Nothing Special   »   [go: up one dir, main page]

Lateral Torsional Buckling of Rectangular Reinforc

Download as pdf or txt
Download as pdf or txt
You are on page 1of 335

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/44000554

Lateral Torsional Buckling of Rectangular Reinforced Concrete Beams

Article in ACI Structural Journal · January 2014


DOI: 10.14359/51686431 · Source: OAI

CITATIONS READS
16 7,547

1 author:

Ilker Kalkan
Kirikkale University
71 PUBLICATIONS 819 CITATIONS

SEE PROFILE

All content following this page was uploaded by Ilker Kalkan on 14 April 2015.

The user has requested enhancement of the downloaded file.


LATERAL TORSIONAL BUCKLING OF RECTANGULAR
REINFORCED CONCRETE BEAMS

A Dissertation
Presented to
The Academic Faculty

by

Ilker Kalkan

In Partial Fulfillment
Of the Requirements for the Degree
Doctor of Philosophy in the
School of Civil and Environmental Engineering

Georgia Institute of Technology


December 2009

Copyright © 2009 by Ilker Kalkan


LATERAL TORSIONAL BUCKLING OF RECTANGULAR
REINFORCED CONCRETE BEAMS

Approved by:

Dr. Abdul-Hamid Zureick, Advisor Dr. Bruce R. Ellingwood


School of Civil and Environmental School of Civil and Environmental
Engineering Engineering
Georgia Institute of Technology Georgia Institute of Technology

Dr. Lawrence F. Kahn Dr. George Kardomateas


School of Civil and Environmental School of Aerospace Engineering
Engineering Georgia Institute of Technology
Georgia Institute of Technology

Dr. Kenneth M. Will


School of Civil and Environmental
Engineering
Georgia Institute of Technology

Date Approved: September 28th,2009


To my family…
ACKNOWLEDGEMENTS

The author is grateful to his advisor Dr. Abdul-Hamid Zureick for his help, guidance,

patience and encouragement throughout this project.

The author would like to express his sincere gratitude to Dr. Lawrence Kahn,

whose technical expertise and guidance contributed greatly to the success of the

experimental program. The author is also grateful to the other thesis committee members

Dr. Kenneth M. Will, Dr. Bruce R. Ellingwood and Dr. George Kardomateas for their

time and suggestions.

The experiments of the present study were carried out at the Structural

Engineering and Materials Laboratory of Georgia Institute of Technology. The author

feels grateful to the facility manager Jeremy Mitchell; former research engineer Marcus

Millard; mechanical specialist Michael R. Sorenson; senior facilities manager Andrew

Udell; fellow graduate students Yavuz Mentes, Jong Han Lee, Jonathan Hurff, Victor

Garas, Robert Moser, Brett Holland, Jennifer Dunbeck, Katherine Snedeker and Kennan

Crane; former fellow graduate students Murat Engindeniz and Felix Kim; former

undergraduate assistants Andrew Cao and Luis Fajardo and other people without whose

helping hand the author could not conduct the experiments. Special thanks are due to

fellow graduate students Mustafa Can Kara and Towhid Bhuiyan for their tremendous

help and support during the course of the experiments.

This dissertation is dedicated to the family of the author, whose support and love

made this dream possible. The author feels indebted to his parents Müzeyyen and İsmail

iv
Kalkan and his brother Murat Kalkan for their unconditional love and unlimited support

particularly at difficult times.

Financial support provided for the experiments by Georgia Department of

Transportation (GDOT) is gratefully acknowledged. The author would like to express his

appreciation to Kırıkkale Üniversitesi, Turkey for providing financial support to the

author during the course of his dissertation. Thanks are also due to Dr. Mustafa Y. Kılınç,

Dr. Osman Yıldız and Dr. Orhan Doğan for their support during this study.

v
TABLE OF CONTENTS

Page

ACKNOWLEDGEMENTS iv

LIST OF TABLES x

LIST OF FIGURES xii

LIST OF SYMBOLS xx

LIST OF ABBREVIATIONS xxvii

SUMMARY xxviii

CHAPTERS

I INTRODUCTION 1

1.1 Introduction 1

1.2 Project Objectives 4

1.3 Organization of the Study 4

1.4 Previous Studies 6

1.4.1 Review of Previous Experimental Work 7

1.4.2 Analytical Methods for Predicting Lateral Torsional Buckling 28

1.4.3 Summary 56

II SPECIMENS AND MATERIAL PROPERTIES 57

2.1 Specimens 57

2.1.1 Specimen Descriptions 57

2.1.2 Experiment Design 58

2.2 Concrete Material Properties 62

III EXPERIMENTAL SET-UP, INSTRUMENTATION AND PROCEDURE 69

3.1 Experimental Setup 69

vi
3.1.1 Loading Mechanism 69

3.1.2 Supports 75

3.1.3 Load, Deflection and Strain Measurements 88

3.1.3.1 LVDT Strain Measurements 93

3.1.3.2 Strain Measurements through Electrical Resistance Strain


Gauges 95

3.2 Test Procedure 98

3.3 Summary of the Test Results 99

IV LATERAL BENDING RIGIDITY OF RECTANGULAR REINFORCED


CONCRETE BEAMS AND INFLUENCE OF SHRINKAGE CRACKING
ON THE RIGIDITY 101

4.1 Introduction 101

4.2 Available Lateral Bending Rigidity Expressions 102

4.3 Proposed Lateral Bending Rigidity Expression 105

4.3 Influence of Shrinkage Cracking on the Lateral Bending Rigidity 121

V TORSIONAL RIGIDITY OF RECTANGULAR REINFORCED


CONCRETE BEAMS 131

5.1 Torsional Behavior of Reinforced Concrete Beams 131

5.2 Torsional Rigidity of Rectangular Reinforced Concrete Beams 137

5.2.1 Uncracked Torsional Rigidity Expressions 137

5.2.2 Post-Cracking Torsional Rigidity 141

5.3 Experimental Torsional Rigidities of the Test Beams 145

5.4 Proposed Torsional Rigidity Expression 151

VI CRITICAL MOMENT CALCULATIONS AND INFLUENCES OF


THE INITIAL GEOMETRIC IMPERFECTIONS ON THE LATERAL
STABILITY OF REINFORCED CONCRETE BEAMS 153

6.1 Introduction 153

6.2 Critical Moment Calculations 153

vii
6.3 Influences of Sweep and Initial Twisting Angle on the Lateral
Stability of Reinforced Concrete Beams 156

VII EXPERIMENTAL RESULTS AND OBSERVATIONS AND


CORRELATION OF THE ANALYTICAL AND EXPERIMENTAL
RESULTS 173

7.1 Experimental Results and Observations 173

7.1.1 Crack Patterns of the Specimens 173

7.1.2 Experimental Results 181

7.2 Correlation of the Analytical and Experimental Results 183

VIII SUMMARY AND CONCLUSIONS 195

8.1 Summary 195

8.2 Conclusions 200

8.3 Futures Research 203

APPENDIX A: NOMINAL AND MEASURED DIMENSIONS OF THE


SPECIMENS 205

APPENDIX B: MEASURED INITIAL GEOMETRIC IMPERFECTIONS AND


PERMANENT DEFORMATIONS OF THE SPECIMENS 213

APPENDIX C: MIDSPAN STRAIN DISTRIBUTIONS OF THE BEAMS AT


DIFFERENT LOAD LEVELS 229

APPENDIX D: EXPERIMENTAL LOAD-DEFLECTION PLOTS OF THE


SPECIMENS 249

APPENDIX E: METHOD FOR THE EVALUATION OF THE CENTROIDAL


DEFLECTIONS AND ROTATION OF A BEAM 265

APPENDIX F: CRITICAL MOMENT CALCULATIONS OF THE SPECIMENS 273

APPENDIX G: CONSTRUCTION DETAILS OF THE SPECIMENS 283

G.1 Compression Reinforcement 283

G.2 Tilt-up and Lifting Mechanisms for the Specimens 284

APPENDIX H: DETERMINATION OF THE EXPERIMENTAL BUCKLING


LOADS OF THE SPECIMENS 289

viii
REFERENCES 297

VITA 304

ix
LIST OF TABLES

Page

Table 1.1: Nominal dimensions of the beams tested by Hansell and Winter (1959). 7

Table 1.2: Results of the tests by Hansell and Winter (1959). 8

Table 1.3: Beams tested by Sant and Bletzacker (1961). 15

Table 1.4: Results of the tests by Sant and Bletzacker (1961). 16

Table 1.5: Specimens tested by Massey and Walter (1969). 21

Table 1.6: Beams tested by Konig and Pauli (1990). 23

Table 1.7: Midspan initial imperfections of the beams tested


by Konig and Pauli (1990). 23

Table 1.8: Loads, midspan deformations and rotations at failure of the beams
tested by Konig and Pauli (1990). 23

Table 2.1: Specimens of the present experimental program. 58

Table 2.2: Mechanical properties of concrete. 64

Table 3.1: Experimental results of the specimens. 100

Table 4.1: Experimental and calculated cracking moments of the second set of
specimens. 113

Table 4.2: Descriptions of the shrinkage specimens from the concrete mixtures
used in B44 and B36L. 126

Table 5.1: Maximum torsional moments of the specimens at the initiation


of buckling and the cracking torques. 149

Table 5.2: Torsional rigidities of the specimens. 150

Table 7.1: Experimental and analytical critical load values of the specimens. 186

Table 7.2: Experimental–to-analytical load ratios of the specimens. 187

Table 7.3: Measured sweeps and angles of twist at limit load of the
specimens at midspan. 189

Table A.1: Nominal and measured heights of the first set of specimens. 206

x
Table A.2: Nominal and measured heights of the second set of beams. 207

Table A.3: Nominal and measured widths of the first set of specimens
along the span. 209

Table A.4: Widths of the first set of specimens along the depth
of midspan section. 209

Table A.5: Nominal and measured widths of the second set of beams. 211

Table A.6: Nominal and measured span lengths of first set of specimens. 212

Table A.7: Nominal and measured total lengths of the second set of specimens. 212

Table B.1: Measured sweeps of Specimens B18. 214

Table B.2: Measured sweeps of Specimens B30 and B36. 214

Table B.3: Measured sweeps of Specimens B44. 216

Table B.4: Measured sweeps of Specimens B36L. 217

Table H.1: Critical loads from the classical and Meck’s (1977) version of
the Southwell (1932) plots for the second set of beams. 295

xi
LIST OF FIGURES

Page

Figure 1.1: (a) Lateral torsional buckling of a beam subjected to a concentrated


load at midspan; (b) Lateral and vertical deflections and torsional
rotation of the midspan section 2

Figure 1.2: Load-lateral deflection curve of a reinforced concrete beam. 3

Figure 1.3: Comparison of the vertical deflections at yield point of the companion
beams tested by Hansell and Winter (1959). 9

Figure 1.4: Comparison of the lateral top deflections at yield point of the companion
beams tested by Hansell and Winter (1959). 9

Figure 1.5: Comparison of the lateral bottom deflections at yield point of the
companion beams tested by Hansell and Winter (1959). 10

Figure 1.6: Experimental-to-calculated ultimate moment ratios of the beams


Tested Hansell and Winter (1959). 10

Figure 1.7: Loading mechanism used by Hansell and Winter (1959). 12

Figure 1.8: Undeflected and deflected configurations of the loading


mechanism. 12

Figure 1.9: Direction of the torsional moments induced by the eccentric


application of the load. 13

Figure 1.10: Web sidesway buckling failure of the specimen due to the lateral
restraining forces in the loading mechanism. 15

