mau the load carrying capacity of the steel cannot be pred.icted. accurately

1
1

1
1

by the tangent modulus theory and the actual load-deflection

could be unstable after

yielding occurs.
(s/d,

Specimens RW2 (s/db = 5.33) and TW2


critical value (s/ d,

3.33) had

s/d,

ratios less than the

5.8); therefore, according to Mau, the load deftection history will

closely follow the material stress-strain curve and the yield plateau will have a negligible
effect on the load-carrying capacity. The tangent modulus theory (modified Euler equation)
can be used to accurately pred.ict the inelastic buckling load.

fer= (kl/r)2

105

patl

,r2 Et

(7. 18)

1
1
where fer is the critical stress, Et is the tangent modulus of the reinforcement, l is the
unsupported length of the longitudinal reinforcement (spacing between transverse ties), k is
a constant that depends on the end restraint conditions, and r is the radius of gyration for
the longitudinal bar (for a reinforcing bar with a circular cross-section, r

0.25db). Mau

suggests using a 'k' value of 0.5 (representing fixed end conditions) to calculate the buckling
loads. The required spacing to restrain longitudinal reinforcement buckling is:

(Et=

s = ::::.
k y-;

1rdb

(Et

(7. 19)

2 y-;

where fa is the stress of the longitudinal steel, and db is the diameter of the longitudinal
reinforcement.
Prior to rebar yielding, Et

E8

= 29,000

ksi (200 GPa), f 8

::;

f 11

63 ksi ( 434

MPa; maximum value) and db = 3/8 inch (9.5 mm); therefore! the maximum spacing of
transverse steel is 12.6 inches (320 mm). All of the specimens tested had spacing less than

12.6 inches (maximum spacing of 3 inches; 76 mm); therefore, it is concluded that yielding
of longitudinal reinforcement would occur prior to buckling.
Allowable spacings for several values of steel stress and appropriate tangent moduli
are tabulated in Table 7.3. The tangent modulus values were interpolated from the stressstrain curve for the #3 (9.5 mm) reinforcing hars shown in Fig. 5.2. Table 7.4 compares
the analytically computed and experimentally observed buckling values (drift levels), for
each specimen. The steel stress at which buckling was expected was interpolated directly
from Table 7.3. The extreme fiber compressive strain corresponding to this steel stress
was than interpolated from the analytical moment-curvature relations plotted in Figs. 7. 1
through 7.4. The top displacement corresponding to this extreme fiber compressive strain
was than interpolated from the analytical monotonic lateral load versus top displacement
relations that were discussed in section 7.6. The lateral drift level corresponding to this top
displacement is listed in Table 7.4 for comparison with the observed drift level.
According to Mau, specimens RW 1 and TW 1 could not have their buckling loads accurately predicted using the tangent modulus theory because they had s/db ratios larger

106

1
1

1
1

1
1
1

1
1
1

1
1
1
1
1
1
1

1
1
1

1
1
1
1
1
1

1
1
1

1
1

than the critical value. Mau suggested that in such cases, the actual load-deflection relation
would be unstable after yielding occurs. If it was assumed that the tangent modulus theory
was valid for these two specimens, the analytically predicted buckling stress would be the
yield stress of the reinforcement (63 ksi). Based on the analytical moment-curvature plots,
the outermost longitudinal reinforcing hars in specimens RWl and TWl yielded in compression at an extreme fiber compressive strain of 0.0025. At this compressive strain level,
the cover concrete is still intact (crushing and spalling typically begin at compressive strains
of approximately 0.004) and it is believed that this cover provides support to the reinforcement which suppresses buckling. When the concrete cover has spalled from the confined
core, the longitudinal hars are no longer supported, and buckling occurs. it is assumed
in this study that the concrete cover no longer provides lateral support to the longitudinal
reinforcement when the compressive strains reach 0.006 (this is the strain level at which significant vertical splitting and crushing of the concrete cover were observed experimentally).
Based on this compressive strain, the top displacement for specimens RWl and TWl were
interpolated from the analytical monotonic lateral load versus top displacement relations.
The analytically predicted buckling drift levels were determined to be 1.93% and 0.91 % far
specimens RWl and TWl, respectively. These analytically predicted drift levels correspond
quite well with the experimentally observed drift levels and indicate that specimen TWl is
much more susceptible to buckling, than any of the other specimens.

Specimens RW2 and TW2 had s/ db ratios less than the critical value; therefore, the tangent modulus theory can be used to accurately predict the inelastic buckling loads for these
two specimens. Based on Table 7.3, specimen RW2 and TW2 are expected to experience

1
1
1

1
1

buckling of longitudinal reinforcement when the compressive stress in the longitudinal steel
reaches 72 and 83 ksi (496 and 572 MPa), respectively. The moment-curvature analyses
indicate that the extreme fiber compressive strains corresponding to these steel stresses are
0.025 and 0.045, respectively. The top displacements corresponding to these compressive
strains were calculated in the same manner as described for the monotonic lateral load ver-

107

1
sus top displacement relations. The drift levels corresponding to these displacements were
5.5% and 2.92% for specimens RW2 and TW2, respectively. it is important to note that
these drift levels are the drift levels at which local buckling is expected to occur between two
adjacent supports (between two adjacent hoops). If a different type of failure mode governs
the behavior of the specimen (such as an out-of-plane failure of the entire boundary zone or
a buckling failure that occurs over more than one spacing of the transverse reinforcement),
localized buckling of longitudinal reinforcing steel will not occur.
The analytically computed buckling drift level for specimen RW2 seems to be overpredicted because buckling of the longitudinal reinforcement in RW2 occured over more than
one spacing of the transverse reinforcement. It is concluded that at such large compressive
strain levels (0.025), there is enough damage to the confined core that some transverse
hoops can no longer provide a '4fixed" end restraint and thus buckling will occur between
the nearest two hoops that still have restraint capabilities. Specimen RW2 experienced this
type of buckling failure prior to the drift level at which localized buckling is expected.
Specimen TW2 has an analytically computed buckling drift level of2.92% which reveals
good correlation with the observed drift level of 2.5%. However, specimen TW2 experienced
an out-of-plane failure of the entire web boundary element prior to local buckling of the
longitudinal reinforcement. Based on the discussion of pedestal rotation in Section 7.6.2,
the drift level at which the observed buckling (out-of-plane failure) actually occurred was
approximately 1.8%. Conclusions for the buckling behavior of specimen TW2 are similar
to those reached for RW2, except that TW2 experienced an out-of-plane failure. The high
level of compressive strains (0.045) caused instability of the relatively narrow confined core
that eventually led to the out-of-plane failure occurring prior to the drift level at which
local buckling was predicted. Minimum wall thicknesses (as recommended in ACI 318-89)
should be ensured to prevent premature stability failures.

1
1
1
1
1
1
1
1
1
1
1
1

Buckling of Flange Reinforcement:


There is concern that walls with unsymmetrical cross-sections may be susceptible to

108

1
1

1
1
1
1
1

1
1
1
1
1
1

1
1
1
1
1

buckling of longitudinal fl.ange reinforcement (boundary and uniform web steel). Tension
in the fl.ange causes cracking of the concrete and yielding of the reinforcement, and then
when the loads are reversed, the flange is placed in compression, and buckling may occur.
UBC-94 has incorporated provisions that require confinement of the entire effective flange
width anda portion of the web (at the web-fl.ange intersection).
The vertical spacing of the transverse reinforcement (hoop and two cross-ties at each
spacing) in the web-flange intersection of specimen TWl was 3 in. (76 mm; 8db), The
uniformly distributed web steel was spaced at 7.5 in. (191 m} and no confinement was provided for this steel. Experimentally observed results indicated no adverse effects caused by
the lack of confinement in the web-flange intersection. Spacing of transverse reinforcement
(hoops only) in the web-flange intersection of specimen TW2 was increased to 4 in. (102
mm; 10. 7db) and still no adverse effects were observed. Based on the results of this study
it is concluded that buckling of flange reinforcement is not a concem (assuming that the
flange is suffi.cietly wide such that relatively small compressive strains are expected) and
that current provisions regarding confinement in the flange are overly conservative.

7 .8

Effectiveness of Transverse Reinforcement


Confinement effectiveness of the transverse reinforcement is evaluated in this section.

Transverse steel is evaluated for its ability to provide confinement of the concrete core in
the boundary zones. Based on Fig. 7.30, the following formula was used to compute the
confinement pressure exerted on the concrete by the transverse reinforcement:
fr = Atr fs

Ds

(7.20)

where Atr is the effective area of the transverse steel, f s is the transverse steel stress, D
is the distance from center to center of the transverse hoop (see Fig. 7.30), and s is the
vertical spacing of the transverse hoops. For the rectangular wall specimens, the confinement
effectiveness was determined in each principle direction (parallel and perpendicular to the

109

web of the structural wall). Table 7.5 lists the confinement pressures as well as the stress and
strain values used to calculate the pressures. Depending on the configuration of transverse
reinforcement in the boundary regions, Atr ranged from 2A., to 5A.,, where A., is the area
of the transverse steel (0.0276 in 2 ; 17.8 mm 2 ).
For the T-shaped walls, the confinement effectiveness was only evaluated for the web
boundary element because compressive strains in the ftange were relatively small and the
transverse steel did not experience significant tensile strains. The confinement effectiveness in the web boundary element was evaluated for both principle directions (as were the
rectangular walls) and the results are listed in Table 7. 5. The confinement effectiveness for
each rectangular specimen was evaluated at lateral drift levels of 1%, 1.5% and 2%. The
T-shaped walls were evaluated at these drift levels as well as O. 75 % drift. The value of steel
strains listed in Table 7.5 were interpolated from the strain histories of the transverse steel
plotted in Figs. 6.36 through 6.39. The stress values corresponding to these strains were
based on the stress-strain curve for the transverse confining steel (Fig. 5.2).
It is evident from the results listed in Table 7. 5 that the confining pressure increases as
transverse reinforcement stress increases (drift level increases); however, once the transverse
steel has yielded, little additional confining pressure is added because of the relatively
ftat post-yield stress-strain response of the 3 / 16 inch diameter wire used as transverse
reinforcement. Determination of which of the orthogonal directions governs the behavior of
the confined region of the wall is accomplished by comparison of the confining pressures in
the two orthogonal directions. The direction that results in the smaller confining pressures
is the limiting (governing) direction. It is evident that the boundary regions of all four
specimens are controlled by the confining pressures developed parallel to the web of the
specimens (long direction).

It is necessary to determine Atr for each direction prior to calculating the confining
pressure. For evaluation in the short direction (perpendicular to the web ), Atr is taken
as 2A., for all four specimens. For evaluation in the long direction (parallel to the web),

110

1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1

1
1
1

1
1
1
1

1
1
1
1

Atr is taken as 4A., for specimens RWl and TWl; however, it is taken as 2A., and 5A., for

specimens RW2 and TW2, respectively. These values are listed in parentheses in Table 7.5
beside the confining pressures.
Due to the smaller spacing oftransverse steel in specimens RW2 and TW2, the confining
pressures are considerably higher in the short direction; however, the pressures in the long
direction are similar to those of RWl and TWl due to less Atr (RW2) and larger spacing
between cross-ties (TW2). Eventhough the goveming confining pressures in specimens RW2
and TW2 are similar to those in specimens RWl and TWl, the overall behavior of RW2 and
TW2 was much better because the closer spacing of the transverse reinforcement postponed
buckling in these two specimens (as discussed in Section 7. 7).

1
1
1
1

1
1
1

1
1
1
1

111

1
Chapter - 8
VERIFICATION OF DISPLACEMENT-BASED DESIGN
FOR REINFORCED CONCRETE STRUCTURAL WALLS

8.1

1
1

1
1

Introduction
Development ofa rigorous new analytical model for predicting the response of reinforced

concrete structural walls was not an objective of this research; therefore, the analytical
studies presented in this research were based on the use of basic principles and existing
modeling techniques. The main objective of this research was to evaluate the effectiveness
of using a displacement-based methodology (such as that outlined in Appendix A) for
designing reinforced concrete structural walls with symmetrical and unsymmetrical crosssections.
This chapter is dedicated to verifying the results of the displacement-based design
procedure by comparing them with the experimentally observed behavior. In addition,

1
1

issues related to estimating effective flange widths for structural walls are discussed. Sections
8.2 and 8. 3 are aimed at verifying the results of the displacement-based design procedure for
the rectangular (symmetrical) walls and the T-shaped (unsymmetrical) walls, respectively.
Section 8.4 addresses the issue of effective flange widths for structural walls.

8.2

Strain Distributions
Verifi.cation that design procedures yield accurate, conservative results is of extreme

importance when attempting to implement these procedures into recommended design practice. The following sections present results of experimentally measured strain distributions
and compare them with analytically computed distributions (using techniques presented in
Appendix A).

112

1
1
1
1
1
1
1
1

1
1
1

RWl:
Figures 8. 1 and 8.2 plot the experimentally measured strain distributions (at the peak
displacement of the first eyde at each drift level) versus the analytically predicted strain
distributions at the base of wall specimen RWl. Figures 8.1 A&B plot the corresponding
experimentally measured steel strains (using strain gage data from longitudinal reinforcement) under positive (northward displacements) and negative (southward displacements)

1
1
1
1
1
1
1
1
1
1

loading, respectively. The analytically predicted strain distributions for three different estimates of maximum roof drift (at 1.0% drift, the design drift level of 1.5%, and 2.0%) are
plotted in each figure, along with the experimentally measured strain profi.les for the same
three drift levels. It is noted that the analytical strain distributions which are plotted in the
fi.gures are the results of hand calculations using Appendix A equations (the results were
verified using the computer program BIAX).
Due to symmetry of the rectangular walls, the analytically predicted strain distributions
under positive and negative loading are mirror images of each other; however, the experimentally measured strains experience some variation between the two loading directions,
thus the results for both loading directions have been plotted for completeness. It is evident
from the figures that the experimentally measured steel strains do not plot linearly (as assumed in the analytical predictions). The longitudinal boundary steel experiences "strain
relaxation" which causes strains to be lower in the boundary reinforcement than they are
in the reinforcement located closer to the neutral axis. The strain relaxation is caused by
the steel yielding in compression and then being subjected to large tensile demands when
the loads reverse. The longitudinal steel nearer to the neutral axis never experiences large
compressive strains (fc

0.0005); therefore, it yields in tension and continues to develop

larger strains as displacement (drift) levels increase.


In general, the analytical predictions yield accurate and slightly conservative estimates
of the maximum expected compressive strains for each drift level. The measured tensile

strains are larger than the predicted strains in the central portion of the wall web; however,

113

1
1
1

the maximum tensile strains are closely estimated even though the analytical relation is
based on monotonic behavior and the experimental results are for cyclic behavior.
Figures 8.2 A&B plot the experimentally measured concrete strains ( using displacement
data from the LVDT's) under positive (northward displacements) and negative (southward

1
1

displacements) loading, respectively. The measured concrete strains provide an average


strain value (measured over a 9 in., 229 mm, gage length at the bottom of the wall);
therefore, the effects of strain relaxation are not as pronounced as they were in the measured
steel strains (steel strains were more susceptible to localized strain effects due to the use

of extremely small strain gages). The measured concrete strains alsa include anchorage
slip at the wall-pedestal interface; however, as noted in section 7.6. 1, anchorage slip was
extremely small for the hooked hars.

lt is evident that the measured concrete strains

(Figs. 8.2 A&B) show bet ter correlation with the predicted strain distributions than do the
measured steel strains (Figs. 8. 1 A&B). The plots shown in Figs. 8.2 A&B indicate that
the proposed design procedure provides a simple, accurate method which tends to yield
slightly conservative estimates of the normal strain distribution along the length of the wall

.xj

web for wall specimen RW 1.


RW2:
Figures 8.3 and 8.4 plot the experimentally measured and analytically predicted strain
distributions (at the peak displacement of the first cycle at each drift level) at the base of
wall specimen RW2. Figures 8.3 and 8.4 are presented in an identical manner to Figs. 8. 1
and 8.2. Figures 8.3 A&B indicate lower measured strains and better correlation (less strain
relaxation) with predicted strains than did Figs. 8. 1 A&B. A majority of the strain gages
mounted on the longitudinal reinforcement broke during negative loading at 2.0% lateral
drift; therefore, an experimental strain distribution is not plotted for this loading, at 2.0%
drift.
The measured concrete strain distributions plotted in Figs. 8.4 A&B show excellent
correlation with the analytically predicted distributions. Specimen RW2 was instrumented

114

1
1

1
1
1
1
1

1
1
1

with 7 LVDT's mounted along the wall length (as opposed to 5 used for specimen RWl);

in Figs. 8.2 A&B. For the design drift level of 1.5%, the analytically predicted strains are

1
1
1

therefore, the measured strain distributions plot a little more "smoothly" than those plotted

extremely accurate (and slightly conservative) for compressive and tensile strains in both
loading directions. At 2.0% drift, the LVDT's located near the ends of the wall indicate
that measured compressive strains are slightly larger than the predicted strains. This is
attributed to the vertical splitting and crushing of the concrete cover (LVDT's were mounted
to cover concrete) at this drift level. Concrete within the confined core would not be subject
to these effects, and therefore would indicate smaller strains.

General
Based on the plotted results of experimentally measured and analytically predicted

strains for the rectangular test specimens, it is concluded that the proposed procedure

sections. It is noted that the analytical procedure shows better correlation (less variation)

1
1
1
1
1

provides accurate, slightly conservative results for structural walls with symmetrical cross-

for concrete strains than it does for steel strains. The comparison indicates that detailing
requirements for confinement and buckling can be established for cyclic loads using a simple
procedure based on monotonic loads.
The ability of the proposed procedure to predict accurate, conservative results for
structural walls with unsymmetrical cross-sections is discussed in the following section.

8.3 T-Shaped Walls


The information presented in this section is aimed at verifying the use of the proposed displacement-based design procedure for structural walls with unsymmetrical crosssections. Experimentally measured strain distributions for the two T-shaped test specimens
are compared with the analytically predicted distributions for various estimates of roof drift.

1
1
1

TWl:

115

1
1
Specimen TWl was not designed using a displacement-based design procedure, but was
formed from two rectangular wall cross-sections. The two rectangular wall sections were
designed using a displacement-based design procedure; however, as noted by Wallace and
Moehle (1992) and Wallace (1994), the behavior of T-shaped walls is significantly different

1
1

than rectangular walls.


Figures 8.5 and 8.6 plot the experimentally measured web strain distributions (at the
peak displacement of the first eyde at each drift level) versus the analytically predicted
strain distributions at the base of specimen TWl. The figures are arranged in an identical
manner to Figs. 8. 1 and 8.2 (for specimen RWl). Figures 8.5 A&B plot the experimentally
measured web steel strains under positive (flange in compression) and negative (flange in
tension) loading, respectively. Due to the unsymmetrical shape of the wall's cross-section,
the analytically predicted strain distributions are drastically different under the two loading
directions.
Under positive displacements, the compressive strains in the flange are relatively low
(~ 0.004), the depth ofthe compression zone is relatively small (rv 3 inches; 76 mm) and the
tensile strains near the end of the web are extremely high. The analytically predicted strain
distributions plotted in Fig. 8.5A indicate that the proposed procedure yields conservative
compressive strain estimates (in the flange) as well as conservative tensile strain estimates
( near the free end of the web), under this loading condition.

However, since the wall

possesses substantial deformation capacity (Fig. 7.3), accurate assessment of the strain
distribution is not critical.
When the loading is reversed, the web is put in compression and tensile strains develop
in the flange reinforcement. This loading condition results in a large depth of compression
zone ( approximately half the wall length) and approximately equal tensile and compressive
strains in the flange and web, respectively. The analytically predicted strain distributions
are directly dependent on the assumed effective flange width; therefore, an accurate assessment of the effective flange width is essential. For the design of the test specimens (Chapter

116

1
1
1

1
1

1
1
1

1
1
1
1

1
1

1
1
1
1
1

1
1
1
1
1

1
1

1
1
1
1

4), the entire fl.ange (48 in.; 1.22 m) was assumed to be effective (discussion of experimentally measured effective fl.ange widths is presented in Section 8.4), and it was assumed that
ali of the longitudinal steel in the Hange had reached yield at the design drift level. Figure
8.58 plots the strain distributions for the design drift level (1.5%) and two lower drift levels.
Due to the poor behavior of specimen TWl, failure occurred prior to reaching the design
drift level under negative loading. lnadequate transverse reinforcement in the web boundary resulted in buckling of the longitudinal boundary steel at approximately 1.25% drift,
prior to reaching the design drift level of 1.5% (this behavior was expected as discussed in
Section 4.3). As can be seen in Fig. 8.58, the experimentally measured strain distributions
show poor correlation with the analytically predicted distributions for loading in this direction. At failure, not all of the longitudinal reinforcement in the fl.ange had reached yield;
therefore, the design assumption that all the steel was yielding resulted in a conservative
design with respect to the measured strain distribution.
Figures 8.6 A&B plot the experimentally measured web concrete strains (using data
from the LVDT1s at peak displacement during the first cycle at each drift level) under
positive (flange in compression) and negative (flange in tension) loading, respectively. The
analytically predicted strain distributions plotted in Fig. 8.6A indicate good correlation
with the measured concrete strains for the positive loading condition ( similar to the results
plotted in Fig. 8.5A). Under negative displacements (flange in tension), Fig. 8.6B shows
poor correlation between experimental and analytical results for the reasons noted in the
previous paragraph.
it is concluded that in order for good correlation to occur between experimental and
analytical results, assumptions used in the design process must accurately represent the
behavior of the wall at the design level. For this to occur, an accurate assessment of the
effective fl.ange width is needed (for the desired design drift level). in general, a conservative
estimate of the effective fl.ange width is needed so that adequate transverse reinforcement is
provided for concrete confinement and to suppress buckling of longitudinal reinforcement.

117

1
On the other hand, an overly conservative design procedure results in costly construction.
The poor correlation shown in Figs. 8.5B and 8.6B (for loads acting in the negative direction) indicate that inadequate transverse reinforcement is provided at the web boundary;
therefore, large compressive strains can not be developed (buckling occurs prematurely).
The behavior of specimen TWl reinforces the need to adequately consider the geometry of
the wall and its influence on performance.
The following section discusses the results obtained from specimen TW2 which was
designed using the displacement-based design procedure outlined in Appendix A, and accounts for the influence of section geometry.

TW2:
Figures 8. 7 and 8.8 plot the experimentally measured web strain distributions ( at the
peak displacement of the first cycle at each drift level) versus the analytically predicted
strain distributions at the base of specimen TW2. Figures 8. 7 A&B plot the experimentally
measured web steel strains (at peak) under positive {flange in compression) and negative
(flange in tension) loading, respectively. Figure 8. 7A indicates that the analytical approach
yields accurate, slightly conservative strain estimates for loads acting in the positive direction.
Figure 8. 7B indicates that the analytical procedure is slightly more conservative for
loads acting in the negative direction (flange in tension) as opposed to the positive direction
(flange in compression). The analytically predicted strain distributions assume that all of
the longitudinal reinforcement in the flange is yielding. Experimentally measured strains
indicate that all of the longitudinal reinforcement in the flange had yielded at approximately
1.5% lateral drift. At a lateral drift of 1.0% the flange steel has not yielded; therefore,
the analytical strain distribution plotted in Fig. 8. 7B results in conservative estimates of
expected normal strains. it is noted that the effective flange width increases as displacement
(drift) levels increase, and that using a higher estimate of the effective flange width yields
conservative results when determining detailing requirements. Estimating effective fl.ange

118

1
1
1

1
1

1
1

1
1
1
1
1

1
1

1
1
1
1
1

1
1
1
1

1
1
1

1
1

widths is discussed. in more detail in Section 8.4.


Figures 8.8 A&B plot the experimentally measured. web concrete strains under positive
(flange in compression) and negative (flange in tension) loading, respectively. Figure 8.8A
indicates excellent correlation (and slightly conservative analytical pred.ictions) for loads
acting in the positive direction. The LVDT's used to measure the displacements at the base
of the wall ( at the free end of the web) reached their maximum stroke capacities at a drift
level of approximately 1. 75%; therefore, strain distributions at 2.0% drift were not plotted.
(the steel strains at 2.0% are plotted in Fig. 8.7).
Figure 8.8B indicates that the analytical procedure yields conservative strain pred.ictions for the negative loading direction (flange in tension), and that the correlation between
experimentally measured. concrete strain and analytically predicted. strain (Fig. 8.8B) is
better than the correlation shown in Fig. 8. 7B (for steel strains). it is concluded. that
the concrete strains (measured. using the LVDT's) are not influenced as much by the localized. effects (such as concrete cracking) that have an impact on the measured. steel strains.
The effective tension reinforcement in the flange has not yielded at O. 75% or 1.0% drift as
was assumed. in the analytical proced.ure; therefore, the analytically pred.icted. strains are
relatively conservative at these drift levels.

General
Based. on the plotted. results of experimentally measured. and analytically pred.icted.
strains for the T-shaped. test specimens, it is concluded. that the proposed. displacementbased. design proced.ure yields accurate and conservative strain estimates for unsymmetrically shaped. structural walls.
Use ofa good conceptual design is imperative to ensure adequate performance of structural walls. Design based. on independent response of web and flange (such as assumed. for
specimen TWl) has the potential for brittle behavior, as witnessed. in this study.
When the flange of an unsymmetrically shaped. wall is in compression, the wall possesses

119

1
substantial ductility and minimum amounts of transverse reinforcement in the fl.ange results
in good behavior. When the flange is in tension, proper detailing of the web boundary region
is essential to ensure good performance. An accurate estimate of the effective tension fl.ange
width is essential in comparison with experimental studies; however, in general, a quick
procedure that is simple enough for design, yet tends to be conservative, is needed so that
adequate detailing is provided.
The predicted compressive strains in the end of the wall web are directly dependent on
the tensile force developed in the fl.ange (amount of effective tension reinforcement and the
stress developed in the effective reinforcement). it was assumed in these analyses that the
entire fl.ange was effective and that all of the effective reinforcement was yielding (this yields
conservative results). Experimental results confirmed this assumption at the design drift
level of 1.5%. However, for drift levels less than 1.5%, the tensile stresses developed in the
fl.ange reinforcement are less than yield, and dependent on the distance from the web as well
as the drift level. The plots comparing experimentally measured and analytically predicted
strain distributions for the fl.ange in tension (figures designated with a 'B') indicate that
the analytically predicted strains are over estimated (conservative) when the flange steel is
not yielding in tension (drift levels less than 1.5%). When the fl.ange steel is yielding (drift
levels larger than 1.5%), the procedure yields accurate and slightly conservative results.

lf the fl.anges of the structural walls tested in this study were increased in length, it
would be imperative to accurately assess the effective fl.ange width. For fl.exural-strength
purposes, it would be conservative to use a slightly low estimate of the effective fl.ange
width; however, when evaluating detailing requirements and shear demands for structural
walls, use of a slightly high estimate of the effective fl.ange width would be conservative.
The following section discusses various provisions for estimating the effective fl.ange widths
of structural walls.

120

1
1
1

1
1
1

1
1

1
1

1
1
1
1

1
1
1

1
1
1
1
1

8.4 Effective Flange Widths


The results presented. in the previous sections clearly show that the use ofa displacementbased. design methodology (such as that outlined in A ppendix A) provides design engineers
with a flexible, effective proced.ure for designing reinforced. concrete struct ural walls with
symmetrical and unsymmetrical cross-sections. However, effectiveness of the proced.ure for
design of unsymmetrical walls is enhanced. with an accurate estimate of the effective flange
wid th. The following subsections address the issue of estimating the effective flange widths
for structural walls, and discuss the experimentally measured. effective fl.ange widths of the
walls tested. in this st udy.

Methods For Estimating Flange Widths

{1) ACI 318-89

'

widths of structural walls; however, provisions are included. for estimating the effective flange

1
1
1
1
/1

1
1
1

The ACI 318-89 building code does not make recommendations for effective fl.ange

widths of T-beams (ACI 318-89, Section 8. 10), which can be extrapolated to structural walls
if the span of the beam is taken as twice the height of the cantilever wall. The provisions
for beams are used. to compute an effective flange width that can be used. for evaluation of
both positive and negative flexural-strength. In general, the effective width in compression
will not significantly infl.uence wall behavior ( strength or ductility); therefore, the effective
flange width estimate should be based. on providing a reliable value for the case when the
flange is in tension. Based. on the extrapolation of the ACI 318-89 provisions for T-beams,
the limits placed. on the effective flange widths of structural walls are given as:

beff = min (l/4 = hw/2; tw

+ 16t;

tw

+ sc)

(8.1)

where 1 is the length of the equivalent beam (2hw), hw is the height of the cantilever wall
specimen, tw is the thickness of the wall web, t is the thickness of the wall flange, and sc
is the clear spacing between webs of adjacent walls. If the structural wall only hasa flange

121

1
on one side of the wall (L-shaped), then the limits placed on the effective fl.ange width are:
bej/= min (tw

+ l/12 =

tw

+ hw/6;

tw

+ 6t;

tw

+ sc/2)

(8.2)

(2) Paulay
Paulay (1986) suggests that effective fl.ange widths should be determined assuming that
the forces in the web spread out ata 2;1 and l;l slope away from the web, for tension and
compression, respectively. lf this method is employed, the effective tension fl.ange widths
would be approximately half the width of the compressive fl.ange widths, which complicates
the design process. Paulay noted that the estimates of effective fl.ange widths represent a
compromise due to the impossible task of determining a unique effective fl.ange width when
walls are subjected to inelastic deformations. As the rotations at the base of the wall (in the
plastic hinge region) increase, the amount of fl.ange steel that becomes effective increases;
therefore, the effective fl.ange width appears to be directly dependent on the deformation
level imposed on the wall.

(3) Pantazopoulou and Moehle


Pantazopoulou and Moehle (1990) conducted a study on reinforced concrete beamcolumn connections with slabs connected monolithically to the supporting beams. The
primary objective of the study was to determine the effect of the slab on the stiffness and
strength of the longitudinal beam in frames subjected to lateral loads. The effective fl.ange
width was linked to a compatibility requirement which stated that the elongation occuring
in the plastic hinge region at the web-fl.ange intersection must be developed in the fl.ange
also. Because the elongation is the product of the end rotation and the beam depth, it was
concluded that as end rotation and beam depth increased, the width of the effective fl.ange
would increase. The results of the analytical model were compared with experimentally
observed behavior ofa quarter-scale frame structure tested by Qi (1986). Experimental
results indicated that for an interior beam (slab on both sides) the effective slab width on
each side of the beam web was equal to approximately 1.6 beam depths at yield, increasing

122

1
1
1
1

,1

'1

1
1
1
il

1
1
1
1

1
1

1
1

1
1
1

1
1

to over three beam depths at ultimate. Exterior beams (slab on one side) also indicated an
effective slab width on each side of the beam web of 1.6 beam depths at yield; however, the
width did not increase as deformations increased. it was concluded that the effective flange
width was limited by the torsional or weak-axis flexural capacity of the transverse beam.

lf the transverse beam was flexible, or if there was a lot of pullout, the effectiveness of the
flange was reduced. The analytical model predicted that the slab contribution increases
linearly until drift levels reach approximately 1.5% of the story height, after which very
little additional contribution is noted. The dependence of effective slab widths on the
deformation level (lateral drift) is similar to the case of estimating an effective flange width
for walls.

(4) Sittipunt and Wood


Sittipunt and Wood (1993) tested two C-shaped structural walls with two 36 in. (0.91
m) long parallel webs and a 60 in. (1.52 m) long intersecting flange. This flange length
was purposely chosen to be longer than the effective width defineci by ACI 318-89, Section

1
1

effective flange widths. All of the vertical steel in the flange was effective in providing tensile

1
1
1
1

1
1

(76 mm) thick flange.