Figure 1.11: Experimental-to-predicted buckling moment ratios of the beams in the


first three specimen groups, B36, B30 and B24. 17

Figure 1.12: Loading frame used by Sant and Bletzacker (1961). 18

Figure 1.13: Deflected configuration of the loading mechanism


used by Sant and Bletzacker (1961). 19

Figure 1.14: Cross-sectional details of the specimens tested by


Konig and Pauli (1990). 22

Figure 1.15: Ratio of the buckling load of each Specimen to the buckling load
of Specimen 2. 25

xii
Figure 1.16: Loading mechanism used by Konig and Pauli (1990). 27

Figure 1.17: Stress-strain curve of normal-strength concrete and the tangent


moduli of elasticity at different stress levels. 31

Figure 1.18: Secant modulus of elasticity corresponding to the extreme


compression fiber strain. 34

Figure 1.19: (a) Stress distribution in the compression zone of the beam section
in the plastic state; (b) Stress-strain curve of concrete;
assumed by Siev (1960). 40

Figure 1.20: Strain distribution at midspan section and the corresponding


reduced modulus of elasticity. 42

Figure 1.21: Definition of the variables in the expressions proposed


by Massey (1967). 45

Figure 1.22: Effect of the vertical location of the applied load with respect to
the shear center of beam section. 48

Figure 1.23: Definition of area and perimeters in Equations (1.30) - (1.32) 54

Figure 2.1: First set of specimens (B36, B30, B22, and B18) 59

Figure 2.2: Second set of specimens (B44, B36L) 60

Figure 2.3: Congested reinforcement in B36 63

Figure 2.4: Application of self-consolidating concrete 63

Figure 2.5: Stress-strain curves of concrete in the first set of beams 67

Figure 2.6: Stress-strain curves of concrete in B44 67

Figure 2.7: Stress-strain curves of concrete in B36L 68

Figure 3.1: Undeflected and deflected configurations of the loading frame


and loading cage. 70

Figure 3.2: Loading frame 71

Figure 3.3: Undeflected and deflected shapes of the Gravity Load Simulator 72

Figure 3.4: (a) Undeflected; and (b) Deflected configurations of the Gravity Load
Simulator. 72

Figure 3.5: Ball-and-socket joint 73

xiii
Figure 3.6: Vertical orientation of the loading cage (a) Before the test;
(b) After buckling. 74

Figure 3.7: Minor-axis rotations at the supports 75

Figure 3.8: Lateral deformations and torsional rotations restrained


at the supports. 76

Figure 3.9: In-plane support conditions: (a) In analysis models;


(b) Hinged-hinged; (c) Roller-hinged; (d) Roller-roller. 77

Figure 3.10: Roller supports at the beam ends in the second set of experiments. 81

Figure 3.11: Behavior of the beams with: (a) Roller supports;


(b) Hinged supports in lateral direction. 82

Figure 3.12: (a) Support frame in the first set of experiments; (b) A ball roller
in contact with the beam. 83

Figure 3.13: Bending of the ball roller 84

Figure 3.14: Lateral support frame in the second set of experiments. 84

Figure 3.15: Rigid caster in contact with the specimen. 85

Figure 3.16: Support frame in the second set of tests: (a) B44-1; (b) B44-2. 87

Figure 3.17: Distortion in the cross-section at the beam end. 87

Figure 3.18: Deviation of the Initial Orientations of the Potentiometers. 89

Figure 3.19: Coupling between the in-plane and out-of-plane deflection


measurements for a lateral string potentiometer with varying distances
from the specimen. 90

Figure 3.20: Lateral torsional buckling (a) with; (b) without distortions in the
cross-sectional shape of the beam. 91

Figure 3.21: Lateral deflection potentiometers in the first set of experiments. 92

Figure 3.22: Lateral deflection potentiometers in the second set of experiments. 92

Figure 3.23: Strain measurement using LVDT’s in the first set of tests. 93

Figure 3.24: Strain measurement through strain gages in the second set of tests. 94

Figure 3.25: 2-element cross strain gage on the side face of B44-1. 96

xiv
Figure 3.26: Longitudinal strain gages along the depth of north face of Specimen
B44-2 at midspan. 97

Figure 3.27: Strain gages on an aluminum strip to measure the longitudinal strain
in the tension zone. 98

Figure 4.1: Spring models defining (a) Branson’s (1963); (b) Bischoff’s (2005)
effective moment of inertia expression. 109

Figure 4.2: In-plane deflections of Beams B44-1 at midspan. 110

Figure 4.3: In-plane deflections of Beams B44-2 at midspan. 110

Figure 4.4: In-plane deflections of Beams B44-3 at midspan. 111

Figure 4.5: In-plane deflections of Beams B36L-1 at midspan. 111

Figure 4.6: In-plane deflections of Beams B36L-2 at midspan. 112

Figure 4.7: Moduli of elasticity corresponding to the fibers in the compression


zone of a beam section. 114

Figure 4.8: Proposed spring model for the lateral bending behavior
of reinforced concrete beams. 117

Figure 4.9: Shrinkage cracking in B30 prior to the test. 123

Figure 4.10: Length changes of specimens with and without SRA


from the concrete mixture of B44. 126

Figure 4.11: Delta rosette for principal strain measurement at a point. 128

Figure 4.12: Principal strains on the side face of B36L-2. 129

Figure 4.13: Principal strains on the side face of B36L-3. 129

Figure 5.1: Torque-twist curve of a reinforced concrete beam


with shear reinforcement. 132

Figure 5.2: Components of the axial torque on the failure surface of a concrete
beam according to the skew-bending theory. 134

Figure 5.3: Comparison of the coefficients βc calculated from different equations. 142

Figure 5.4: Thin-walled tube space truss model. 144

Figure 5.5: Experimental torque-twist curve of Specimen B44-1. 146

Figure 5.6: Experimental torque-twist curve of Specimen B36L-1. 146

xv
Figure 5.7: Approximation of the torque-twist curve of B44-2
into a series of line segments. 148

Figure 6.1: Lateral centroidal deflections of B44-1 and B44-2 at midspan. 158

Figure 6.2: Lateral centroidal deflections of B36L-1 and B36L-2 at midspan. 158

Figure 6.3: Longitudinal strain distributions in a cross-section from major-axis


and minor-axis bending moments. 161

Figure 6.4: Extreme compression fiber strains of B44-1 and B44-2 from
. major-axis bending 163

Figure 6.5: Extreme compression fiber strains of B36L-1 and B36L-2 from
. major-axis bending 163

Figure 6.6: Extreme top strains on the convex faces of B44-1 and B44-2 caused by
. minor-axis bending 165

Figure 6.7: Top strains on the convex faces of B36L-1 and B36L-2 caused by
. minor-axis bending 167

Figure 6.8: Rotation of the major and minor axes of a section due to twist. 169

Figure 6.9: (a) Southwell (1932) Plot; (b) Load-Deflection Plot for
Specimen B44-1. 169

Figure 7.1: Typical crack pattern on the convex faces of the specimens
after buckling 174

Figure 7.2: Typical crack pattern on the concave faces of the specimens
after buckling 174

Figure 7.3: Flexural cracks on the concave face of B44-3 at midspan


before buckling 175

Figure 7.4: Vertical cracks on the convex face of B44-2 at midspan


after buckling 176

Figure 7.5: Directions of the shear and principal stresses due to the shear forces. 178

Figure 7.6: Directions of the shear and principal stresses due to the torsional
moments. 179

Figure 7.7: Diagonal tension cracks on the convex face of B36L-1 after buckling. 180

Figure 7.8: Diagonal tension cracks on the concave face of B18-2 after buckling. 180

xvi
Figure 7.9: Diagonal tension cracks continuing on the top surface of B44-2
after buckling 181

Figure 7.10: Maximum compressive strains in the first set of beams,


illustrated on the stress-strain curve of concrete. 182

Figure 7.11: Maximum compressive strains in B44 at the instant of buckling. 182

Figure 7.12: Maximum compressive strains in B36L at the instant of buckling. 183

Figure 7.13: Experimental to analytical critical load ratios of the specimens


according to different formulae. 188

Figure A.1: Height measurement points along the lengths of the beams. 208

Figure A.2: Width measurement points along the lengths of the specimens. 210

Figure A.3: Length measurement depths of the specimens. 212

Figure B.1: Imperfection measurement points on Beams (a) B18; (b) B30 and B36;
(c) B44 and B36L. 215

Figure B.2: Sweep at midheight of B18-1. 219

Figure B.3: Sweep at midheight of B18-2. 219

Figure B.4: Sweep at midheight of B30. 220

Figure B.5: Sweep at midheight of B36. 220

Figure B.6: Sweep at midheight of B44-1. 221

Figure B.7: Sweep at midheight of B44-2. 221

Figure B.8: Sweep at midheight of B44-3. 222

Figure B.9: Sweep at midheight of B36L-1. 222

Figure B.10: Sweep at midheight of B36L-2. 223

Figure B.11: Permanent lateral deformation at midheight of B30. 223

Figure B.12: Permanent lateral deformation at midheight of B44-1. 224

Figure B.13: Permanent lateral deformation at midheight of B44-2. 224

Figure B.14: Permanent lateral deformation at midheight of B44-3. 225

Figure B.15: Permanent lateral deformation at midheight of B36L-2. 225

xvii
Figure B.16: Permanent torsional rotations of B30. 226

Figure B.17: Permanent torsional rotations of B44-1. 226

Figure B.18: Permanent torsional rotations of B44-2. 227

Figure B.19: Permanent torsional rotations of B44-3. 227

Figure B.20: Permanent torsional rotations of B36L-1. 228

Figure C.1: Loads and lateral deflections corresponding to the strain distributions
in Figures C.2 to C.4. 230

Figure C.2: Midspan strain distributions of B44-1 at the initial stages of loading. 231

Figure C.3: Midspan strain distributions of B44-1 close to buckling. 232

Figure C.4: Midspan strain distributions of B44-1 after buckling. 233

Figure C.5: Loads and lateral deflections corresponding to the strain distributions
in Figures C.6 and C.7. 233

Figure C.6: Midspan strain distributions of B44-2 at the initial stages of loading. 234

Figure C.7: Midspan strain distributions of B44-2 at the initiation of buckling. 235

Figure C.8: Loads and lateral deflections corresponding to the strain distributions
in Figures C.9 to C.11. 235

Figure C.9: Midspan strain distributions of B44-3 at the initial stages of loading. 236

Figure C.10: Midspan strain distributions of B44-3 at different load levels. 237

Figure C.11: Midspan strain distributions of B44-3 close to buckling. 238

Figure C.12: Loads and lateral deflections corresponding to the strain distributions
in Figures C.13 and C.14. 239

Figure C.13: Midspan strain distributions of B36L-1 at the initial stages of loading. 240

Figure C.14: Midspan strain distributions of B36L-1 close to buckling. 241

Figure C.15: Loads and lateral deflections corresponding to the strain distributions
in Figures C.16 to C.18. 242

Figure C.16: Midspan strain distributions of B36L-2 at the initial stages of loading. 243

Figure C.17: Midspan strain distributions of B36L-2 at different load levels. 244

xviii
Figure C.18: Midspan strain distributions of B36L-2 prior to and after buckling. 245

Figure C.19: Depthwise strains along the midspan section of B44-1 at the
initial stages of loading. 246

Figure C.20: Depthwise strains along the midspan section of B44-1 at the
final stages of loading. 247

Figure C.21: Depthwise strains along the midspan section of B44-1 during
unloading. 248

Figure D.1: Out-of-plane deflections of B18-2 at midspan. 250

Figure D.2: Out-of-plane deflections of B22-1 at midspan. 250

Figure D.3: Out-of-plane deflections of B22-2 at midspan. 251

Figure D.4: Out-of-plane deflections of B30 at midspan. 251

Figure D.5: Out-of-plane deflections of B36 at midspan. 252

Figure D.6: Out-of-plane deflections of B44-1 at midspan. 252

Figure D.7: Out-of-plane deflections of B44-2 at midspan. 253

Figure D.8: Out-of-plane deflections of B44-3 at midspan. 253

Figure D.9: Out-of-plane deflections of B36L-1 at midspan. 254

Figure D.10: Out-of-plane deflections of B36L-2 at midspan. 254

Figure D.11: Lateral centroidal deflections of Beams B44 at midspan. 255

Figure D.12: Lateral centroidal deflections of Beams B36L at midspan. 255

Figure D.13: In-plane deflections of Beam B18-2 at midspan. 258

Figure D.14: In-plane deflections of Beams B22 at midspan. 258

Figure D.15: In-plane deflections of Beams B30 at midspan. 259

Figure D.16: In-plane deflections of Beams B36 at midspan. 259

Figure D.17: In-plane deflections of Beams B44-1 at midspan. 260

Figure D.18: In-plane deflections of Beams B44-2 at midspan. 260

Figure D.19: In-plane deflections of Beams B44-3 at midspan. 261

xix
Figure D.20: In-plane deflections of Beams B36L-1 at midspan. 261

Figure D.21: In-plane deflections of Beams B36L-2 at midspan. 262

Figure D.22: Experimental torque-twist curve of Specimen B44-1. 262

Figure D.23: Experimental torque-twist curve of Specimen B44-2. 263

Figure D.24: Experimental torque-twist curve of Specimen B44-3. 263

Figure D.25: Experimental torque-twist curve of Specimen B36L-1. 264

Figure D.26: Experimental torque-twist curve of Specimen B36L-2. 264

Figure E.1: Potentiometer configuration in the tests. 266

Figure E.2: Angle of twist calculations. 269

Figure E.3: Centroidal deflection calculations. 272

Figure G.1: Reinforcement in Specimen B36. 285

Figure G.2: Lifting mechanism in the first set of beams. 285

Figure G.3: Use of spreader beams to lift the beams. 287

Figure G.4: The lifting point in the second set of beams


connected to the spreader beam. 288

Figure H.1: Southwell (1932) Plots for Specimen B44-2. 292

Figure H.2: Meck’s (1977) Version of the Southwell (1932) Plots


for Specimen B44-2. 293

Figure H.3: Massey’s (1963) Version of the Southwell (1932) Plots


for Specimen B44-2. 294

xx
LIST OF SYMBOLS

Ac Gross area of the cross-section (Figure 1.24)

Ae Area bounded by the centerline of the effective wall

Ao Area bounded by the shear flow zone

As Total cross-sectional area of the longitudinal reinforcement

At Cross-sectional area of one leg of a stirrup

A1 Area bounded by the centerline of a closed stirrup (Figure 1.24)

A2 Area of the rectangle formed by the lines connecting the centroids of the
corner longitudinal bars (Figure 1.24)

B Lateral bending rigidity

E Modulus of elasticity

Ec Elastic modulus of concrete

Eit Initial tangent modulus of elasticity

Eo Overall modulus of elasticity [= (Esec+Ec)/2]

Reduced modulus of elasticity   4  Ec  Etan   


2
Er Ec  Etan


Es Modulus of elasticity of steel

Esec Secant modulus of elasticity

Etan Tangent modulus of elasticity

Etano Tangent modulus of elasticity corresponding to the extreme compression fiber


strain (Figure 1.17)

ECw Warping rigidity

EIx In-plane bending rigidity

EIy Out-of-plane bending rigidity

xxi
G Modulus of rigidity

Gc Modulus of rigidity of concrete

G’c Reduced modulus of rigidity of concrete   Esec 2  1    

GcrCcr Post-cracking torsional rigidity (Figure 5.1)

Go Overall modulus of rigidity of concrete

Gs Modulus of rigidity of steel

(GC) Torsional rigidity

Icr Cracked moment of inertia about the major axis

Ie Effective moment of inertia

Ig Moment of inertia of the gross cross-section about the major axis

Ix Moment of inertia about the major axis

Iucr Uncracked moment of inertia about the major axis considering the contribution
of the longitudinal reinforcement

Iy Moment of inertia about the minor axis

J Torsional constant

L Unbraced length

Mx In-plane bending moment

My Out-of-plane bending moment

Ma Maximum in-plane bending moment along the span of a beam

Mc Calculated ultimate moment

Mcr Critical moment

Mcra Cracking Moment

ML Limit moment

Mex Experimental ultimate moment

P Applied load

xxii
Pan Analytical ultimate or critical load

Pcr Critical load

PL Limit load

T Torsional moment

Ta Applied torque

Tb Maximum torsional moment in the beam at the instant of buckling

Tcr Cracking torque

Tmax Maximum torsional moment in a beam

Tn Torsional strength of a reinforced concrete member

Tnp Torsional strength of a plain concrete member

a Moment lever arm of a section

b Beam width

bs Width of the longitudinal reinforcement layer (Figure 1.18)

b1 Width of the area bounded by the centerline of a closed stirrup (Figure 1.18)

b2 Width of the rectangle formed by the lines connecting the centroids of the
longitudinal reinforcing bars (Figure 1.24)

c Neutral axis depth from the compression face

cu Neutral axis depth at the ultimate flexural moment level

cv Percent coefficient of variation

d Depth of the centroid of the tension reinforcement from the compression face

d1 Depth of the area bounded by the centerline of a closed stirrup (Figure 1.18)

e Vertical distance of the point of application of load from the centroid of the
beam section

f Stress

fc Concrete stress

f’c Compressive strength of concrete

xxiii
fr Modulus of rupture of concrete

ft Splitting tensile strength of concrete

h Beam height

kcr Lateral bending rigidity of the cracked portion of a concrete beam (Figure 4.1)

keq Lateral bending rigidity of a concrete beam (Figure 4.1)

kucr Lateral bending rigidity of the uncracked portion of a concrete beam (Figure 4.1)

n Modular ratio of steel to concrete [=Es/Ec]

pe Perimeter of the area bounded by the centerline of the effective wall

po Perimeter of the area bounded by the shear flow zone

p1 Perimeter of the area bounded by the centerline of a closed stirrup (Figure 1.24)

p2 Perimeter of the rectangle formed by the lines connecting the centroids


of the corner longitudinal bars (Figure 1.24)

s Spacing of the stirrups

ti Wall thickness of a tube

ts Thickness of the longitudinal reinforcement layer (Figure 1.18)

uc Out-of-plane deflection of the centroid of midspan section

uo Sweep at midheight of a beam at midspan

ut Out-of-plane deflection at the top of midspan section

uto Sweep at the top of a beam at midspan

u(z) Centroidal sweep at a distance z from the end of a beam

v Lateral deflection of the centroid of midspan section in the direction of the


major axis of the twisted configuration of a beam (Figure 6.8)

vc In-plane deflection of the centroid of midspan section

vo Initial imperfection at the center of a beam in the direction of the major axis
of the initial configuration of midspan section

w Self-weight per unit length of a beam

xxiv
wcr Critical self-weight per unit length of a beam causing buckling

y Depth of the center of gravity of the transformed beam section from


the compression face (Equation 4.6)

ΣIsy Moment of inertia of the longitudinal reinforcement about the minor axis of the
beam section

β Coefficient for St. Venant’s torsional constant

ε Strain

εc Strain of the compression fibers at an arbitrary depth from the compression


face created by the major-axis bending (Figure 1.17)

εco Extreme compression fiber strain from major-axis bending

εcl Compressive strain on the concave face of the beam originating from the
lateral bending moment only (Figure 1.17)

εcr Cracking strain of concrete

εo Strain at peak stress

εto Strain at the centroid of the tension reinforcement from major-axis bending

εtl Tensile strain on the convex face of the beam originating from the lateral
bending moment only (Figure 1.17)

εy Yielding strain of steel

θ Twist

θcri Twist at the initiation of diagonal cracking (Figure 5.1)

θcrp Twist at the end of diagonal cracking (Figure 5.1)

θu Twist at ultimate torque (Figure 5.1)

μ Mean value

ν Poisson’s ratio

νc Poisson’s ratio of concrete

ρl Volumetric ratio of the longitudinal reinforcement   As Ac 

ρt Volumetric ratio of the transverse reinforcement   At  p1  Ac  s 

xxv
ρto Total volumetric ratio of the transverse and longitudinal reinforcement [= ρl + ρt]

σ Standard deviation

σc Extreme compression fiber stress

φc Angle of twist of the beam at midspan

φo Initial twisting angle of a beam at midspan

φult Twisting angle of a beam at midspan at the instant when the limit load is
reached

φ(z) Twisting angle at a distance z from the end of a beam

xxvi
LIST OF ABBREVIATIONS

DEMEC Demountable mechanical strain gage

HRWR High range water reducing admixture

LVDT Linear variable differential transducer

NA Neutral axis

OC Conventionally vibrated ordinary concrete

SCC Self-compacting concrete

SRA Shrinkage reducing admixture

WWR Welded wire reinforcement

xxvii
SUMMARY

The study presents the results of an investigation aimed at examining the lateral stability

of rectangular reinforced concrete slender beams. A total of eleven reinforced concrete

beams having a depth to width ratio between 10.20 and 12.45 and a length to width ratio

between 96 and 156 were tested. Beam thickness, depth and unbraced length were 1.5 to

3.0 in., 18 to 44 in., and 12 to 39.75 ft, respectively. The initial geometric imperfections,

shrinkage cracking conditions and material properties of the beams were carefully

determined prior to the tests.

Each beam was subjected to a single concentrated load applied at mid-span by

means of a gravity load simulator that allowed the load to always remain vertical when

the section displaces out of plane. The loading mechanism minimized the lateral

translational and rotational restraints at the point of application of load to simulate the

nature of gravity load.

Each beam was simply-supported in and out of plane at the ends. The supports

allowed warping deformations, yet prevented twisting rotations at the beam ends.

In the experimental part of the study, reinforced concrete beams with initial

imperfections (sweep) failed under loads lower than the critical loads corresponding to

the geometrically perfect configuration of the respective beams. The maximum load

carried by an imperfect beam is known as the limit load (PL). In the present study, the

limit load (PL) and the critical load (Pcr) were distinguished.

In the first part of the analytical investigation, a formula was developed for

determining the critical loads corresponding to the lateral torsional buckling of

xxviii
rectangular reinforced concrete beams. The effects of shrinkage cracking and inelastic

stress-strain properties of concrete and the contribution of longitudinal reinforcement to

the lateral stability are accounted for in the critical load formula. The second part of the

investigation focused on developing a formula for the estimation of limit loads of

reinforced concrete beams with initial lateral imperfections. The proposed limit load

formula was obtained by introducing the destabilizing effect of sweep as a reduction term

to the critical load equation.

Finally, the experimental results were compared to the proposed analytical

solution and to various lateral torsional buckling solutions in the literature. The

formulation proposed in the present study was found to agree well with the experimental

results. The good correlation with the experimental results and the incorporation of the

geometric and material nonlinearities into the formula makes the proposed solution, given

below for a simply supported rectangular reinforced concrete beam loaded with a

concentrated load at midspan, practical for design purposes:

Pcru  

4  M cr uto  48  Ec  I y  (1)
L sin(ult )  L3

where PL is the limit load; L is the unbraced length of the beam; uto is the sweep at the top

of the beam at midspan; Ec is the elastic modulus of concrete; Iy is the second moment of

area of the beam section about the minor axis; φult is the angle of twist of the beam at

midspan corresponding to the limit load (PL). Mcr is the critical moment corresponding to

the geometrically perfect configuration of the beam, obtained from Equation (2):

xxix
4.23  e Bo 
M cr   1  1.74     B   GC o (2)
L 

L  GC o  o

where Bo is the lateral bending rigidity, obtained from Equation (3); (GC)o is the torsional

rigidity, calculated from Equation (4); e is the vertical distance of the load application

point from the centroid of the midspan cross section.

 
 3 
  b c  1    Esec  Ec 
Bo     (3)
 12   
2
 M cra   c  
  2


 1      1 
  M cr   h  

Esec  Ec  b3  h  b 
(GC )o    1  0.63    (4)
4  1     3  h 

where b and h are the width and height of the beam, respectively; c is the depth of the

neutral axis from the compression face; Mcra is the cracking moment; ω is a constant,

which has a value of 1 in the absence of restrained shrinkage cracks in concrete and a

value of 2/3 in the presence of restrained shrinkage cracks and υ is Poisson’s ratio of

concrete. Esec is the secant modulus of elasticity of concrete corresponding to the extreme

compression fiber strain at midspan at the instant when Mcr is reached.

xxx
INTRODUCTION

1.1 Introduction

Due to the increasing use of slender structural concrete beams in long-span bridges and

other structures, lateral stability is becoming an important criterion in the design of

structural concrete girders. Lateral-torsional buckling of long-span precast concrete

girders is a matter of concern, particularly during bridge construction.