Based on the analytical studies performed to

8.10, for a 3 in.

compare with experimental results, Sittipunt and Wood concluded the following, regarding

resistance when the flange was in tension (at drift levels of approximately 2.0%); therefore,
it was concluded that the effective tension flange width could be as high as ten times the
thickness of the flange (larger than the limiting value of six flange thicknesses listed in Eq.
8.2). it was also concluded that use ofa too small effective width can lead to a significant

underestimation of the strength of the wall, that may lead to the selection of inadequate
transverse reinforcement in the boundary regions opposite the flange (as noted by Wallace
and Moehle; 1992). it was also concluded that when the flange is in compression, using
an effective flange width which is smaller than the actual effective flange width yields a
conservative estimate of the wall's strength; however, this may lead to underestimation of
tensile strain in the longitudinal web reinforcement.

123

1
(5) 1994 Uniform Building Code
The Uniform Building Code (UBC-94) has recently incorporated. a displacement-based.
design approach for determining the detailing requirements need.ed. at wall boundary regions.
According to UBC-94, Section 1921.6.5.2, the effective flange widths to be used. in design of
unsymmetrical cross-sections (I, L, C and T-shaped walls) shall not be assumed to extend
further from the face of the web than (1) one half the distance to an adjacent shear wall
web, or (2) 10 percent of the total wall height.

(6) Summary
Recommendations for effective fl.ange widths give considerable variation. Extrapolation
of ACI 318-89 equations for T-beams is based. on the beam length (wall height) or the
number of slab thicknesses (fl.ange thicknesses). Paulay and UBC-94 also recommend an
~ffective flange width that varies with wall height. However, Pantazopoulou and Moehle
recommend an effective flange width that varies with wall length (beam depth). The increase
in effective flange width with drift level is well established.; however, incorporating this
trend into the design process has been hampered since design using expected. maximum
displacement response has not been considered. in building codes until UBC-94.
Relatively few experimental studies have been conducted. to investigate effective flange
widths for unsymmetrical walls.

The study by Sittipunt and Wood observed. that the

effective flange width for a C-shaped. wall could reach 0.28hw at 2% lateral drift; whereas,
effective flange widths of 0.31hw at 1.5% lateral drift were observed. for the T-shaped walls
tested as part of this study.
Extrapolation of ACI relations for T-beams (Eq. 8-1, 8-2) yields effective flange widths
of 0.50hw and 0.17hw, for T-walls and C-walls, respectively. Paulay recommends using an
effective tension flange width of 0.5hw, and UBC-94 recommends using O. lhw, Based on
these observations, it is apparent that Eq. 8.2, and particularly UBC-94, yield low effective
fl.ange width estimates for C-shaped and L-shaped. walls; however, it should be noted that

124

1
1
1
1
1

1
1

1
1

1
1

1
1

i
1

1
1

1
1

the 0.28hw value reported by Sittipunt and Wood is for 2.0% lateral drift which may be
beyond what is expected for a reasonably configured structural wall building.
For T-shaped walls, ACI 318-89 and Paulay values of 0.5hw appear to give an upper
bound estimate; therefore, these approaches may be quite conservative for stiff buildings.
The following subsection discusses the experimentally measured effective fl.ange widths for
the unsymmetrically shaped walls tested in this study.

1
1
1
1
1
1
1

1
1

1
1

Experimentally Measured Effective Flange Widths


Effective fl.ange widths of the structural walls tested in this study were determined
to compare with the estimates obtained using the methods described in the preceeding
subsection. The measured effective fl.ange widths were determined as:
bef!

( E~';{~b)

.
(48 n.}

(8.3)

!11

where f sis the tensile stressin a given longitudinal bar (both #2 and #3 hars were located
in the fl.ange), Ab is the area of the longit udinal bar, f 11 is the yield stress of the reinforcement
("'63 ksi; 434 MPa for #2 and #3 hars), and 48 inches is the actual length of the fl.ange.
in essence, Eq. 8.3 is calculating the average tensile stress accross the fl.ange and dividing
by the yield stress of the reinforcement to determine the ratio of effective width (belf) to
actual width ( 48 in., 1.2m).
Table 8. 1 presents the effective fl.ange widths for specimens TWl and TW2 as determined by Eq. 8.3. The fl.ange widths are expressed in absolute lengths (inches), as a
percentage of the wall length (lw = 48 in.), asa percentage of the wall height (hw

144 in.),

and as a number of fl.ange thicknesses on each side of the wall web. it is evident from Table
8. 1 that the entire fl.ange does not become effective in tension until a drift level of approximately 1.5%. Specimen TWl never reached this drift level due to a premature buckling
failure at approximately 1.25% drift. Based on the experimentally measured fl.ange widths,
it appears that the widths listed in Eq. 8. 1 provide a conservative upper-bound limit. At
a drift level of 2.0%, specimen TW2 had an effective fl.ange width of tw

125

llt (0.34hw)

1
compared. to the

liniting

value of tw

+ 16t

(0.4 7hw ), given in Eq. 8. 1. UBC-94 provisions

would recommend using an effective fl.ange width for specimens TWl and TW2 of tw

2.6t (O. lOhw ), which would drastically underestimate the participation of the flange. it
should be noted. that lateral drift levels of 2. 0% are not very common for buildings utilizing structural walls for lateral load resistance; however, even at lower drift levels, UBC-94
provisions for effective flange widths are extremely low.
The T-shaped. walls tested. in this study included. a fl.ange that was equal to the wall
length or one-third of the wall height. However, unlike T-beams, concentrated. reinforcement
was used. at the flange boundaries. The distributed flange reinforcement consisted. of #2
hars at 7.5 in. (191 mm), whereas the longitudinal boundary steel was comprised. of 8-#3
hars. The total area of tension reinforcement outside of the web-fl.ange intersection was 2. 15
in 2 (16-#3 hars and 8-#2 hars). If only uniformly distributed. fl.ange reinforcement was
used., a fl.ange width of 165 in. (4.2 m) would be needed. to accomodate this reinforcement.
Based. on this comparison, the test can be considered. to be demanding. It also points
out the importance of properly considering the influence of flange boundary reinforcement
on wall behavior. Flange longitudinal boundary reinforcement that is within an effective
flange width must be included. as effective tension reinforcement; however, it may be unsafe
to exclude this steel if it lies just beyond the effective fl.ange width estimate.
General
Based. on the experimental and analytical results of this study and the recommendations
of previous researchers, the following conclusions are made regarding effective flange widths:

1
1

1
1
1
1
1

1
1
1

(1) use ofa conservative estimate of the effective flange width is essential for safe design
of structural walls, (2) when evaluating (estimating) the flexural-strength and stiffness of

a structural wall a conservatively low estimate of the effective flange width should be used.;

and detailing, (3) the effective fl.ange width should be selected to provide a conservative

1
1

126

however, slightly overpred.icting the fl.exural-strength is not critical if adequate ductility is


provided; therefore, it is recommended. to use the same flange width for strength, stiffness

1
1
1
1
1
1

1
1
1

design with respect to detailing requirements and shear-strength, (4) the limits set farth
in Eq. 8. 1 provides a reasonable upper-bound estimate of the effective flange width far
T-shaped walls and need not be exceeded, (5) the limits set forth in Eq. 8.2 may slightly
underestimate an upperbound effective flange width for L-shaped walls. As an alternative,
one-half of the values of Eq. 8.1 for T-shaped walls could be used (hw/4; tw

tw

sc/2).
The potential to incorporate effective flange widths that vary with drift level should be
considered. To do this, analytical studies similar in scope to those by Pantazopoulou and
Moehle should be conducted far unsymmetrical walls. lmportant parameters should include
wall length, wall height, flange width, reinfarcing ratios (concentrated longitudinal boundary
reinforcement for web and flange, as well as distributed web and flange reinfarcement), axial
load and deformation leveL The influence of shear should also be considered.

1
1

1
1

1
1
1

1
1

+ 8t;

127

1
1
Chapter - 9
SUMMARY AND CONCLUSIONS

9.1

1
1

Summary
The performance of buildings which utilize reinforced concrete structural walls for

lateral load resistance has been documented to be very good even during severe seismic
events.

Previous studies have shown that buildings possessing a significant number of

strategically placed structural walls have performed extremely well during seismic events,
even though detailing of many of the walls did not meet current U.S. design practice.
U.S. codes have recently undergone a major change in philosophy for design of structural
walls. The UBC-91 used a strength-based approach for the design of structural walls with
nominal details; whereas, UBC-94 has recently incorporated a displacement-based design
methodology to assess detailing requirements. ACI 318 is currently considering use of a
displacement-based design for walls.
The experimental and analytical study described in this report had many objectives.
The primary objective was to evaluate the effectiveness of using a displacement-based design procedure for designing reinforced concrete structural walls with symmetrical and unsymmetrical cross-sections. Secondary objectives included: (1) addressing the design (and
behavior) of "flanged" structural walls, (2) addressing the issue of effective flange widths
for unsymmetrical wall cross-sections, (3) evaluating the accuracy of simplified analytical
models for predicting the behavior of structural walls, and ( 4) reviewing key provisions of
the recently adopted code (UBC- 1994).

128

1
1

1
1
1

1
1
1
1
1

1
1
1

1
1
1
1
1
1

9.2

Conclusions
The following subsections discuss conclusions regarding the behavior of reinforced con-

crete structural walls subjected to simulated seismic loads. Topics which are addressed
include: (1) general conclusions regarding the behavior of structural walls, (2) the use ofa
displacement-based design methodology for design of reinforced concrete structural walls,
(3) design of RC structural walls with unsymmetrical cross-sections, and (4) analytical
techniques used to model the behavior of RC structural walls.

9.2.1

General Behavior

The behavior of all four test specimens was flexurally dominant, as expected for structural walls with an overall aspect ratio, hw/lw, greater than two. Lateral load versus top
displacement plots, base moment versus base rotation plots, and strain histories were ali
characteristic of flexurally dominant response.
Shear effects were slightly higher for the T-shaped specimens than for the rectangular

specimens; however, shear played only a minor role in the behavior of all four test specimens.

1
1

significantly influence specimen behavior. lt is concluded that exclusion of slippage would

1
1
1

Slippage of longitudinal reinforcement from the pedestal was very small and did not

not adversely effect the results of the simplified analytical models.

9.2.2

Displacement-Based Design of Structural Walls

The experimental and analytical studies described in this report have shown that
displacement-based design methodologies are extremely ftexible and effective for evaluating
structural wall behavior. Using a displacement-based approach results in wall designs that
are directly related to the building configuration, as well as the wall aspect ratio, the wall
axial load level, the wall cross-sectional configuration and the wall reinforcing ratios.

1
1

Normal Strain Distributions:

129

1
A displacement-based. design proced.ure (Wallace, 1994; 1995) determines the expected.
normal strain distribution along the wall length, from which detailing requirements are established.. The pred.icted. strain distributions were compared. to strain distributions that
were measured. using strain gages and displacement transducers and were found to be accurate and slightly conservative for walls with both rectangular and T-shaped. cross-sections.
Therefore, wall normal strain distributions derived from expected displacement response,
equilibrium requirements, and monotonic load-deformation (moment-curvature) relations
provide a reliable technique for establishing detailing requirements at wall boundaries.
Limiting Compressive Strain:
Detailing requirements for the end regions of a structural wall are directly related. to the
maximum expected. compressive strain at that location. A maximum expected. compressive

1
1

strain of 0.004 is incorporated. into the design proced.ure proposed by Wallace (1994), as
a division between areas that require special transverse reinforcement and those that do
not. UBC-94 recommends using a limiting strain of 0.003. Experimental results indicate
compressive strain levels of approximately 0.005 at the base of the wall prior to observing
vertical splitting, and strain levels of approximately 0.006 prior to spalling of the concrete
cover. Therefore, UBC-94 requirements appear to be overly conservative and use of 0.004
is recommended.. For compressive strains less than 0.004, use of minimum reinforcing steel
(as recommended by current codes) in areas where the compressive strain is expected. to be
less than 0.004, results in adequate behavior.
Failure Modes:
When maximum expected. compressive strains exceed 0.004, special detailing requirements are recommended. for concrete confinement and suppressing buckling of longitudinal
reinforcement. The proposed. displacement-based design procedure incorporates a simple
relationship to determine the amount of transverse reinforcement required. As expected.
compressive strain levels increase from 0.004 to 0.008, the percentage of transverse rein-

1
1

1
1
1
1

forcement increases from 20% to 100% of that required. by current codes (ACI 318-89).
This strain dependent relation often results in wall designs that require considerably less

130

'

1
1

1
1

1
1
1
1

1
1

1
1

transverse reinforcement (more economical) than required by current codes (ACI 318-89,
UBC-91).
The proposed displacement-based design procedure lets design engineers predict the
governing mode of failure (confinement, buckling, shear) so that premature failures can be
avoided and good behavior can be ensured. Provisions for ensuring adequate shear-strength
are discussed in the following subsection.

The proposed displacement-based design procedure provides guidelines that reasonably predict the maximum expected shear demand at the fl.exural capacity of the structural
wall. Design of uniformly distributed web reinforcement is directly related to the maximum
expected shear demand. Degradation of shear-strength was not witnessed in any of the
walls tested in this study. it is concluded that if the maximum shear stress is kept below

6/Fc psi (0.5/Fc MPa), that shear-strength degradation will not be a concern even for
large deformation demands (if walls have similar aspect ratios and axial load levels). Determination of shear demand from the expected fl.exural capacity, and a limiting shear stress

6/Fc (psi) are recommended to avoid the possibility of poor performance due to shear.

of

nated with a standard hook and fully developed in the boundary region at the end of the

1
1
1
1

'1
1

UBC-94 requires that uniformly distributed horizontal web reinforcement be termi-

wall. Specimens tested in this study used horizontal web reinforcement that were not fully
developed in the boundary regions, but were terminated with a 90 degree hook into the
boundary region. Based on experimentally observed results, use of 90 degree hooks on horizontal web reinforcement into the boundary region resulted in adequate specimen behavior.

If possible, these hooks should not be placed into the edge of the boundary region where
the greatest level of deterioration would be expected.

9.2.3

Structural Walls with Unsymmetrical Cross-Sections

Use of a good conceptual design that adequately considers the effects of the crosssectional shape of the structural wall is critical. Conclusions regarding the displacement-

131

1
based. design of structural walls with unsymmetrical cross-sections are discussed in the

following subsections.
Normal Strain Distributions:
A displacement-based. design methodology can be used. to effectively design structural
walls with unsymmetrical cross-sections as well as walls with symmetrical cross-sections.
Comparison of analytically computed. and experimentally measured. strain distributions
yielded. good correlation. Pred.icting accurate strain distributions is much more important
when the fl.ange of an unsymmetrically shaped. wall is in tension. This requires a reasonable

1
1

estimate of the effective flange width.


Effective Flange Widths:
An accurate assessment of the effective flange width is an essential component in evaluating the need for special detailing (transverse reinforcement) at the boundaries of unsymmetrically shaped. structural walls, as well as shear-strength requirements. Stable, hysteretic
behavior of unsymmetrically shaped. structural walls can be achieved. if the wall is well detailed., such as specimen TW2 tested. in this study. Effective flange widths recommended.
by ACI 318-89 (extrapolated from provisions for T-beams) provide an upper-bound estimate which need not be exceed.ed for T-shaped. walls. Values for L-shaped and C-shaped.
walls should be taken as half of those values used. for T -shaped. walls. U se of the same
effective fl.ange width in tension and compression is recommended. to simplify wall design.
Provisions in UBC-94 significantly under-estimate the fl.ange participation and could lead
to poor behavior.

1
1
1
1

1
1

Shear:
Shear demands are considerably higher on unsymmetrically shaped. structural walls
than they are on symmetrical walls due to the increase in flexural-strength when the flange is
in tension. To accurately assess the shear demand foran unsymmetrically shaped. structural
wall, it is recommended. that a moment-curvature analysis be conducted. (based. on probable
material properties) to estimate wall fl.exural-strength. Based. on the wall's flexural-strength
and an assumed. lateral force distribution over the height of the wall, a reasonable estimate

132

1
1

1
1
1

1
1
1

of the expected shear demand can be determined far which adequate shear-strength can be
provided.
Detailing Requirements:
Detailing requirements at the free web boundary are much more critical than the detailing provided in the web-flange intersection or the flange. When transverse reinfarcement
at the critical locations is designed using the proposed displacement-based design proce-

dure, stable hysteretic behavior is achieved. Detailing of transverse reinfarcement within

adequate behavior was observed. Compressive strains were low in the flange; therefare,

1
1
1
1
1

1
1
1
1
1

the flanges of the walls tested in this study did not meet current code provisions; however,

confining steel was not needed and buckling was not a problem even though the flange reinfarcement was subjected to cyclic loads. Far unsymmetrical walls, it is unlikely that special
transverse reinfarcement is required at the web-flange intersection; therefare, UBC-94 requirements far confining the web-flange intersection and the entire effective flange width
are overly conservative and, in general, would not result in improved wall behavior.

9.2.4

Analytical Modeling of Structural Wall Behavior

Using simple moment-curvature relations and well established theories to account for
the distribution of elastic and inelastic defarmations over the height of the wall, it is possible
to accurately predict the behavior of slender reinfarced concrete structural walls. The exclusion of shear defarmations from analytical models is not critical far slender walls similar
to those tested in this study; therefare, design based solely on flexural theory is recommended far structural walls whose behavior is expected to be flexurally dominant (hw/lw
~

2). Neglecting the contribution of shear defarmations on displacement response results

in a conservative estimate of required flexural deformations to achieve a specified drift level


far a displacement-based design methodology, and thus errs on the conservative side.
Accurate prediction of structural wall behavior requires a firm understanding of all
possible failure modes. Buckling of longitudinal reinforcement played a role in the behavior
of all four specimens tested in this study. It is concluded that the ultimate behavior of

133

1
many structural walls will be governed by buckling of longitudinal reinforcement. Use of
the modified Euler equation with an effective length factor (k) of 0.5, provides a reasonable approach for predicting the spacing of transverse reinforcement necessary to suppress
buckling.

1
1

9.3 Suggested Future Research


The research presented in this report has successfully shown that the use ofa displacementbased design methodology (such as that recommended by Wallace; 1995) provides design
engineers with a simple, flexible procedure for designing reinforced concrete structural walls
with symmetrical and unsymmetrical cross-sections, whose behavior is flexurally dominant.
Additional research is needed that assesses the effectiveness of using a displacement-based
design methodology to design reinforced concrete structural walls with openings ( "pierced"
and "coupled" walls) and walls with lower aspect ratios. To do this, the influence of shear
on the deformation capacity of structural walls is required; therefore, detailed experimental
and analytical studies that address the shear-strength of lower rise walls (hw/lw .$ 1.5) are
needed. Important parameters would appear to be: the wall configuration ( aspect ratio and
cross-sectional shape), material properties

u;,

/11), reinforcing ratios (p, p', and p11 ), and

the deformation (displacement, drift) level.


A comprehensive study of specimens with varying flange widths is also needed to develop guidelines for determining the effective flange widths of structural walls. Guidelines
that link effective flange widths of structural walls to the expected drift level and/ or aspect

1
1
1
1

1
1
1

ratio of the wall should be considered.


Future studies that consider biaxial loadings of flanged structural walls are recommended to help assess the effect of the web steel when the lateral loads act parallel to
the flange. Studies of complete building systems may be helpful in evaluating the biaxial
loading response of structural walls.

134

1
1
1
1

1
1

1
1
1
1

1
1
1

REFERENCES
1. Abrams, D. P. (1991), "Laboratory Definitions of Behavior for Structural Components
and Building Systems," ACI Special Publication 127, Vol. 127, No. 4, pp. 91-152,
April 1991.
2. ACl-318 (1989), "Building Code Requirements for Reinforced Concrete," American
Concrete lnstitute, Detroit, Michigan, 1989.

3. Aktan, A. E., and Bertero, V. V. (1985), "Reinforced Concrete Structural Walls: Seismic Design for Shear," Journal of Structural Engineering, ASCE, Vol. 111, No. 8, pp.
1775-1791, August 1985.
4. Ali, A., and Wight, J. K. (1990), "Reinforced Concrete Structural Walls with Staggered
Opening Configurations Under Reversed Cyclic Loading," Report No. UMCE 90-05,
Department of Civil Engineering, University of Michigan, Ann Arbor, April 1990.
5. Ali, A., and Wight, J. K. (1991), "Reinforced Concrete Structural Walls with Staggered

Door Openings," Journal of Structural Engineering, ASCE, Vol. 117, No. 5, pp. 1514-

6. Applied Technology Council (1978), "ATC-03-06: Tentative Provisions for the Devel-

1531, May 1991.

opment of Seismic Regulations for Buildings," Applied Technology Council, Palo Alto,
California, 1971.

7. Aschheim, M., and Moehle, J. K. (1992), "Shear-Strength and Deformability of Rein-

1
1

Report No. UCB/EERC-92/04, Earthquake Engineering Research Center, University

1
1
1
1

forced Concrete Bridge Columns Subjected to lnelastic Cyclic Displacements," EERC

of California at Berkeley, February 1992.


8. Buchanan, G. R. (1988), Mechanics Of Materials, Holt, Rinehart and Winston, ine.,
New York, NY, 720 pages, copyright 1988.
9. Carvajal, O., and Pollner, E. (1983), "Muros de Concreto Reforzados con Armaduro
Minima," Baletin Technico, Universidad Central de Venezuela, Facultad de lngenieria,
Ano 21 (72-73), Enero-Diciembre, 5-36 (in Spanish).
135

1
, / 10. Corley, W. G. (1966), "Rotational Capacity of Reinforced. Concrete Beams," Journal
of the Structural Division, ASCE, Vol. 92, No. ST5, pp. 121-146, October 1966.
11. DeVries, R. A., Moehle, J. P., and Hester, W. (1991), "Lap Splice Strength of Plain
and Epoxy-Coated. Reinforcements," Report No. UCB/SEMM-91/02, Structural Engineering Mechanics and Materials, University of California at Berkeley, January 1991.
12. EERI (1986), "The Chile Earthquake of March 3, 1985," Earthquake Spectra, Vol. 2,
No. 2, April 1986.
13. Filippou, F. C., Popov, E. P., and Bertero, V. V. (1983), "Effects of Bond Deterioration on Hysteretic Behavior of Reinforced. Concrete Joints," EERC Report N o.

1
1

1
1
1

UCB/EERC-89/19, Earthquake Engineering Research Center, University of Califor-

nia at Berkeley, August 1983.


14. Goodsir, W. J. (1985), "The Design of Coupled. Frame-Wall Structures for Seismic
Actions," Research Report 85-8, Department of Civil Engineering, University of Canterbury, Christchurch, New Zealand, August 1985.
15. Hiraishi, H. (1983), "Evaluation of Shear and Flexural Deformations of Flexural Type
Shear Walls," Proceedings, Fourth Joint Technical Coordinating Committee, U.S.Japan Cooperative Earthquake Research Program, Building Research Institute, 1983.
16. Kabeyasawa, T., Hiraishi, H., and Kumagai, H. (1994), "Tests and Analyses of HighStrength Shear Walls in Japan," Proceedings, Second U.S.-Japan-New Zealand-Canada
Multilateral Meeting on Structural Performance of High-Strength Concrete in Seismic
Regions, Honolulu, Hawaii, November 1994.
17. Mau, S. T., (1990), "Effect of Tie Spacing on Inelastic Buckling of Reinforcing Bars,"
ACI Structural Journa~ Vol. 87, No. 6, pp. 671-677, Nov.-Dec. 1990.
18. Meigs, B. E., Eberhard, M. O., and Garcia, L. E. (1993), "Earthquake-Resistant Systems For Reinforced. Concrete Buildings: A Survey Of Current Practice," Report No.
SGEM 99-9, Department of Civil Engineering, University of Washington, Seattle, October 1993.
136

1
1

1
1

1
1
1

1
1
1

1
1
1
1
1
1
1

1
1
1

19. Moehle, J. P. (1992), "Displacement-Based Design Of Reinforced Concrete Structures


Subjected to Earthquakes," Earthquake Spectra, Earthquake Engineering Research Institute, Vol. 8, No. 3, 1992.
20. Moehle, J. P., and Wallace, J. W. (1989), "Ductility and Detailing Requirements of
Shear Wall Buildings," Proceedings, Fifth Chilean Conference on Seismology and Earthquake Engineering, Santiago, Chile, pp. 131-150, August 1989.
21. New Zealand Standard Code of Practice for Design of Concrete Structures (1982), NZS
3101, Standard Association of New Zealand, Wellington, New Zealand, 1982.
22. Newmark, N. M., and Hall, W. J. (1982), "Earthquake Spectra and Design," Engineering M onographs on Earthquake Criteria, Structural Design, and Strong M otion
Records, Earthquake Engineering Research Institute, 1982.

23. Oesterle, R. G., Fiorato, A. E., Johal, L. S., Carpenter, J. E., Russel, H. E., and Corley,
W. G. (1976), "Earthquake Resistant Structural Walls - Tests of Isolated Walls," Report
to the National Science Foundation, Construction Technology Laboratories, Portland

Cement Association, Skokie, IL., 315 pp., November 1976.

24. Oesterle, R. G., Aristizahal-Ochoa, J. D., Fiorato, A. E., Russel, H. E., and Corley, W.

II," Report to the National Science Foundation, Construction Technology Laboratories,

1
1
1

1
1
1

G. (1979), "Earthquake Resistant Structural Walls - Tests of Isolated Walls - Phase

Portland Cement Association, Skokie, IL., 325 pp., October 1979.


25. Oesterle, R. G. (1986), "Inelastic Analysis of In-Plane Strength of Reinforced Concrete Shear Walls," PhD Dissertation, Northwestern University, Evanston, Illinois, June
1986.
26. Pantazopoulou, S. J., and Moehle, J. P., (1990), "Truss Model for 3-D Behavior of R.C.
Exterior Connections," Joumal of Structural Engineering, ASCE, Vol. 116, No. 2, pp.
298-315, February 1990.
27. Park, R., Priestley, M. J. N., and Gill, W. D. (1982), "Ductility of Square-Confined
Concrete Columns," Journal of the Structural Division, ASCE, Vol. 108, No. ST4, pp.
137

929-950, April 1982.


28. Park, R., and Paulay, T. (1975), Reinforced Concrete Structures, John Wiley and Sons,
New York, 1975.
29. Paulay, T., and Priestley, M. J. N. (1992), "Seismic Design of Reinforced Concrete and
Masonry Buildings," John Wiley and Sons, New York, 744 pp., 1992.
30. Paulay, T., Priestly, M. J. N., and Synge, A. J. (1982), "Ductility in Earthquake
Resisting Squat Shear Walls," Proceedings ACI Journal, Vol. 79, No. 4, pp. 257-269,

1
1
1
1

August 1982.
31. Paulay, T. (1986), "The Design of Ductile Reinforced Concrete Structural Walls for
Earthquake Resistance," Earthquake Spectra, Vol. 2, No. 4, pp. 783-823, October
1986.
32. Paulay, T. (1991), "Seismic Design Strategies for Ductile Reinforced Concrete Structural Walls," Proceedings Int. Conf. on Buildings with Load Bearing Concrete Walls
in S eismic Zones, French Assoc. for Earthquake Engineering and the French Assoc.

for Construction, Paris, France, AFPS, Saint-Remy-Les-Chevreuse, France, 397-421,


1991.
33. Priestley, M. J. N., and Limin, H. (1990), "Seismic Response of T-Section Masonry
Shear Walls," Proceedings, Fifth North American Masonry Conference, University of
Illinois at Urbana-Champaign, pp. 359-372, June 1990.
34. Priestley, M. J. N., Verma, R., and Xiao, Y. (1994), "Seismic Shear Strength of Reinforced Concrete Columns," Journal of Structural Engineering, ASCE, Vol. 120, No. 8,
pp. 2310-2329, August 1994.
35. Qi, X., (1986), "The Behavior ofa RC Slab-Column Subassemblage Under Lateral
Load Reversals," CE299 Report, Structural Engineering and Structural Mechanics,
Department of Civil Engineering, University of California, Berkeley, 1986.
36. Qi, X., and Moehle, J. P. (1991), "Displacement Design Approach for Reinforced Concrete Structures Subjected to Earthquakes," Report No.
138

UCB/EERC-91/2, Earth-

1
1
1
1

1
1
1
1

1
1
1

1
1
1
1
1

1
1

quake Engineering Research Center, University of California, Berkeley, January 1991.


37. Riddell, R., Wood, S. L., and De La Llera, J. C. (1987), "The 1985 Chile Earthquake,
Structural Characteristics and Damage Statistics for the Building Inventory in Vina
del Mar," Structural Research Seri.es No. 594, Univ. of Illinois, Urbana Ill., 1987.
38. Saatcioglu, M., Razvi, S. R., (1992), "Strength and Ductility of Confined Concrete,"
Journal of Structural Engineeri.ng, ASCE, Vol. 118, No. 6, pp. 1590-1607, 1992.

39. Salmon, C. G., and Johnson, J. E., (1990), "Steel Structures: Design and Behavior,"
Third Edition, Harper & Row, 1990, 1086 pp.

40. SAP90, Computers and Structures, Inc., Berkeley, California, 1990.


41. Shimazaki, K., and Sozen, M. A., (1984), "Seismic Drift of Reinforced Concrete Structures," Research Reports, Hazama-Gumi, Tokyo, 1984.

42. Shiu, K. N., Daniel, J. I., Aristizabel-Ochoa, J. D., Fiorato, A. E., and Corley, W.

out Openings," Report to the National Science Foundation, Construction Technology

43. Sittipunt, C., and Wood, S. L. (1993), "Finite Element Analysis of Reinforced Con-

1
1
1

1
1
1
1

G. (1981), "Earthquake Resistant Structural Walls - Tests of Walls with and with-

Laboratories, Portland Cement Association, Skokie, IL., 120 pp., July 1981.

crete Shear Walls," Report to the National Science Foundation, Department of Civil
Engineering, University of Illinois at Urbana-Champaign, December 1993.
44. Thomsen IV, J. H., and Wallace, J. W. (1994), "Lateral Load Behavior of Reinforced
Concrete Columns Constructed Using High-Strength Materials," ACI Structural Journa~ Vol. 91, No. 5, pp. 605-615, Sept.-Oct. 1994.

45. Thomsen IV, J. H., and Wallace, J. W. (1994), "T-Shaped Shear Walls: Design Requirements and Preliminary Results of Cyclic Lateral Load Testing," Proceedings, Fifth
U.S. National Conference on Earthquake Engineering, Chicago, Illinois, Vol. II, pp.
891-900, J uly 1994.
46. "Uniform Building Code (1991)," International Conference of Building Officials, Whittier, California, 1991.
139

1
47. "Uniform Building Code (1994)," Intemational Conference of Building Officials, Whittier, California, 1994.
48. Wallace, J. W. (1992), "BIAX: Revision 1 - A Computer Program for the Analysis of
Reinforced Concrete and Reinforced Masonry Sections," Report No. CU/CEE-92/4,
Structural Engineering, Mechanics and Materials, Clarkson University, 1992.
49. Wallace, J. W. (1994a), "A New Methodolog;y for Seismic Design of Reinforced Concrete Shear Walls," Joumal of Structural Engineering, ASCE, Vol. 120, No. 3, pp.
863-884, March 1994.
50. Wallace, J. W. (1994b ), "Displacement-Based Design of Reinforced Concrete Structural Walls," Proceedings, Fifth U.S. National Conference on Earthquake Engineering,
Chicago, Illinois, Vol. II, pp. 191-200, July 1994.
51. Wallace, J. W. (1995), "Seismic Design of Reinforced Concrete Structural Walls: Part
I: A Displacement-Based Code Format," Jov.mal of Strv.ctv.ral Engineering, ASCE,
Vol. 121, No. 1, January 1995.
52. Wallace, J. W., and Moehle, J. P. (1989), "The 3 March 1985 Chile Earthquake: Structural Requirements for Bearing Wall Buildings," EERC Report No. UCB/EERC-89/5,
Earthquake Engineering Research Center, University of California, Berkeley, July 1989.
53. Wallace, J. W., and Moehle, J. P. (1992), "Ductility and Detailing Requirements of
Bearing Wall Buildings," Jov.mal of Strv.ctv.ral Engineering, ASCE, Vol. 118, No. 6,
pp. 1625-1644, June 1992.
54. Wallace, J. W., and Moehle, J. P. (1993), "An Evaluation of Ductility and Detailing
Requirements of Bearing Wall Buildings Using Data From the March 3, 1985 Chile
Earthquake," Earthqv.ake Spectra, Earthquake Engineering Research Institute, Vol. 9,
No. 1, February 1993.
55. Wallace, J. W., Moehle, J. P., and Martinez-Cruzado, J. (1990), "lmplications for the
Design of Shear Wall Buildings Using Data From Recent Earthquakes," Proceedings,
140

1
1
1
1

J
1

1
1
1
1
1
1
1
1

1
1

1
1

1
1
1
1
1
1

Fourth U.S. National Conference on Earthquake Engineering,

Paln

Springs, California,

Vol. 2, pp. 359-368, May 1990.