Bridge girders are laterally supported by diaphragms and the bridge deck after the

completion of a bridge. Nonetheless, lateral stability of the precast bridge girders should

be assured also during fabrication, lifting, transportation and erection stages.

Accordingly, precast concrete girders should be designed to remain stable even under the

most unfavorable loading and support conditions of the transitory phases of construction.

Lateral instability of a beam arises from the compressive stresses in the beam

resulting from flexure due to transverse loading. The compression zone of the beam tends

to buckle about the minor axis of the overall cross-section of the beam while the tension

zone tends to remain stable. When the load reaches a certain “critical” value, the beam

buckles out of plane and twists (Figure 1.1) as a result of the differential lateral

displacements of the compression and tension zones.

For assessing the stability, the critical moment of a concrete girder should be

evaluated for the loading and support conditions of different phases of construction. A

beam free from initial geometric imperfections does not undergo out-of-plane deflections

and rotations before reaching a critical moment value. When the maximum moment in

the beam reaches the critical moment value, the beam experiences sudden excessive out-

1
Figure 1.1 – (a) Lateral torsional buckling of a beam subjected to a concentrated load at

midspan; (b) Lateral and vertical deflections and rotation of the midspan section

of-plane deformations and torsional rotations. This type of buckling is known as

bifurcation instability and the moment at which the beam loses its stability and

experiences rapid and excessive deformations at a constant load level is known as the

critical moment (Mcr).

A beam having initial geometric imperfections, on the contrary, does not bifurcate

at the limit load. The beam undergoes deformations and rotations throughout the whole

2
course of loading, even prior to buckling. The moment carried by the beam reaches an

ultimate value, called the limit moment (ML), beyond which greater lateral deformations

and rotations take place while the moment-carrying capacity of the beam slowly

decreases (Figure 1.2). This type of instability is known as limit load instability.

Figure 1.2 – Load-lateral deflection curve of a slender reinforced concrete beam

In reinforced concrete beams, the difference between the critical moment (Mcr)

and the limit moment (ML) is more pronounced since cracking in an imperfect concrete

beam due to the lateral displacements prior to buckling decreases the moment-carrying

capacity of the beam significantly.

3
ACI 318-05 (2005) does not include an analytical method for the calculation of the

critical moment of a concrete beam. The only provision regarding the stability is given in

Section 10.4, which limits the ratio of beam span to beam width, L/b, to less than 50.

AASHTO LRFD (2005) specifies in Section 5.5.4.3 that: “Buckling of precast

members during handling, transportation, and erection shall be investigated.” However,

no analytical method is given for the calculation of the critical moment of a reinforced

concrete beam

1.2 Project Objectives

The research described herein investigates the lateral stability of rectangular reinforced

concrete beams experimentally and analytically. The analytical study was carried out to

develop an analytical method to estimate critical moments of rectangular reinforced

concrete beams. In the experimental part of the study, a total of eleven slender

rectangular reinforced concrete beams were tested to produce experimental data for

supporting the analytical methods proposed for examining the lateral-torsional buckling

of reinforced concrete beams.

Attention is given to the effects of the initial geometric imperfections and

shrinkage on the lateral stability of reinforced concrete beams.

1.3 Organization of the Study

Section 1.4 summarizes the previous studies in the literature on lateral stability of

reinforced concrete beams. Chapter II introduces the specimens of the experimental

program and mechanical properties of the concrete mixtures used in the specimens.

4
Chapter III presents the experimental setup used to test the beams and summarizes the

test procedure.

Chapter IV summarizes the previously developed formulae concerning the lateral

bending rigidity of rectangular reinforced concrete beams and introduces the new lateral

bending rigidity equation proposed in this study. The chapter also presents the spring

systems used to model a reinforced concrete beam when developing the flexural rigidity

expressions. Finally, the effect of restrained shrinkage cracking on the lateral bending

rigidity of a concrete beam is examined in the last section of Chapter IV, where a

modification to the proposed lateral bending rigidity expression is introduced to account

for the reduction in the rigidity due to the presence of possible shrinkage cracks in

concrete.

In Chapter V, the torsional rigidity expressions for rectangular reinforced concrete

beams available in the literature are presented. Later, the slopes of the experimental

torque-twist curves of the specimens are compared to the analytical values obtained from

the torsional rigidity expressions given in the chapter. The torsional rigidity expression

giving the closest agreement with the experimental results is modified to account for the

possible inelastic material behavior of concrete at the time of buckling.

Chapter VI presents the critical moment calculations of reinforced concrete

beams. In Section 6.2, effects of the initial geometric imperfections on the ultimate

moment and the out-of-plane deformations and twisting rotations of an imperfect

reinforced concrete beam are explained and modifications to the critical moment

expression are proposed to account for the effects of geometric nonlinearities.

5
In Chapter VII, the crack patterns of the specimens and some experimental

results are presented, and the analytical critical load values obtained from the formulae

given in Chapter VI are compared to the experimental buckling loads of the specimens to

determine the degree of correlation between the analytical and experimental results.

Finally, conclusions of the study are summarized in the last chapter.

In the present study, limit moments of the specimens were taken as the greatest

moments in the experimental load-deflection plots of the specimens. There are some

other methods given in the literature for obtaining the buckling moments of beams by

using the experimental data. The methods developed by Southwell (1932), Meck (1977)

and Massey (1963) and their applications to reinforced concrete beams are explained in

Appendix H.

1.4 Previous Studies


This section reviews the previous studies on lateral torsional buckling of rectangular

reinforced concrete beams. The experimental studies in the literature are presented in

Section 1.4.1. Next, the analytical methods in the literature for predicting the critical

loads of reinforced concrete beams are explained in Section 1.4.2. Finally, the

contributions of the previous studies to the field of lateral stability of reinforced concrete

beams are summarized in Section 1.4.3, where the factors that remained uninvestigated in

the literature are also emphasized to support the need for the present research.

6
1.4.1 Review of Previous Experimental Work
Hansell and Winter (1959) studied the lateral stability of reinforced concrete beams both

experimentally and analytically. The main goal of the experimental study was to

investigate any possible reductions in the flexural capacities of reinforced concrete beams

with increasing L/b ratios. Hansell and Winter (1959) tested five different groups of

beams, namely B6, B9, B12, B15 and B18. Two companion beams for each group of

specimens were made and tested to failure. Nominal dimensions of the beams are

presented in Table 1.1. All specimen groups except B6 violated the slenderness criterion,

given in the 1956 Edition of ACI Building Code, which limited the L/b ratio to less than

32 for reinforced concrete beams.

Table 1.1 – Nominal dimensions of the beams tested by Hansell and Winter (1959)

Height, h Width, b Length, L


Specimen d/b ratio L/b ratio
(in.) (in.) (ft)
B18 13 2.5 18 4.5 86.4
B15 13 2.5 15 4.5 72.0
B12 13 2.5 12 4.5 57.6
B9 13 2.5 9 4.5 43.2
B6 13 2.5 6 4.5 28.8

All specimens tested by Hansell and Winter (1959) failed in in-plane bending

compression failure after the yielding of tension reinforcement and developed their

ultimate flexural strength prior to lateral torsional buckling. Experimental ultimate

moments of the specimens are presented in Table 1.2 together with the calculated

ultimate flexural moments. The experimental-to-calculated ultimate moment ratio of each

test beam is also given in the table. The experimental ultimate moments are in good

7
Table 1.2 – Results of the tests by Hansell and Winter (1959)

Experimental Calculated Experimental-


Ultimate Ultimate to-Calculated
Specimen Failure Mode
Moment, Mex Moment, Mc Moment Ratio
(in-kips) (in-kips) Mex/Mc
B6-1 Flexure 216 196.7 1.10
B6-2 Flexure 199 196.7 1.01
B9-1 Flexure 201 196.7 1.02
B9-2 Flexure 205 196.7 1.04
B12-1 Flexure 193 197.0 0.98
B12-2 Flexure 199 197.0 1.01
B15-1 Flexure 192 195.9 0.98
B15-2 Flexure 198 195.9 1.01
B18-1 Flexure 190 196.2 0.97
B18-2 Flexure 196 196.2 1.00

agreement with the calculated moment values. The mean and the coefficient of variation

of Mex/Mc are 1.01 and 3.5%, respectively.

Hansell and Winter (1959) also reported the midspan vertical, lateral top and

lateral bottom deflections of the specimens at the onset of yielding of the flexural

reinforcement. These are shown in Figures 1.3, 1.4 and 1.5, respectively. The test results

of the identical (companion) beams are also shown in the figures. It is to be noted in these

figures that the vertical deflections corresponding to the onset of steel yielding and the

ultimate moments (Figure 1.6) are in close agreement among the companion beams while

the lateral top and bottom deflections show significant variations among the companion

beams. For instance, out-of-plane deflections of the companion beams in Specimen

Groups B12, B15 and B18 display considerable variation.

Hansell and Winter (1959) loaded the specimens at quarter points to have constant

in-plane flexural moment over the middle part of the span. Under the loading and support

conditions reported by Hansell and Winter (1959), the bottom portions of the beams at

8
Figure 1.3 – Comparison of the vertical deflections at yield point of the companion
beams tested by Hansell and Winter (1959)

Figure 1.4 –Comparison of the lateral top deflections at yield point of the companion
beams tested by Hansell and Winter (1959)

9
Figure 1.5 – Comparison of the lateral bottom deflections at yield point of the companion
beams tested by Hansell and Winter (1959)

Figure 1.6 – Experimental -to- calculated ultimate moment ratios of beams tested by
Hansell and Winter (1959)

10
midspan were subjected to tensile stresses from major-axis bending while the top portions

were subjected to compressive stresses. In lateral torsional buckling, the compression

zone of a beam is prone to undergo greater lateral deflections than the tension zone due to

the stabilizing effect of the tensile stresses from in-plane bending. Nevertheless, the

midspan lateral bottom deflections of some test specimens of Hansell and Winter (1959)

exceeded the lateral top deflections.

Figure 1.7 shows the mechanism used by Hansell and Winter (1959) to convey

the load from the head of a universal testing machine to the test beam. Hansell and

Winter (1959) used a loading ball for the rotational freedom and a roller assembly to

provide lateral translational freedom at the loading point. When the test beam deflected

out of plane, the parts of the loading mechanism below the roller assembly were

supposed to move with the rollers in the lateral direction (Figure 1.8), preventing any

lateral restraint to the test beam. Furthermore, the loading cage around the beam was

expected to rotate with the beam about the loading ball, preventing any torsional restraint

to the beam at the loading point. The specimens were loaded using a universal testing

machine. The load was transmitted to the loading points (quarter points of the span)

through a steel beam connected to the specimen at each loading point, through the

loading mechanism shown in Figure 1.7.

An examination of the loading fixture (Figure 1.8) used by Hansell and Winter

(1959) reveals that the steel beam transmitting the load to the specimen does not displace

in the lateral direction while the beam deforms out of plane. Therefore, the line of action

11
Figure 1.7 – Loading mechanism used by Hansell and Winter (1959)

Figure 1.8 –Undeflected and expected deflected configurations of the loading mechanism

12
of the vertical load, initially passing through the shear center of the beam section,

becomes eccentric with respect to the shear center as the specimen deforms out of plane.

The eccentricity of the applied load creates larger and larger torsional moments in the

beam as the applied load increases in the course of the test.

Figure 1.9 depicts the position of the line of the applied load relative to the beam

when the beam undergoes lateral deflections and torsional rotations. The roller

assemblies allow free out-of-plane deflections in the beam at the loading points while the

steel beam remains stationary in the lateral direction. Hence, the line of action of the

applied load stays in its original position, rendering the applied load eccentric relative to

the shear center, which results in torsional moments in opposite direction to the torsional

rotations from instability. The accidental torsions constitute a restraint to lateral torsional

buckling.