56. Wallace, J. W., and Thomsen IV, J. H. (1993), "Seismic Design of Reinforced Concrete
Structural Walls," Report No. CU/CEE-99/16, Department ofCivil and Environmental Engineering, Clarkson University, Potsdam, New York, August 1993.
57. Wallace, J. W., and Thomsen IV, J. H. (1995), "Seismic Design of Reinforced Concrete
Structural Walls: Part II : Applications," Journal of Structural Engineering, ASCE,
Vol. 121, No. 1, January 1995.
58. Wood, S. L. (1989), "Minimum Tensile Reinforcement Requirements in Walls," ACI

Stroctural Journal, Vol. 86, No. 4, pp. 582-591, September-October 1989.


59. Wood, S. L. (1991), "Performance of Reinforced Concrete Buildings During the 1985
Chile Earthquake: lmplications For the Design of Structural Walls," Earthquake Spec-

tra, Vol. 7, No. 4, November 1991.


60. Wood, S. L., Wight, J., and Moehle, J. P. (1987), "The 1985 Chile Earthquake, Obser-

vations on Earthquake-Resistant Construction in Vina del Mar," Stroctural Research

61. Yan, W., and Wallace, J. W. (1993), "Analytical Studies of Four Shear Wall Buildings

partment of Civil and Environmental Engineering, Clarkson University, Potsdam, New

Series No. 592, Univ. of Illinois, Urbana Illinois, 1987.

Using Data From Recent Califomia Earthquakes," Report No. CU/CEE-99/15, De-

York, 141 pages, July 1993.

1
1
1

1
1
1

141

1
Table 3.1: Prototype Wall Design Parameters
Rectangular

Cross-Section

Base Level

As

10-#11

us

10-36 mm

A's

10-#11

A"s

Floors 4-6
6-#11

us

6-36 mm
6-#11

us

Flange in Comp.
10-#11

us

10-36 mm
30-#11, 16-#5

T-Shaped (E-W)

(N-S)
Flange in Tension
30-#11, 16-#5
30-36, 16-16 mm
10-#11

us

10-#11

us

10-36 mm
10-#11

us

10-36 mm

6-36 mm

30-36, 16-16 mm

10-36 mm

10-36 mm

#5@ 14"

#5@ 14"

#5@ 7"

#5@ 7"

20-#11, 48-#5

16 mm@ 18 cm

20-36, 48-16 mm

0.00508
0.0152

16 mm

p
p'
p"

us

T-Shaped

35.5 cm

16 mm

35.5 cm

0.00508

0.00305

0.00508

16 mm@ 18 cm
0.0168 1

0.00508

0.00305

0.0168 1

0.00508

0.00277

0.00277

0.00554

0.00554

0.00508

0.0040

0.00089

0.00873

0.00539

41.1 (104.4)

9.23 (23.4)

90.4 (230)

55.8 (142)

c (in.; cm)
4

19.2 (48.8)

49.0 (124.5)

19.2 (48.8)

0.0041

0.0011

0.0075

0.007

c (in.; cm)

42.9 (109)

11.4 (29.0)

74.1 (188)

72. 9 (185.2)

19.2 (48.8)

34.6 (87.8)

31.2 (79.2)

c,ma:c

c (in.; cm)
1

c,ma:c
1

c (in.; cm)
1
2
3
4

lncludes Contribution of Flange Reinforcement


Includes Contribution of Web Reinforcement (Perpendicualr to the Fla.nge)
Based on Proposed Displacement-Ba.sed Design Approach (Hand Calculations)
Based on Sectional Analysis Program (BIAX)

142

1
1
1

1
1
1

1
1

1
1
1

1
1
1
1
1
1

1
1
1

Table 3.2: Shear-Strength Requirements For Prototype Walls


Vcode

1
1

1
1
1
1

Specimen

Vn

251 (1,116)

899 (3,999)

Eq. A17

669 (2,977)

3.44

M-<t> Analysit

694 (3,087)

3.57

251 (1,116)

1,409 (6,267)

Eq. A17

669 (2,977)

3.44

M-<t> Analysis 3

1,389 (6,178)

7.13

T-Wall (E-W):

251 (1,116)

899 (3,999)

Eq. A17

669 (2,977)

3.44

M-</> Analysis

856 (3,807)

4.41

Based on Even Distribution of Base Shear ta Six Walls


Based on ACI 318-89, Equation 21-6 (Eq. A16)
Based on Critical Case of Flange In Tension

1
1
1

E
1

:::

kips (kN)

1
1

Ve~ad

kips (kN)

T-Wall (N-S):

Ve:q,ected

kips (kN)

Rect angtlar:

143

1
1
1

Table 3.3: Prototype Wall Design Forces 1

UBC-91

Computer (3-D)

Dynamic Response

Simplified

Equivalent Static

S pectrum Analysis

Hand Analysis

Analysis

(Rw=12)

Rectangular Prototype Wall Forces (N-S Direction)


Axial Load; kips

1,229 2

1,297

1,299

(kN)

(5,467)

(5,769)

(5,778)

Shear; kips

351 3

256

229

(kN)

(1,561)

(1,139)

(1,019)

Moment; kip-ft

16,848

9,676

8,801

(kN-m)

(22,846)

(13,121)

(11,934)

T-Shaped Prototype Wall Forces (N-S Direction)


Axial Load; kips

1,178 2

1,154

1,158

(kN)

(5,240)

(5,133)

(5,151)

Shear; kips

351 3

388

352

(kN)

(1,561)

(1,726)

(1,566)

Moment; kip-ft

16,848

17,794

16,201

(kN-m)

(22,846)

(24,129)

(21,969)

~ All Forces Are Based on Factored Loads


3

Based on Tributary Area (Fig. 3. 1)


Based on Even Distribution of Base Shear To Six Walls

1
1
1

1
1
1

1
1
1
1
1

1
1
144

1
1
1
1
1

Table 3.4: Prototype Wall Design Displacements

1
1
1
1
1
C

1
1

Simplified

Level

(1)

Analysis
(2)

Ali Members 2

Col. (3) x

100% 19

./23

All Members
50% lg

(3)

(4)

(5)

6.71

9.49

9.24

La ter al Displacements (inches)


Roof - (6th Floor)

1
1
1
1
1

Floor

9.23

(5th Floor)

5.37

7.59

7.32

(4th Floor)

3.98

5.63

5.38

(3rd Floor)

2.61

3.69

3.48

(2nd Floor)

1.38

1.95

1.80

(1 st Floor)

0.45

0.64

0.55

Fundamental Period (N-S Direction)


T (sec)
1
2

1.02

0.598

0.846

0.825

Using Displacement-Based Design Philosophy (Hand Calculations)


Columns (3)-(5) Are Based On 3-D Elastic Response Spectrum Analyses (R10 =l)
Displacements And Period Of Column (3) Are Multiplied By v'2 To Account For Cracking

Earthquake Eng Res Ctr Ubrary


Unlv of calif. Berkeley
1301 S. 46th St. RFS 453

RlchmOnd, CA

948Q4-4698 USA

(51 O) 665-3419

145

Table 4.1: Design Parameters of Model Wall Specimens


Specimen

As

RWl
8-#3

us

8-9.5 mm

A'8

8-#3

us

8-9.5 mm
8-#3

us

24-#3, 8-#2

24-#3, 8-#2

24-9.5, 8-6.4 mm

24-9.5, 8-6.4 mm

8-#3

us

8-9.5 mm

8-9 ..5, 2-6.4 mm

#2@ 7.5"

#2@ 7.5"

#2@ 7.5"

#2@ 5 . .5"

6.4 mm@ 191 mm

6.4 mm @ 191 mm

p
p'

0.00458

0.00458

191 mm
0.01579 2

0.00458

0.00458

0.004.58

0.00.509

p''

0.00327

0.00327

0.00327

0.0044.5

0.00603

0.00603

0.014

0.0137

10 ..58 (269)

10 ..58 (269)

24.57 (624)

24.04 (611)

4.8 (122)

4.8 {122)

17..5.5 (446)

17.02 (432)

0.00.5

0.00.5

0.01.5

0.012

c (in.; mm)
(in.; mm)
4

c,max

6.4 mm

Cl

6.4 mm

Cl

140 mm
0.01579 2

c (in.; mm)

8.8 (224)

8.8 (224)

28.2 (716)

21.2 (.537)

c'

4.8 (122)

4.8 (122)

20.7 (.526)

14.1 (3.59)

1
2
3
4

(in.; mm)

For Critical Condition When Flange Is in Tenion


lncludes Contribution of Ali Flange Reinforcement (Based on be//=48 in.)
Based on Proposed Displa.cement-Based Design Approach (Eq. All)
Based on Sectional Analysis Program (BIAX)

8-#3, 2-#2

8-9.5 mm

c,max

8-#3

TW2 1

8-9.5 mm

A"a

us

TW 1

RW2

1
1
1

1
1
1
1
1
1

1
1

146

1
1
1
1
1
1
1

1
1
1
1

1
1
1
1

Table 4.2: Shear-Strength Requirements For Model Walls


Vn

Ve:cpected

tlc::l!cccd

kips (kN)

kips (kN)

psi (kPa)

RWl

61.95 (276)

32.64 (145)

170 (1,172)

2.69

RW2

61.95 (276)

32.64 (145)

170 (1,172)

2.69

TW1 3

61.95 '(276)

68.06 (303)

354 (2,444)

5.60

TW2 3

75.55 (336)

69.44 (309)

362 (2,496)

5.72

Based. on ACI 318-89, Equation 21-6 (Eq. A16)


Based. on M - </> analysis
L\ ,.,, : .
, ,
j-,,c..c
Based. on Critical Case of Flange in Tension

1
1
1

1
1
1
147

d) """'-+

'1H

Specimen

1
1

1
1

Vezpected

i,N'(

l.d,

;. ,c .

...

Table 5.1

Concrete Compressive and Tensile Strengths

Comp. Strength

Comp. Strength

Rupture Strength

at 28 Days

at Test Date

at Test Date

(psi; MPa)

(psi; MPa)

(psi; MPa)

Pedestal

5,464; 37. 7 (2) 1

7,586; 52.3 (1)1

979; 6. 75 (2) 2

#1

1st Story

3,723; 25.7 (2)

4,580; 31.6 (2)

653; 4.50 (3)

(RWl)

2nd Story

5,334; 36.8 (3)

5,871; 40.5 (1)

787; 5.43 (4)

3rd Story

4,828; 33.3 (2)

5,632; 38.8 (2)

725; 5.00 (3)

4th Story

7,109; 49.0 (2)

8,462; 58.4 (2)

Pedestal

5,173; 35. 7 (2)

6,967; 48.0 (3)

890; 6. 14 (2)

#2

1st Story

3,838; 26.5 (2)

4,925; 34.0 (2)

653; 4.50 (3)

(TWl}

2nd Story

5,388; 37.2 (3)

6,071; 41.9 (3)

787; 5.43 (4)

3rd Story

4,775; 32.9 (2)

6,419; 44.3 (2)

725; 5.00 (3)

4th Story

3,360; 23.2 (2)

4,156; 28.7 (2)

Pedestal

4,174; 28.8 (2)

5,005; 34.5 (2)

729; 5.03 (2)

#3

1st Story

4,678; 32.3 (2)

6,331; 43.7 (3)

817; 5.63 (2)

(RW2)

2nd Story

5,606; 38. 7 (2)

6,632; 45. 7 (3)

950; 6.55 (2)

3rd Story

5,217; 36.0 (2)

5,918; 40.8 (3)

830; 5. 72 (2)

4th Story

5,164; 35.6 (3)

5,987; 41.3 (3)

754; 5.20 (2)

Pedestal

4,465; 30.8 (2)

5,022; 34.6 (2)

729; 5.03 (2)

#4

1st Story

5,677; 39.1 (2)

6,048; 41. 7 (3)

817; 5.63 (2)

(TW2)

2nd Story

5,446; 37.6 (2)

6,413; 44.2 (3)

950; 6.55 (2)

3rd Story

5,199; 35.9 (2)

5,894; 40.6 (3)

830; 5. 72 (2)

4th Story

5,164; 35.6 (3)

5,600; 38.6 (3)

754; 5.20 (2)

Specimen

Story Level

Numbers in parentheses indicate the number of 6" x 12" cylinders tested


Numbers in p~rentheses indicate the number of rupture beams tested

148

1
1
1

1
1
1

1
1
1
1
1
1
1
1
1
1
1
1
1

1
1
1
1
1
1

1
1
1
1
1
1
1
1

Table 5.2 Concrete Mix Proportions

Specimens: RWl, RW2, TWl, TW2


(Quantity / yd 3 )

Materials

Coarse Aggregate
1,224 lbs.

(3/8" Peastone)
Fine Aggregate

1,894 lbs.

(Sand)
Portland Cement

580 lbs.

(Type II)
Air-Entraining Agent

5.8 oz.

DAREX II A.E.A.
Water Reducing Agent

17.4 oz.

WRDA with Hycol


ASTM C-494, Type A

Water

325 lbs.

Total

4,023 lbs.

1
1
1

1
1

149

1
1
1

Table 5.3 Reinforcement Properties


#3 Deformed. Bar

#2 Deformed. Bar

3/ 16 in. Diameter

Longit udi nal

Uniformly Distributed.

Smooth Wire

Boundary Steel

Web Steel

Transverse Steel

f yi ks~ (MPa)

63 (434)

65 1 (448)

63 (434)

f ui ksi (MPa)

"J93 (641)

"J85 (586)

70 (483)

f fi ksi (MPa)

"J92 (634)

"J80 (552)

70 (483)

29,000 (200,000)

29,000 (200,000)

29,000 (200,000)

1,500 (10,344)

fy

0.0022

0.003

fsh

0.016

fu

0.10

0.08

0.06

f/

0.12

0.10

0.07

Properties

ksi (MPa)

Es;
2

Esh ;

1
2

ksi (MPa)

Yield stress calculated. using 0.2% offset method


Tangent Modulus at the Onset of Strain Hardening

0.003

1
1
1

1
1
1
1

1
1

1
1
1
1

1
150

1
1

1
1

Table 5.4 Chronology Of S pecimen Construction and Testing


Date

Event

Sept. 18, 1992

Placed Concrete for Pedestal of Specimen RWl

Oct. 7, 1992

Placed Conc. for Pedestal of Specimen TW 1

Dec. 16, 1992

Placed Conc. for First Stories of Specimens RWl and TWl

Jan. 7, 1993

Placed Conc. for Second Stories of Specimens RWl and TWl

Jan. 28, 1993

Placed Conc. for Third Stories of Specimens RWl and TWl

May 7, 1993

Placed Conc. for Fourth Story of Specimen TWl

May 9, 1993

Placed Conc. for Fourth Story of Specimen RWl

Oct. 16-17, 1993

Tested Specimen RWl

Nov. 4, 1993

Placed Conc. for Pedestals of Specimens RW2 and TW2

Nov. 18, 1993

Placed Conc. for First Stories of Specimens RW2 and TW2

Nov. 23, 1993

Placed Conc. for Second Stories of Specimens RW2 and T\Y2

Dec. 3, 1993

Placed Conc. for Third Stories of Specimens RW2 and TW2

Dec. 15, 1993

Placed Conc. for Fourth Stories of Specimens RW2 and TW2

Jan. 28-29, 1994

Tested Specimen TWl

Feb. 22 - Mar. 7, 1994

Tested Specimen TW2

April 14-16, 1994

Tested Specimen RW2

1
1
1

1
1
1
1
1

1
1

151

Table 6.1: Loads, Top Displacement and Shear Stress of Specimen RWl
Top Displacement
Cycle

Data Point

Number

Numbers

lnitial Cycle

Drift
inches

(%)

Repeat Cycle

North
Load

Shear Stress

South
1

North

South

Load

Shear Stress

Load

Shear Stress

Load

Shear Stress

(kips)

J1ic (psi)2

(kips)

J1ic (psi)

(kips)

J1ic (psi)

(kips)

J1ic (psi)

l and 2

51-115

0.144

0.1

7.92

0.61

7.48

0.58

7.87

0.61

7.41

0.57

3 and 4

116-179

0.36

0.25

15.36

1.18

14.67

1.13

14.62

1.13

14.03

1.08

5 and 6

180-243

0.72

0.5

25.15

1.94

26.84

2.07

26.13

2.01

25.99

2.00

7 a.nd 8

244-307

1.08

0.75

30.32

2.33

30.34

2.34

29.16

2.24

29.38

2.26

......

9 and 10

308-371

1.44

.oo

30.55

2.35

31.67

2.44

29.90

2.30

31.66

2.44

l'-'

11 and 12

372-435

2.16

1.50

31.20

2.40

32.75

2.52

30.65

2.36

32.31

2.49

13 and 14
153

436-499

2.88

2.00

31.83

2.44

33.35

2.57

30.84

2.37

31.94

2.46

500-515

4.32

3.00

30.55

2.35

C11

1 Shear Streu Bued on Grosa Area (192 in 2 )


2
/~
4,580 pi
3
Specimen Failed Before Reaching Peak Diaplacement in Firat Half Cycle

------------------Table 6.2: Loads, Top Displacement and Shear Stress of Specimen RW2
Initial Cycle

Top Displacement
Cycle
Number

......

C1'

C,.)

Data Point
Number

inches

Repeat Cycle

Drift

North

South

(%)

Shear Stress

I;oa.d

North

South

Load

Shear Stress 1 Load

Ji'c (psi)2

Shear Stress

Load

Shear Stress

(kips)

(kips)

Ji'c (psi)

(kips)

Ji'c (psi)

(kips)

Ji'c (psi)

1 and 2

21-85

0.144

0.1

9.46

0.62

10.30

0.67

9.74

0.64

9.70

0.63

3 and 4

86-149

0.36

0.25

15.03

0.98

17.49

1.15

15.55

1.02

16.53

1.08

5 and 6

150-213

0.72

0.5

23.55

1.54

25.28

1.65

23.77

1.56

24.45

1.60

7 and 8

214-277

1.08

0.75

28.46

1.86

29.86

1.95

27.79

1.82

29.48

1.93

9 and 10

278-341

1.44

1.00

30.31

1.98

31.01

2.03

29.74

1.95

30.80

2.02

11 and 12

342-405

2.16

1.50

31.97

2.09

32.24

2.11

31.70

2.08

31.81

2.08

13 and 14

406-469

1.44

1.00

25.00

1.64

25.48

1.67

24.91

1.63

26.35

1.73

15 and 16

470-533

2.16

1.50

31.31

2.05

31.24

2.05

32.10

2.10

32.67

2.14

17 and 18

534-597

2.88

2.00

34.19

2.24

33.96

2.22

33.83

2.21

34.01

2.23

19 and 20

598-661

3.60

2.50

35.50

2.32

35.13

2.30

31.62

2.07

33.77

2.21

1 Shear Streu Ba.ed on Gro Area (192 in 2 )


2 !'C
6,330 psi

Table 6.3: Loads, Top Displacement and Shear Stress of Specimen TWl
lnitial Cycle

Top Displacement
Cycle
Number

1 and 2

.....

c:.,

,.:..

Data Point
Numbers

inches

Drift

North

(%)

Load

Shear Stress 1

(kips)
9.94

1-65

0.144

0.1

Repeat Cycle
South

North

South

Shear Stress

Load

Shear Stress

Load

Shear Stress

Jfic (psi)2

Load
(kips)

Jfic (psi)

(kips)

Jfic (psi)

(kips)

Jfic (psi)

0.74

12.19

0.90

9.88

0.73

12.29

0.91

3 a.nd 4

66-129

0.36

0.25

19.82

1.47

22.55

1.67

20.14

1.49

22.21

1.65

5 and 6

130-193

0.72

0.5

30.04

2.23

41.67

3.09

30.20

2.24

40.25

2.99

7 and 8

194-257

1.08

0.75

38.02

2.82

48.91

3.63

37.74

2.80

47.79

3.55

9 and 10

258-321

1.44

1.00

40.18

2.98

57.32

4.25

40.10

2.97

56.38

4.18

113

322-385

2.16

1.50

42.42

3.15

65.37

4.85

124

386-401

2.88

2.00

43.93

3.26

1 Shea.r Stres Ba.sed on Cross-Sectional Area of Web (192 in 2 )


2
= 4,930 psi
3 Specimen Experienced Brittle Buckling Failure During Southward Loading of First Cycle
4 Half Cycle Performed to Straighten The Buckled Boundary Reinforcement in The Web

f:

-----------------Table 6.4: Loads, Top Displacement and Shear Stress of Specimen TW2
lnitial Cycle

Top Displacement
Cycle
Number

.....
C11
C11

Data Point
Number

inches

Drift

North

(%)

Shear Stress 1

Load
(kips)

JPc (psi)2

Repeat Cycle
South

North

South

Load

Shear Stress

Load

Shear Stress

Load

JPc (psi)

Shear Stress

(kips)

(kips)

JPc (psi)

(kips)

JPc (psi)

1 and 2

21-85

0.144

0.1

15.48

1.04

16.99

1.14

15.59

1.04

16.91

1.13

3 and 4

86-149

0.36

0.25

23.67

1.58

33.78

2.26

23.32

US6

33.16

2.22

5 and 6

150-213

0.72

0.5

32.34

2.17

48.80

3.27

32.06

2.15

47.75

3.20

7 and 8

214-277

1.08

0.75

35.60

2.38

55.42

3.71

35.16

2.35

57.46

3.85

2.48

9 and 10

278-341

1.44

1.00

37.05

65.11

4.36

36.75

2.46

63.22

4.23

11 and 12

342-405

2.16

1.50

39.47

2.64

74.53

4.99

37.25

2.49

72.49

4.85

13 and 14

406-469

1.44

1.00

34.22

2.29

42.40

2.84

34.16

2.29

42.73

2.86

US and 16

470-533

2.16

1.50

39.81

2.67

69.24

4.64

39.05

2.61

69.88

4.68

17 and 18

534-597

2.88

2.00

41.85

2.80

81.26

5.44

40.71

2.73

78.53

5.26

19

598-629

2.88

2.00

40.48

2.71

77.87

5.21

20 and 21

63().-693

3.60

2.50

42.65

2.85

81.56

5.46

41.41

22
23 3

694-725

3.60

2.50
2.704

41.06

2.75

68.60

4.59

1
2
3
4

726-741

3.89

41.49

2.78

2
Shear Stres Baaed on Croe-Sectional Area of Web (192 in )
/~ = 6,050 psi
Only Half Cycle Performed Because Web Boundary Element Wu Beginning To Buckle Out-of-Plane
Couldn't Achieve Full Displa.cement Because Actuator Rea.ched Ma.ximum Stroke

2.77

78.87

5.28
-

aa

1
1
Table 7.1

Monotonic Top Displacement Components of Rectangular Walls

Column Top Displacements 1

Cracking
Specimens

Yielding

el

el

tlip

top

el

pl

tlip

top

(in.)

(in.)

(in.)

(in.)

(in.)

(in.)

(in.)

(in.)

3.007

3.638

1
1
1
1
1
1
1
1
1
1

Modified Kent-Park (Park, Priestly and Gill, 1982)

RWl

0.0897

0.4893

0.0255

(95.0%)

(5.0%)

0.106

0.5556

0.0255

(100%)

(95.6%)

0.0897

0.4893

0.0255

(100%)

(95.0%)

(5.0%)

0.106

0.5556

0.0255

(100%)

(95.6%)

(4.4%)

(100%)

RW2

RWl
RW2

1
2

0.5148

0.5811

0.4893

2.478

0.0395

(16.3%)

(82.4%)

(1.3%)

0.5556

2.89

0.193

(79.4%)

(5.3%)

0.4893

3.592

0.3312

(11.1%)

(81.4%)

(7.5%)

0.5556

4.025

0.4258

(11.1%)

(80.4%)

(8.5%)

(4.4%)
(15.3%)
Saatcioglu and Razvi (1992)
0.5148

0.5811

4.412

5.006

Drift levels can be calculated as top/144


Numbers in parentheses represent the percentage of top caused. by that particular component

156

1
1
1

1
1
1
1

1
1

1
1
1

1
1
1

1
1
1
1
1

Table 7.2 Monotonic Top Displacement Components of T-Shaped Walls


Cracking
Specimens

TW1 3

TW1 4

TW2 3

TW2 4

TW1 3

TW1 4

TW2 3

TW2 4

1
2
3
4

Column Top Displacements 1

Yielding

el

el

alip

top

(in.)

(in.)

(in.)

(in.)

0.0557

0.375

0.0255

0.4005

(100%) 2

(93.6%)

(6.4%)

0.2268

0.6676

0.0255

(100%)

(96.3%)

(3.7%)

0.1002

0.5078

0.0255

(100%)

(95.2%)

(4.8%)

0.2374

0.702

0.0255

(100%)

(96.5%)

0.077

0.4139

0.0255

(100%)

(94.2%}

(5.8%)

0.2329

0.673

0.0255

(100%)

(96.3%)

(3.7%)

0.1085

0.5088

0.0255

(100%)

(95.2%)

(4.8%)

0.2352

0.6856

0.0255

(100%)

(96.4%)

(3.6%)

el

pl

(in.)
(in.)
Modified Kent-Park (Park, Priestly and Gill, 1982)

0.693

0.5333

0.7275

0.6987

0.5343

0.711

top

(in.)

(in.)
3.802

0.375

3.088

0.3385

(9.9%)

(81.2%)

(8.9%)

0.6676

0.438

0.0255

(59.0%)

(38.7%)

(2.3%)

0.5078

3.444

0.422

(11.6%)

(78.8%)

(9.6%)

0.702

1.064

0.0255

(59.4%)

(1.4%)

0.4139

3.136

0.345

(10.6%)

(80.5%}

(8.9%)

0.673

1.624

0.0255

(29.0%)

(69.9%)

(1.1%}

0.5088

3.437

0.422

(11.6%)

(78.7%}

(9.7%}

0.6856

2.70

0.0845

(19.8%)

(77.8%)

(2.4%}

(3.5%)
(39.2%)
Saatcioglu and Razvi (1992)
0.4394

dip

1. 131

4.374

1.792

3.895

2.323

4.368

3.471

Drift levels can be calculated as top/144


Numbers in parentheses represent the percentage of 6top caused by that particular component
Displacements Calculated for the Case of Flange in Compression
Displacements Calculated for the Case of Flange in Tension

157

1
Table 7.3 Allowable Spacing of Transverse Reinforcement

Is

Smaz

(ksi)

(ksi)

(in)

63 1

29,000

12.63

63 2

1
2

3.000

63 3

1,500

2.874

65

1,295

2.629

70

925

2.141

72

797

1.959

75

652

1.737

78

541

1.551

80

463

1.420

83

402

1.297

85

344

1.185

Elastic (Pre-Yielding) Region


Yield Plateau Region
Beginning of Strain Hardening Region

1
1
1

1
1
1

1
1

158

1
1
1
1
1
1

1
1
1
1

Table 7.4

1
1

1
1

Transverse Steel

Steel

Extreme Fiber

Computed

Observed

Spacing

Stress 1

Comp. Strain2

Drift Level3

Drift Level

s (in.)

fa (ksi)

(%)

(%)

RWl

3.0

63 4

0.0025 (0.006) 5

1.936

2.0

RW2

2.0

72

0.025

5.5

2.5

TWl

3.0

63 4

0.0025 (0.006)5

0.91 6

1.25

TW2

1.25

83

0.045

2.92

2.5

Specimen

2
3
4
5

Computed and Observed Drift Levels For Buckling

Ec

Interpolated From Table 7.3 Based on Actual Spacing of Transverse Reinforcement


Interpolated From Moment-Curvature Analyses Based on Steel Stress Values
Drift Level At Which Local Buckling is Expected to Occur
Buckling Expected During Yield Plateau Region
Buckling Will Not Occur Until Concrete Cover Has Spalled (ec ;::: 0.006)
Buckling Drift Calculated For Extreme Fiber Compressive Strain of 0.006

1
1

1
1
1
1
1

1
1

(in/in)

159

1
1
1

Table 7 .5 Conflnement Effectiveness


Lateral
Specimen

RWl

RW2

TWl

TW2

1
2
3

Steel

f r, Confinement Pressure (ksi)

Steel
1

Drift

Strain

(%)

(in/in)

1.0%

Stress

Perpendicular To Web

Parallel To Web

(ksi)

(Short Direction)

(Long Direction)

0.0008

23.2

0.139 (2A.,) 3

0.130 (4A.,) 3

1.5%

0.0015

43.5

0.261 (2A.,)

0.244 (4A.,)

2.0%

0.0025

63

0.378 (2A.,)

0.353 (4A.,)

1.0%

0.0012

34.8

0.313 (2A.,)

0.146 (2A.,)

1.5%

0.0034

63

0.567 (2A.,)

0.265 (2A.,)

2.0%

0.005

63

0.567 (2A.,)

0.265 (2A.,)

0.75%

0.0012

34.8

0.209 (2A~)

0.195 (4A.,)

1.0%

0.0016

46.4

0.278 (2A.,)

0.260 (4A.,)

1.5%

0.0025

63

0.378 (2A.,)

0.353 (4A.,)

0.75%

0.0007

20.3

0.292 (2A.,)

0.136 (5A.,)

1.0%

0.00085

24.65

0.355 (2A.,)

0.165 (5A.,)

1.5%

0.0012

34.8

0.501 (2A.,)

0.233 (5A.,)

2.0%

0.002

58.0

0.835 (2A.,)

0.389 (5A.,)

Read Directly From Transverse Steel Strain Histories Plotted in Chapter 6


Read Directly From Stress-Strain Curve For Transverse Steel Based on Steel Strain
Value of Atr Used To Calculate Confining Pressure

1
1
1

1
1

1
1
1

1
1
1

1
160

1
1

1
1
1
1

1
1

Ta.ble 8.1: Experimentally Measured Effective Flange Widths

Specimen TWl
bef1

5.5

0.49

17.05

35.5

11.8

1.63

29.1

9.7

1.25

21.96

45.8

15.3

2.25

24.08

50.2

16.7

2.51

33.17

69.1

23.0

3.65

39.85

83.0

27.7

4.48

44.05

91.8

30.6

5.00

48.00

100.0

33.3

5.5

48.34

100.0

33.6

5.54

% oflw 2 % of hw 3

beJ/

0.5%

7.91

16.5

0.75%

13.99

1.0%
1.5%

1
1

1
2
3
4

of t4

Drift Level

2.0%

2.5%
Effective
Effective
Effective
Effective

Specimen TW2

Flange
Flange
Flange
Flange

Width
Width
Width
Width

Based on Eq. 8. 3
Expressed as a Percentage of Wall Length
Expressed as a Percentage of Wall Height
in Terms of Flange Thicknesses on Each Side of Web

1
1
1
1
1
1
1

1
1

% of lw 2 % of hw 3

161

oft4

1
1
6

35 tt.

30 ft. = 180 ft.

1
1

l
1

30 ft.

100 ft.

35 ft.

il

~
16 ft.

16in.

Fig. 3. 1 Plan View of Prototype Building

16 ft. = 192 in.

@ 14 in.

13 in.
1

(o)

4 @ 6 in.
1

3 in.
1 f--l

Bose Level

1
1
1
1

1
1
1
1

1
1
(b)

Top Three Stories

Fig. 3.2 Cross-Sectional Views of Rectangular Prototype Wall

162

1
1
1
1
1
1
1
1

15000 -,--,---,--,--~-,--,-~,------.-~.--~---~.---------

r,j

C.

0.7

-c

=
=
~
-<=

5000

o
ACI Load Cases
-5000

---------.----,--.---,----,--,-----,-----,--,-----,-----,--.----....,..---.-----

100000

200000

300000

400000

Moment (in-kips)
Fig. 3.3A P-M Interaction Diagrams For Rectangular Prototype Wall (Base Level)

1
1
1
1
1
1
1
1

<t> =

""-"'

10000

10000

4> = 0.7
5000

--=
~

<

ACI Load Cases


-5000

-------.----.---,--...----,.---.-----------------.------.-----

100000

200000

300000

400000

Moment (in-kips)
Fig. 3.3B P-M Interaction Diagrams For Rectangular Prototype Wall (Fourth Story Level)

163

1
1
1
1
1

-=:::::::::J

192 in.

(o)

WALL CROSS-SECTION

Et= 0.0147

TENSi ON

COMPRESSION

E: c,max

(b)

WALL STRAIN PROFiLE:

411"

0.0040

1
1

P=0.1 OAgf c

Fig. 3.4 Calculated. Strain Profile For Rectangular Prototype Wall

1
1

<I> u = O. 0000966

........._

300000

'-"'

200000

.:.=
1

-.....=
=
e

Q}

=
~

100000

*
0.0000

0.0001

Buckling Expected
0.0002

0.0003

Curvature (rads/in)
Fig. 3.5 Moment-Curvature Relation For Rectangular Prototype Wall

164

1
1
1

1
1

1
1

1
1
1
1
1
1
1
1

16 ft. = 192 in.

3 in

4 @

8 in

..; n.lYI
10 in

5 in

8 @ 1 4 in.

112 in.

8 in.

in.Y

#4 hoops & cross- tie


#5 bors

O 5 in.

@ J 4 in

#4 hoop

12 in.

15 @ 7 in.

#5 bors

7 in.

4.5 in

#4
4@

hoops & cross-tie


@

12 in.

in.

10-#11 bors

J 1----l
16 in.

Fig. 3.6 Cross-Sectional View of T-Shaped. Prototype Wall

1
1
1
1

4@

3 in.

-r

165

3 in
--

16 ;n

192

I 1s

n.

in.

192 in.

( o)

WALL CROSS-SECTION

TENSION

r-E't

------

COMPRESSION

90.4"

= 0.00981

41.4"

tc

= 0.004

49"

_J

Ec.mox= 0.00873
(b)

WALL STRAIN PROFiLE (FLANGE iN TENSION):

P=0.05Agf'c

Et= 0.0176

COMP. - - - TENSION

9.23"-

L
( c)

1
1
1
1

1
1

1
1
1
1
1

f-

tc.mox = O. 00089

WALL STRAIN PROFiLE (FLANGE iN COMPRESSION):

P=0.05Agf'c

Fig. 3. 7 Calculated Strain Profiles For T-Shaped Prototype Wall

166

1
1
1
1

1
1

1
1

1
1
1

<P =

20000

10000

--=
~

<

ACI Load Cases


-10000

--+---.---.--.....-.,.---.--,---.----+--...--...--,---,.--,---.....-~----,,---.---.....--.----,---1

-800000

-400000

1
1

1
1
1
1
1
1
1
1
1

1.0

400000

800000

1200000

1600000

Moment (in-kips)
Fig. 3.8 P-M Interaction Diagrams For T-Shaped. Prototype Wall (Base Level)
800000 ----------.------------------,.---..------------....-....---.--.

<f>u=0.0000966

.......

600000

.......

--

___ T
C

- - - -

400000

200000

- - - -

Confined
Unconfined

0.0000

0.0001

0.0002

0.0003

0.0004

Curvature (rads/in)
Fig. 3.9 Moment-Curvature Relations For T-Shaped. Prototype Wall (Loads Parallel To Web)
167

1
1
400000

4> u = 0.0000966

-- -

1
1
1

300000

200000

100000

- - - -

Confined
Unconfined

0.0000

0.0001

0.0002

1
0.0003

Curvature (rads/in)
Fig. 3. 10 Moment-Curvature Relations For T-Shaped Prototype Wall (Loads Parallel To Flange)

9.24 in.

1
1

1
1

72 feet

1
1
I16 feetI
Rectangulor Woll
Profile

9.676 k-ft

8,801 k-ft

Moment Profile

Moment Profile
Static Anolysis

Dynomic Anolysis
(Rw=12)

Displocement Profile
Dynomic Anclysis

(50% lg. Rw=l)

Fig. 3.11 Calculated Moment and Displacement Profi.les For Rectangular Prototype Wall

168

1
1

-~-~--~~------~---~
,,~

<,,<o

vv ....

Vc:9 __

>

/
l<L,,
A-

.....

O)

c.c

..

4
.d'
ti""'

Fig. 4.2 3-Dimensional View of T-Shaped Wall

.4''

Fig. 4.1 3-Dimensional View of Rectangular Wall

1
1
1
1
1
1

~48in.~
__[_5 in.

FOURTH FLOOR -

T
@

3.75 in.

THIRD FLOOR -

SECOND FLOOR -

......
rn-

---

....

FIRST FLOOR

8 @ 7.5 in.

T4.5

144 in.

in.

... -

t
GROUND LEVEL -

...

16 @ 3 in.

...

...
...
...

---

-----

..........

1O to 16 in.
-

...

--

-1--- 76

__L

..__

----

in. - -.....

Fig. 4.3 RWl: Profile View Showing Reinforcement Locations

170

1
1
1

'

27 in.

J_

1
1
1
1
1

1
1
1

1
1
1
1

1
1

O. 75 in

1
1
1

2 in

6 in

7. 5 in

6 in

2 in

O. 75 in

.L 0.75

4 in

in

2.5 in

"L o.75
#2 bors

in

7.5 in.

SECTION A-A
3

2 in.

6 in.

7.5 in.

6 in.

2 in.

0.75 in

'--,------+----+-----------........-------------

}[ ~: :: : : y
\.... 8-#3 bors

#2 bors

7.5 in.

:;:n I

.L 0.75

lop splice

2.5 in

"L 0.75

6 in.

in

in

SECTION B-B
O. 75 in.

'--il

2 in.

2.25 in.

3.75 in.

14

2.25 in.

1
1

1
1
1

0.75 in.

- -----+----+------------..___.:...;,.;..;_---..;;;._;::...=_;.;.;.., -

3
1

2 in. O. 75 in.
l ---

:]7 :lop s;lice :: ~n.; :~ ][


6

#2 bars

SECTION

3. 75 in.

Fig. 4.4 RWl: Cross-Sectional Views Showing Reinforcement Details

171

..C- 0.75

in

2.5 in

"L 0.75

in

.
....

1
7

'

i
9W:::::~

-~

. ...... .
:

' ..

'..f

1
1

:.;~:,:.

i. =::?::

, . _.,.' ~:; . .

1
1
1

1
\.,~ .

Fig. 4.5 RWl: Photograph of First Story Reinforcement

1
1
1
1
1

1
1

1
1
.
_f\~
.--. -~ .- -.,:._.. .,_.. J

Fig. 4.6 RWl: Photograph of Second Story Reinforcement

172

1
1

1
1
1

48 in.

1
1

(o)

WALL CROSS-SECTION

Et

0.0213

TENSION

COMPRESSION

~f-3.56"

~
Ec

Ec,mox

(b)

'1
1

1
1

P=0.10Agf' c

Fig. 4.7 RWl: Calculated Strain Profile at 1.5% Lateral Drift

t
1
1

WALL STRAIN PROFiLE:

_JJ

= 0.004
= 0.00603

6000

<Pu = 0.00057

5000
~

flJ

-...=
c..

4000

..._...

._
=

=
~

3000

Local Buckling
Expected

2000
1000

Unconfmed

Confmed: Modified Kent-Park

Confmed: Saatcioglu & Razvi

o
0.0000

0.0004

0.0008

0.0012

Cunrature (rad/in)
Fig. 4.8 RWl: Moment-Curvature Relations
173

0.0016

1
O. 75 in - --=-3-=@--=2=-i;:.;.n4-......:6:........;.;.;.in-------3_@_7_.5_i_n-----+--6_in_'!--3_@_2___,in ,... O. 75 in

4 in

1[

:~ 1

: E:~ll

3/16 in. hoops @ 2 in.

~-#3 bars ' {


#2 bars

7.5 in.

SECTION A-A

Fig. 4.9A RW2: Cross-Sectional View at Base of Wall

.r 0.75 in

2.5 in

L. 0.75 in

1
1
1
1

1
1

1
1

1
1

1
1
1
Fig. 4.9B RW2: Photograph of First Story Reinforcement

174

1
1
1

1
1

1
1
1

48 in.
O. 75 in ....,j

0.75 n.
3

3 O 2 in .

3 O 7.5 in

6 in

6 in

3 O 2 in

...L 0.75

2 in.
3/16 in. hoops
cross-ties @ 3 in.

#2 bors O 7.5 in.

6 in

1
1
1
1

7 5 in.

#2

7.5 in.

8-/13 bars
@

2 in.

3/16 in. hoops &


cross-ties @ 3 in.
-----,

0.75 in.

4 in.

SECTION A-A

Fig. 4.10 TWl: Cross-Sectional View at Base of Wall

1
1

1
1

6 in.

bars

48 in.

- O. 75 in

175

in

2.5 in

0.75 in

1
1

1
1
1
1

Fig. 4.11 TWI: Photograph of First Story Reinforcement

,,

1
1

1
1
1
1
Fig. 4.12 TWI: Photograph of Web B.E. Reinforcement at Base of Wall

176

'

1
1

1
1

12000

10000

..-..
fil

C.

<Pu = 0.00057

8000

!21

-,._.....=

e
e

6000

4000
Unconfmed

2000

0.0000

0.0005

0.0010

0.0015

0.0020

48 in. - - - - - - - 0.75 i""'"

J 2

in

in

e 7.5

in

6 in

J 2 in

- 0.75 in

0.75 in.

_r-0.75 in

.3 O 2 in.

2.s

in

Lo.75 in
4 in.

Confmed: Saatcioglu & Razvi

Fig. 4. 13 TWl: Moment-Curvature Relations

1
1

Cu"ature (rad/in)

Confmed: Modified Kent-Park

',,
'
1

:&

J/16 in. hoops


O 4 in.

12 bors O 7.5 in.

.3 O 5.5 in.

12 bors O 5.5 in.

48 in.
4 in.

J/16 in. hoop &


cross-tie O 1.5 in.
4 O 4 in.

.3/16 in. hoop


O 1.25 in.

0.75 in.

SECTION A-A

---
4 in.

Fig. 4.14 TW2: Cross-Sectional View at Base of Wall


177

1
1

1
1
1
1
Fig. 4. 15 TW2: Photograph of First Story Reinforcement

'

1
1

1
1
1
1
Fig. 4.16 TW2: Photograph of Web B.E. Reinforcement at Base of Wall
178

1
1

1
1
1
1
1
1
1

CJ

48 '"

c.....--------'
tJ
(o)

WALL CROSS-SECTION

TENSION _,___ COMPRESSION

1
1
t

24.04""

~ - - 7.02"" 1

17.02""

e,.. =

(b)

WALL STRAIN PROFiLE:

Ec.mo,

0.0137

P0.10Agf'c

Fig. 4. 17 TW2: Calculated. Web Strain Profile at 1.5% Lateral Drift

1
1

r-Et = 0.0137

1
1

4 in.

48 in.

1
1

12000
10000

<l>u = 0.00057

,.-..
~

C.

!a1

8000

-...=

6000

4000

cw
=

=
~

Unconfmed

2000

..

Confmed: Modifed Kent-Park

Confmed: Saatcioglu & Razvi

-&.--.~,--.....--.--.-~---r--r-----~---r--r--~r-,---r--,--.---

0.0000

0.0005

0.0010

0.0015

Curvature (rad/in)
Fig. 4. 18 TW2: Moment-Curvature Relations
179

0.0020

1
:

-c

1
1
1
1

r ll

......

6000

rll
rll
~

00

4000

't

Specimen RW1:

2000

0.000

0.001

0.002

--

Pedestal

_,.__

lnSto~

-+-

2nd Sto~

0.003

1
1
0:004

Concrete Strain (in/in)


Fig. 5. lA RWl: Measured Concrete Stress-Strain Relations

-c

1
1

r ll

......

6000

rll
rll
~

00

4000

Specimen RW2:

....
~

t
C
=
U

-il-

-..-

-+-

2000

-e-+-

0.000

0.001

0.002

1st Sto~
2nd Stoy

3rd Sto~
4thSto~

0.003

Concrete Strain (in/in)


Fig. 5. lB RW2: Measured Concrete Stress-Strain Relations
180

Pedesla1

0.004

1
1,

1
1

!
1
1

6000

4000

Specimen TW1:

2000

1
1

0.000

0.001

0.002

----

-atr-

Pedesa1
1st Stoy

-+-

2nd Stoy

0.003

0.004

Concrete Strain (in/in)


Fig. 5.lC TWl: Measurecl Concrete Stress-Strain Relations

'1
1

1
1
1
1

f;IJ

C..

6000

_J

f;IJ
f;IJ

~
..

....
....

4000

2000

rJ1

Specinen

TW2:

~
..
CJ

----

-+-

Pedestal
1st Stoy

-atr-

2nd Stoy

-e-

3rd Stoy

-+-

0.000

0.001

0.002

4th Stoy

0.003

Concrete Strain (in/in)

Fig. 5.lD TW2: Measured Concrete Stress-Strain Relations

181

0.004

1
1
1
1

80
60

40

#3 deformed rebar

20

0.00

0.02

- - - -

#2 deformed rebar

3/16 in. diameter wire

0.04

0.06

1
0.10

0.08

Steel Strain (in/in)


Fig. 5.2 Measured Steel Stress-Strain Relations

'

1
1

1
1

1
Fig. 5.3A RWl: Pedestal Reinforcing Cage

182

1
1

1
i
1
1

1
1
1

Fig. 5.3B TW2: Pedestal Reinforcing Cage

'1
'1
,1

1
1
1
1
1

Fig. 5.4 TWl: Pedestal Formwork Prior to Casting ~ncrete

183

1
1
1

1
1

1
1
1

Fig. 5.5 RWI: Reinforcing Steel in First Story

t
1

1
1

1
1
1
1
Fig. 5.6 RWI: Reinforcing Steel in Second Story
184

1
1
1

1
1
1
1

1
1
1
1

1
1
1
1

Fig. 5.7 RWl: 'Slip' Formwork Prior to Casting Second Story

1
1
1
1
1

Fig. 5.8 TWl: Reinforcing Steel in First Story


185

00
O)

...,,
Cn"fH

Fig. 5.10 TWl: Reinforcing Steel in Second Story

Fig. 5.9 TWl: Wall and Slab Formwork Prior to Casting First Story

1
1
1
1
1
1
1
1
1
1
1
1
1

Fig. 5.11 TWl: Typical Slab Reinforcement Prior to Casting Second Story

1
1,
1
1
1
1

Fig. 5.12 TWl: Reinforcing Steel at Top Story


187

1
1
1
(+)
N ort h

1
1
1

South(-)

-....
------------

Hydroulic Actuotor

12

in

Lood Cell

High-Strength
Post-Tensioning
Cobes

144 in.

Rectonguor

Woll Specimen
Reoction Woll
-----..8

::

in.---..

Pedesto

H .

HrT;e-OownsH

Strong F"loor

27 in.

Fig. 5.13 Schematic of Test Setup (For Rectangular Wall)

188

1
1
1

1
1
1
1
1
1

1
1

1
1
1
1
1
1
1
1
1

1
1
1
1
1
1
1
1

1
1

Fig. 5. 14 Photograph of Test Setup (For Specimen RWl)

Fig. 5.15 RWl: Load Transfer Assembly (Side View)

189

1
1
1
1
1
1
1

1
1
1
Fig. 5.16 RWl: Load Transfer Assembly (End View)

1
1

1
1
1

il
Fig. 5. 17 RWl: Hydraulic Actuator Used to Apply Cyclic Lateral Loads

190

1
1

1
1
1
1
1
1
1

Fig. 5. 18 RWl: Attachment of Lateral Support to Load Transfer Assembly

1
1
1
1

1
1

1
1
1

Fig. 5.19 RWl: Tube-on-Tube Connection of Lateral Support to Reaction Wall


191

V"""
\..,/

......
(O
1-..::>

,,,,,,.

......

,.

.,

.,

.. "

..
'{

_...

'

~
,

' .

...
,

Fig. 5.21 lnstrumentation Used To Measure Pedestal Movement

Fig. 5.20 Steel Reference Frame Used For lnstrumentation

1
1
1

'i

-~- _:. ~

1
1

1
1

Fig. 5.22A RWl: Wire Potentiometers Used To Measure Shear Deformations

1
1
1

1
1
1
1
1
1
1

~-,:-;:r;~~
'" .. :~<1.-

il

'.

. . . . . . . . . .I....,;,-:,

'{~ .t

tf;~
._c, :.-~:t4

Fig. 5.22B TWl: Wire Potentiometers Used To Measure Shear Deformations

193

1
1
1
1
1

1
1
Fig. 5.23 RWl: lnstrumentation Located on First Story Level

1
1
1

1
1
1

1
1
Fig. 5.24 RW2: Instrumentation Located on First Story Level

194

1
1

,,

1
1
1
1
1
1
1

1
1
1
1
1
1

Fig. 5.25A TW2: LVDT's Used To Measure Axial Deformations Along Web

1
1

1
1

1
1

Fig. 5.25B TW2: LVDT's Used To Measure Axial Deformations Along Flange

195

48

1
1

n.

---

-----

---->

7 @ 7.5

-a---

r,.

-=

..

-a- -----

'(.)'

"-"

,_

_,_

----

---e-

f,

I "\

L"'>.

......,,

--.........

n.

-----

f\

... -

16 @ 3

n .

""'

---- ---

'<.J

-a-

'

........

.........

rll-l:; ,~1-

v_;y

3@ 7.5 in.

--

t ('\

f >

0.75 in.

---=--top of pedestal

-1~ i~lrlh .
0.75

2 in

n.

2 in.

single gage at this location


0 two gages at this location
e three gages at this location

11

13

15

17 19 21 23

10

12

14

16

18 20 22 24

Fig. 5.26 RWl: Strain Gage Locations

196

1
1
1
1
1
1

1
1
1
1

1
1
1

1
1
1
1
C

---

------

----

>

\ol

----

''
@

7.5

n.

---

----~
---------a-----

- --

'<J

'<..,)'

--------~
-----sI'

\ :,1

---

"-"

----

!:"51

1
1
1

--

---

24

n.

---

---

,<""}.

,<""}.

'<..,)'

'<.,;)/

,<""}.

.._,

---~~

---

-1-

~ ~

top of pedestol

single goge ot this locotion


0 two goges ot this locotion

three goges ot this locotion

1
1
1
1

I
---

-----------

1
1

n.

~--

1
1
1

48

N
3

j[U:

11

13

15

17 19 21 23

16

18 20 22 24

E: : n

10

12

14

Fig. 5.27 RW2: Strain Gage Locations

197

Flange
4 in.

1
1

44 in.

--

- -l. J

7 @ 7.5 in.

- ... ...

... ...

....
.....

-----_J:

a-

...,
-

""'
....
...

--

~P

- ... ......

; ..
--- -... "" .. ......

--

...,,

l'

li' 1

16 O 3 in .

,....

'U'

'<J

... ...

.. ....-

...
...

<:.

r P

... 1..

1 ~ ~ Pli!

top of pedestal

1
1

single goge ot this locotion


two goges ot this locotion
three goges ot this locotion

Fig. 5.28A TWl: Strain Gage Locations in Web

22

--1
_,n_.-I

--1
_22_in.____...,
,1

,1

-......

--:._

c~

(l

------

---.........
--

'-

""""""

""' ......

-=-

.. '"'"'..
.. '"'"'
e..
:

""' ... -

"" '

16 @ 3 in.

.. -..

'-

... 5

---......... u:

... '-

---~D
"--7 @ 7.5 in.

1
1

Web
4 in.

j ~

l'l

,1

.~

P'I

top of

pedesto

rJI :I; ,: I . ,.5 ,.. I; ,;_IJ lh

0.75 in.

0.75 in.

3 O 2 in.

3 O 2 in.

single goge ot this locotion


0 two goges ot this locotion

Fig. 5.28B TWl: Strain Gage Locations in Flange

198

1
1
1
1

1
1
1

1
1
1

1
1

1
1
1
1
1
1

1
1
1

1
1

1
1
1
1
1

10

11

25

26

33

13

15

17 19 21 23

16

18 20 22 24

34

Gages Mounted at First


and Second Story Level

35

36

37

38

39

40

41

42

43

44

45

46

47

48

Gages Mounted at First


Story Level Only

Fig. 5.28C TWl: Longitudinal Strain Gage Locations

199

Flonge
4 n.

1
1

44 in.

1
1

--

- -

..no

1)

10

5.5 in.

,..,.
....,

--

.n.
'V'

,...

-.

.....

.....

- -

- - ,...
-- b

,.....

...,,

'-"

'' ~

'W

t~
V

U'

--
~

,_

t1'

38

1.25 in.

.-

.
--.

,~

top of

pedesto

sngle

goge ot this locotion


two goges ot this locotion
three gages ot this ocotion

Fig. 5.29A TW2: Strain Gage Locations in Web


Web
4 in.

22
1-_n_.
-il___
22_in._I

-.. ----

-----

--~-

(,O

......

7.5 in.

cp

......
.......

......
--......
-~

q;

1
12

@ 4 in .

. --....
'""-""

.. Bu

n 1

. -.. J
. -. --

---

...

'w'l'

''

'P'II

top of pedestol

singe

goge at this locotion


two gages ot this location

Fig. 5.29B TW2: Strain Gage Locations in Flange

200

1
1
1
1
1
1

1
1
1
1
1
1

1
1
1
1

1
1

10

11

25

26

13

31

32

33

34

35

36

15

17 19 21 23

16

18 20 22 24

Gages Mounted at First


and Second Story Level

1
1
1
1
1
1
1
1
1

1
1

37

38

39

40

41

42

43

44

Gages Mounted at First


Story Level Only

45

46

47

48

49

50

Fig. 5.29C TW2: Longitudinal Strain Gage Locations

201

1
1

1
1
1

1
1
1
1
Fig. 5.30 RW2: Embedded Concrete Strain Gage in Boundary Element

1
1
1

1
1
1
1
Fig. 5.31 TW2: Embedded Concrete Strain Gages in Web Boundary Element
202

1
1

1
1
1

Specimen: RWI

1
1
1

1
1
1
1
1
1

-3

Total# Cycles = 14.5

Time
Fig. 5.32 RWl: Lateral Drift Routine

J
Specimen: RW2

1
1
1

1
1
1

-2

-3

Total# Cycles = 20

Time
Fig. 5.33 RW2: Lateral Drift Routine
203

1
1
4
3

~
Specimen: TWI

1
1
1

Total# Cycles = 11.5

Time
Fig. 5.34 TWl: Lateral Drift Routine

4---.---------------------,
3

d::
.....

-.

--

~
-.
(1)

Specimen: TW2

1
1
1
1
1

o -Pt-A~~---+++-+-+--l--++H-+-+-+-+-+-t-+-+-+-t-+-+--+--r-+-+--+--

-1

-2

-3

Total# Cycles = 22.5


-4-"-------~----~-----~

Time
Fig. 5.35 TW2: Lateral Drift Routine
204

1
1
1
1

1
1

1
1
1
1

Pavg= O. lOAgf~ = 90 kips

--<=

30

1
1

1
1

100

200

300

400

500

Data Point Number


Fig. 6.1 RWl: Axial Load History

120

1
1
1

90

60

Pavg= 0.07Agf~ = 85 kips

30

1
1
1
1

200

400

600

Data Point Number


Fig. 6.2 RW2: Axial Load History

205

800

-=

1
1
1
1

Pavg= 0.09Agf~ = 158 kips


l 00

-<

50

100

200

300

400

Data Point Number


Fig. 6.3 TWl: Axial Load History

1
1
1
1
1

-=!

--<=

1
1
1
1

100

Pavg= 0.075A8 f ~ = 164 kips

50

200

400

600

Data Point Number


Fig. 6.4 TW2: Axial Load History

206

800

1
1
1

1
1

1
1
1
1
1
1
1
1
1
1

Lateral Drift (%)


-1.4

-0.7

O.O

0.7

1.4

-2

-1

Top Displacement (in.)


Fig. 6. 5A RW 1: La ter al Load vs. Top Displacement (Initial Cycles)

Lateral Drift (%)


-2.8

-1.4

O.O

1.4

2.8

1
1
1
1
1
1
1

Vu,max = 2.57-(i;"

-4

-2

Top Displacement (in.)


Fig. 6.5B RWl: Lateral Load vs. Top Displacement (Ali Cycles)

207

1
1

Lateral Drift (%)


-1.4

...i

O.O

-0.7

0.7

1.4

P = 0.07Agfc'

20

1
1

~
.._.,

"'C

1
1

1
-2

-1

Top Displacement (in.)


Fig. 6.6A RW2: Lateral Load vs. Top Displacement (Initial Cycles)

Lateral Drift (%)


-2.8

-1.4

O.O

1.4

2.8

1
1

1
1

-f
~

1
1

-20
Vu,max = 2.32K

-4

-2

Top Displacement (in.)


Fig. 6.6B RW2: Lateral Load vs. Top Displacement (All Cycles)
208

1
4

1
1
1
1
1
1
1

1
1
1

Lateral Drift (%)


-1.4

-60

O.O

-0.7

0.7

1.4

-+---.-..---.....---.--...---.---,---...-----.-----.----.----.----

-2

-1

Top Displacement (in.)


Fig. 6. 7A TWl: Lateral Load vs. Top Displacement (lnitial Cycles)

-2.8

-1.4

Lateral Drift (%)


o.o

1.4

2.8

40

1
1
1

1
1
1

-40

Vu,max = 4.85-{r';

-4

-2

Top Displacement (in.)


Fig. 6. 7B TWl: Lateral Load vs. Top Displacement (All Cycles)
209

Lateral Drift (%)


-1.4

-0.7

O.O

0.7

r: C

P = 0.075Agf c'

30

1.4

-30

1
1
1
1
1
1
1
1

-2

-1

Top Displacement (in.)


Fig. 6.8A TW2: Lateral Load vs. Top Displacement (Initial Cycles)

Lateral Drift (%)


O.O
-1.4
-2.8
1.4
2.8
100 ---.---.-----.---.......-----,-----,---.-----,-----,

-f
~

-50
Vu,max

-4

= 5.46K
2

Top Displacement (in.)


Fig. 6.8B TW2: Lateral Load vs. Top Displacement (All Cycles)

210

1
1

1
1
1
1

1
1
1

1
1

1
1
1
1
1
1
.

-s.
C:

:.i2 40

1
1
1
1
1

North (+) Displacements

South (-) Displacements

Q)

e
.:::

en

C:

~ 20

Q)

en

o
1

Fig. 6.9

1
1

1
1
1
1
1

Specimen RW1

60

-m

13

11

Measured Secant Stiffnesses

Specimen RW2

80

i-

RW:

7
Cycle Number

60

:.i2

North (+) Displacements

South (-) Displacements

Q)

e 40
.:::

en

C:

ca

(.)

Q)

en 20

o
1

9
11
13
Cycle Number

15

Fig. 6.10 RW2: Measured Secant Stiffnesses


211

17

19

1
1
Specimen TW1

North (+) Displacements

South (-) Displacements

1
1

20

o
5

11

7
9
Cycle Number

1
1
1

12

Fig. 6. 11 TWl: Measured. Secant Stiffnesses

Specimen TW2

140

120

:::

c:

[ 100

:e
=
:.i2

~
Q)

80

CJ)

60

North (+) Displacements

South (-) Displacements

CJ)

1
1
1
1

:::

ca
(,)

Q)

1
1

40

20

o
1

9
11
13
Cycle Number

15

Fig. 6. 12 TW2: Measured. Secant Stiffnesses


212

17

20

1
1

1
1

8000
1

1
r:.

C.

1
1

-.....=

1
1

1
1
1
1

1
1

1
1
1

1
1
1

1
1

4000

..._,..

=
QJ

e
o
QJ
r:.

-4000

_j

=
-8000
-0.03

-0.02

-0.01

0.00

0.01

0.02

0.03

Base Rotation (rads)


Fig. 6.13 RWl: Base Moment vs. Base Rotation

8000
~

r:.

-..._,.......=
C.

4000

=
QJ

e
o

QJ
r:.

-4000

=
-8000
-0.03

-0.02

0.00

-0.01

0.01

0.02

Base Rotation (rads)


Fig. 6.14 RW2: Base Moment vs. Base Rotation
213

0.03

10000

'

....
1

1
.....

-5000

1
1
1
1
1
1

1
-0.015

-0.010

-0.005

0.000

0.005

0.010

0.015

Base Rotation (rads)

Fig. 6.15 TWl: Base Moment vs. Base Rotation

1
~

Q.

10000

~
1

-..._,.....
C

5000

o --+---------..,.;,,.__~

:;

-5000

<IJ

==
rl}

-10000

-0.03

-0.02

-0.01

0.00

0.01

0.02

Base Rotation (rads)


Fig. 6.16 TW2: Base Moment vs. Base Rotation
214

0.03

1
1
1
1
1
1
1
1

1
1

1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1

Lateral Drift (%)


o

-1

-2

South

North

4
-

(J.)

.,,o

E
::,

....o
Cf.)

----+-

o
-3

0.2S% Drift

....,.._

--+-

O.S004 Drift

--+-

0.7S%0rift

-2

o
l
Lateral Displacement (in.)
-1

1.0% Drift
.s4 Drift

2.004 Drift

Fig. 6.17 RWl: Lateral Displacement Profiles

Lateral Drift (%)


o
-1

-2

South
4
-

(J.)

.,,o

E
::,

....o
Cf.)

--.----

-+-

o
-4

-3

0.2S%Drift

--+-

O.S004 Drift

l.5%Drift

0.7S%Drift

--+-

2.004 Drift

l .004 Drift

--M--

2.S% Drift

o
l
2
Lateral Displacement (in.)

-2

-1

Fig. 6.18 RW2: Lateral Displacement Profiles

215

1
1

Lateral Drift (%)


o

1
1

-1

-2

5 --,--,----,----,~-,--,---,~-..---,---,---.---,------------------

North

South

..D

E
::s

+-

en

--e---

-il-

0.25% Drift

_._

0.50% Drift

-+-

_._

0.75% Drift

__.._ 2.0% Drift

1.0% Drift
1.5% Drift

o------------------~---

-2

-3

-1

Lateral Displacement (in.)


Fig. 6.19 TWl: Lateral Displacement Profiles

1
1
1
1

1
Lateral Drift (%)
-2

-1

-rr--r--r--r-r--------------------------r----r--.--,--.--...........--..

North

..D

E
::s

co

1
1

+en

-il-

0.25% Drift

_._

O.SOoA> Drift

-+-

_._

0.75% Drift

__.._ 2.004 Drift

-e-

1.0% Drift

--M--

1
1

1.,% Drift

2.,% Drift

o -+--~~--,--~--,------.----
-4

-3

-2

..

Lateral Displacement (in.)


Fig. 6.20 TW2: Lateral Displacement Profiles
216

1
1

1
1

1
1
1
1
1
1

20

-20

1
1

-40----------------~----------------------------0.012

1
1
1

-0.008

-0.004

0.000

0.004

0.008

0.012

Shear Distortion (rads)


Fig. 6.21A RWl: First Story Shear Distortions

1
1
1

1
1
1
1

-.....=
..
~

-20

-0.012

-0.008

-0.004

0.000

0.004

0.008

Shear Distortion (rads)


Fig. 6.21B RWl: Second Story Shear Distortions
217

0.012

1
1

1
1

-=
...=
;.
c

-20

-0.012

1
-0.008

-0.004

0.000

0.004

0.008

0.012

Shear Distortion (rads)


Fig. 6.22A RW2: First Story Shear Distortions

1
1

1
1

1
1
-=
;.