Figure 1.9 –Direction of the torsional moments induced by the eccentric application of
the load

13
In the loading fixture used by Hansell and Winter (1959), the loading ball right

above the specimen, the socket plate, the lower roller block and the roller assembly

(Figure 1.7) move with the specimen in the lateral direction when the specimen

undergoes out-of-plane deflections. The loading ball, moving with the specimen in lateral

direction, applies lateral forces to the socket plate (Figure 1.10). If there is rolling friction

in the roller assembly, the lateral translation of the roller assembly and the socket plate is

restrained and the socket plate applies reaction forces to the loading ball, which restrains

the lateral deflection of the top portion of the beam. Significant friction forces in the

roller assembly can cause the top portion of the beam to be more stable than the bottom

portion, which has no lateral translational restraint. In this case, the bottom portion

undergoes greater lateral deformations than the top portion (Figure 1.10) and the beam

experiences a different type of buckling called the web sidesway buckling. Hansell and

Winter (1959) stated that all rolling surfaces in their setup was cleaned and oiled prior to

each test to minimize the rolling friction in the loading fixture and the lateral translation

restraint to the top portion of the beam.

Considering the good agreement between the experimental ultimate moments and

the analytical values calculated according to Eq. (A.1) in 1956 Edition of the ACI

Building Code, Hansell and Winter (1959) concluded that there were no reductions in the

experimental ultimate moments of the beams due to the slenderness effects.

Sant and Bletzacker (1961) tested four different groups of beams, denoted B36,

B30, B24 and B12 whose nominal dimensions are specified in Table 1.3. Three identical

beams of each of the first three groups, B36, B30 and B24 and two identical beams of the

fourth group, B12 were tested to failure. Table 1.4 summarizes the test results. The mean

14
Figure 1.10 –Web sidesway buckling failure of the specimen due to the lateral restraining

forces in the loading mechanism

Table 1.3 – Beams tested by Sant and Bletzacker (1961)

Specimen Number of Height, h Width, b Length, L


d/b ratio L/b ratio
Group Samples (in.) (in.) (ft)
B36 3 36 2.5 20 12.45 96
B30 3 30 2.5 20 10.20 96
B24 3 24 2.5 20 8.13 96
B12 2 12 2.5 20 3.78 96

value of the test results of identical beams are included in the table.

15
Table 1.4 – Results of the tests by Sant and Bletzacker (1961)

Test Moment,
Group Test Specimen Failure Mode
Mtest (in-kips)
B36-1 Stability 1620
B36-2 Stability 1845
I
B36-3 Stability 1350
µ* 1605
B30-1 Stability 2040
II B30-2 Stability 2160
B30-3 Stability 1402
µ 1867
B24-1 Stability 1260
III B24-2 Stability 1350
B24-3 Stability 1440
µ 1350
B12-1 Flexure 300
IV B12-2 Flexure 210
µ 255
* - Mean value of the test moments of the beams in the same group

Test results show considerable variation. For instance, the experimental buckling

moment of Specimen B30-2 is 54% larger than the experimental moment value obtained

by testing its companion, B30-1. In Fig. 1.11, the experimental-to-predicted buckling

moment ratios of the beams in specimen groups B36, B30 and B24 are shown to reveal

the variation in the test results of companion beams. Since B12-1 and B12-2 did not

experience lateral torsional buckling, they are not included in the figure.

Sant and Bletzacker (1961) used a steel loading ball to provide rotational freedom

and a rolling mechanism to provide lateral-translational freedom at the point of

application of load (Figure 1.12). The specimens were loaded through a hydraulic load

cylinder, placed right above the beam and connected to the beam through threaded rods.

16
Figure 1.11 – Experimental-to-predicted buckling moment ratios of the beams in the
first three specimen groups, B36, B30 and B24

The loading ball was located on the head of the load cylinder in a ball-and-socket joint.

Finally, the roller assembly was placed above a load cell, which was located adjacent to

the top surface of the socket plate of the ball-and-socket assembly.

A ball-and-socket joint allows free angular motion of the connecting parts relative

to each other. When a beam experiences torsional rotations in the test setup used by Sant

and Bletzacker (1961), the beam and the load cylinder, connected to it, rotate relative to

the top portion of mechanism above the loading ball. The loading ball rotates in the

socket with the specimen and cylinder, preventing any rotational restraint to the beam at

loading point. As illustrated in Figure 1.13, the load cylinder is also free to rotate relative

17
Figure 1.12 – Loading frame used by Sant and Bletzacker (1961)

18
Figure 1.13 – Deflected configuration of the loading mechanism
used by Sant and Bletzacker (1961)

to the top portion of the mechanism, since the cylinder is placed between the specimen

and the ball-and-socket joint. Therefore, the cylinder ceases to be oriented vertically once

the specimen experiences torsional rotations. The deviation of the applied load from the

vertical axis induces lateral restraining force to the loading mechanism, which prevents

the rollers in the rolling mechanism to move freely in the lateral direction. The lateral-

translational restraint at the loading point increases the buckling load of a beam and

decreases the out-of-plane deformations and the torsional rotations.

The verticality of the applied load in a lateral-torsional buckling test is crucial,

particularly in the case of geometrically imperfect beams. Theoretically, a geometrically

19
perfect beam experiences little or no out-of-plane deformations and torsional rotations

prior to buckling. On the contrary, a beam with initial geometric imperfections undergoes

lateral deformations and torsional rotations throughout the entire course of loading.

Therefore, the vertical orientation of the load applied by the loading mechanism used by

Sant and Bletzacker (1961) is lost at the very early stages of the test of an imperfect

beam. The inclination of the applied load with respect to the vertical axis continuously

increases in the course of loading, introducing greater and greater lateral restraining

forces to the roller mechanism. This lateral-translational restraint affects the experimental

results.

Massey and Walter (1969) tested five small-scale beams with the details given in

Table 1.5. The table also includes the experimental buckling loads of the specimens. The

specimens were simply-supported in plane and out of plane and subjected to a

concentrated load at mid-span. Massey and Walter (1969) used a special method of

loading. The specimens were loaded through a water tank connected to the beam at the

centroid of the mid-span section. Using dead weights hung from the specimen is a proper

method of loading in lateral-torsional buckling experiments for two reasons. First, a dead

weight hanging from the beam travels with the beam and does not induce any lateral-

translational and rotational restraint to the beam at the load application point. Secondly,

the vertical orientation of the dead weight does not change regardless of the rotations in

the beam, since the gravitational forces are always vertical. Loading a beam with dead

weights is also a quite economical and convenient method of loading. Nevertheless, this

method of loading has a limitation preventing it to be applicable to large-scale beam tests,

particularly if water is used as the means of loading. Water has a low unit weight (0.0624

20
kip/ft3). Therefore, large volumes of water are needed to load large-scale beams up to the

failure.

Table 1.5 – Specimens tested by Massey and Walter (1969)

Effective Experimental
Width, b Length, Tension
Specimen Depth, d Buckling
(in.) L (ft) Reinforcement
(in.) Load, Pcr
1 12 1 10 ½ x ½ square bar 3.77
2 12 1 12 ½x½ 3.68
3 15 ¾ 12 1x¼ 2.20
4 15 ¾ 12 ¾ x¼ 1.42
5 12 ¾ 14 ¾ x¼ 0.60

Due to the limitations of the loading method, Massey and Walter (1969) tested

small-scale beams, which are also easier to fabricate and to test, compared to the large-

scale ones. Nevertheless, due to their relatively small lateral-flexural and torsional

rigidities, the experimental results of the beams with smaller scales are more sensitive to

the parameters associated with the test setup (tolerance errors). For instance, restraints

from the loading mechanism, accidental deviations from vertical and eccentricities of the

applied load have more pronounced influences on the behavior and test results of a small-

scale beam, as opposed to a beam with larger scale.

Konig and Pauli (1990) carried out an extensive experimental study in which they

tested six reinforced and prestressed concrete beams. The first five beams were T-shaped

and they were designed in a way that each specimen was distinct from the other four

specimens in an individual parameter, influencing the lateral stability of concrete beams.

The sixth beam was totally different from the other five specimens in cross-sectional

shape, dimensions and reinforcement (Figure 1.14). In the following discussion, the

21
Figure 1.14– Cross-sectional details of the specimens tested by Konig and Pauli (1990)

specimens are introduced by emphasizing the individual parameters whose effects were

examined. Then, the influence of each parameter will be discussed in the light of the

experimental results obtained by Konig and Pauli (1990). The nominal dimensions and

the flexural reinforcement of the test specimens are presented in Table 1.6 and the initial

geometric imperfections are tabulated in Table 1.7. Finally, the experimental buckling

loads and deformations of the specimens are given in Table 1.8.

22
Table 1.6 –Beams tested by Konig and Pauli (1990)

Top
Span Beam
Flange Tension Compression
Specimen Length, Height,
Width, Reinforcement Reinforcement
L (ft) h (in)
b (in)
1 59 10.4 51.2 6 M25 4 M12 & 4 M8
2 59 10.2 51.2 6 M25 4 M12 & 4 M8
3 59 14.2 51.2 6 M25 4 M12 & 4 M8
4 59 10.2 51.2 6 M25 4 M25 & 4 M8
5 59 10.2 51.2 14 M12.5 strands 4 M12 & 2 M12.5
6 84 14.2 53.1 24 M12.5 strands 4 M12 & 4 M8

Table 1.7 – Midspan initial imperfections of the beams tested by

Konig and Pauli (1990)

Angle of twist, φo
Specimen Initial Sweep at Midheight, uo (in.)
(radian %)
1 0.79 0
2 0.12 0.30
3 0.24 0.30
4 0.10 0.15
5 0.63 0.30
6 0.43 0.40

Table 1.8 – Loads, midspan deformations and rotations at failure of the beams

tested by Konig and Pauli (1990)

Mid-span Deformation (in.) Midspan Angle


Critical Load,
Specimen of Rotation
Pcr (kips) Lateral Top Vertical (%)
1 42.7 6.4 4.6 0.5
2 44.5 3.3 2.4 0
3 57.0 5.6 4.6 1.0
4 53.4 1.9 3.7 0
5 45.1 7.2 2.8 0
6 50.9 8.8 5.5 0

23
Specimens 1 and 2 had the base nominal dimensions, reinforcement and cross-

sectional details. They only differed in the initial geometric imperfections. Specimen 3

was identical to the first two specimens, except the width of the top flange. The top

flange of the third specimen was made stockier than the first two specimens by increasing

the breadth from 10.4 in. (25 cm) to 14.2 in. (35 cm). Specimen 4 had heavier

compression reinforcement than the first three beams. The M12 bars in the top flanges of

the first three beams were replaced with M25 bars in the fourth specimen while keeping

the nominal dimensions identical to the first two specimens. Specimen 5 was reinforced

with prestressing strands instead of rebars to examine the influence of prestressing of

reinforcement on the lateral stability of concrete beams. The M8 bars in the top flanges

and the M25 bars in the bottom portions of the first four specimens (Figure 1.14) were

replaced with M12.5 strands in the fifth specimen. Specimen 6 was tested to investigate

the lateral stability of prestressed concrete beams with I-section to observe the influence

of the cross-sectional shape on the stability.

Specimen 2 had a smaller initial lateral deformation, sweep, than the first

specimen. Accordingly, the test results of the second specimen are closer to reflect the

behavior of an initially-perfect beam with the base dimensions. Therefore, the results of

each of the five specimens are compared to the experimental values of Specimen 2

(Figure 1.15) when discussing the influence of an individual parameter on the lateral

stability of concrete beams.

Specimen 1 with the greater sweep (0.79 in. at mid-height of the mid-span

section) failed at an applied load, 4 % lower than the buckling load of Specimen 2, whose

sweep was measured as 0.12 in. Although the midspan sweep of one of the specimens

24
Figure 1.15 – Ratio of the buckling load of each specimen to the buckling load of
Specimen 2

was almost seven times greater than the sweep of the other beam, the reduction in the

buckling load was only 4 %, implying that the influence of the initial imperfections was

not quite significant.

The loading and support conditions reported by Konig and Pauli (1990) suggests

that the top flanges of the specimens were subjected to compressive stresses while the

bottom portions were under tension in the tests. Specimen 3, which had a wider top

flange than the first two beams, buckled under an applied load of 57 kips, which is 28 %

greater than the buckling load of the second specimen. The significant increase in the

failure load depicts the major stabilizing effect of a wider compression flange on the

beam.

25
The test results of Specimen 4 indicated that the stabilizing effect of the

compression reinforcement was two-fold. First, the buckling load increased to 53.4 kips,

corresponding to an increase of 20 % with regard to the second specimen. Secondly, the

lateral-top deflection of Specimen 4 at failure was smaller than the top deflections of the

first three beams. Both the increase in the failure load and the decrease in the lateral-top

deflection were bound up with the increase in the out-of-plane rigidity of the top flange.

The M12 bars in the top flanges of the first three beams were placed 3.5 in. away from

the weak axis of the section (Figure 1.14). Owing to the distance from the minor axis, the

reinforcing bars significantly contributed to the lateral bending rigidity. The use of M25

bars in Specimen 4 in replacement of the M12 bars increased the resistance of the beam

to lateral-torsional buckling by further constraining the top flange from deforming out of

plane.

The buckling load of Specimen 5 was only 1.3 % greater than the buckling load of

Specimen 2. Accordingly, Konig and Pauli (1990) concluded that the stabilizing effect of

prestressing was not as pronounced as the effects of the top flange width and the

compression reinforcement. However, the type of reinforcement was not the only

difference between Specimen 5 and Specimen 2. Specimen 5 had a significantly larger

sweep than Specimen 2. The buckling load of Specimen 5 might have been reduced by

the major sweep, causing the experimental results not to reflect the actual degree of

stabilization provided by prestressing. To evaluate the influence of prestressing on the

lateral stability of concrete beams, more experimental results are needed.

Konig and Pauli (1990) used the loading mechanism illustrated in Figure 1.16 in

their experiments. A water tank was connected to the beam at the one-third points of the

26
Figure 1.16 – Loading mechanism used by Konig and Pauli (1990)

span. Steel plates were attached to the tank. The use of steel plates in addition to the

water tank reduced the need for excessive volumes of water in the tests and enabled the

researchers to use a tank of smaller capacity to attain the buckling loads of the beams.

The water tank was connected to a loading cage through cables. The loading cage

transmitted the load to the top of the beam. A pivot bearing, joining the beam and the

loading cage, provided the beam with the rotational freedom at the point of application of

load.

The loading mechanism used by Konig and Pauli (1990) was clearly superior to

the mechanisms used by Hansell and Winter (1959) and Sant and Bletzacker (1961).

When introducing the loading mechanism used by Massey and Walter (1969), the

27
efficiency of dead weights, hung from the specimen, in minimizing the lateral-

translational and rotational restraining forces at the loading point was discussed. Konig

and Pauli (1990) also overcame the need for a spacious water tank in the setup by using

steel plates. By considering the nature of the dead loads, the test results obtained by

Konig and Pauli (1990) can be considered reliable to be used in the analytical studies.

The beams tested by Konig and Pauli (1990) were simply supported in and out of

plane at the ends. The boundary conditions of a simply-supported beam are explained in

the third chapter of the present text in detail. One of the conditions that need to be

fulfilled to achieve the simple support conditions is the absence of a major restraint from

the supports against the displacement in longitudinal direction. The lateral supports used

by Konig and Pauli (1990) allowed the longitudinal displacements at the beam ends. The

top flanges of the beams were supported laterally through ball-bearings. The bottommost

portions of the beam ends were also supported in lateral direction to preserve the integrity

of the support sections. Sliding pads were placed between the beam and the bottom

supports. The ball bearings at the top and the sliding pads at the bottom minimized the

longitudinal friction forces from the lateral supports and allow the ends to rotate in plane

with no major restraint.

1.4.2 Analytical Methods for Predicting Lateral Torsional Buckling


One of the first investigations on the lateral stability of reinforced concrete beams was

conducted by Marshall (1948). The analytical study aimed at developing critical load

expressions for a laterally-unsupported beam under three different loading conditions:

1. Concentrated load at mid-span;

2. Uniformly distributed load throughout the span;

28
3. Equal and opposite bending moments at the beam ends;

Marshall (1948) obtained the critical load equations (1.1) and (1.2) for the loading cases 1

and 2, respectively. Equation (1.3) gives the critical moment of a beam subjected to the

equal and opposite end moments (loading case 3):

16.93
Pcr  2
 B   GC  (1.1)
L

28.6
wcr  3
 B   GC  (1.2)
L

8.47
M cr   B   GC  (1.3)
L

where Pcr, wcr and Mcr are the critical concentrated load, the critical unit load and the

critical moment of a laterally unsupported beam, respectively; L is the unbraced length of

the beam; B and GC are the out-of-plane flexural and the torsional rigidities of the beam,

respectively. Marshall (1948) proposed the use of the following lateral flexural and

torsional rigidity expressions for rectangular reinforced concrete beams:

b3  d
B  2.5  10 6
 (1.4)
12

b3  d
GC  0.9  10 6
 (1.5)
3

where b and d are the width and the effective depth of the beam, respectively. The

multipliers 2.5x106 and 0.9x106 in Equations (1.4) and (1.5) are the modulus of elasticity

and the modulus of rigidity of concrete, respectively. Marshall (1948) assumed the

29
modulus of elasticity and the modulus of rigidity to be constant for the concrete fibers

throughout the length and depth of the beam at the time of buckling. This assumption

disregards the inelastic lateral-torsional buckling behavior of reinforced concrete beams.

Figure 1.17 illustrates the stress-strain curve of a normal-strength concrete. The moduli of

elasticity corresponding to different stress values are shown on the curve. The first

portion of the curve up to the proportional limit stress (0.4.fc’ for normal-strength

concrete) is linear. The slope of this line represents the initial tangent modulus of

elasticity (Eit), and it is calculated according to Equation (1.6), given in ACI 318-05

(2005) for normal-weight concrete.

Eit  57000  f c' (1.6)

where Eit and fc’ are the initial tangent modulus of elasticity and the compressive strength

of concrete in psi, respectively.

If all the compression fibers throughout the depth and length of a beam are

stressed below the proportional limit (elastic limit in many cases) of concrete at the

instant of buckling, the beam experiences elastic lateral-torsional buckling. In the case of

elastic buckling, the initial tangent modulus of elasticity of concrete provides a good

estimate for the rigidity of all the compression fibers in the beam. The constant modulus

of elasticity value, 2.5x106, proposed by Marshall (1948) corresponds to a concrete

compressive strength of 1920 psi according to Equation (1.6). 1920 psi is a low concrete

strength compared to the compressive strength values encountered in today’s practice.

Therefore, the constant modulus of elasticity value proposed by Marshall (1948) will

30
Figure 1.17 – Stress-strain curve of normal-strength concrete and the tangent moduli of
elasticity at different stress levels (from Nawy 2005)

result in low estimates when computing the load associated with the elastic lateral

torsional buckling of a reinforced concrete beam.

In the case of inelastic lateral-torsional buckling, however, some compression

fibers of the beam are stressed beyond the proportional limit of concrete at the time of

buckling. Figure 1.17 indicates that the slope of the stress-strain curve reduces as the

stress and strain increase beyond the proportional limit [tan(α1)>tan(α2)>tan(α3)].

Particularly, if the concrete is stressed to more than 0.7fc’, it loses its rigidity to a major

extent and the modulus drops drastically vanishing at ultimate stress. Depending on the

stresses reached at the initiation of buckling, the modulus of elasticity values for highly-

stressed fibers at the outermost parts of the compression zone can be significantly lower

than the initial tangent modulus of elasticity of concrete (Eit>Et2>Et3), reducing the

overall modulus of the beam used in the evaluation of the lateral bending rigidity.

31
To summarize, the modulus of elasticity value used in the calculation of the

flexural rigidity about the minor axis varies significantly along the length and depth of

the beam at the instant of buckling. Therefore, a constant value of the elastic modulus

should not be used when computing the lateral flexural rigidity.

The multiplier 0.9x106 in Equation (1.5) is the assumed modulus of rigidity value

of the concrete. By using a Poisson’s ratio of 0.3, this value can be calculated from

Equation (1.7):

E
G (1.7)
2  (1   )

where E and G are the modulus of elasticity and the modulus of rigidity, respectively: ν is

the Poisson’s ratio. Similar to the elastic modulus term in the lateral bending rigidity

expression, Marshall (1948) proposed the use of a constant modulus of rigidity value in

the torsional rigidity calculations, which disregards the inelastic material behavior of

concrete and the variation of the modulus of rigidity along the length and depth of the

beam at the time of buckling.

The use of the effective depth d in Equations (1.4) and (1.5) suggests that

Marshall (1948) assumed that all fibers of the beam from the compression face to the

centroid of the tension reinforcement contribute to the resistance of the beam against

lateral-torsional buckling. Only the portion of the beam below the centroid of the tension

reinforcement is neglected in the critical load calculations. Using d in the critical load

calculations is based on a very general assumption that the concrete above the centroid of

the tension reinforcement remains uncracked until buckling. Depending on the strain

distribution in the tension zone of the beam, the flexural cracks may propagate upward

32
close to the compression zone before buckling, rendering the cracked zone ineffective in

resisting buckling. Particularly, in the case of inelastic lateral-torsional buckling, many

tension fibers in the beam reach strains higher than the cracking strain of concrete in

tension. Therefore, extension of the flexural cracks in the tension zone should be well-

established to determine the portion of the beam providing rigidity against buckling. The

use of d in the critical load calculations may overestimate the portion of the beam

effective in resisting lateral buckling at the time of failure.

Marshall (1948) used uncracked, elastic and homogeneous material assumption in

the critical load calculations. Consequently, the rigidity expressions given in the study do

not reflect the true behavior of reinforced concrete beams, especially if the buckling takes

place close to the ultimate flexural load levels. Marshall (1948) also inferred that the

stability criteria based on L/b ratio only is not factual and the lateral stability of a beam

should be evaluated based on d/b ratio as well as the L/b ratio. The study included the

stability analysis of both singly- and doubly-reinforced concrete beams. Marshall (1948)

did not investigate the effects of initial geometric imperfections on the lateral stability of

reinforced concrete beams.

Hansell and Winter (1959) conducted an analytical study to investigate the lateral

stability of an initially perfect rectangular reinforced concrete beam. In their study,

Hansell and Winter (1959) found that the secant modulus of elasticity corresponding to

the extreme compression fiber strain at the instant of buckling reflects the material

behavior of concrete in the compression zone, and therefore, the secant modulus of

elasticity should be used as the material rigidity term when evaluating the critical

moments of rectangular reinforced concrete beams. Secant modulus of elasticity is the

33
slope of the line on the stress-strain curve connecting the origin to the point

corresponding to the extreme compression fiber strain (Figure 1.18). The modulus of

rigidity used in the assessment of the torsional rigidity of a beam is calculated from the

secant modulus of elasticity according to Equation 1.7.

Figure 1.18 – Secant modulus of elasticity corresponding to the extreme compression

fiber strain

Hansell and Winter (1959) conservatively assumed that the concrete below the

neutral axis is fully cracked at the time of buckling and its contribution to the resistance

to lateral-torsional buckling should be disregarded. By only taking the compression zone

of the section into account, Hansell and Winter (1959) obtained the following lateral

bending and torsional rigidity expressions for rectangular reinforced concrete beams:

34
b3  c
B  Esec  (1.8)
12

Esec  b3  c  b 
2

GC     1  0.35    (1.9)
2  1     3  d 

where c is the neutral axis depth, b is the beam width, d is the effective depth to the

centroid of reinforcement and Esec is the secant modulus of elasticity corresponding to

the extreme compression fiber strain at the instant of buckling.