-20

-40-4--,,---,--,.-...--,,-,.--,--,--r-~--l------------r--r-----------i

-0.012

-0.008

-0.004

0.000

0.004

0.008

Shear Distortion (rads)


Fig. 6.22B RW2: Second Story Shear Distortions

218

0.012

1
1
1

1
1

1
1
1
1

1
1
1

-=
.

....u

..J

-40

-0.012

1
1

0.000

0.004

0.008

0.012

Fig. 6.23A TWl: First Story Shear Distortions

1
1
1
1

-0.004

Shear Distortion (rads)

-0.008

1
40

-=
.

....u
=
..J

-40

-0.012

-0.008

-0.004

0.000

0.004

0.008

Shear Distortion (rads)


Fig. 6.23B TWl: Second Story Shear Distortions

219

0.012

1
1
1

1
1

"C

=
c
-=
.....
~

O-+-~~~~~~~~~~

.
~

-40

1
-0.012

-0.008

-0.004

0.000

0.004

0.008

0.012

Shear Distortion (rads)


Fig. 6.24A TW2: First Story Shear Distortions

1
1

1
~

-,._,

C.

40

"C

=
-=
= -40
~

o---~~~~~~~~~~--~~~~~~~~~-

-80

1
1
1

--+---.---..--.....---,--..----,,--,--,---,--__,..;--,------,,---,----,----r-----......,------..----,,--,..----.----

-0.012

-0.008

-0.004

0.000

0.004

0.008

Shear Distortion (rads)


Fig. 6.24B TW2: Second Story Shear Distortions
220

0.012

1
1

1
1

1
1
1

U1

V1
V2

1
1

,..

1
1

d~ = V(h - V2)2

d~

+ (l + U2)2

= V(h + Vl)2 + (l -

Ul)2

1
1

1
1

1
1

d, d2

d~, d

= undeformed dimensions of two wire potentiometers

2 = deformed dimensions of two wire potentiometers


h = height of 'X' configuration
1 = length of 'X' configuration

Fig. 6.25 Variables For Estimating Shear Distortions


221

1
0.06

,,-......
C

---...
-......
C

0.04

"-"

~
,...

en

0.02

O)

......
O)

1
1
1

0.25% Drift
0.5% Drift
0.75% Drift
1.00/o Drift
1.5% Drift
2.00/o Drift

,...

c..>

0.00

North

-0.02

48

36

24

12

South

Distance Along Web (in.)


Fig. 6.26A RWl: Web Concrete Strain Profile (Pos. Displacements)

---...
-c::
-......
C

0.04

"-"
~

,...

en

0.02

O)

......
O)

0.25% Drift
0.50% Drift
0.75% Drift

1. 00/o Drift
1.5% Drift

2. 00/o Drift

,...

c..>

c::
o

0.00

North

South

-0.02

24

12

36

48

Distance Along Web (in.)


Fig. 6.26B RWl: Web Concrete Strain Profile (Neg. Displacements)
222

1
1

0.06
,,-......
C

1
1

1
1
1
1

0.06

--.---------r-----------r---------,----,-----,

1
1

1
1
1
1

1
1
1
1
1

1
1

0.50% Drift

0.75% Drift

l.0%Drift

l.5%Drift

2.0%Drift

North .,...-....- -..l South


-0.02 -4----.----,-------r---r---.-----,---...----,

1
1
1
1

0.25% Drift

12

24

36

48

Distance Along Web (in.)


Fig. 6.27A RW2: Web Concrete Strain Profile (Pos. Displacements)

0.06 ---,-----,----,------,----,----,-----,----r----,

0.04

0.02

0.25% Drift

0.75%Drift

1.00/o Drift

l.5%Drift

0.50% Drift

North

..

..--- -...-

2.00/o Drift

South

-0.02 --1-----,----,------.---,----,-------r----r-----t

12

24

36

48

Distance Along Web (in.)


Fig. 6.27B RW2: Web Concrete Strain Profile (Neg. Displacements)
223

1
0.03

,,-....

c::
-=
-..._.,c::

-.....

0.02

cd
-

Cl'J

0.01

1
1

0.25% Drift
0.50% Drift
0.75% Drift

1.00A, Drift
1.5% Drift

O)
~

O)
-

(J

C:

1
1

0.00

Questionable Data (see discussion in section 6.6)

North

Il

1111

South

-0.01

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 6.28A TWl: Web Concrete Strain Profile (Pos. Displacements)

0.015

,,-....
C:

-=
-..._.,c::
-.....=
Cl'J

0.010

0.005

0.25% Drift
0.50%Drift

1
1

0.75% Drift
1.0% Drift
1.5% Drift

.S:l
O)
-

1
1

(J

c::

0.000

North

1111

-0.005

24

12

36

48

Distance Along Web From Flange (in.)


Fig. 6.28B TWl: Web Concrete Strain Profile (Neg. Displacements)

224

1
1

1
1
1
1
1

1
1
1
1
1
1

1
1
1

1
1
1
C

0.010

-C:
C

-il-

__..,_

-......

0.005

o
......

0.000

"-""

0.2So/o Drift
O.So/o Drift
0.7So/o Drift

--+---e-

00

-+-

1.So/o Drift

---*-

2.00/o Drift

~
u

1.0% Drift

-0.005

Ol)

-0.010
-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.28C TW 1: Flange Concrete Strain Profile (Pos. Displacements)

0.015

.........

"-""

-......
C

0.010

00

o
......
~
u

0.005

0.25%Drift
0.500A, Drift
0.75%Drift
1.0% Drift
1.5% Drift

0.000

00

-a
~

-0.005
-24

-12

12

Distance Along Flange From Web (in.)


Fig. 6.28D TWl: Flange Concrete Strain Profile (Neg. Displacements)

225

24

1
1
1

0.03

-'c
-......ca
C:

0.02

'-'

C:

rr.

0.01

0.25%

0.50%
0.75%

1.00/o
1.5%

~
......

C:

0.00

North
-0.01

12

24

South

48

36

Distance Along Web From Flange (in.)


Fig. 6.29A TW2: Web Concrete Strain Profile (Pos. Displacements)

0.015
~

C:

0.010

'-'

0.005

-'c
-......ca

C:

rr.

0.000

1
1
1

0.25%
0.50%
0.75%
1.0%

~
......

~
-

-0.005

C:

North

-0.010

-0.015

24

12

South

36

48

Distance Along Web From Flange (in.)


Fig. 6.29B TW2: Web Concrete Strain Profile (Neg. Displacements)
226

1
1
1
1
1

1
1
1
1
1

1
1
1
1

0.015 ---,----,---.....----,----,----.,..-..----,..--.......-----

0.010

0.005

0.15%

l .OOA,
1.5%

2.00A,

)(

-0.005

2.5%

--+-------.....----...---+-----------____.
-24

1
1
1
1
1

O.SOOA

1
1

0.25%

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.29C TW2: Flange Concrete Strain Profile (Pos. Displacements)

0.03

-.---

-+-

0.02

-e-

0.2.S%

O..So/o

-+-

1..S%

-*-

2.00/0

---

o.,.so/0
1.00/0

2 ..S%

1
1
1

-0.01

--1---......---......---------+---,------,----,------

-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.29D TW2: Flange Concrete Strain Profile (Neg. Displacements)
227

1
1

1
1
1
1

2000

\I
\/ \ I
~Ir

North B.E.

SouthB.E.

100

200

300

400

500

Data Point Number

Fig. 6.30 RW2: Concrete Strain Gage Histories

1
20000

10000

1
1
1

LVDT

Inside Gage
Outside Gage

1
1
o

100

200

300

400

Data Point Number


Fig. 6.31 TW2: Web B.E. Concrete Strain Gage Histories

228

500

1
1
1
1
1
1

0.06 -----,.---.,........------,-----,---.,..-----,----,-------,

0.04

0.5% Orift

0.75% Drift

1.0% Orift

1.5% Drift

0.02

1
1
1
1
1
1

2.0% Drift

Noth ...----9 South


-0.02

----1------..------.-----r---~---,----,-------

12

1
1

48

36

Distance Along W eb (in.)


Fig. 6.32A RWl: Web Steel Strain Profile at Base (Pos. Displacements)

0.06 - - - - - - - . . . - - - - - - - - - . . - - - - - - - - - - - .

0.04

North
-0.02

0.50% Drift

o. 75% Drift

1.00/o Drift

l.5%Drift

0.02

1
1

1
1

24

2.00/o Drift

...----~--~ South

----1------------..-----r---.,------,.,----,-----

12

24

36

48

Distance Along W eb (in.)


Fig. 6.32B RWl: Web Steel Strain Profile at Base (Neg. Displacements)

229

0.009

,,-....

-C:
-

0.006

0.003

-.._,,

-......
C

..

-o
o
......

C/'J

Il

South

0.50% Drift
0.75% Drift
1.0% Drift
1.5% Drift

C/'J

North

0.25% Drift

2.0% Drift

0.000

-0.003 - + - - - - - - ~ - - - - - . - - - r - - - ~ - - . , - - - - . - - - ;

12

24

36

"48

Distance Along Web (in.)


Fig. 6.32C RWl: Web Steel Strain Profile at First Story Level (Pos. Displacements)

0.009

,,-....

-=

North

0.006

South

""'-..

-......=
C

"-"
~
..

0.003

00

1
1

0.25% Drift
0.50%Drift

0.75% Drift

1
1

1.0% Drift
1.5% Drift

00

.......

o
o
......

.
il

2.0% Drift

0.000

-0.003

---+-------.....------,----,----,-..--.,..-----,.----

12

24

36

48

Distance Along Web (in.)


Fig. 6.32D RWl: Web Steel Strain Profile at First Story Level (Neg. Displacements)
230

1
1
1
1
1
1
1
1
1
1

1
1

1
1

1
1

1
1
1

0.06 ---r---......,.---------,-----------.....---

,-...
C::

~
"-"
c::

-e

0.04

C/':J.

-0.02

1
1

0.500/o Drift

O. 75% Drift

l.0%Drift
l.5%Drift
2.0%Drift

South

--+--------------------------

12

24

36

48

Distance Along Web (in.)


Fig. 6.33A RW2: Web Steel Strain Profile at Base (Pos. Displacements)

1
1

1
1
1

North ..4---..9..

1
1
1
1

0.25% Drift

+
*

0.02

+-

1
1
1

0.06 - - - - - - - - - - - - - - - - - - - - - -

,-...

C::

-~c::
-

0.04

"-"

cd
..

0.02

+-

0.25% Drift
0.50% Drift

o. 75% Drift

1.0%Drift

1.5% Drift

C/':J.

North
-0.02

..

..4--- l South

--+--------------------,---...,......-----1
o

12

24

36

48

Distance Along Web (in.)


Fig. 6.33B RW2: Web Steel Strain Profile at Base (Neg. Displacements)
231

1
0.015 ------....-------,----,----,----,-----r---,

,,,-....
C::

~
......
'-"'
c::
......
Cd
..
......

--.----.-

0.010

02s.o Drift
O.SOO.o

Drift

o.1s.o

Drift

-e-+-0.005

el).

1.00.o

Drift

U%Drift

-*-

2.00.o

-M---

2.S%Drift

Drift

North ...------~~~ South


-0.005

---L-----..-------,----,----,----T---.----

12

24

48

36

Distance Along Web (in.)


Fig. 6.33C RW2: Web Steel Strain Profile at First Story Level (Pos. Displacements)

--__._
--.-

-e-+--

0.2S% Drift
0.SOO.o

Drift

1
1
1
1

1
1
1

0.1S% Drift

1
1

UWo Drift
U%Drift

--*-

2.0%Drift

-M---

2.S% Drift

1
-0.005

North

"4111------..~-

South

---L-----..------------r---,----.----

12

24

36

48

Distance Along Web (in.)


Fig. 6.33D RW2: Web Steel Strain Profile at First Story Level (Neg. Displacements)

232

1
1
1
1

0.06

-il-

_._

-+-

1
1
1
1
1
1
1
1

1
1
1
1

1
1
1

-+-e--

0.04

0.2S04 Drift
O.S0%Drift
0.7S% Drift
1.004 Drift
1.So/o Drift

0.02

(l)
(l)

00

0.00

North
-0.02

41111111...- -..- South

-4----..-------------------,----,-----

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 6.34A TWI: Web Steel Strain Profile at Base (Pos. Displacements)

0.009 - - - - - - . . . - - - - - - - - - - - - - . - - - - , . - - - - - - - - ,

-il0.006

_._

0.ljo/o Drift

-+-e--

0.7'04 Drift

-+-

0.003

o.,% Drift

1.004 Drift
1-'% Drift

North 41111111---..-...- South


-0.003

--4-----.--------,----,---r-----r---------

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 6.34B TWI: Web Steel Strain Profile at Base (Neg. Displacements)
233

0.015

----------r-----------r-----r-----,----,-----,

-il-

025% Drift
0.5% Drift

.............

_._

-e-

1.0/c, Drift

-+-

1.5% Drift

-*-

2.0% Drift

,,,,-....,
C

-:l.-

0.010

~
;..

0.005

.......

0.75% Drift

C/'J.

Q)

0.000

J,~~~~~~~~~~:;;==ii=====::::::::::::==-t----===

-0.005

----------or------,----,----,------,----,--------1

North
O

<11111-------

12

South

24

36

Fig. 6.34C TWl: Web Steel Strain Profile at First Story Level (Pos. Displacements)

North
,,,,-....,

-C:
C

0.000

-0.005

- -

South

;..

.......

C/'J.
,,_
Q)
Q)

.......

C/'J.

0.25%Drift

0.5% Drift

-0.010

-0.015

il

12

24

1
1
1
1
1

0.75% Drift

1.0% Drift
1.5% Drift

36

48

Distance Along Web From Flange (in.)


Fig. 6.34D TWl: Web Steel Strain Profile at First Story Level (Neg. Displacements)
234

-------,----or------,---....,----,----..---,-----

1
1
1

48

Distance Along Web From Flange (in.)

0.005

1
1
1

1
1

1
1
1

0.009

,.-....

0.006

--

0.003

---

............

1
1
1

'-'
..

......

00
(J.)
(J.)

......

(J.)

Ol)

1
1
1

-+-e-

0.7'%Drift

-+-+-

1.$04 Drift

1.004 Drift

2.004 Drift

0.000

C
cd

-0.003
-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.34E TWl: Flange Steel Strain Profile at Base (Pos. Displacements)

0.009

---.---

,.-....

-C:
C

'-'

0.006

-+-e-

C
cd

-+-

..

......

--......
00

0.003

0.2,4 Drift
0.$% Drift
0.7'%Drift
1.004 Drift
.,4

Drift

Yield Strain

(J.)
(J.)

00
(J.)

1
1

0.$%Drift

00

1
1
1

1
1
1

-..-

C
cd

0.2,,,.. Drift

Ol)

0.000

-~

-0.003
-24

-12

12

Distance Along Flange From Web (in.)


Fig. 6.34F TWl: Flange Steel Strain Profile at Base (Neg. Displacements)
235

24

0.0000

(l.)
(l.)
-+-o-

-----,----r------,.----,---...,----r------,----,

-0.0005

-0.0010

1
-il-

-0.0015

r:r
(l.)

o.o

-0.0025

_..._

0.25% Drift

-+-

0.1S% Dri.ft

O.So/o Dri.ft

-e-+-

-0.0020

1.0% Drift
1.So/o Dri.ft

-*-

2.0% Drift

--L-----...--------+---...,----r---------

-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.34G TWl:

Flage

Steel

Strai

Profile at First Story Level (Pos.

Displacemets)

0.009

....-....

-il-

...........

'-"

0.006

_..._

0.2S% Dri.ft

-+-

0.7S%Drift

-e-+-

C
c

s-

-+--

r:r

I'
1
1

1.0'/e Drift

Drift

0.003

-0.003

-------.----.---------+----r----,------,.---;

-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.34H TWl: Flage Steel Strai Profile at First Story Level (Neg. Displacemets)
236

0.S% Dri.ft

.s.

1
1

1
1
1
1
1
1

1
1

1
1
0.06

1
1
1
1

-il-

-.-

.....C:

C:

-.....

,...._,,

--C:
cd

....,_

-+-

0.7'% Drift
1.0% Drift
U%Drift
2.0%Drift

Q)

Q)
....,_

r/).

0.00

North

South

-0.02

12

36

24

48

Distance Along Web From Flange (in.)


Fig. 6.35A TW2: Web Steel Strain Profile at Base (Pos. Displacements)

0.030

1
1
1

0.,004 Drift

r/).

-*-

0.02

1
1
1
1

-+-e--

0.04

0.2,% Drift

-il-

-.-

-.....C:,...._,,
C:

--C:
cd

....,_
Cl"1

0.020

0.010

0.2.5'1o Drift
0 ..500/o Drift

-+-e--

0.7.5% Drift

-+-

1..So/o Drift

-*-

2.004 Drift

1.004 Drift

Q)

Q)
....,_

Cl"1

0.000

North
-0.010

~ South

-24

12

36

Distance Along Web From Flange (in.)


Fig. 6.35B TW2: Web Steel Strain Profile at Base (Neg. Displacements)
237

48

0.020 - - . - - - - - . - - - - . - - - - - - . - - - - - - - - - - - - - - -

0.015

0.010

0.005

-il-

02S% Drift

-lr-

O.SOOo

-+-e-

0.7S%Drift

Drift

l.0'%Drift

-+-

1.So/o Drift

-*-

2.00o

--M--

2.S%Drift

Drift

1
1

North '4----ll., South


-0.005

--+------.---......,...---...---.------.----------

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 6.35C TW2: Web Steel Strain Profile at First Story Level (Pos. Displacements)

0.005

North
,,-....

-t:=
-....=

0.000

"'-"'

c'd
~

-0.005

r.rJ

.,.....

(l)
(l)

....

00

-0.010

-0.015

-il-

0.2.5% Drift

-.k-

0..50% Drift

-+-e-

0.7'% Drift

-+-

1..5% Drift

-*--

2.0%Drift

--M--

2..5% Drift

South

1
1
1
1
1
1

1.0% Drift

--4--------------.----r----r---------r-----;

1
1
1

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 6.35D TW2: Web Steel Strain Profile at First Story Level (Neg. Displacements)
238

1
1
1

'1
1
1
1
1
1
1
1

0.009 - - , - - - ~ - - - , - - - - - - - , - - - - - - - - - - - - - - - - -

0.006

-il-

0.2.5% Drift

-.-

0.,00,4 Drift

-+-e-+-

0.003

-*-

0.7.5% Drift
1.Oo/o Drift
1..5% Drift
2.0% Drift

- l ( - 2 ..5% Drift

-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.35E TW2: Flange Steel Strain Profile at Base (Pos. Displacements)

1
0.020 - - - - - - - - - - - - - - . - - - - - - - - - - - - - -

1
1
1
1
1
1
1

-il-

0.2,o/o Drift

-+-e-+-

.ooo Drift

-.-

l . .5%Drift
2.0%Drift

-0.005

---4------------..----------,.---....-----,.----

-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.35F TW2: Flange Steel Strain Profile at Base (Neg. Displacements)

1
1

239

0.001 - - - - - - . . . . . . - - - - - - - - - - - - - - - - - - -

0.2.5% Drift

-k- 0..500.4 Drift

-0.003

-+-e-

0.1S% Drift

-+-

U%Drift

-*-

2.00/o Drift

1.0% Drift

-M-

2..SO/o Drift
---1-----------------------------

-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.35G TW2: Flange Steel Strain Profile at First Stoy Level (Pos. Displacements)

0.009 - - - - - - - - - - - - - - - - - - - - - -

0.006

0.003

-il-

0.2.5%

-k-

0..500/o

-+-e-

0.7.50/o

-+-

UO/o

-*-

2.0%

-M-

2..5%

1.()0,4

Yield

-0.003

---1----------------------~---

-24

-12

12

24

Distance Along Flange From Web (in.)


Fig. 6.35H TW2: Flange Steel Strain Profile at First
240

Stoy

Level (Neg. Displacements)

1
1
1
1
1
1
1

1
1
1
1
1
1
1
1
1

1
1
1
1
1

20000

IL
--

10000

2L

1
1

4L

10000
--

1
1
1
1

SL

BL

40000
--

20000

9L

1
1
1

1
1
1
1

,.

20000

12L
--

10000

100

13L

200

300

400

500

Data Point Number


Fig. 6.36 RWl: Reinforcing Steel Strain Gage Histories

241

600

1
1

1
40000

16L
--

20000

1
1
1

17L

1
1

20L

20000

21L

1
1
1
1
1

23L

20000

24L

1
1
1

1L37

2000

100

200

300

400

Data Point Number


Fig. 6.36 Cont.'

242

500

600

1
1
1

6000

']
_j

9L37

3000
--

13L37
~

1
1
1
1
1
1

17L37

W7.5NW

1000
--

W7.5SW

1
1

W30NE

3000

1
1

1
1
1

--

W30E

100

200

300

400

Data Point Number


Fig. 6.36 Cont.'

243

500

600

6000 --,--..--,--,--,--,---,---,--~,----,--.....---.---.--..--.-~------__.....,--H3NW

3000

--

H3NE
T3N

H9NW

4000
--

H9NE

1
1
1
1
1
1
1
1

1
1
H18NW

1
1
1

1000
--

1
1

H33NW

1000
--

-1000

H18NE

H33NE

---+--.--..,...........,-~-,--,----,-..---.---.---.--,,-....-.......--.-.........-......--,,--.--..---.--..--..---,-...--,--

100

200

300

400

Data Point Number


Fig. 6.36 Cont.'

244

500

600

1
1
1
1
1
1
1

H3SW

2000

--

H3SE

H9SW

2000

--

H9SE

1
1
1

Hl8SW

1000

1
1

H33SW

500

1
1
1

100

200

300

400

Data Point Number


Fig. 6. 36 Cont.'
245

500

600

1
1
1
1

lL

12000

1
1

1
1

10000

1
20000

8L
--

10000

1
1

9L

1
20000

12L

1
1

10000

100

200

300

400

500

600

Data Point Number


Fig. 6.37 RW2: Reinforcing Steel Strain Gage Histories

246

700

1
1
1

1
1
1
1

1
1
1
1
1
1
1
1

1
1
1
1
1
1

2000

16L
17L

20L

20000

40000 ---.---,--,-.....,......,........,.....,..-,--,-......,.....,..-,--,---.--,.--.--,.--.--,.----------23L

20000

20000

1L37
--

10000

100

8L37

200

300

400

500

Data Point Number


Fig. 6.37 Cont.'

247

600

700

1
1

1
1
1

1
20000

10000

,,-._

'U.:J
~

.......

o --t~~~~f-TI~-'Cr~c-t-i~7!-'f:::~~~T7!-t-r~~~~
-1 0000 --;.--,--.--,--..,........,.......,..---,-.......--,.--,--.---.----.,..---.---,---,----,--,.--,-,--,--.....--.--,---,..--,--,-,--..--,--...,..........__,.......,.___,.

'-"'

C
......
~

1
1

......

r..rJ.

2000
1000

1
1

1
1

W30E

4000

W30SE

2000

W30NE
vA-J.

100

200

1
300

400

500

Data Point Number


Fig. 6. 37 Cont.'
248

600

700

1
1
1

1
1
1
1

8000

1
1
1

-4000

--

H4S

2000
H8SE

1000
,,-.....

"w
~

...._..,
...-

H4SE

4000

-..,._
-..,._
C

--

H8SW

o
-1000
2000

1
1

'fJJ
~
~

'fJJ

-1000

1
1

2000

1
1
1

H12SE

1000

~
Hl2SW

1000

o
-1000

100

200

300

400

500

Data Point Number


Fig. 6.37 Cont.'
249

600

700

1
9000

'

1
1

H4NE

6000

-3000

1
1
1

H4NW

H4N

o
-3000

3000
2000

1
1
1
1
1
1
1

H8NE

1000
,.-._

\O
~

......

o
-1000

'._,t

-....=
cd

2000

-....

1000

C/'j
(1)
(1)

C/'j

H12NE

o
-1000

2000
H12NW

1000

o
-1000

100

200

300

400

500

Data Point Number


Fig. 6.37 Cont.'
250

600

700

1
1
1
1
1

1
1
1

1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1

2L

2000
--

4000

3L

--r---r---r---r---r-----------r--r--'"-----------------.------.-----~

7L

2000

-2000

~--,---,--,---,,--,--,-----;-----;--r-------r---r----.----.----.-----___J

4000

--r---r------r---r--r--r--r--r--r-------------------------------.---.-------~

-2000

--t--,--,,---,,--.--r--r---r---r---r------r--.---.------.----.---....---.----.-----l

100

200

300

Data Point Nunber


Fig. 6.38 TWl: Reinforcing Steel Strain Gage Histories

251

400

1
1

1
1

4000
17L

2000

--

-2000

1
1
1
1

4000
21L
2000
,..-....
\O

- - - 23L

>(

......

-2000

-......

4000

..-

2000

'-"

C:

C'd
...

al

1
1
1

25L

r.r:

--

al
......

r.r:

o
-2000

1
1

15000
10000

30L

- - - 31L

5000

o
-5000

200

100

Data Point Nurrber


Fig. 6.38 Cont.'
-

252

300

400

1
1

1
1

1
1
1
1
1
1

1
1
1
1
1
1

1
1
1
1
1

30000
34L

20000
--

10000

35L

---- 38L

o
-10000
30000
20000

39L

10000

- - 42L

//

'
~
?(
,......

o
-10000

"-"

-......
C:

30000

r:/"J

~
~

......

20000

43L

10000

- - 46L

r:/"J

o
-10000
30000
20000

--

10000

o
-10000

200

100

Data Point Number


Fig. 6.38 Cont.'
253

300

400

1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1

8000
34L37

4000

- - 38L37

o
1

- -- -

A__ _ , , " )

-4000
20000
42L37

10000

j\

--

....-....
\O
~
~

......
..._...

....C:

o
-10000

Cd

3000

cr..

2000

o
o
.....

1000

.....

..-

W7.5SW

cr..

o
-1000

--

W7.5W

----

W7.5NW

2000

--

1000

\~\

W15SE
W15NE

o
-1000

100

200

Data Point Number


Fig. 6.38 Cont.'
255

300

400

1
1

1
1
1
1
1
1
1

3000
W22.5SW

2000

- - W22.5W

1000

----

o
-1000
2000
H3SE

1000

..-...

'U.:J
>(

_...

- - H3SW

o
-1000

'-"'

C
Cd

2000

--

1000

1
1

.,..._

CJj_

o
o
.,..._

CJj_

o
-1000
6000

4000

-2000

----

T9S

o
-2000

200

100

Data Point Number


Fig. 6.38 Cont.'
256

300

400

1
1

1
1

1
1
1

1
1

1
1

HiSSE

1000

--

Hl8SW

1
1
1
1

H33SE

2000
--

H33SW

1
1
1
1
1

H3NE

2500 - ,

o
-2500

--

H3NW

--.~~~~~=--::;.;a.~:;..._~~~~~~~::___;_~

_l_,;..._.......-._-,..-....--....---....---.------.---,----.---.-----.-----.----.-----.---

1000 ------.-----.----.---.---...-~-....---.----,---.--....---,---.--......-....--~
H3ENE

500
--

H3WNW

1
1

100

200

Data Point N umber


Fig. 6.38 Cont.'
257

300

400

1
1
1
1
1
1

2L

7L

2000

1
1

9L

100

200

300

400

500

600

700

Data Point Number


Fig. 6.39 TW2: Reinforcing Steel Strain Gage Histories
258

800

1
1
1
1
1
1

1
1
1
1

1
1

23L

2000
--

24L

a
1

1
1

25L

10000
--

26L

100

200

1
1
1
1
1
1

10000

300

400

500

Data Point Number


Fig. 6.39 Cont.'

259

600

700

800

1
1
24000

1
1
1
1

31L

12000

38L

20000

10000

a
3 0000
20000

--,--.........--,-.......--.--,--,--,-.....--,..-,---.---r-........--,-......--.--.--,--,--..,......--,-_,........_,.....................--.--,,--.---,--..,.........._.__,............

-=

--

41L

j
1
o-----------------------------~----~
44L

10000

100

200

1
1

45L

20000

300

400

500

Data Point Number


Fig. 6.39 Cont.'

260

600

700

800

1
1
1
1
1

1
1

24000

49L

12000

1
1
1
1
1

1
1

1
1

1
1
1

1
1
1

100

200

300

400

500

Data Point Number


Fig. 6.39 Cont.'

261

600

700

800

1
1
26L37

2000

1
1
1

38L37

16000

- - 44L37

8000

49L37

1
WS.SE

2000
--

WS.SNE

1
1
1

Wl6.5NW

2000

--

100

Wl6.5SW

200

300

400

500

Data Point Number


Fig. 6.39 Cont.'

262

600

700

800

1
1
1

1
1

6000
4000

W27.5SE

o
-2000
4000
_j

H3.75SE

2000

\O

,....~

-2000

- - H3.75SW

"-"'

c::
......

cd

3000

en
,......

2000

.-

.......
(1)
(1)

.......

en

T4.SS

--

H4.SS

1000

1
1

-1000

1
1
1

2000

1
1

--

2000

'1
1
1
1
1
1
1

W27.5NE

4000
H8.75SE

--

o
-2000

100

200

300

400

500

Data Point Number


Fig. 6.39 Cont.'
263

600

700

800

1
1
1
1

8000
T9S

4000

- - - H9S

1
1
1

o
-4000
4000
Hl2.5SE

2000
,.,-.....

'

--

H16.5SW

1
1
1

,.......

-2000

cd

3000

rJJ

2000

..._,,
C:

,...
~

.....o
o

H4NE
H4NW

1000

----

rJJ

o
-1000

'1

1000
H4ESE

500

--

o
-500

100

200

300

400

500

Data Point Number


Fig. 6.39 Cont.'
264

600

700

800

1
1
1
1

600'0 ~

1
1

1
1
1
1
1
1

50001

..-..
fil

C.

4000

--=-=
=
e

3000

c-

:;

2000
Concrete Stress-Strain Relations

1000

:A

Modifed Kent-Park

Saatcioglu &. Razvi

-+-~-,-~....-~~--.-~---~-.----------~----------~-----a

0.0000

0.0003

0.0006

0.0009

0.0012

Cu"ature (rad/in)
Fig. 7. 1 RWl: Moment-Curvature Relations

1
1

1
1
1
1
1
1

5000

3000
2000

Concrete Stress-Strain Relations

1000

:A

Modified Kent-Park

Saatcioglu &. Razvi

o -+------,-~------------------~-----------------------------------
0.0000

0.0003

0.0006

0.0009

Cu"ature (rad/in)
Fig. 7.2 RW2: Moment-Curvature Relations
265

0.0012

1
12000

10000 --

8000

1
1

6000
4000
Concrete Stress-Strain Relations

2000

*.

Modified Kent.Park

Saatcioglu &. Razvi

0------------.--------------,---,~-.---.----,----,.-----r----.----l

0.0000

0.0004

0.0008

0.0012

0.0016

Cunrature (rad/in)
Fig. 7.3 TWl: Moment-Curvature Relations

12000 -...----.-------.-..........-----------.----.----.-.....--.---......----,.----,----,.---,

10000

8000

flJ

-...=
C.
1

"-"

=
e
:;=

,.-..,
~

1
1

6000

'

4000

Concrete Stress-Strain Relations

2000

Modified KentPark

Saatcioglu &. Razvi

---~,---,--,---,~-,--,---,-~~-r-----~-r----~~---,---;

0.0000

0.0004

0.0008

0.0012

Cunrature (rad/in)
Fig. 7.4 TW2: Moment-Curvature Relations

266

0.0016

1
1

1
1

1
1
1

1
1
1

1200

.......
fil
C.

:;
,,_..
c,

=
=
-=
~

1
1

1
1

1
1

600
Concrete Stress-Strain Relations

300

<

1
1

1
1
1

900

_J

Modified Kent-Park
Saatcioglu &. Razvi

-300

2000

4000

6000

8000

Moment (in-kips)
Fig. 7.5 RWl: P-M Intera.ction Curves

1200

.......
fil

-,,_..
=
=
--=

900

C.

.:.=

c,

600

Concrete Stress-Strain Relations

300

<

Modified Kent-Park
Saatcioglu &. Razvi

o
-300

4000

2000

6000

Moment (in-kips)
Fig. 7.6 RW2: P-M Intera.ction Curves
267

8000

--=
<
~

2000

1
1
1
1

1000

Concrete Stress-Strain Relations

--~~--~-----:,~-=,...::;..~~----~~~~~--~~-----,

-1000

Modifed Kent-Park

Sattcioglu &. Razvi

-----.----,,...-.....---+-----,----,-----.-"t'--.----,------,-~-.,...---------;

-12000

12000

24000

36000

Moment (in-kips)
Fig. 7.7 TWl: P-M Interaction Curves

1
1

1
1
')

2000

1000

--=
~

<

0 --~.::.,..~~'"'"'"""'~---=..-:::=~--~~C=o=nc=re=~==Stre===s=s-=S=tran===&=el=ab=o=ns=--J

-1000

..

Modifed Kent-Park

Saatcioglu & Razvi

--t--------------,.-..,....--,----r------r-------------r-~--.----

-12000

12000

24000

Moment (in-kips)
F .g. 7. 8 TW2: P-M Interaction Curves

36000

1
1

1
'I

268

1
1

Unconfned

Confined - Modified Kent-Park


Confined - Saatcioglu & Razvi

1
1

\
\

1
1
1

0.00

O.Ol

0.02

0.03

0.04

0.05

Concrete Strain (in/in)


Fig. 7.9A RWl: Analytical Concrete Stress-Strain Relations

1
1
1
1
1
1
1
1

Unconfined

,-_
c:

Confined - Modified Kent-Park

Confined - Saatcioglu & Razvi

'-"
c:
c:

~
.....
en
<1.)
.....

~
u
C

0.00

O.Ol

0.02

0.03

0.04

Concrete Strain (in/in)


Fig. 7.9B RW2: Analytical Concrete Stress-Strain Relations

269

0.05

1
14

fm

Q.

10

.g
C:

"'

:,

m o
.c

(..)

1
1

Analytically Comp. Moment

:i:1
C:
C

Experimentally Meas. Moment


mm Nominal Moment

12

'[

1
1

C:
G)

E
o
:E

4 '

o
RW1

RW2

TW1+

TW1-

TW2+

TW2-

Specimen

Fig. 7. lOA Experimental, Nominal and Analytical Moment Capacities

1
1

Experimentally Measured Shear

Nominal Shear Capacity

1
1
20

o
RW1

RW2

TW1+

TW1-

TW2+

TW2-

1
1

Specimen

Fig. 7.10B Experimental and Nominal Shear Capacities


270

...

1
1
1

ispecimen RW1

:I

1
1

Base Rotation Component


+ Experimentally Measured

1
1

'--~~-'-~~----~~~..._~~---~~.........~~----'--~~---~~--

1
1
1

Top Oisplacement On.)

Fig. 7.11 RW 1: Base Rotation Contribution To Measured Top Displacement

1
1
1
1
1

-2

Specimen RW2: 1

Base Rotation Component


+ Experimentally Measured

1--~~......-~~---~~-------~~...._~~---~~---'~~~-'-~~-'

-2

Top Oisplacement Qn.)