The use of the neutral axis depth in the equations is appropriate in the case of

inelastic lateral-torsional buckling. When a beam fails in inelastic buckling, many

portions along the depth of the tension zone exceed the cracking strain of concrete.

Hence, the flexural cracks in the tension zone propagate towards the compression zone

rendering an important portion of tension zone ineffective at the time of buckling. On the

other hand, a considerable portion of the tension zone can be still effective in resisting the

instability failure in the case of elastic lateral-torsional buckling, particularly if the

buckling takes place at the early stages of loading when only some fibers in the outermost

portion of the tension zone reach the cracking strain of concrete. Thus, use of the neutral-

axis depth results in low buckling load estimates for slender concrete beams subject to

elastic lateral-torsional buckling.

Another noteworthy detail in the rigidity expressions proposed by Hansell and

Winter (1959) is the use of a constant c value for different sections along the span. When

the in-plane bending moment is constant throughout the span of a beam, the neutral axis

depth (c) of each section along the span is the same. Nevertheless, the neutral axis depths

of different cross-sections of a beam are different in the case of non-uniform in-plane

35
bending moment along the span. For design purposes, Hansell and Winter (1959)

proposed the use of a single c value for the entire beam independent of the in-plane

bending moment distribution in the span. Hansell and Winter (1959) recommended the

use of the rigidity values corresponding to the beam section with maximum bending

moment along the span (for example the midspan section of a beam subjected to a

concentrated load at midspan), since the rigidities in a concrete beam are minimum at the

maximum moment locations.

Siev (1960) identified three different states of a reinforced concrete beam along

the loading history; uncracked elastic, cracked elastic and cracked plastic states. A

different lateral-bending rigidity expression was developed for each state. Nonetheless,

Siev (1960) advocated the use of a single torsional rigidity expression for reinforced

concrete beams independent of the state of the beam at the time of buckling.

In his study, Siev (1960) analyzed a beam simply supported in and out of plane at

the ends and subjected to a concentrated load at mid-span. Under the specified loading

and support conditions, the largest in-plane bending moments occur at mid-span of the

beam, while the end portions of the beam are subjected to minor bending moments.

Therefore, few or no flexural cracks form in the tension zone of the beam near the end

supports. Since the largest torsional moments are resisted by the end portions of the

beam, Siev (1960) included the contribution of the tension zone to the torsional rigidity

and proposed the following rigidity expression:

Ec  b3  h  b 
GC     1  0.63    (1.10)
2  1     3  h 

where h is the overall depth of the beam and ν is the Poisson’s ratio.

36
In the uncracked state, Siev (1960) considered the reinforced concrete as a

homogeneous material and disregarded the contribution of the flexural and shear

reinforcement to the lateral-flexural rigidity. The uncracked flexural rigidity (Bu) is given

by Equation (1.11):

b3  h
Bu   Ec (1.11)
12

The second state of a reinforced concrete beam was identified as the cracked

elastic state. In the cracked elastic state, flexural cracks form and propagate in the tension

zone of the beam while the concrete in the compression zone is still linearly elastic. Siev

(1960) approximated the stress-strain curve of concrete into a linearly elastic and a plastic

portion. The lateral-bending rigidity of the beam in the cracked elastic state was obtained

by dividing the out-of-plane bending moment to the out-of-plane bending curvature

induced by the moment. The curvature of the beam was determined from the stresses and

strains in the cross-section. Siev (1960) considered the fact that the neutral axis of a

cross-section of the beam deviates from horizontal in the presence of biaxial moments,

namely the in-plane and out-of-plane bending moments. Based on a linear stress-strain

relationship in the compression zone of the section and a rotated neutral axis due to the

presence of lateral bending moments in addition to the major-axis bending moments, Siev

(1960) developed the following lateral-flexural rigidity expression for the cracked elastic

state of the beam:

M c  Ec  b 2 bo2 
Bc      (1.12)
c a  6  c 4 d  c 

37
where M is the in-plane bending moment; σc is the extreme compression fiber stress

corresponding to M; bo is the horizontal distance between the centroids of the reinforcing

bars and a is the internal moment arm of the section. As a result of assuming a triangular

stress distribution in the compression zone of the section, a is equal to d-c/3.

The lateral-flexural rigidity in the cracked elastic state (Bc) is a function of the in-

plane bending moment (M) the extreme compression stress (σc) and the neutral axis depth

(c) corresponding to M. Therefore, the rigidity value at the time of buckling can only be

calculated by knowing the critical moment and the stress and strain distributions in the

section corresponding to the critical moment. The evaluation of the critical moment based

on the rigidity expressions given by Siev (1960) requires an iterative approach. First, an

initial value of M is assumed and the lateral-flexural rigidity corresponding to the initial

value of M is calculated. Subsequently, the critical moment is computed from the

calculated lateral-flexural and torsional rigidity values, using Equation (1.13).

C1
M cr   B C (1.13)
C2  L

where C1 and C2 are the constants corresponding to the loading and support conditions of

the beam, respectively. The iterations are then continued until the moment value

converges.

Finally, Siev (1960) proposed a lateral-flexural rigidity expression for the cracked

plastic state of the beam. In the plastic state, some fibers in the compression zone of the

beam are strained beyond the elastic limit of concrete (Figure 1.19). Since an elastic-

perfectly plastic stress-strain behavior was assumed for concrete, a uniform stress

38
distribution is reached within the outermost portion of the compression zone. Siev (1960)

derived the following lateral-flexural rigidity expression for the plastic state of the beam:

b 2  M c p  ce
Bp   (1.14)
12   c  a c  ce
p
2

where cp and ce are the depths of the plastic and elastic portions of the compression zone,

respectively (Figure 1.19); εc is the strain at the extreme compression fibers.

Sant and Bletzacker (1961) also carried out an analytical study on the lateral

torsional buckling of rectangular reinforced concrete beams. The following lateral-

flexural and torsional rigidity expressions were proposed:

b3  d
B  Er  (1.15)
12

Er b3  d
GC   (1.16)
2  1    3

where Er is the reduced modulus of elasticity of concrete corresponding to the extreme

compression fiber strain.

Equations (1.15) and (1.16) are different from the rigidity expressions adopted by

Hansell and Winter (1959) mainly in two aspects. First, Sant and Bletzacker (1961)

assumed that only the concrete above the centroid of the tension reinforcement

contributes to the resistance of a beam against buckling. Therefore, the effective depth (d)

is used in the rigidity expressions instead of the neutral axis depth (c) presuming that the

concrete below the centroid of the tension reinforcement is ineffective at the time of

39
Figure 1.19 – (a) Stress distribution in the compression zone of the beam section in the

plastic state; (b) Stress-strain curve of concrete; assumed by Siev (1960)

(* σe = elastic limit stress)

buckling due to cracking. The portion of a beam effective in resisting lateral torsional

buckling is determined according to the strain distribution in the section and in the span

of the beam at the onset of buckling. The fibers in the tension zone strained beyond the

cracking strain of concrete are not taken into consideration in critical load calculations.

According to the moment levels reached prior to buckling and the cross- sectional and

material properties of the beam, the use of c or d or a value between them might be more

appropriate to account for the effective portion of the beam in buckling resistance.

40
However, the use of c in critical load calculations results in lower buckling load estimates

than the use of d, proposed by Sant and Bletzacker (1961).

In addition, Sant and Bletzacker (1961) argued that the reduced modulus of

elasticity reflects the material rigidity of concrete at the time of buckling. The reduced

modulus of elasticity expression used by Sant and Bletzacker was first derived by

Considère (1891) and Engesser (1895) and later supported by the experimental and

analytical studies on inelastic column buckling by Von Karman (1910). The application

of the reduced modulus theory to lateral torsional buckling is briefly explained in the

following discussion.

In the double modulus theory for lateral-torsional buckling, the strain distributions

along the depth as well as the width of the midspan section of a beam at the time of

buckling are established as in Figure 1.20. For simplification, the beam is assumed not to

experience out-of-plane bending deformations prior to buckling. Accordingly, the beam

is only strained as a result of the in-plane bending deformations at the onset of buckling.

When the beam loses its stability and bends out of plane, the fibers in the concave half of

the section are compressed further. On the other hand, the out-of-plane deformations after

buckling introduce tensile stresses and strains to the fibers in the convex half of the beam.

In other words, the compressive strains resulting from the in-plane and out-of-plane

bending moments add up in the concave side of the compression zone, while the tensile

strains caused by the lateral bending cancel the compressive strains resulting from the

vertical bending in the convex side of the compression zone of the section. The formation

of the tensile strains in the compression zone of the beam is named as strain reversals. As

shown in Figure 1.20, the further loading of the fibers in the concave side of the section

41
Figure 1.20 – Strain distribution at midspan section and the corresponding reduced

modulus of elasticity

takes place along the line tangent to the curve at point A. On the other hand, the

unloading of the fibers in the convex side takes place along a line parallel to the initial

straight portion of the stress-strain curve. Therefore, the modulus of elasticity of the

loading fibers is the tangent modulus of elasticity, Etan corresponding to the compressive

strain of the fibers at the onset of buckling, while the modulus of elasticity valid for the

unloading fibers is the initial tangent modulus of elasticity of concrete, Ec.

The reduced modulus (double modulus) theory is based on the presumption that

the load increase in the beam due to additional compressive strains in the concave side of

the beam is equal to the decrease in the load due to tensile strains developed in the

42
convex side of the beam after buckling. Thus, the load-carrying capacity of the beam is

constant during buckling. Based on the constant load assumption, the following

expression is developed for the reduced modulus of elasticity:

4  Ec  Etan
Er  (1.17)
 
2
Ec  Etan

Since the tangent modulus of elasticity depends on the strain of the fibers at the

onset of buckling, the reduced modulus of elasticity is a function of the strain in the fibers

at the initiation of buckling. Equation (1.17) was developed for the buckling of a column,

subjected to equal concentrated loads at the ends. In column buckling, the axial strain is

assumed to be constant across the width and the length of the column. Therefore, all

fibers in the concave side of the column have the same tangent modulus of elasticity at

the onset of buckling. On the other hand, the strains resulting from the in-plane bending

vary throughout the depth and the length of a beam, subjected to a concentrated load at

mid-span. Consequently, the tangent and reduced moduli of elasticity change along the

depth and the length of the beam. Therefore, the use of a constant tangent modulus of

elasticity along the depth of the concave half of the compression zone does not actually

reflect the material rigidity of the beam at the time of buckling. However, the rigidity and

critical load calculations considering the variations in the reduced modulus of elasticity in

the section and in the span of the beam are not practical and quite time-consuming. Thus,

Sant and Bletzacker (1961) proposed the use of the smallest reduced modulus of

elasticity corresponding to the most-strained compression fibers in the beam, which are

the extreme compression fibers of the midspan section, in the case of midspan loading.

As indicated in Figure 1.20, the smallest reduced modulus of elasticity corresponds to the

43
outermost side of the compression zone of the midspan section because the slope of the

loading line (the tangent modulus of elasticity) reduces as the strain in the fibers

increases.

The double modulus theory makes use of the assumption that strain reversals take

place in the convex part of the compression zone of the beam. The strain measurements

taken by Sant and Bletzacker (1961) validated the presence of the strain reversals, i.e. the

formation of the additional tensile strains, in the convex side of the compression zone

after buckling.

In contrast to the previous researchers, Massey (1967) included the contribution

of the longitudinal reinforcement to the lateral-bending and torsional rigidities and the

contribution of the shear reinforcement to the torsional rigidity of a reinforced concrete

beam and proposed the Equations (1.18) and (1.19):

b3  c
B  Esec   Es  I sy (1.18)
12

1   b12  d1  At  Es
GC  Gc'    b3  h    Gs  Gc'   bs3  ts  (1.19)
3 2 2 s

where h is the height of the section; ΣIsy is the moment of inertia of the longitudinal steel

about the minor axis of the section; bs and ts are the width and the thickness of the

longitudinal reinforcement layer, respectively, as illustrated in Figure 1.21; γ is a constant

defined by Cowan (1953); b1 and d1 are the breadth and the depth of the cross-sectional

area enclosed by a closed stirrup, respectively (Figure 1.21); s is the spacing of the

stirrups; Ao is the cross-sectional area of one leg of the stirrup; β is the coefficient for St.