Fig. 7. 12 RW2: Base Rotation Contribution To Measured Top Displacement

271

1
1
1

1
1

Specimen TW1 : 1

1
1

Base Rotation Component

+ Experimentally Measured
4

~~~-'--~~__..._~~--'--~~......._~~-'-~~---J'--~~....-~~--1

-2

Top Displacement On.)

1
1
1

Fig. 7. 13 TW 1: Base Rotation Contribution To Measured Top Displacement

Specimen TW2: 1

1
Base Rotation Component

+ Experimentally Measured
4

'--~~......-~~-'-~~~...-~~....-~~--..~~~...-~~-'-~~---

-2

Top Displacement On.)

Fig. 7.14 TW2: Base Rotation Contribution To Measured Top Displacement

272

1
1
1
1

u-------~-~~------, ),"_ l

O.mr:, .

....

T
1-..:)

c:J

Wall Specimen

Shear

Moment

Curvature

Rotation

Fig. 7.15 Shear, Moment, Curvature and Rotation Distributions For Cantilever Wall Specimen

'"~

1
0.8 . - - - - - - - - - - - - - - - - , . - - - - - - - - - - - - - - .

-c

0.6

ispecimen RW1

;., 0.4

:I

1
1

Q)

E 0.2

Q)
(.)

Q.

.sa
o

o r---------------....--------------1

5-0.2
-0.4

Flexural Component

+ Experimentally Measured

-0.6

-0.8 '-----...1..---.....l-----'------l.---....._--...l.-..--___J_------J
0.8
-0.6
-0.4
-0.2
O
0.2
0.4
0.6
-0.8
First Story Displacement (in.)

Fig. 7.16 RWl: Flexural Contribution To Measured First Story Displacement

1
0.8 . - - - - - - - - - - - - - - - , - - - - - - - - - - - - - - - - - ,

Specimen RW2: 1

-0.8

-0.6

-0.4

-0.2
O
0.2
First Story Oisplacement (in.)

0.4

__

-0.8 .....__ _.....__ ___.___ ___,__ _ _..___ _.......__ _.......__ _

-0.6

0.6

Fig. 7. 17 RW2: Flexural Contribution To Measured First Story Displacement

274

1
1
1
1
3

Flexural Component
+ Experimentally Measured
-

0.6

!
ii:

1
1
1

0.8

1
1
1
1

1
1
1

0.8

----------------r---------------,

0.6
1

Specimen TW1 : 1

1
1

Flexural Component

1
1
1

+ Experimentally Measured

-0.6

-0.8 ~ - - - - - - - - - - - - - - - - - - - - " - - - - - - - - - - - '


-0.8
-0.4
-0.2
O
0.2
0.4
-0.6
0.6
0.8

First Story Displacement On.)

Fig. 7. 18 TW 1: Flexural Contribution To Measured. First Story Displacement

1
0.8 , - - - - - - - - - - - - - - - - , - - - - - - - - - - - - - - - - ,

1
1
1

1
1
1
1

0.6
""":"-

Specimen TW2: 1

0.4

.;.

Q)

E
0.2
Q)

u
m

15.

a
~

o 1----------------~--------------
-0.2

ciS

f!

-0.4

Flexural Component

+ Experimentally Measured

-0.6

-0.8 l....---....L..-----'----L-----L----A-----'----------0.8
0.6
0.4
-0.4
-0.2
O
0.2
-0.8
-0.6
First Story Displacement On.)

Fig. 7.19 TW2: Flexural Contribution To Measured. First Story Displacement

275

Compression

Tension

\ - - s\ip

h;;:;144 in.
t,.:)

-l

O')

Uniform Bond
stress, ub

r-

jd---\

Fig. 7.20 Parameters For Calculating Slippage


of Longitudinal

Reinforcenent

Fig. 7.21 Parameters For Calculating Top


Disp\acenent Caused by S\ippage

------------------~

Lateral Drift ( 0/o)

1
1
1

1
1
1
1

40

-2.8

-1.4

O.O

1.4

2.8

--r-~~----r-~~---,----~--,r--~---,-~~--r-~~-,--~--.~~--,

RWl: P = 0.1 OA 8 f ~

-...=
.

.;;

Experimental

-20

-4

-2

Modified Kent-Park

Saatcioglu & Razvi

Top Displacement (in.)

1
1
1

Fig. 7.22 RWl: Monotonic Lateral Load vs. Top Displacement Envelopes

Lateral Drift ( 0/o)


40

-2.8

-1.4

O.O

1.4

2.8

--r-~~----r-~~--,--~~,--~--r-~~---.---~~-.-~--,,--~---,

RW2: P = 0.07Agf~

1
1
1

-...=
.

.;;

Experimental

-20

Modified Kent-Park

Saatcioglu & Razvi

-4

-2

Top Displacement (in.)


Fig. 7.23 RW2: Monotonic Lateral Load vs. Top Displacement Envelopes

277

-1

Lateral Drift (lo)


80

-2.8

O.O

-1.4

1.4

2.8

--~~-,-~~--~-,-~~-.-~---,,--~-r-~~-,-~--.

TWl: P = 0.09Agf~

-f:
~

Experimental

-40

-4

-2

Modifed Kent-Park

Saatcioglu &. Razvi

Top Displacement (in.)


Fig. 7.24 TWl: Monotonic Lateral Load vs. Top Displacement Envelopes

Lateral Drift (%)


-2.8
-1.4
O.O
1.4
2.8
100--T----,,--~.....-~------~~--~--~---------

50

-100

Experimental

Modifed Kent-Park

Saatcioglu &. Razvi

--------.----,-..----.----,.---,-----r--------.-------.---..----.----

-4

-2

Top Displacement (in.)


Fig. 7.25 TW2: Monotonic Lateral Load vs. Top Displacem~nt Envelopes
278

1
1
1
1
1
1

1
1

1
1
1
1
1

TW2: P=0.075Agf~

-5 O

1
1

..

'I
1
1
1
1

1
1

Lateral Drift (%)


-0.70

..-..

0.00

0.35

0.70

s_
:;

RWI: P=O.lOAgf~
50

'-"

,:

-f
~

o --+-------------w

-50

-1 00

--t----,---.--,---,---......---,---,----,----.---+--,---,.---,,---,---,----,---......---,---...---

O.O
0.5
1.0
Top Displacement Due To Pedestal Rotation (in.)

-1.0

1
1
1

-0.35

lQQ-,-......-.-.-...,..-,-~-r--r-------r--r-r-~-.--.-.----..-----------

-0.5

Fig. 7.26 RWl: Top Displacement Due To Peclestal Rotation

Lateral Drift (%)


0.70
-0.35
0.00
0.35
-0.70
100--......-.------------~---,-._...-.-~----..-.--------.-.-~

RW2: P = 0.07Agf~
..-..
~

C.

1
1

:;

1
1
1
1
1

50

'-"

"C

-f
~

O-+----------~----------,

-50

O.O
0.5
1.0
Top Displacement Due To Pedestal Rotation (in.)

-1.0

-0.5

Fig. 7.27 RW2: Top Displacement Due To Peclestal Rotation


279

Lateral Drift ( 0/o)


-0.70

-0.35

0.00

0.35

0.70

TWl: P = 0.09Agfc

-1.0

-0.5

O.O

0.5

1.0

Top Displacement Due To Pedestal Rotation (in.)


Fig. 7.28 TWl: Top Displacement Due To Pedestal Rotation

Lateral Drift ( 0/o)


0.70
-0.35
0.00
0.35
-0.70
lOQ~-......-,-----.-----....-r--.--...-----.-------.----.-___,-,---,-.,~-.-.......,

1
1
1
1
1

1
1

TW2: P = 0.075Agfc
50

-=

1
1
1

-
.:

O ----------~~

. -50

-1.0

-0.5

O.O

0.5

1.0

Top Displacement Due To Pedestal Rotation (in.)


Fig. 7.29 TW2: Top Displacement Due To Pedestal Rotation

280

1
1
1
1

1
1

1
1
1

Al

_J

1
1

1
1
1
1

1-D-i

Asfs
Asfs

_j

Asfs

1~ 0 ~I

Confinement
Pressure

Asfs

fr

Asfs

fr

Asfs

l ll l

Confinement
Pressure

1
1

Section

Section A-A

Fig. 7.30 Evaluation of Confinement Effectiveness

1
1

1
1
1

281

8-8

1
1
1
1
1

0.06

0.04

Experimental Drifts

Analvtical Drifts

--e-

1.00/o

--e-

1.00/o

-+-

1.So/o

-8-

l ..So/o

-*-

2.00/o

---$-

2.00/o

0.02

1
1

North ..---~~ South


-0.02

--l-----.-----,----------------.----,------,r------1

12

24

48

36

Distance Along Web (in.)


Fig. 8.lA RWl: Measured. Steel Strain vs. Anal. Strain Distribution (Pos. Disp.)

Experimental Drifts

Analllical Drifts

--e-

--e-e-

-+_._

.oeo

l.S%

2.0%

---$-

1.00/o
l.5'1'o
2.00/o

0.02

..

North ...--- .~ South


-0.02

--+-----r----,-----,-----,----,----~---,-----

1
1
1

0.06

0.04

12

24

36

48

Distance Along Web (in.)


Fig. 8.lB RWl: Measured Steel Strain vs. AnaL Strain Distribution (Neg. Disp.)

1
1
1

282

1
1
1
1

0.06
Anal:rical

--e---e-

0.04

Exnerimental Drifts

Drifts

-e-

1.00o

-+-+-

1.!5%

~ 2.0%

1.0%

U%
2.00.4

0.02

1
1
1
1

1
1

1
1

North

1
1
1

South

-0.02 - + - - - - - - - - - - - - . - - - - - - - - - - - - - - - - - '

12

24

36

48

Distance Along Weh (in.)


Fig. 8.2A RWl: Measured Concrete Strain vs. AnaL Strain Distribution (Pos. Disp.)

0.06

0.04

AnaMical Drifts

Exnerimental Drifts

--e---e-

1.004

-e-

1.0%

1.!5%

-+-

U%

~ 2.004

-*-

2.0%

0.02

....---..

North ....---.. South


-0.02

---------------..---------------.----4

12

24

36

48

Distance Along Weh (in.)


Fig. 8.2B RW 1: Measured Concrete Strain vs. Anal. Strain Distribution (Neg. Disp.)
283

f,,.' vnr
il

, ti.}

1
1
0.06

1
Analytical Qti~

Ex3rimental Drifts

---

0.04

...-+-

-*-

1.0%

-e--

l.004

U%

-B-

1.5%

2.0%

--&-

2.00/o

1
1

0.02

North ....---IIJ~
-0.02

South

-------r----r------,.--~--.....---......----.----

12

24

36

48

Distance Along Web (in.)


Fig. 8.3A RW2: Measured Steel Strain vs. Anal. Strain Distribution (Pos. Disp.)

1
1
1
1

0.06 - - - - - - - - - - - - - - - - - - - - - - - - - - - - -

---

-+-

1
1

Analytical Qtifts

Experimental Qtifts
1.0%
1.5%

-e--

1.004,

-8-

1.5%

--$-

2.004

1
1

_....
a.)
a.)

ij5

0.00

North
-0.02

lJ

South

--+---------------------------

12

24

36

Distance Along Web (in.)


Fig. 8.3B RW2: Measured Steel Strain vs. AnaL Strain Distribution (Neg. Disp.)
284

48

0.06 - - - . - - - - - - - - - - - - - - - ~ - - - - - - - - . - - - - - -

1
1
1
1
1

1
1
1
1
1

Experimental Drifts

Analytical Drifts

---

1.00/o

2.00...

-+-

.,o/o

-6-8-

--k-

2.00o

--$-

1.0%

North
-0.02

Il

South

--+---------------------------'
o

12

24

48

36

Distance Along Web (in.)


Fig. 8.4A RW2: Measured Concrete Strain vs. AnaL Strain Distrihution (Pos. Disp.)

0.06

--r----,----r""--------~--,----------

Experimental Drifts

0.04

---

-+-*-

1
1
1
1
1
1
1

Uo/o

North

l.OOo

Analytical Drift

-6-

1.00o

.,o/o

-a-

.,%

2.00o

--$-

2.00o

.r.---~...,. South

-0.02 - - ; - - - - - , - - - ~ - - - - - ~ - - - - - - - - - - - - '

12

24

36

48

Distance Along Web (in.)


Fig. 8.4B RW2: Measured Concrete Strain vs. Anal. Strain Distribution (Neg. Disp.)
285

1
0.06
Analyical
~

C
......

=
......
C
......

0.04

Drifts

-v--

r\

1.00/o

-B-

l.So/o

-e-+-

~ 2.00/o

-*-

'.._;'

...C'd

1
1
1

Ex3rirnen11l Drifts
1.0%

U%
2.QOA,

0.02

Cl"J
.....
(1.)
(1.)

Cl':.

0.00

North

South

-0.02

12

48

36

24

Distance Along Web From Flange (in.)


Fig. 8.5A TWl: Measured Steel Strain vs. Anal. Strain Distribution (Pos. Disp.)
0.015
f;xoerimmta

0.010

-+-e-+-

C
......

=
......
C
......

0.005

'.._;'

...

C'd

Drifts

--6-

0.7.SIYo

1.0%

-e-

l.OOA,

1.2.5%

-B-

1..50/o

0.000

Cl"J
.....
(1.)
(1.)

-0.005

Cl"J
-0.010

North ...---l South

-0.015

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 8.5B TWl: Measured Steel Strain vs. Anal. Strain Distribution (Neg. Disp.)

286

Analvtical 12rill!

0.7.S%

1
1
1
1
1
1

1
1

0.03

--,----,----,------r----,------..-----...---

1
1
1
1

1
1
1
1
1
1
1
1

1
1
1

Analytical Drifts

Experimental Drifts

--6-

0.1S%

-+-

-8-

l.OOAt

--e-

-8-

.sAt

-+-

0.1S%

1.0%

Questionable Data (see discussion in section 6.6)

North ...---111 South

~.01

-+---,---......-----r~---,--~----,---------~
O

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 8.6A TWl: Measured Concrete Strain vs. AnaL Strain Distribution (Pos. Disp.)

0.015
~

Experimntal Drifts

0.010

"-"

0.005

C:

-=
-.....

0.000

.....oo

~.005

C:
C'd
-.

r.r

-.

Analytical Drifts

0.7S%

--6-

0.7S%

1.0%

-e-

1.0-At

1.25%

-9-

1.5%

(.,)

C:

~.010

Noth ...---~ South


~.015

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 8.6B TWl: Measured Concrete Strain vs. AnaL Strain Distribution (Neg. Disp.)

1
1

287

1
1
1
1
1
1

0.06

..-...

C:
......
C:
......
,,,_.,
C:
......
c:j
-

0.04

Ex~rimental D:rifts

Analytical Drifts

-e-+-

1.0%

-e-

1.0-e

U%

-8-

l.S%

2.0%

2.0-e

-*0.02

--

C/"J.
(1..)
(1..)

C/"J.

0.00

North

-0.02

12

South

24

36

48

Distance Along Web From Flange (in.)


Fig. 8. 7A TW2: Measured Steel Strain vs. Anal. Strain Distribution (Pos. Disp.)

0.020 - - - - - , - - - - - - - - - - , - - - - - - - - - - - - -

Analyyal Drifts
1.0-At

0.010

1.5%
2.0-e

Experimental Drifts

-e-+-*-

1.0%
U%
2.0%

--

(1..)
(1..)

-0.010

..

-------,---...---------------~

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 8. 7B TW2: Measured Steel Strain vs. Anal. Strain Distribution (Neg. Disp.)
288

1
1
1
1

North -4111l--- 9a,. South


-0.020

1
1
1
1

1
1

1
1
1

0.03

Analytical Drifts

1
1

0.02

1
1
1
1
1
1

0.7S%

-e-

.ooo

-8-

1.S%

Experirnental Drift

-+-e-+-

0.7'%
1.0%
1.S%

0.01

1
1
1
1
1

-,----,-----r------,~--r----.---------

North
-0.01

_....4---..l

South

--;----r------,-----------------1

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 8.8A TW2: Measured Concrete Strain vs. AnaL Strain Distribution (Pos. Disp.)
0.015
~

-..._...
-=
-....
t::
t::

Analytical Drifts
0.010

0.7'%

1.0-o

0.005

1.,%

Experimenta Prifts

-+-e-+-

0.7'%
1.0%
1.S%

e'd
...

'(/').

....

0.000

(1)
(1)
...

u
t::

-0.005

-0.010

North -,..4---..l South

-0.015
O

12

24

36

48

Distance Along Web From Flange (in.)


Fig. 8.8B TW2: Measured Concrete Strain vs. AnaL Strain Distribution (Neg. Disp.)

289

Appendix A
DISPLACEMENT-BASED DESIGN

This appendix describes the displacement-based design procedure that was used to
design the wall specimens tested in this study. A detailed outline of the methodology and
its implementation are presented; however, the complete description of the procedure and
its applications are available elsewhere (Wallace and Thomsen, 1993).

1
1
1

1
1

1
1

1
OVERVIEW OF DISPLACEMENT-BASED DESIGN
Unlike current code procedures which are strength-based, a displacement-based approach compares directly the expected displacement capacities and displacement demands
for the building. Therefore, reinforcing details are directly related to expected building
response. The overall process can be summarized as follows:
(1)

Ensure minimum level of building strength

(2)

Characterize earthquake demands at building site

(3)

Characterize global (building) deformations

(4)

Characterize local ( structural element) deformation capacity

(5)

Relate global and local deformations

(6)

Establish detailing requirements

The -following sections-describe -the six various topics of the displacement-based approach as they pertain to structural wall systems in regions of high seismic risk.

290

1
1
1

1
1

1
1

1
1
1

1
1

1
1
1
1

1
1
1

1
1

1
1

BUILDING STRENGTH DEMANDS


A building must possess a minimum amount of flexural-strength to ensure that damage
does not occur during minor and moderate intensity earthquakes. Current U.S. codes (Uniform Building Code, 1991) are used to prescribe lateral forces that are used to proportion
structural members.

In this particular study, a design base shear value was calculated using UBC-91, equation 34-1. The base shear was calculated based on the following values: w=l 75 psf, Z=0.4,
l=l, S=l.2, T=0.02(hn) 3 / 4 =0.4 sec, and Rw=l2. The structural walls were conservatively
assumed to resist all of the lateral force. The base shear was distributed over the height
of the building in accordance with UBC-91, equation 34-8, andan overturning moment at
the base of each wall was then calculated. Tributary axial loads were approximated and
longitudinal reinforcement was designed to resist UBC-91 specified load combinations. It
was assumed that both boundary and web longitudinal reinforcement contribute to wall
flexural-strength. The computer program BIAX (Wallace, 1992) was used to verify wall
flexural-strengths.
The required flexural-strength decreases over the height ofa building; therefore, the
amount of longitudinal reinforcement can be curtailed as the height above the base increases.
During severe earthquakes a large amount of inelastic deformation is expected. It is common
practice to target these inelastic deformations to take place in a well detailed portion at the
base of the structural wall (critical section) referred to as a plastic hinge region. However,

1
1

due to higher mode effects, adequate flexural-strength should be provided over the wall

2lw) is recommended (Paulay 1986, 1991). Figure Al shows an example ofa typical wall

Based on this strength envelope longitudinal reinforcement can be terminated at various

291

height such that large inelastic deformations do not occur at levels where special detailing
is not provided. To ensure that the plastic hinge occurs at the base of the wall a linear
strength (moment) envelope with an offset (the greater of lw or hw/6, but need not exceed

and a flexural-strength envelope utilizing the recommended offset at the base of the wall.

levels above the base.

1
CHARACTERlZATION OF EARTHQUAKE GROUND MOTIONS
Studies reported by Newmark and Hali (1982) and Shimazaki and Sozen (1984) examine the maximum inelastic and elastic displacement responses of single-degree-of-freedom
oscillators. These studies find that if the initial period exceeds the characteristic ground
period (approximately the period at which the constant acceleration and constant velocity
regions coincide, or approximately 0.5 sec for firm soil sites) the maximum displacement is
nearly independent of strength. If the initial period is less than the characteristic ground
period, inelastic displacements are expected to be larger than the elastic displacements.
Newmark and Hail (1982) present an approach to estimate inelastic displacements from the
elastic displacement spectrum; therefore, elastic displacement spectra are suitable tools for
estimating maximum displacement response.
Displacement response spectra can be developed for a particular building site, or generalized spectra such as those presented in ATC-3-06 (1978) can be used. In this section,
a generalized spectrum will be used to present the fundamentals of the displacement-based
approach. According to ATC-3-06, spectral acceleration for elastic response is
Sa

_ l.2AvS
-

Tf

(Al)

except Sa need not exceed 1.0 for soil types 1 (8=1.0) and 2 (8=1.2), and need not exceed
0.8 for soil type 3 (8=1.5) (It is noted that subsequently a fourth soil condition was added
to ATC-03-06.J In Eq. (Al), T is the fundamental period of vibration of the structural
system (in the direction under consideration), Sis a factor to account for soil characteristics,
and Av is a factor to account for seismicity. The two-thirds exponent was included in Eq.
(Al) as a safety factor for taller buildings and longer period structures for which higher

1
1
1
1
1

1
1
1

1
1
1

modes may significantly influence internal forces. Since displacements are not likely to be
influenced significantly by higher modes, and because the approach being developed is based

292

1
1

1
1
1
1
1
1

on displacements, the two-thirds exponent is dropped. With the modified exponent, the
elastic spectral displacement can be computed in terms of the elastic spectral acceleration
as:
(A2)
Assuming values Av=0.4 and S=l.2 (for firm soil sites), an expression for elastic spectral
displacement is obtained from Eq. (Al) and (A2) as follows:
10T 2

(in.)

O.O :5 T :5 0.585 sec

(A3a)

Sd = 6T

(in.)

0.585 sec < T

(A3b)

Sd = 25T 2

(cm)

O.O :5 T :5 0.585 sec

(A3c)

Sd = 15T

(cm)

0.585 sec < T

(A3d)

sd

1
1

The period range for which Eq. (A3a) and (A3c) are applicahle represents a region of

1
1
1
1
1
1

United States. Equations (A3a) and (A3h) are plotted in Fig. A2.

1
1

constant acceleration, whereas Eq. (A3h) and (A3d) represent a region of constant velocity. The elastic displacement response spectrum as given hy Eq. (A3) may he considered
representative of fi.ve percent damped spectra for strong ground motions on firm soil in the

The relationships plotted in Fig. A2 can he modified to incorporate the effects of inelastic response hased on the procedure presented hy Newmark and Hall (1982). [It is noted
that other procedures are availahle such as that proposed by Qi and Moehle (1991)]. The
procedure derives inelastic acceleration and displacement spectra from the elastic spectrum
given a displacement ductility factor. Figure A2 also includes an inelastic displacement
spectra for a displacement ductility ratio of five, and a simplified spectrum (Sd

6T in.).

Figure A2 reveals that the linear spectrum (Sd = 6T in.) gives an envelope of the inelastic
spectra; therefore, the linear spectrum may he used to provide an estimate of the maximum
elastic and inelastic displacement for all periods (the estimate will tend to he conservative
for periods less than approximately 0.3 sec). A linear spectrum (Sd = 6T in.) is used in all
suhsequent calculations for spectral displacement.

293

1
.

ESTIMATING GLOBAL BUILDING RESPONSE


Estimates of building displacement can be obtained from a displacement spectrum
given an estimate of the fundamental period. For structural wall buildings, the fundamental
period based ona cracked section stiffness can be estimated as (Wallace and Moehle, 1992)

(A4)
in which n is the number of floors, w is the unit floor weight including tributary wall weight,
h 8 is the mean story height, Ec is the concrete modulus of elasticity, hw is the wall height,
and p is the ratio of wall area to floor plan area for the walls aligned in the direction the
period is calculated (p= I: Aw/ Ar, where Aw = lwtw, lw is the wall length, tw is the wall
thickness, and Ar is the floor plan area ofa typical floor of the building). Equation (A4)
assumes all walls have the same cross-section ( the equation can be modified to account for
walls with different cross-sections). The validity of Eq. (A4) was verified by comparison
with measured periods for shear wall buildings (Wallace and Moehle, 1992; Yan and Wallace,
1993).
For a shear wall building, roof displacement can be approximated by 1.5 times the
spectral displacement ( to account for the difference between the displacement of a singledegree-of-freedom oscillator and the building system the oscillator represents). Therefore,
the roof drift (roof displacement divided by building height, 6u/hw) can be computed by
multiplying Eq. (A3b) or (A3d) by 1.5 and dividing by the building (wall) height, hw,
(A5)
The roof drift can be expressed in terms of the wall aspect ratio and the ratio of wall area
to floor plan area by substituting the period Tas given by Eq. (A4) for Tin Eq. (A3), and
substituting the result into Eq. (A5).

(A6)

294

1
1
1
1
1

1
1
1

1
1

1
1
1
1
1

1
1
1

1
1
1
1

1
1

1
1

1
1
1
1

For typical values of w=l 75 psf (875 kg/m 2 ), h5 =108 in. (275 cm), g=386.4 in./sec 2 (981
cm/sec 2 ), and Ec=3,500 ksi (250,000 kg/cm 2 ), Eq. (A6) can be expressed as Eq. (A 7).
u
hw

O. 00023 hw
lw

{I

(A7)

VP

Equation (A 7) is plotted in Fig. A3 for several wall aspect ratios and is valid for elastic or
inelastic response. The roof drift ratio obtained should tend to be conservative because the
period T given by (A4) is based on half the gross-section stiffness (and the "typical" values
are selected to produce a high drift estimate).
The roof drift ratio provides an estimate of the global deformation demands on the
building. The global deformations of the structure are related to local deformations imposed
on the wall cross-section in the following section.

RELATING GLOBAL AND LOCAL DEFORMATIONS


The deformation imposed on individual walls as a result of the global building deformations can be evaluated using well established procedures to account for the distribution
of elastic and inelastic deformations over the wall height. Based on the model of Fig. A4,
the displacement at the top of the wall can be computed as

(AS)
where, 611 is the displacement resulting from elastic deformations, 8phw is the displacement
resulting from inelastic deformations, hw is the wall height, lw is the wall length, <t, 11 is
the yield curvature (curvature at first yield of the wall boundary reinforcement), <l>u is
the ultimate curvature, lp and 8p are the plastic hinge length and rotation, respectively.
Based on this relation and assumptions for yield curvature (0.0025/lw) and plastic hinge
length (0.5lw), the deformations imposed ona wall can be derived in terms of the ultimate

1
1

curvature times the wall length (Wallace and Moehle, 1992).


<Pulw =

2Z:: +

0.0025 [ 1 - 1 hw]

295

2~
hw

(A9)

1
If Eq. (A 7) is substituted in Eq. (A9), the deformation imposed at the base of the wall can
be expressed directly in terms of the building configuration as Eq. (AlO)

hw]

= 0.0025 [ 1 - -1 2 lw

</>.lw

hw

0.00046lw

(AlO)

Equation (AlO) describes the deformation (ultimate curvature) imposed on the wall crosssection. The need to provide concrete confinement can be evaluated by comparing directly
the deformations imposed on the wall cross-section with the available deformation capacity
of the wall cross-section. Wall deformation capacity is estimated in the following section.

1
1

1
ESTIMATING WALL DEFORMATION CAPACITY
The deformation capacity ofa wall cross-section can be estimated using the model of
Fig. A5. The wall has uniformly distributed reinforcement plus boundary steel. Axial load
is centered on the wall web. The longitudinal tension and compression reinforcement is
assumed to develop a stress of af1J and "'Y f y, respectively, to account for possible material
overstrength and strain hardening (o=l.50 and ""(=1.25 are used in the subsequent analyses).
Based on equilibrium of the wall cross-section, the following relations can be derived for
rectangular, T- and L-shaped walls (Eq. A11, Fig. A5a) and barbell-shaped walls (Eq.
A12, Fig. A5b)

fc,ma:r

fc,maz

[(

__E_,]

+ p" - :1.p') ~ + lutu/c


( 0.85/31 + 2p" o/t-)
Q

/~

</> l

(AH)

111

1
1

1
1
1

distributed steel reinforcing ratio, /31 is a factor defined in ACI 318-89 section 10.2. 7.3, P

1
1
1
1

296

(p + p") el,__ p''11;.. _

where

fc,max

le.

le.

( 0.85/31

0.85b(a _
tulu

+ 2p" o/t-)

is the extreme fiber concrete strain, p =

tw)

__E_,]

+ lutulc.

A/t 111 l 111

</> l

111

(A12)

is the tension steel rein-

forcing ratio, p' = A~/t111 l 111 is the compression steel reinforcing ratio, p11

A~ /twlw is the

1
1
1
1

1
'l.
1
1

1
1
1
1
1

is the axial load, fy is the nominal yield stress of the steel, f~ is the stress in the compression steel, tw is the web thickness, b is the length of the boundary element, and a is the
thickness of the boundary element (Fig. A5b). For T- and L-shaped walls, the maximum
extreme fiber compression strain arises when the flange of the T or L-shaped section is in
tension; therefore, an effective flange width must be determined. Paulay (1986) suggests an
approximate approach to estimate the effective flange width (for slender walls, the entire
flange is typically effective). All reinforcement within the effective flange width should be
included in calculating the tension reinforcing ratio p. For Eq. (A12), the neutral axis is
assumed to be within the wall web. The stress in the compression reinforcement can be
assumed to be at yield; however, this should be verified. A similar expression can be derived
for the neutral axis within the boundary element and an iterative approach should be used
to verify the stress in the compression reinforcement and to converge to a unique solution.
It should also be noted that Eq. (A11) and (A12) are valid only where the depth of the
neutral axis is not greater than one-half the wall length, lw.
Equations (A11) and (A12) can be used directly with Eq. (AlO) to determine the
maximum compressive strain that will develop at the wall boundary. At first glance, the
equations appear to be too detailed for design; however, all terms in the equations are readily
available and no detailed analysis or design is required. In addition, for symmetrically
reinforced walls, the equations can be simplified. Application of the equations to preliminary
design and in the evaluation of existing construction are discussed elsewhere (Wallace and
Thomsen, 1993).
The need to provide concrete confinement at the wall boundary can be established by
examining the maximum concrete strain and the distribution of concrete strain along the

1
1
1

wall cross-section. This topic is considered in the next section.

METHODOLOGY OF DETAILING REQUIREMENTS


The need to provide concrete confinement at the boundaries of structural walls can be

297

evaluated by substituting Eq. (AlO) into Eq. {AH) or (A12). The equations indicate that
the maximum concrete compressive strain for a structural wall depends on general response
quantities ( the ground motion characteristics at the building site and the fundamental
period of the building) and the building and wall attributes (wall reinforcing ratios, wall
axial stress, material properties, wall aspect ratio, and ratio of wall area to floor plan area).
This section is organized into four subsections: (1) maximum concrete compressive strain,
(2) required zone of confinement, (3) transverse steel spacing requirements for buckling and

1
1

(4) design for shear.

(1) Maximum Concrete Compression Strain

Figure A6 plots the computed extreme fiber compression strain for symmetrically reinforced walls (for ;

= 4000 psi, f 11

= 60 ksi, p = p1 = O.Ol, p"

0.0025), and reveals that

extreme fiber compression strain (1) increases with the level of axial stress (Fig. A6a), (2)
increases with wall aspect ratio (Fig. A6b), and (3) decreases as the ratio of wall area to
floor plan area increases (Fig. A6a and A6b). Figures A6a and A6b provide a convenient
means of evaluating the need to provide transverse reinforcement for concrete confinement.
For example, given an axial load of P = 0.lOAwJ;, Fig. A6a (or A6b) reveals that where
the wall area to floor area ratio in one direction of the building exceed.s O. O1 and the wall
aspect ratio is less than fi.ve, the extreme fiber compression strain is 0.0045 or less. Since
structural elements are capable of extreme fiber compression strains of 0.004 without significant deterioration in strength, the analysis indicates that there is a large class of buildings
where special detailing at the wall boundaries is not required.
Figure A 7 plots computed extreme fiber strain for unsymmetrically reinforced wall
cross-sections for h 10 /l 10 = 5 and P = .D.lOAwf~ (for ;

4000 psi, f 11 = 60 ksi, p

O.Ol to 0.02, p' = O.Ol, and p" = 0.0025). (For example, a T-shaped wall where all of
the flange reinforcement is assumed to be effective as tension reinforcement; therefore,

1
1
1
1
1
1
1
1

p exceeds p'.) Figure A 7 indicates that the computed extreme fiber compression strain

298

1
1
1
1
1
1

increases as the difference in the tension and compression steel reinforcing ratios increase.
Therefore, design of walls with T- or L-shaped cross-sections are more likely to require
special transverse reinforcement for concrete confinememt and to suppress buckling of the
longitudinal reinforcement (in the stem of the wall).
Current US code formats generally result in the use of relatively few shear walls; therefore, ratios of wall area to floor plan area of O. 5 to 1. 0% are common. From a review of Fig.
A6 and A 7 it is apparent that the ratio of wall area to floor plan area plays a major role
in the expected performance of shear wall buildings. At ratios less than 1%, Fig. A6 and

A 7 indicate a substantial increase in extreme fiber compression strain. In addition, where

tially result in significant torsional effects; therefore, well confined boundary elements are

relatively few walls are used to resist lateral loads, damage to one of the walls could poten-

necessary to ensure adequate behavior. Conversely, Fig. A6 and A 7 indicate a substantial

1
1
1
1

1
1
1

reduction in extreme fiber compression strain as the ratio of wall area to floor plan area
is increased from 0.5 to 1.5%; therefore, as an alternative to current US design tendencies,
construction utilizing a greater number of walls might allow for greater design flexibility by
alleviating the need for well confined boundary elements.

(2) Required Zone of Conflnement


The depth of the wall cross-section that requires confinement can be determined from
Eq. (A11) (or Eq. (A12) for barbell-shaped walls) if it is noted that

fc,max

= ct/>u,

where

c is the depth of the compression zone. The resulting equation gives the depth of the
compression zone as a function of the wall length (lw ). Given the extreme fiber compressive
strain and the depth of the compression zone, the portion of the wall cross-section requiring
confinement can be estimated if it is assumed that concrete confinement should be provided
at locations where the concrete compressive strain exceeds a limiting value, for example
0.004 (recommended by Paulay, 1986, 1991; Wallace and Moehle, 1989, 1992).
Figures AS and A9 plot the required length of confinement for a limiting compression

299

strain of 0.004 for symmetrically (Fig. A8) and unsymmetrically (Fig.


walls with rectangular, T- and L-shaped cross-sections (for

:=4,000

A9) reinforced

psi, f = 60 ksi,

p"=0.0025). Figures A8 and A9 provide a direct approach to determine the required depth of
confinement for the wall cross-section, and also reveal that concrete confinement is unlikely
for symmetrically reinforced walls with low to moderate levels of axial stress. For walls that
require concrete confinement, the amount of transverse steel required depends on the level of
concrete compressive strains. This topic is discussed in the following section ("Application
of Detailing Requirements" ).
(3) Transverse Steel Spacing Requirements for Buckling
The amount of transverse steel required at the critical portion ofa wall cross-section
depends not only on the need to confine the concrete, but also on the tendency of the
longitudinal reinforcing steel to buckle.

The tangent modulus theory (see Salmon and

Johnson, 1990, for a review of buckling models) can be used to estimate the required
spacing of transverse steel to prevent buckling of compression steel. Recent research on
buckling (Mau, 1990) indicates that use of the tangent modulus theory (with an end restraint
coefficient provided by the transverse steel of 0.5) is appropriate for transverse spacing-tolongitudinal reinforcing bar diameters expected at wall boundaries. Therefore, a restraint
coefficient of 0.5 is used in this paper, and results in Eq. (A13).

Smaz

In Eq. (A13),

Smax

= 1.6

db{f;

(A13)

is the maximum spacing of the transverse steel to suppress buckling

of the longitudinal reinforcing steel, db is the reinforcing bar diameter, Et is the tangent
modulus of the reinforcing steel, and / 8 is the reinforcing bar stress. Based on Eq. (A13),
required spacing can be estimated fora given-level of compressive strain (and an assumed
steel stress-strain relationship ). Therefore, the quantity of transverse steel required to both
confine the concrete and restrain buckling of the reinforcing steel can be evaluated in terms
of the strain distributions described by Eq. (AlO) and (A11) for walls with rectangular,

300

1
1
1
1
1
1

1
1
1
1

1
1
1
1
1
1
1
1
1

'

1
1
1
1

1
1

'1
1
1

1
1

1
1
1
1
1

T- and L-shaped cross-sections, and Eq. (AlO) and (A12) for walls with barbell-shaped
cross-sections. Use of Eq. (AlO) through (A12) should allow for increased design flexibility
because they enable the designer to evaluate minimum spacing requirements based on the
expected building performance (wall deformations); whereas, current code provisions are
indiscrimina te.

(4) Design For Shear


For bearing wall buildings, the ratio of wall area to fl.oor plan area is usually sufficient to
preclude wall shear failure modes. However, frame-wall interaction results in an increase in
the slope of the moment diagram over the lower levels of the wall, and thus increases the level
of shear stress that must be resisted by the wall compared with cantilever walls. In addition,
for walls with coupling beams, a greater portion of the lateral ~hear is resisted by the wall
with increased axial load. Therefore, greater attention to design for shear is required for
frame-wall and coupled wall systems compared to bearing wall systems. Additional research
is needed to address the effects of shear on frame-wall systems constructed with rectangular
walls. However, until additional information is available, a reduced level of maximum design
shear stress is recommended. For example, a maximum shear stress of

6J'H psi (0.5J'H

MPa) as recommneded by Aktan and Bertero (1985) could be used compared with current
UBC-91 values of 8J'H psi (0.67

JH MPa) for all walls sharing a common lateral force and

lOJ'H psi (0.83/H MPa) for an individual wall.

APPLICATION OF DETAILING REQUIREMENTS


Based on the computed compressive strain distribution using Eq. (AlO) through (A12)
and requirements for buckling using Eq. (A13), transverse reinforcement required to provide concrete confinement and to restrain buckling of the longitudinal reinforcement can be
evaluated. This section is organized into two subsections: (1) requirements for transverse
reinforcement, and (2) requirements for shear. The first subsection discusses the requirements for concrete confinement and buckling as well as depth and height of the required zone

301

of confinement. Simplified design expressions have alsa been developed and are available
elsewhere (Wallace, 1995 ).

(1) Requirements for Transverse Reinforcement


The following sections specify required transverse reinforcement based on maximum
compression strain less than 0.004 and greater than 0.004, respectively.

(a)

fc,max ~

0.004:

Far ext.reme fiber compression strains less than O. 004, no special transverse reinforce-

1
1
1
1
1

ment is required far concrete confinement; therefore, the following relations are recommended to provide minimum transverse reinforcement based on current ACI 318-89 requirements.

(i) Boundary Reinforcement:


Minimum transverse reinforcement as specified in Chapter 7, Section 7. 10.5, of the
ACI 318-89 building code should be used. in particular, tie spacing should not exceed
16 longitudinal bar diameters, 48 tie bar or wire diameters, or the least dimension of

the compression member. in addition, a maximum spacing of 12 in. (30 cm) is alsa
suggested for the plastic hinge region (at the base of the wall).
(ii) Web Reinforcement:
Minimum transverse reinforcement as specified in Chapter 21 of the ACI 318-89 building code shouldbe used. in particular, reinforcing ratios and spacing limits as specified
in ACI Section 21.5.2.1 should be used (longitudinal and transverse reinforcing ratios
of at least 0.0025 anda maximum spacing of 18 in. (45 cm)).

302

1
1
1
1
1
1
1
1
1
1
1

E
1
1
1
1
1
1
1
1
1
1

1
1

{b)

fc,max

> 0.004:

For extreme fiber compression strains greater than 0.004, special transverse steel is
required for concrete confinement and to prevent buckling of the longitudinal reinforcing
hars. As discussed previously, special transverse reinforcement for concrete confinement
should be provided over the portion of the wall cross-section where the compressive strains
exceed 0.004. A depth {lw/10) of confined zone is recommended to ensure a minimum
depth of confined cross-section for cases where the strain gradient is large ( and thus, only
a small portion of the wall cross-section would require special transverse reinforcement ).
The following paragraphs present recommendations for (i) concrete confinement and (ii)
buckling of longitudinal reinforcement.

(i) Concrete Confinement:


Required transverse reinforcement for concrete confinement for a given wall crosssection can be evaluated using readily available computer programs capable of generating load-deformation relations. In computing the load-deformation relations, a
model that accounts for the variation in confinement effectiveness for rectangular crosssections should be used (for example, Saatcioglu and Razvi, 1992); however, such a
process does not lend itself well to a design environment. One possible approach would
be to base required transverse reinforcement on current ACI provisions (Eq. 21-3 and
21-4) for the entire depth of the compression zone where strains exceed 0.004. Although
convenient, this approach is restrictive because minor changes in building confi.guration
= 0.004

can lead to significant changes in required transverse reinforcement (at

1
1
1
1
1

0.004, relatively large quantities of transverse reinforcement are required). To over-

no special transverse reinforcement is required, whereas for

Ec,max

Ec,max

just greater than

come this-shortfall, the following equations are recommended to provide a flexible, yet
simple design approach to determine the required transverse reinforcement for concrete
confinement (it should be noted that minimum reinforcement, as required for strains
less than 0.004, must always be provided).

303

1
A~h

. .,

[cc 0.005
- 0.003]

A h

"

< A h

"

(A14)

Where A~h is the required. area of transverse reinforcement and A sh is the greater of
(Eq. 21-3 and 21-4 of the ACI 318-89 Building Code):
A.,h =

0.3

- (-A
)
(s hfyhJ;)
Ach
9

!'
== 0.09 shc-c
fyh

(A15a)

(A15b)

Where A 9 is the gross area of the portion of the wall cross-section requiring confinement, Ach is the area of the confined. core defineci. to the outside of the transverse
reinforcement, he is the cross-sectional dimension of the confined. core measured. centerto-center of the transverse reinforcement, and Jyh and s are the yield stress and the
spacing of the transverse reinforcement, respectively.
Equation (A14) is based. on observation that compressive strains in excess of 0.008
are more likely to occur in buildings with low ratios of wall area to floor plan area
(Wallace, 1995), which are representative of current design practice for shear wall
buildings in the United. States. For buildings with relatively few walls, damage to
one wall could result in substantial increases in deformation demands for other walls
(due to torsional response or redistribution). Therefore, for strains in excess of 0.008,
the use of current ACI code provisions, which are based on equating the pre- and
post-spalling loads, would appear to be a reasonable approach. A linear variation of
required transverse reinforcement between strains of 0.004 and 0.008 (with a 20% offset
at a strain of 0.004) is used to provide a simple means to vary the required transverse
reinforcement.
(ii) To Restrain Buckling of Longitudinal Reinforcement

0.004 <

Ec

< 0.006:

Smax

= 8db or 8 in. (20 cm).

304

1
1

1
1
1
1
1

1
1
1
1
1
1
1

1
1
1

This spacing requirement is based on an assumed tangent modulus at the onset of


strain hardening (far Grade 60 steel) of E ~ 1,500 ksi (10,000 MPa), and fa = f 11 (60
ksi, 410 MPa). Based on these assumptions, use of Eq. (A13) results in

1
1
1

1
1
1
1
1
1
1

smax

= Sdb, A

spacing limit of 8 in. (20 cm) is also suggested to discourage use of few, large diameter
hars at wall boundaries (to increase spacing of transverse reinforcement where buckling
controls).
0.006 :::;

le :::;

O.Ol:

Smax

= 6db or 6 in. (15 cm).

This spacing requirement is based on the observation that between compressive


strains of 0.006 and O.Ol, the tangent modulus remains essentially constant. For U.S.
Grade 60 steel, a representative tangent modulus far this range of strain (at the upper
limit of fc=0.01) is approximately 1,200 ksi (8,300 MPa) and the maximum steel stress
is approximately 80 ksi (550 MPa). Use of Eq. (A13) results in

Smax

6.2db,

spacing limit of 6db or 6 in. (15 cm) is recommended. it is noted that this spacing
limit is similar to recommendations used in the New Zealand code (NZS 3101, 1982)
of 6 in. (15 cm), 6db, or one-half the wall thickness.

O.Ol $ fc:

Smax

4db or

4 in. (10 cm).

Computed strains in excess of O.Ol should be discouraged; therefare, a maximum


spacing of 4 in. (10 cm) or 4db is suggested. This spacing limit is consistent with
current ACI building code requirements (ACI 318-89).

( c) General Requirements

1
1
1

.,

Where special transverse steel is required (c > 0.004), it should be extended over a
height of the wall equal to the greater oflw or hw/6, but need not exceed 21w as recommended
by Paulay ( 1986) based on expected plastic hinge lengths far shear walls. In addition,
spacing of crossties or legs of overlapping hoops should not be spaced more than 14 inches
(35 cm) in the direction perpendicular to the longitudinal axis of the structural member
as specified in Section 21.4.4.3 of ACI 318-89. To ensure adequate lateral restraint of

305

1
longitudinal hars against buckling, ties satisfying the requirements of ACI 318-89 Section
7. 10.5. 1 should be used [at least U.S. #3 (0.375 in., 10 mm) ties for longitudinal hars U.S.
#10 {1.27 in., 32 mm) and smaller, and at least U.S. #4 (0.5 in., 12 mm) ties for larger
hars].

(2) Requirements for Shear


Although the proposed format may allow for the variation of wall shear-strength with
the computed maximum compression strain or displacement ductility, until a comprehensive
review of this problem is undertaken, current design provisions for wall shear-strength are
used. Based on current design provisions (ACI 318-89, Eq. 21-6) the wall shear-strength is
computed as:

{Al6)
where ..\=0.083 if /~ is in MPa and ..\=l if /~ is in psi, Acv is the concrete area bounded by
the web thickness and the length of the section in the direction of shear force considered,
and Pn is the ratio of distributed shear reinforcement on a plane perpendicular to Acv (as
defined by ACI 318-89, pages 301-302).
An estimate of maximum shear force expected to develop in the wall is required to ensure adequate shear-strength is provided. The maximum wall shear force can be estimated
as {Paulay, 1991):
Vexpected

Wv

Mo
--Vcode
Mcode

{A 17)

where wv is a factor to account for the distribution of lateral force over the height of the
wall at maximum shear, M 0 is the flexural capacity accounting for overstrength factors,
M code

is the code required flexural-strength, and Vcode is the code specified shear force for

a given wall.
Code required flexural-strength is based on a distribution of lateral forces that increases
approximately linearly with wall height; however, due to higher mode effects, greater wall
shear is likely to develop. To account for these effects, values of wv equal to 4/3 and 5/3

306

1
1
1
1
1
1

1
1
1
1
1
1
1

1
1
1

1
1
1
1
1

are recommended far buildings with 10 ar fewer stories and buildings with more than 10
stories, respectively. In addition, because the provided flexural reinfarcement is likely ta
exceed that required, and ta account far potential overstrength and strain hardening effects
of the flexural reinfarcement, an overstrength ratio (M 0 / Mcode) should be computed. The
overstrength ratio can be computed fara given wall cross-section using a RC section analysis
program; however, a value of 2.0 is reasonable for symmetrically reinforced walls. A value

of 2.0 is based on multiplier of 1. 15 far excess reinfarcement, 1.25 far strain hardening of

1
1

stress computed using Eq. (Al 7) should be limited ta 6v7'c psi (0.5v7'c MPa) as suggested

tension reinforcement, and 1.4 ta account far the capacity reduction factor. The wall shear

by Aktan and Bertero (1985).

1
1

1
1

1
1
1

1
1

1
1

307

1
1
required envelope
linear envelope
code prescribed
envelope

1
1

[offset

///////////////

Mu,max

Flexural Strength Envelope

Wall Elevation

1
1
1

Fig. A.1 Requirements For Wall Flexural-Strength vs. Wall Height

il
Elastic Spectrum t=Ss
lnelcstic Spectrum J.Ld=S
Simplified Spectrum
Sd=6T

.::::,5. O
~

::e 4.0
w

/,"

o.. 3.0

,,,,,,/ , '
,,,,,,
,,,,.,,
,,,,.,,
,,,,

U)
c

2.0

/
/

-~ 1.0

U)

. o.o

OO

__ .,,- ,,

,,

, ,"

,'

,"

,,,"

,,'

-t---::=.-~--~---,--~--~---~-------------~--~

0.2

0.4

0.6

0.8

PERIOD (SEC)
Fig. A.2 Elastic and lnelastic Displacement Spectra

308

1.0

1
1
1
1
1
1
1

1
1

1
1
1

1
1

1
1
1
1
1
1
1

,........,
-

2.5
(.!)
I

------

1....

o:::

o 1.0

\
'\

''

''

'

''

....

.......

.......

.......

-.. -..

u..

~0.5

o.o-+----------------.--------------------------...-..--~
1.0
2.0
3.0
4.0
5.0
o.o
WALL AREA (iN ONE DIRECTION) %
TOTAL FLOOR PLAN AREA
Fig. A.3 Estimate of Roof Drift Ratio

r-w

Roof
Displ.

I .,

r
l
hw

1
1

1
1
1

10

7
5
3

1
1

2.0

~1.5

h./. h./. h./.


h./.

Load

Wall Elevation

Pu Py
Displacement

Curvature

Fig. A.4 Relationship Between Global and Loca! Deformations


309

1
p

p"

p's_ lw

1,..-:-:-.-.-.--.-:-:-:-.-.----.---:I
1w

_ _ _ _ _ _..,.

----- TENSION - - - - COMPRESSION

~ l-2c - - -

WALL

-1

~~
C ----- C

CROSS-SECTION

WALL STRAIN

-----

WALL EQUILIBRIUM

o.sst; r

tw
~ //ezuzznzzf zzzzzz~
~b

TENSION

WALL CROSS-SECTION

. f

lw

Fig. A. 5a Equilibrium Requirements for Rectangular Wall Cross-Section

-1

1
1

1"
s

1
1

1
1
1

COMP.

STRAIN DISTRIBUTION
1

-b.lw-2C 19-q~-~lb~c
tfttt ft

l
WALL EQUILIBRIUM

tcca tCc ~

Fig. A.5b Equilibrium Requirements for Barbell-Shaped Wall Cross-Section

310

1
1
1

1
1

1
1
1
1
1

1
1
1
1

ev
:i

o.o 1O - - . . . . . . - - - - , - - - . - - - - - - . - - - - - - - - - - h.,,/1.,,=5
\

~ 0.008

P/Awf'c=0.20
P/Awf'c=O. 1O
P/Awf'c=0.05

\
\
\

E-4

rl

r:.t:
~

0.006

r:c
0.004

''

~
:::

'

~
0.002
~

''

''

''

' ' ...


' ... ....

....... ........

--

1
1

.................

0.000 -+---......----,---------.-----------
0.00
o.o 1
0.02
0.03
0.04
0.05
lfALL AREA (iN ONE DIRECTION)

TOTAL FLOOR PLAN AREA


Fig. A.6a Extreme Fiber Strain for Constant Wall Aspect Ratio

P=0.10Awf'c

1
1

1
1
1

--- --- --- ----- - - -

------- hw/lw= 1 O
---- hw/lw=7
- - - hw/lw=5
hw/lw=3

0.000 ......-----.----------,----.....---.----.------t
0.00
0.01
0.02
0.03
0.04
0.05
lfALL AREA (IN ONE DIRECTION)

TOTAL FLOOR PLAN AREA


Fig. A.6b Extreme Fiber Strain for Constant Wall Axial Load

311

1
o.o 1O -.----~,---,------y-----------.--'4-)

z-

\\
\

~ 0.008

\
\

E-t

t:'ll

''

er:: 0.006
r::
~

.....
rz..

r::

:s
r::

P=O.lOA,,f'c
h,,/1,,=5
''
-------- p-p'=0.01
''
''
- - - p-p'=0.005
''
''
p-p'=O.O
' ',
... ......

',, ......

0.004

.... .......
.......

.... ..... ......


.._

........

-- ... ......... ... ......

- .._

er::

~ 0.002

---

--

0.000 -+---------...---..--------.------
0.00
0.01
O. 2
0.03
0.04
0.05
WALL AREA (iN ONE DIRECTION)

TOTAL FLOOR PLAN AREA


Fig. A. 7 Extreme Fiber Strain for Unsymmetrically Reinforced Walls

0.30 ...,..___ _ _ _ _ _ _ _ _ _ _ _ ____...__.........___

''

''

~ 0.20
C!,

''

''

''

h,,/1,,=5
P=0.20A,,f'c
P=O.lOA,,f'c
P=0.05A,,f'c
''

f;3

z.....

0.1 O

:z..

u O.00

''

' ' ...


'

''

' ' ...

' ' ...

......

... ...

... ...

...

' ... ',

' ... ...


''
' ... ...
' ... ...
''
... ...
... ...
... ...
'
... . .
......________.,___...........
' _______.__...._ _

0.000

~-----4

0.005

0.01 O

0.015

0.020

0.025

WALL AREA (iN ONE DIRECTION)

TOTAL FLOOR PLAN AREA


Fig. A.8 Required Length of Concrete Confinement for Constant Wall Aspect Ratio

312

1
1
1

1
1

1
1
1
1
1

1
1
1
1
1

1
1

1
1
1

..J

'

=
e-. 0.30

1
1

1
1

CJ

:l..t
~
~
~
~

z
~
z
o
t.)

0.40

0.20

' ' ...

' ...

P=O.lOA.,,f' 0 h,,/1,,=5
------- p-p'=0.01
- - - p-p'=0.005
p-p'=O.O

...............
......................
........

' '

......

...............

.........

......... ......

...... ......

...... ......

...... ......

...... .........

...... ......

o. oo -t----r-~r--T"-r--~--,--"'l"'-T---r--"'1'--,--...,...-,--.....----__,;:::,.---
......

0.000

0.005

0.01 O

0.015

0.020

0.025

WALL AREA (iN ONE DIRECTION)


TOTAL FLOOR PLAN AREA
Fig. A.9 Required. Length of Concrete Con.finement for Unsymmetrically Reinforced Walls

1
1
1

.......

0.1 O

1
1
1

313

1
Appendix B
OBSERVED BEHAVIOR

This appendix provides a detailed description of the experimentally o bserved behavior and damage of each individual test specimen. Sections B. 1 through B.4 describe the
observed behavior of specimens RWl, RW2, TWl and TW2, respectively. The numerous
photographs included in this appendix are intended ta provide a visual "history" of the
damage that was experienced by each specimen at progressively higher drift levels.
As mentioned in Chapter 6, each specimen was loaded axially with approximately

O. 10A 9

J: and then subjected ta reversed cyclic lateral loads applied at the top of the spec-

imen. All four specimens were subjected ta the same initial displacement cycles (2 cycles
each at lateral drift ratios of 0.1, 0.25, 0.5, O. 75, and 1.0%); however, displacement cycles
at drift ratios beyond 1.0% varied depending on the specimen's behavior up ta that point

1
1
1

1
1
1

1
1
1

(Figs. 5.32 through 5.35).


Each specimen was clearly marked with north, south, east and west directions ta facilitate the description of the test and the observed behavior. Displacements in the northward
direction are considered positive (Fig. 5. 13) and the lateral load necessary ta cause northward displacements are alsa taken as positive. Displacements in the south direction are
considered negative and the lateral load causing southward displacements are likewise negative. The fianges of the two T-shaped walls are located at the north end of their web;
therefore, positive (northward) displacements place the fiange in compression and negative
(southward) displacements place the fiange in tension. All four specimens had lateral support at the top of the wall ta prevent any out-of-plane movement in the east-west direction.

1
1
1

1
J

The lateral load versus top displacement relations far each of the four specimens are
plotted in Figs. 6.5 through 6.8. These plots are discussed in detail in Chapter 6; how-

314

1
1

1
1

ever, they will be referred to in this appendix to help describe the experimentally observed

behavior.

B.1

'1

Specimen RWl

Observed Behavior

Specimen RWl was tested under reversed cyclic lateral loading with an average axial
load of 90 kips (0.10A 9 /~). Figure 6.1 plots the axial load history of specimen RWl for the
duration of the test. The fluctuation shown at the beginning of the plot is the result of
temporarily relieving the axial load to readjust the post-tensioning cables. The readjustment

1
1

was completed prior to applying any lateral load. The testing began with two complete
cydes ata lateral drift level of O. 1%. These two cydes did not cause any noticeable damage;
however, they provided a good estimate of the initial wall stiffness and they provided an

opportunity to check that all instrumentation was working properly.

1
1
1

cracks formed at the lowest two construction joints and flexural cracks formed over the first

1
1

1
1
1
1

The first eyde at 0.25% lateral drift produced the first visible damage. Horizontal

story height. These flexural cracks extended into the wall 6 to 18 inches (152 to 457 mm)
and were evenly spaced at approximately 3 inches (76 mm). This spacing corresponds to
the vertical spacing of the transverse reinforcement provided in the boundary zones. All
cracks were marked with a dark marker so that they were easily distinguished. Each crack
was also labeled with a data point number that corresponded to the data point in the data
acquisition program to provide a "memory" of when the crack occurred. The second eyde
at this drift level continued to produce new flexural cracks. A photograph of specimen RWl
at the completion of the two cydes at 0.25% lateral drift is shown in Figure B. 1.
The two cycles at 0.5% lateral drift resulted in flexural cracks forming over the bottom
two stories (six feet). The flexural cracks still exhibited a uniform vertical distribution
of approximately 3 inches (76 mm). Flexural cracks that were originally horizontal were
now starting to extend in a diagonal fashion toward the center of the wall, indicating that
shear was playing a role in specimen behavior. These "shear-flexure" cracks were inclined

315

1
at approximately a 45 degree angle and extended over more than half of the wall length.
The presence of compression struts (outlined by the "diagonal" cracks) over the bottom two
stories was apparent. The influence of shear on specimen behavior is discussed in detail in
Chapter 7. Figure B.2 shows

aphotograph of RWl at the completion of two cycles at 0.5%

lateral drift.
Flexural and shear cracks continued to form on the bottom two stories and began forming on the third story during the first cycle to O. 75% lateral drift. Yielding of longitudinal
boundary reinforcement was also observed during this cycle. it can be seen in Fig. 6.5
that the wall experiences a substantial decrease in stiffness at O. 75% lateral drift due to the
yielding of the longitudinal boundary steel. This is the first drift level shown on the plot
where the second cycle does not follow the first cycle, indicating that the specimen is experiencing inelastic behavior. The second cycle at this drift level resulted in some extensions
of existing cracks but the formation of new cracks slowed substantially. Figure B.3 shows a
photograph of RWl at the completion of two cycles to O. 75% lateral drift.
The first cycle to a lateral drift level of 1.0% resulted in the beginning of vertical
splitting and minor crushing at the wall boundaries. Vertical splitting over the bottom 9
inches (229 mm) and some minor crushing and spalling was observed at the north boundary.
The south boundary experienced splitting over the bottom 6 inches (152 mm); however,
crushing and spalling had not yet occurred. Flexural and shear cracks were now spread
over the three lower stories and some cracks extended as far as 36 inches (914 mm) along
the length of the wall. The second cycle at this drift level caused some minor extensions
of existing cracks; however, once again the formation of new cracks slowed significantly. It
seems apparent that most new cracks are formed during the first cycle at a new drift level;
whereas, the repeat cycle(s) result in extensions and widening of existing cracks. Figure
B.4 shows an overall photograph of specimen RWl after two cycles at 1.0% lateral drift and

1
1

1
1

1
1

1
1

1
1
1
1

Fig. B.5 shows a close-up photograph of the north boundary zone that experienced some

minor crushing and spalling.

316

1
1

1
1
1

1
1
1
1
1

1
1
1
1

Cycles to 1.5% lateral drift resulted in very few new cracks; however, existing cracks
continued to grow in length and width. Maximum crack widths near the hase of the wall
were measured at 3/32 inch (2.4 mm). The first cycle at this drift level caused major vertical
splitting over the hottom 12 inches (305 mm) at hoth houndaries. One of the vertical cracks
passed through the epoxy holding on one of the LVDT's, knocking it loose. Crushing and
spalling of the south houndary element hegan and the north houndary element was now
spalling its cover over the hottom 6 inches (152 mm). Figure B.6 shows a photo of the north
houndary element after two cycles to 1.5% lateral drift.
The first cycle at 2.0% lateral drift caused extensive crushing and spalling at hoth
houndaries.

The two outermost longitudinal houndary hars in each houndary element

huckled during this cycle. The hars in the north houndary element huckled hetween the
hoops located at 6 and 9 inches (152 and 229 mm) ahove the pedestal. Strain gages on
the hoops indicated yielding of the transverse steel at this drift level. The hars in the
south houndary element huckled hetween the hoops at 3 and 6 inches (76 and 152 mm).
Buckling of the longitudinal steel seems to have occurred immediately following spalling of
the concrete cover (huckling is discussed in detail in Chapter 7). The cracks located near
the hottom of the wall are no longer growing in length; however, a majority of the cracks
were measured hetween 1/16 and 3/32 inch (1.6 and 2.4 mm) in width. The second cycle at
this level continued to cause more extensive spalling of the concrete cover. Both houndary

elements have experienced spalling over the lower 10 inches (254 mm) and approximately 4

longitudinal steel in each houndary has experienced huckling hetween two different spacings

1
1
1
1
1

inches ( 102 mm) along the length of the wall. The extensive spalling has revealed that the

of the transverse steel. Figures B. 7 and B.8 show a closeup of the huckling in the south
houndary element andan overall photo of RWl upon completion of the two cycles at 2.0%
lateral drift, respectively.
Specimen RWl failed during the first eyde at an attempted drift level of 3.0% (the failure actually occurred at a top displacement corresponding to a drift level of approximately

317

1
2.5%). The north houndary element had spalled its entire cover allowing huckling of all
eight longitudinal houndary hars and two pairs of the vertical weh hars. The horizontal
weh steel lost its anchorage in the houndary element and the entire specimen lost its load
carrying capacity (Figure 6.5). The test was terminated after one half of the first eyde at
this drift level. Figures B.9 and B. 10 show an overall anda close-up photograph of specimen

RW 1 at the completion of the test, respectively. The extensive damage is clearly visihle in
these photographs.
The lateral load-top displacement curve shown in Fig. 6.5, and the hase moment-hase
rotation curve shown in Fig. 6. 13 reveal that specimen RWl exhihited stahle hysteretic
hehavior and that no noticeahle drop in load capacity was experienced up to a lateral drift
level of 2.0%. The specimen was designed to withstand a lateral drift level of 1.5% and the

'I
1

1
1

first drop in load capacity did not occur until approximately 2.5% lateral drift.

1
B.2

Specimen RW2 - Observed Behavior


The major difference hetween specimen RWl and RW2 was the spacing and configu-

ration of the transverse houndary reinforcement. Specimen RWl had one hoop and two
cross-ties at a vertical spacing of 3 inches (76 mm); whereas, RW2 had only a single hoop
(no cross-ties) at a spacing of 2 inches (51 mm). These two configurations yielded similar
values of

Ash

(defineci and discussed in detail in Section 4.2); however, the primary pur-

pose for the change in configuration was to attempt to delay huckling of the longitudinal
houndary reinforcement (spacing was decreased from 8db in RWl to 5.33db in RW2; where
db is the diameter ofa longitudinal boundary bar). in general, the behavior of specimen
t-X

RW2 was very similar to specimen RW 1 <)Ccept that the tighter spacing of the transverse
boundary reinforcement did delay the huckling of the longitudinal steel.
Specimen RW2 was tested with an average axial load of 85 kips (0.07 A 9 J~). Figure 6.2
plots the axial load history of specimen RW2 for the duration of the test. During the test
it was noticed that the axial load had decreased slightly; therefore, the load was increased

318

'I
,1
1
1
1

1
1
1
1

1
1
1
1
1
1
1

1
1
1

back to its intended level as shown in the plot. The oscillation of axial load was due to
pressure losses within the axial jacks caused by the lateral cycling of the specimen. lnitial
la ter al drift cycles were the same as those used for testing specimen RW 1. Two cycles at
O. 1% lateral drift provided an accurate assessment of the initial, uncracked, wall stiffness.
The first fl.exural cracks appeared during the first cycle at a lateral drift level of 0.25%.
Flexural cracks formed over the bottom 24 inches (610 mm) of the wall and the longest
cracks extended approximately 20 inches (508 mm) along the wall length. These fl.exural
cracks exhibited a uniform vertical crack distribution of approximately 4 inches (102 mm).
The lateral load-top displacement curve shown in Fig. 6.6 reveals a change in stiffness
between drift cycles of O. 1% and 0.25%, resulting from the cracking of the concrete. A few
new fl.exural cracks formed, and existing cracks extended, during the second cycle at this
drift level. Figure B.11 shows a photograph of RW2 after two cycles to 0.25% drift.
The first cycle at a lateral drift level of 0.5% resulted in fl.exural cracks forming over
the bottom two stories and "shear-fl.exure" cracks in the first story extending over 26 inches
(660 mm) along the length of the wall. The cracks exhibited a uniform spacing distribution

of approximately two inches (51 mm), corresponding to the spacing of the hoops in the

1
1
l
1
1

the outermost hars at both boundary elements were approaching yield when in tension and

1
1
1

boundary zone. Strain gages mounted on the longitudinal boundary steel indicated that

compression. The second cycle at this drift level caused some extensions of existing cracks
but not many new cracks. Figure B.12 shows specimen RW2 after completion of two cycles
at 0.5% drift.
Flexural and shear cracks were evident over the bottom three stories after the first
cycle to a lateral drift of O. 75%. The diagonal "shear" cracks extended over 30 inches (762
mm) along the wall length. The.development of compression struts was evident as outlined
by the large shear cracks. The longitudinal boundary steel was yielding in both tension
and compression. Concrete strain gages located within the confined core indicated strains
of approximately 0.002 at peak displacements. Strain gages mounted on the transverse

319

1
boundary reinforeement indieated that the hoops loeated near the bottom of the wall were
experieneing strains of approximately 25% of their yield strains. This seems reasonable eon-

1
1

sidering the hoops do not start to become effective until the eonerete eover begins erushing
and spalling and the eonfined eore attempts to expand outward. The second eyde at this
drift level resulted in very little new damage. Figure B. 13 shows the general eraek pattern
of specimen RW2 at the end of two eydes to O. 75% lateral drift.
Vertieal splitting at the wall boundaries initiated at the first eyde to 1.0% lateral drift.
The splitting was eoneentrated over the lowest three inehes (76 mm) at eaeh boundary.
Some of the strain gages loeated on the longitudinal boundary steel had broken at this

1
1

1
1

level (presumably beeause the vertieal splitting knoeked the lead wires loose). The eonerete
strain gages in the boundary zones indieated strains of approximately 0.003 in tension
and eompression; however, strains ealculated from the LVDT readings indieated strains of
approximately O.Ol in tension and 0.004 in eompression. The measured strain values vary
due to the LVDT's being mounted on the eover eonerete (more damage expected); whereas,
the eonerete strain gages are mounted in the well eonfined eore (less damage expected).
The appearanee of minor eracks in the pedestal indieated that slippage of the longitudinal
boundary steel may eontribute slightly to overall wall behavior. The formation of new eraeks
in the wall has slowed eonsiderably; however, existing eraeks eontinue to extend and widen.
The eonerete strain gage in the north boundary element and many longitudinal steel strain
gages broke during the second eyde to 1.0% drift. Strain gages mounted on transverse
reinforeement indieated strains of approximately 50% of their yield strains. Figure B. 14
shows a photograph of the south boundary element after two eydes to 1.0% lateral drift.
Vertieal splitting of the two boundary zones was quite extensive and some minor erushing and spalling of eonerete eover .was observed during the first eyde at a lateral drift level
of 1.5%. The eonerete strain gage in the south boundary element broke and the transverse
reinforeement had reaehed approximately 75% of its yield strain. Strain gages mounted
on longitudinal boundary steel indieated tensile strains of approximately O. 02 during peak

320

1
1
1
1

1
1
(\

1
1
1
1

1
1
1

1
1
1

1
1
1
1

1
1

displacements. The second eycle resulted in more erushing and minor spalling of the boundary element eover. Figure B.15 shows a photograph of the south boundary element after
eompletion of two eydes to 1.5% lateral drift.
Upon eompletion of two eydes at 1.5% lateral drift, specimen RW2 was subjected to
two additional eydes at 1.0% lateral drift. There was no new damage observed during these
two eycles. Figure 6.6 shows that these two eydes eoincided with eaeh other, as expeeted
when no additional damage is observed. The test then eontinued with two additional eydes
at 1.5% lateral drift. The third and fourth eycles at 1.5% drift eaused minor extensions
and widening of existing eraeks. The eonerete eover at each boundary element was eraeked
extensively and spalling was observed at both boundaries over the bottom three and six
inehes (76 and 152 mm) for the south and north boundary elements, respeetively. Figures
B. 16 and B. 17 show photographs of the north and south boundary elements, respectively.
Lateral drift eydes at 2.0% resulted in more splitting and spalling of the eonerete
eover. The eonfined eore still appeared to be in exeellent shape; eventhough, the eover
was experieneing severe distress. Maximum eraek widths near the base of the wall were
measured at 3/32 ineh (2.4 mm). The LVDT mounted on the south boundary element fell
off due to splitting of the eonerete eover. Figures B.18 and B.19 show photographs of the
north and south boundary elements after two eydes at 2.0% lateral drift, respectively.
The first eyde to 2.5% lateral drift eaused major spalling at both boundaries. The
south boundary element had spalled 9 inehes (229 mm) over the bottom of the wall. The

outermost longitudinal reinforeing hars at eaeh boundary started to buekle. At the north

inehes above the pedestal) and at the south boundary they were buekling between two

boundary, the hars were buekling between the second and third hoops (between 4 and 6

spacings -(between 2 and 4 inehes, and between 4 and 6 inehes). The LVDT at the north
boundary element fell off due to spalling of the eonerete shell. The seeond eyde at 2.5%
drift eaused buekling of six longitudinal hars in the north boundary (Figure B.20). Spalling

1
1
1

oeeurred over the bottom 10 inehes (254 mm) and 7 inehes (178 mm) along the length

321

of the wall.

Concrete within the confined core was crushing and a slight drop in load

capacity was noticed on the lateral load vs. top displacement plot (Fig. 6.6 ). The south

1
1
1

boundary element was not damaged as extensively as the north; however, buckling of most
of the longitudinal steel was assumed due to the swelling of the boundary element. The
buckled longitudinal steel at the north boundary straightened during peak displacement to
the south; however, as the wall returned to zero displacement, the hars "rebuckled". The
test was stopped after two cycles at 2.5% lateral drift. If additional cycles were conducted
the north boundary element would not have been able to provide the necessary compressive
resistance and a failure similar to specimen RWI's would have resulted.

The observed

behavior indicated that the closer spacing of the transverse hoops delayed buckling of the
longitudinal reinforcement and crushing of the concrete within the confined core became a
governing factor in the overall behavior of specimen RW2. Figure B.21 shows a photograph
of the damage at the base of the wall after completion of 2 cycles at 2.5% lateral drift. Figure
B.22 shows an overall photograph of specimen RW2 after the test had been completed.

B.3

Specimen TWl - Observed Behavior

1
1

1
1
1
1

The most obvious difference between specimen TWI and RWI is the cross-sectional
shape of the wall (the existence of a flange resulting in a T-shaped specimen).

Many

differences exist that were discussed in detail in Chapter 4; however, this section will only
describe the experimentally observed behavior; whereas, comparisons with other specimens
are discussed in Chapter 6. Specimen TWI consists of two of specimen RWI joined together
to form a "T" shape. The longitudinal, web and transverse boundary reinforcement are
all identical to that used in specimen RWL As discussed previously, in Chapter 4, poor
behavior was expected from this specimen.
Specimen TWI was subjected to an average axial load of 158 kips (0.09A 9 /~). Figure
6.3 plots the axial load history of specimen TWI for the duration of the test. As can be seen

1
1
1

1
1

in the plot, the axial load was kept essentially constant throughout the duration of the test

322

1
J

1
1
1
1

1
1
1
1

1
1
1

1
1
1

1
1
1
1

(very little oscillation due to lateral movement). The lateral loading history of this specimen
was very similar to that of RWl. Two complete cycles at a lateral drift of 0.1% gave an
initial stiffness estimate. The first flexural crack occurred at the wall-pedestal interface at
the south boundary element during the second eyde to 0.25% lateral drift. The lateral load
versus top displacement plot (Fig. 6. 7) shows that the stiffness in the two directions are
not the same (as compared to the rectangular walls which exhibited similar stiffnesses in
the two directions).
Significant flexure and shear cracks began to form during the first eyde to O. 5% lateral
drift. When the flange was in compression (northward displacements), the south (web)
boundary formed flexure and shear cracks over the bottom two stories. The flexural cracks
exhibited a vertical spacing pattern of approximately 3 inches (76 mm) corresponding to
the spacing of the transverse reinforcement in the boundary element. Some diagonal shear
cracks extended 36 inches (914 mm) along the wall web. When the flange was in tension
(southward displacements), flexural cracks extending the entire flange length were observed
over the lowest story. These cracks also extended approximately 15 inches (381 mm) along
the length of the web. The second cycle at O. 5% caused the formation of a few new cracks
but mostly caused extension of existing cracks. Figures B.23 and B.24 show photographs of
the web and flange of specimen TWl upon completion of two cycles at a lateral drift level
of 0.5%
The first cycle at a drift level of O. 75% caused cracking on the third story level. Under
positive displacement the longitudinal steel in the web boundary yielded in tension and
the longitudinal flange steel was at approximately 25% of its yield strain in compression.
Cracks alsa began forming in the bottom two floor slabs. Negative displacements caused
extensive shear cracks to form in the web over all four stories. Flexural cracking in the flange
was still confined to the lowest two stories. The second cycle at this drift level resulted in
some extensions of existing cracks and the onset of vertical splitting of the web boundary
element when subjected to negative displacement. The wire potentiometer measuring lateral

323

1
displacement of the first floor was knoeked loose during this eyde; however, it was remounted
before starting the next eycle. Figures B.25 and B.26 show photographs of the web and
flange of TWl after eompletion of two eydes to O. 75% lateral drift.
Under positive displacement of the first eyde at 1.0% lateral drift, the "shear" eraeks
extended all the way to the flange. Negative displaeements eaused the formation of many
new "shear" eraeks over the bottom two stories and eaused extensive vertieal splitting
and minor erushing of the web boundary element over the bottom 12 inehes (305 mm).
The LVDT loeated nearest the web boundary element reaehed its maximum stroke and its
readings beeame questionable. The linear potentiometers mounted to the pedestal indicated
that the pedestal was lifting off of the strong floor. The influenee of pedestal rotation on
top displaeement is diseussed in detail in Chapter 7. The seeond eyde at 1.0% lateral drift
eaused more flexural and shear eracking in the upper two stories. Figures B.27 through
J;3.29 show photographs of the web, flange and web boundary element of specimen TWl
after two complete cycles at 1.0% lateral drift, respectively.
Failure of specimen TWl oecurred during the first eyde at 1.5% lateral drift. Under positive displacement, some new cracks developed but the web boundary element was

1
1
1
1
1

1
1
1

experiencing extremely high tensile strains and maximum crack wid ths were measured at
approximately 1/8 inch (3.2 mm). Under negative displacements (approximately 1.4% lateral drift), the web boundary element and about half of the length of the wall experienced a
brittle buckling failure. Concrete cover exploded from the wall and buckling of the longitudinal boundary hars and some vertical web hars was observed. A second eyde at this drift
level was not conducted. Figures B.30 and B.31 show photographs of the web boundary
element and the compression failure damage after the first eyde at 1.5% lateral drift. It can
be seen in these figures that extensive damage extended approximately 36 in. (914 mm)
along the length of the wall.
Specimen TWl was then subjected to positive displacements at a lateral drift of 2.0%.
This was done in an attempt to straighten the hars which had buckled in the previous eyde

324

1
1
1

1
1
1
1
1

1
1
1
1

1
1
1
1

and to see if the hars would fracture. Figure B.32 shows a photograph of the straightened
longitudinal steel after completion of halfa eyde at 2.0% lateral drift. The test was stopped
after this half eyde to prevent longitudinal steel from rebuckling.
The poor behavior of specimen TW 1 was expected (detailed reasons were discussed in
Chapter 4); however, it served as an excellent example of the potentially brittle behavior
that can result from poor conceptual design.

B.4

Specimen TW2 - Observed Behavior


Specimen TW2 was designed using the displacement-based design procedure outlined

in Appendix A in an attempt to eliminate the inadequacies witnessed in the behavior of

specimen TWL When the flange of the T-wall is in compression, very small compressive

1
1

condition occurs when the flange of the T-shaped wall is in tension and the relatively thin

1
1
1

1
1
1
1
1

strains are expected and damage is not likely to occur under this condition. The critical

web must develop large compressive forces to balance the high tensile forces that develop in
the flange longitudinal reinforcement. The flexural strength of a T-shaped wall is greater
than that ofa rectangular wall due to the addition of longitudinal reinforcement in the
flange.

This increase in flexural strength results in an increase in shear demand which

is typically resisted by distributed web steel.

The major difference between TW2 and

TW 1 were the spacing and confi.guration of the transverse boundary reinforcement and the
distribution of web steel, as described in detail in Section 4.4.
Specimen TW2 was subjected to an average axial load of 164 kips (0.075A 9 J;). Figure
6.4 plots the axial load history of specimen TW2 for the duration of the test. Approximately half way through the test (data point number 364) one of the post-tensioning cables
experienced an anchorage failure causing the axial load to drop dramatically. The eyde that
was currently being conducted was completed with the reduced axial load level; however,
the total axial load was reestablished before additional cydes were conducted.

325

1
1
Two cycles at 0.1 % lateral drift yielded elastic response and good estimates of uncracked
stiffness. The first cracks were observed during the first cycle at a lateral drift level of 0.25%.
Under positive displacement, fl.exural and shear cracks formed over the entire first story.
The largest shear cracks extended approximately 36 inches (914 mm) along the web. The
lateral load vs. top displacement plot (Fig. 6.8) revealed a slight change in stiffness due to
the cracking. Under negative displacement, only a single crack formed at the wall-pedestal
interface; therefore, Fig. 6.8 does not reveal any change in stiffness in this direction. The
second cycle at 0.25% drift caused extensions of existing cracks but few new cracks were
formed. Figure B.33 shows a photograph of specimen TW2 after completion of two cycles
at 0.25% drift.
During positive displacements of the first cycle at 0.5% lateral drift, flexural and shear
cracks formed over the bottom two stories. Strain gages on the longitudinal steel in the
web boundary element indicated that strains were at approximately yield. During negative
displacements cracks formed over the lowest one and a half stories. The second cycle at
0.5% drift did not result in any new damage. Figure B.34 shows specimen TW2 after two
cycles at 0.5% lateral drift.
Lateral drift cycles of O. 75% caused cracking to initiate in the third and fourth stories.
Under negative displacements of the first cycle, diagonal shear cracks propogated 18 inches
(457 mm) along the wall web. Under negative displacements of the second cycle, the cracks
formed on the fourth story. Strain gages mounted on the longitudinal reinforcement in the
web boundary element indicated that the hars were yielding in tension. Figure 6.8 reveals
a change in stiffness at this drift level. Figures B.35 and B.36 show photographs of the web
and the fl.ange of specimen TW2 after two cycles to O. 75% lateral drift.
Positive displacements during the first cycle at 1.0% lateral drift, caused existing cracks
to grow in width to approximately 1/16 inch (1.6 mm). Negative displacements resulted
in vertical splitting in the web boundary element. The longitudinal reinforcement in the
web-fl.ange intersection was at approximately 75% of its tensile yield strain; whereas, the

326

1
1

1
1
1
1

1
1
1

1
1

1
1
1
1

1
1

1
1
1
1

1
1
1
1
1
1

1
1

longitudinal steel in the web boundary element was experiencing compressive strains of
approximately 400% of yield. Strain gages mounted on the transverse reinforcement in the
web boundary element were indicating strains of approximately 25% of yield. Minor cracks
formed in the bottom two floor slabs. The second cycle caused some minor crushing of
the web boundary element and flange tension steel was approaching yield under negative
displacements. Figures B.37 and B.38 show an overall photograph of specimen TW2 and a
close-up of the web boundary element after two cycles to 1.0% lateral drift, respectively.
The first cycle of positive displacements at 1.5% lateral drift caused the cracks in the
bottom story to widen to approximately 3/32 inch (2.4 mm). The longitudinal reinforcement
in the web boundary was experiencing tensile strains fi.ve times larger than their yield
strains, whereas, the outermost layer of flange steel was at approximately 75% of their
compressive yield strains (neutral axis depth was approximately lined up with the inside
layer of longitudinal flange reinforcement thus strains in this steel were extremely small).
The LVDT's mounted along the flange indicated nearly uniform concrete compressive strains
of approximately 0.002. One of the concrete strain gages embedded in the web boundary
element broke. Vertical splitting of the web boundary expanded over the bottom 20 inches
(508 mm) during the first cycle of negative displacements. The concrete cover crushed and
spalled over the bottom 9 inches (229 mm). At the peak negative displacement, one of the
axial load post-tensioning cables experienced an anchorage failure. The cycle was completed
(with approximately 50% of the axial load, Fig. 6.4); however, the axial load apparatus was
repaired before the second cycle was performed. During the second cycle at 1.5% lateral

1
1

1
1
1
1

drift, an additional 3 inches (76 mm) of concrete cover spalled at the web boundary element
and knocked off the LVDT at the boundary. Figures B.39 and B.40 show photographs of the
web boundary element after the first and second cycles at 1.5% lateral drift, respectively.
Upon completion of two cycles at 1.5% lateral drift, specimen TW2 was subjected
to two additional cycles at 1.0% lateral drift. There was very little additional damage
observed during these two cycles (some minor extensions of existing cracks). The test then

327

1
continued with two more cycles at 1.5% lateral drift. The third and fourth cycles at 1.5%
drift caused more extensive crushing and spalling of the concrete cover at the web boundary
element. The cover had completely spalled over the bottom 12 inches (305 mm) at the end
of the boundary element and approximately 9 inches (229 mm) along the length of the
wall; however, the confined core appeared to be in excellent condition. The LVDT located
6 inches (152 mm) from the end of the wall was giving questionable readings due to the
extensive damage to the concrete where it was mounted. During negative displacements, the
pedestal was lifting approximately 0.2 inches (5 mm) from the fl.oor indicating substantial
pedestal rotation. Figures B.41 and B.42 show photographs of the web boundary element
after the third and fourth cycles at 1.5% lateral drift, respectively.

The first cycle of positive displacements at 2.0% lateral drift caused many new cr.acks to
form, mainly over the first story level. A large crack (1/16 inch in width) at the wall pedestal
interface, and minor cracking of the pedestal indicate that slippage of the longitudinal
reinforcement from the pedestal may be playing a role in specimen behavior. Under negative
displacements, all of the fl.ange steel is yielding in tension and splitting of the web boundary
element has extended over the bottom 30 inches (762 mm). The wire potentiometer mounted
30 inches (762 mm) above the pedestal to measure base rotations was questionable because
of the extensive vertical splitting. One of the wire pots used to measure shear distortions
fell off because of the damage. The second cycle caused some additional spalling of the
cover at the web boundary. The cover over the last 12 inches along the length of the
web (305 mm) had totally delaminated from the core; however, the confined core remained
in good condition. An additional (third) cycle was performed at this drift level before
proceeding to 2.5% drift. The pedestal continued to form new cracks as did the upper two
stories. Longitudinal steel in the outermost layer of the fl.ange yielded in compression under
positive displacements, and experienced tensile strains of approximately 150% of their yield
strains under negative displacements. The bottom 24 inches (610 mm) of concrete spalled at
the web boundary element. Figures B.43 and B.44 show photographs of the web boundary

328

1
1
1

1
1

1
1

1
1
1
1
1

1
1
1

1
1
1
1
1
1
1

element after the first and third eydes at 2.0% lateral drift, respectively.
The first eyde of positive displaeements at 2.5% lateral drift eaused new eraeks to form
on all four stories. The largest eracks on the first story had widths of approximately 1/8
inch (3.2 mm). Cracks in the pedestal eontinued to form. Under negative displacements
some new eracks formed in the flange on the bot tom three stories. The second cycle at
2.5% drift provided the first visible evidenee of distress within the eonfined eore. Under
negative displacements, cracking and minor erushing in the eon.fined core was observed.
Figure 6.8 reveals a slight drop in load eapacity (in the negative direction only) during this
eyde. A third eyde performed at this drift level resulted in an even larger decrease in the
wall's lateral load capacity. Concrete within the confined core was crushing and pushing out

1
1
1
1

1
1
1
1

through the hoops and ties. Figure B.45 shows a photograph of the web boundary element
after the third eyele at 2. 5% la ter al drift.
The test was stopped midway through the first eycle at a lateral drift level of 3.0%.
Positive displaeements only reaehed 2. 7% because the actuator had reaehed its maximum
stroke. U nder negative displaeements, the entire web boundary element began to experienee
an out-of-plane stability failure at approximately O. 75% lateral drift; therefore, the test was
stopped.

Any further displacements would have eaused a total destruction of the web

boundary element. Figures B.46 and B.4 7 show photographs of the web boundary element
after eompletion of the entire test. It can be seen in these photos that the longitudinal steel
in the web boundary were not buckling loeally between two adjacent hoops; however, they
were buekling globally over a large number of spacings. This out-of-plane buckling resulted
from the high compressive strains on the relatively thin confined wall cross-section (the
wall thickness is only 3.06 inches from eenterline to eenterline of hoops onee the cover has
spalled).

1
1
1

329

1
1
1
1
1

Fig. B.1 RWl: After Two Cycles at 0.25% Drift

1
1
1
1
1
1

1
1
1
1

1
1
Fig. B.2 RWl: After Two Cycles at 0.50% Drift
330

1
1

-------------------

w
.....

Fig. B.4 RWl: After Two Cycles at 1.0% Drift

Fig. B.3 RWl: After Two Cycles at O. 75% Drift

~
~
t...:)

..

,;1:,:;~-:'!:;{)~ : ~
Fig. B.6 RWl: North B.E. After Two Cycles at 1.5% Drift

.. -

Fig. B.5 RWl: North B.E. After Two Cycles at 1.0% Drift

------------------'

---~'

'"i"'' .,

~- \.". -"'.. \'-:.; . ..,. .

,_.,,,-

..
.................4---~. t ..
',

,.

-,.. . ._ - ~- .1\. /
\ 'J

~,

/,:.'

~-~f

.'

~l

'
f

~
~
~

Fig. B.8 RWl: After Two Cycles at 2.0% Drift

Fig. B. 7 RWl: South B.E. After Two Cycles at 2.0% Drift

1
1
1
1
1
1

1
1
1
Fig. B.9 RWl: After Half Cycle at 3.0% Drift (End of Test)
/

'~) ,,./
.
~

:/
,/

.-

__

.
-

1/-J',f .

:,/,-t.

:~../
-

-- .

1
1
1
1
1

Fig. B.10 RWl: North B.E. After Half Cycle at 3.0% Drift (End of Test)
334

1
1
1

E
1
1
1

1
.....
1
1
1

Fig. B.11 RW2: After Two Cycles at 0.25% Drift

1
1
1
1
1

1
1

Fig. B.12 RW2: After Two Cycles at 0.50% Drift


335

1
1
1

1
1
1
1

Fig. B.13 RW2: After Two Cycles at 0.75% Drift

1
1
1
1
1
1
1

Fig. B.14 RW2: South B.E. After Two Cycles at 1.0% Drift
336

1
1
1

1
E
1
1
1
1
1
1

Fig. B.15 RW2: South B.E. After Two Cycles at 1.5% Drift

1
1
1
1
1
1
1
1
1
1

Fig. B.16 RW2: North B.E. After Four Cycles at 1.5% Drift
337

~
~

00

,......,

L ..SI.... ..

.,.%.Nt

Fig. B.18 RW2: North B.E. After Two Cycles at 2.0% Drift

Fig. B.17 RW2: South B.E. After Four Cycles at 1.5% Drift

1
1
1
1
1
1
1
l"

1
1
1

Fig. B.19 RW2: South B.E. After Two Cycles at 2.0% Drift

1
1
1
1
1
1
1
1

Fig. B.20 RW2: Buckling of Long. Bars in North B.E. During Second Cycle at 2.5% Drift
339

1
1

1
1
1
1
1
Fig. B.21 RW2: Damage at Base of Wall After Two Cycles at 2.5% Drift

.......

1
1

1
1
1
1
1
Fig. B.22 RW2; Overall Photograph Upon Completion of Testing
340

1
1

!
1
1
1
1
1
[

t
1
1
1
1
1
1
1
1
1
1
1

Fig. B.23 TWl: Web After Two Cycles at 0.50% Drift

Fig. B.24 TWl: Flange After Two Cycles at 0.50% Drift


341

~
,::..

t,:)

Fig. B.26 TWl: Flange After Two Cycles at O. 75% Drift

Fig. B.25 TWl: Weh After Two Cycles at O. 75% Drift

- - - - - - - - - - - - - - - - - ... '
,,kil

c:,...,
.::..

c:,...,

Fig. B.28 TWl: Flange After Two Cycles at 1.0% Drift

Fig. B.27 TWI: Web After Two Cycles at 1.0% Drift

~
~
~:t''\

1 . \
.,,i ' .';

/\

, .~ . \
SPEC. #.z
/.S 7. .J...,~,...)elt..-..:;:;..,

~-.,

'

'l..,.,;./.;},,,
..........

.,

/~,,

"

-~

c:,..:,

:t

~7 -~.

"'

~.

:~;:.-

,.

.., .

:.f; .. ~

Fig. B.30 TWl: Web B.E. During First Cycle at 1.5% Drift

Fig. B.29 TWl: Web B.E. After Two Cycles at 1.0% Drift

_j

E
1

1
1
1
1
1
1

Fig. B.31 TWl: Web B.E. After First Cycle at 1.5% Drift

1
1

1
1
1
1

1
1
1
1

Fig. B.32 TWl: Web B.E. After Half Cycle at 2.0% Drift

345

1
1
1

1
1
1
Fig. B.33 TW2: After Two Cycles at 0.25% Drift

1
1

Fig. B.34 TW2: After Two Cycles at 0.50% Drift


346

1
1
1
1
1
1
1
1

------------ - - - -

,:.

-..

Fig. B.36 TW2: Flange After Two Cycles at O. 75% Drift

..

-,,

Fig. B.35 TW2: Web After Two Cycles at O. 75% Drift

iMi

<:...:>

00

'i ' ~

.;t~\V::.

,,: . ~~~~)'_f!;"~.~1

.,:t. :.1"

Fig. B.38 TW2: Web B.E. After Two Cycles at 1.0% Drift

Fig. B.37 TW2: After Two Cycles at 1.0% Drift

--- - - - -- - - - - - - - - - - - "

~
::,..

(O

Fig. B.40 TW2: Web B.E. After Second Cycle at 1.5% Drift

Fig. B.39 TW2: Web B.E. After First Cycle at 1.5% Drift

.....J

1
1
1
1

Fig. B.41 TW2: Web B.E. After Third Cycle at 1.5% Drift

1
1
1
1
1
1

1
1
Fig. B.42 TW2: Web B.E. After Fourth Cycle at 1.5% Drift
350

1
1

___________________,

C,71

.....

Fig. B.44 TW2: Web B.E. After Third Cycle at 2.0% Drift

Fig. B.43 TW2: Weh B.E. After First Cycle at 2.0% Drift

C,11

t,.j

Fig. B.46 TW2: Web B.E. During First Cycle at 3.0% Drift

Fig. B.45 TW2: Web B.E. After Third Cycle at 2.5% Drift

y
1
1

1
1
1
1
1
1
1
1
1
1
1

Fig. B.47 TW2: Out-of-Plane Failure of Web B.E. During First Cycle at 3.0% Drift

1
1
1

1
.d

353

Vous aimerez peut-être aussi