Venant’s torsional constant; Es and Gs are the modulus of elasticity and the modulus of

44
Figure 1.21 – Definition of the variables in the expressions proposed by Massey (1967)

rigidity of steel, respectively; G’c is the reduced modulus of rigidity of concrete,

calculated according to Equation (1.20):

Esec
Gc'  Gc  (1.20)
Ec

where Ec and Gc are the modulus of elasticity and the modulus of rigidity of concrete,

respectively.

Massey (1967) modified the lateral-bending rigidity expression developed by

Hansell and Winter (1959) by adding the second term, Es.ΣIsy, corresponding to the out-

45
of-plane bending resistance provided by the longitudinal reinforcement. According to

Equation (1.18), the contribution of the longitudinal reinforcement is only of concern if

the lateral-torsional buckling takes place prior to the yielding of the flexural

reinforcement. When the steel yields, its modulus of elasticity becomes zero and the

second term vanishes. After yielding of the flexural reinforcement, only the uncracked

concrete above the neutral axis provides the out-of-plane flexural resistance.

According to Massey (1967), the torsional rigidity of a reinforced concrete beam

is calculated by the summation of the three different rigidity terms given in Equation

(1.19). The first two terms correspond to the contributions of the concrete and the flexural

reinforcement, respectively. The rectangular concrete section and the thin-walled

rectangular reinforcement layer (the gray area in Figure 1.21), consisting of the

longitudinal reinforcing bars, are considered as the two main components of the non-

homogeneous concrete beam. The last term, on the other hand, is the contribution of the

shear reinforcement and is taken into account only if the stirrups are closed.

The inelastic behavior of the concrete is taken into consideration by the use of the

secant modulus of elasticity and the reduced modulus of rigidity in the lateral-flexural

and torsional rigidity expressions, respectively. Similar to Equation (1.18), yielding of the

longitudinal reinforcement nullifies its contribution to the torsional rigidity of the beam.

Stiglat (1991) investigated the agreement of the critical moment predictions

based on an approximate method proposed by Stiglat (1971) with the experimental results

obtained by Konig and Pauli (1990). The approximate method suggests that the critical

moment calculated for an elastic and uncracked concrete beam should be modified using

the stresses at the extreme compression fibers at the onset of buckling to account for the

46
inelastic material properties. To begin with, an initial critical moment value is calculated,

neglecting the inelastic material properties of concrete. The following initial critical

moment expressions were presented for a simply-supported beam with three different

loading conditions: uniformly-distributed load along the span (Equation 1.21), a single

concentrated load at mid-span (Equation 1.22) and equal concentrated loads at one-third

points of the span (Equation 1.23).

3.54  e 2.5  I y  Ix  I y
M cri    1  1.44     E c Gc  I y  J  (1.21)
L  L J 
 Ix

4.23  e 2.5  I y  Ix  I y
M cri   1  1.74     E c Gc  I y  J  (1.22)
L  L J 
 Ix

3.25  e 2.5  I y  Ix  I y
M cri   1  1.44     E c Gc  I y  J  (1.23)
L  L J 
 Ix

where Mcri is the initial (uncorrected) critical moment; Ix and Iy are the moments of

inertia about the major and minor axes, respectively; J is the torsional constant; L is the

unbraced length of the beam; e is the initial vertical distance of the load from the shear

center of the beam section.

The terms in the parenthesis in each equation correspond to the stabilizing or

destabilizing effect of the vertical location of the load with respect to the shear center.

Figure 1.22 illustrates the deflected and undeflected configurations of a beam with a

concentrated load, applied at the top, at the shear center and at the bottom of the cross-

section. In all three cases, the line of action of the applied load passes through the shear

center prior to torsional rotations. When the beam experiences torsional rotations, the line

47
Figure 1.22 – Effect of the vertical location of the applied load with respect to the shear
center of beam section

of action of the load continues to pass through the shear center in case (b). In cases (a)

and (c), on the other hand, torsional rotations in the beam render the applied load laterally

eccentric with respect to the shear center. A load acting above the shear center creates

torsional moments, in the same direction as the existing torsional rotations due to

instability. Therefore, the applied load increases the rotations in the beam, having a

destabilizing effect. On the contrary, the torsional moments induced by the load applied

below the shear center oppose the torsional rotations due to instability. Consequently, a

load acting below the shear center has a stabilizing effect on the beam.

In Equations (1.21) - (1.23), the term “e” is taken positive, when the load acts

above the shear center. For a positive value of e, the expression in the parenthesis is less

than unity, so the critical moment is reduced due to the destabilizing effect of the load.

48
When the load acts below the shear center, on the other hand, e is negative, increasing

the buckling moment in account for the stabilizing influence of the applied load in the

deflected configuration of the beam.

Stiglat (1991) recommended using a value of 60 % of the uncracked torsional

constant of the beam section in Equations (1.21) - (1.23). Using reduced values for the

torsional constant takes into consideration the zones in the beam which are already

cracked at the onset of buckling.

Next, the critical moment calculated based on the elastic material properties

should be corrected for the material nonlinearities of concrete. The correction is done

using the comparative slenderness parameters. First, the stress at the extreme

compression surface of the most-stressed beam section (for instance, the midspan section

in the case of a concentrated load at midspan) is calculated according to Equation (1.24):

eo
 cri  M cri  (1.24)
Ix

where σcri is the stress corresponding to the extreme compression fibers of the most-

stressed section along the span; eo is the vertical distance of the outermost compression

fibers from the centroid of the section. The comparative slenderness parameter is

obtained using σcri in Equation (1.25):

Ec
v    (1.25)
 cri

where λv is the comparative slenderness parameter. The slenderness parameter, defined by

Stiglat (1991), is a function of the modulus of elasticity of concrete and the maximum

49
compressive stress in the beam. The initial critical moment, calculated with the modulus

of elasticity and modulus of rigidity corresponding to the initial portion of the stress-

strain curve of concrete below the proportional limit stress, is corrected using λv to

account for the reduced modulus of elasticity, valid for the fibers stressed beyond the

elastic limit of concrete. Later, an equivalent stress value, σT, is obtained from the

comparative slenderness parameter using tables presented by Stiglat (1991). Finally, the

critical moment of the beam is calculated by correcting the initial critical moment

according to Equation (1.26):

T
M cr  M cri  (1.26)
 cri

Stiglat (1991) reported that the analytical critical moment values according to the

proposed approximate method were in close agreement with the experimental results

obtained by Konig and Pauli (1990). Moreover, Stiglat (1991) stated that the use of the

reduced torsional constant in Equations (1.21) - (1.23) resulted in conservative critical

moment predictions since all analytical critical moment values were smaller than the

experimental buckling moments.

Revathi and Mennon (2006) modified the effective moment of inertia expression

in ACI 318-05 (2005) Section 9.5.2.3 for the case of out-of-plane bending. The original

form of the expression is given in Equation (1.27):

 M cra 
3
  M 3 
Ie     I g  1      I cr  I g
cra
(1.27)
 Ma    M a  

50
where Ig, Icr and Ie are the uncracked, the cracked and the effective moments of inertia,

respectively; Ma is the maximum moment in the span at the particular applied load level;

Mcra is the cracking moment of the beam.

Equation (1.27) is the weighted average of the uncracked and cracked moments of

inertia of a concrete beam. The uncracked moment of inertia corresponds to the early

stages of loading when the cracking moment of the beam is not exceeded and the entire

beam section contributes to the in-plane bending resistance. When the cracking moment

is exceeded, flexural cracks form in the outermost layers of the tension zone of a beam.

Later, the flexural cracks propagate in the tension zone towards the compression zone and

the moment of inertia of the beam decreases as the applied load increases. When the

flexural cracks render the entire tension zone ineffective, the moment of inertia reaches a

minimum limit, called the cracked moment of inertia. Equation (1.27), which is the

moment of inertia of a concrete beam when the maximum moment in the span is Ma,

reflects the variation in the moment of inertia of a concrete beam from the uncracked

state to the fully cracked state as the flexural cracks propagate in the tension zone.

The out-of-plane bending rigidity expression proposed by Revathi and Mennon

(2006) is given in Equation (1.28):

3
 M cra   b3  h  
    
 0.8  M ult   12  
B  Ec    (1.28)
   M cra    b3  cu
3
 Es  
 1   0.8  M     12     E  I sy   
   ult     c  

where Mult is the ultimate flexural moment of the beam; cu is the neutral axis depth of the

beam at ultimate flexural load; ΣIsy is the moment of inertia of the longitudinal

51
reinforcement about the minor axis; ψ is a multiplier, which is taken 0 for under-

reinforced beams and 1 for over-reinforced beams. The first term in the

parenthesis, b3 h 12 , is the uncracked moment of inertia of a reinforced concrete beam

about the minor axis. The term b3cu 12      Es / Ec  I sy  is the moment of inertia of a

concrete section at the ultimate flexural load. Accordingly, Revathi and Mennon (2006)

proposed a lateral bending rigidity at the time of buckling, which is a weighted average of

the uncracked moment of inertia and the moment of inertia of the beam at ultimate load.

Equation 1.27 uses the maximum moment in the span at a particular load to

average the uncracked and cracked moments of inertia. In the case of lateral bending, the

maximum moment at the time of buckling is the buckling moment of the beam. Hence,

the buckling moment should be known to calculate the out-of-plane flexural rigidity of

the beam at the buckling moment. To avoid an iterative procedure, Revathi and Mennon

(2006) proposed the use of the lateral bending rigidity corresponding to 80% of the

ultimate flexural moment. Although the use of 0.8.Mult is a close approximation in the

case of inelastic lateral-torsional buckling, it can underestimate the lateral bending

rigidity in the case of elastic lateral torsional buckling. When the buckling moment is

much smaller than 0.8.Mult as in the case of elastic lateral torsional buckling, there are less

flexural cracks in the beam and the rigidity of the beam at the instant of buckling is

significantly greater than the value calculated from Equation (1.28).

Revathi and Mennon (2006) also proposed the use of initial tangent modulus of

elasticity, Ec, in the rigidity calculations. As shown in Figure 1.17, the initial tangent

modulus of elasticity is used when concrete is in the elastic range of stress-strain curve.

Since all compression fibers throughout the beam are stressed below the elastic limit, the

52
use of Ec is appropriate when the beam undergoes elastic lateral-torsional buckling. In the

case of inelastic lateral-torsional buckling, nonetheless, some compression fibers in

highly-stressed portions of the beam are strained beyond the elastic range, where the

modulus of elasticity of concrete is smaller than Ec. Consequently, Ec is not applicable for

all compression fibers in the beam, buckling inelastically, contrary to the assumption in

Equation (1.28).

Finally, Revathi and Mennon (2006) proposed the following torsional rigidity

expression:

4   '  Es  A22  Ac
C (1.29)
 1 1 
p22    
 l t 

where Ac is the area of the gross cross-section of the beam; A2 and p2 are the area and the

perimeter of the rectangle connecting the centers of the corner longitudinal bars (Figure

1.23); μ’ is a rigidity multiplier taken as 1.2 for under-reinforced and 0.8 for over-

reinforced sections; ρl and ρt are the volumetric ratios of the longitudinal and transverse

reinforcement, respectively, calculated from Equations (1.30) and (1.31):

As
l  (1.30)
Ac

At  p1
t  (1.31)
Ac  s

where As is the area of the longitudinal reinforcement in the cross-section; At is the cross-

sectional area of one leg of a stirrup; p1 is the perimeter of the centerline of a stirrup

(Figure 1.23); s is the spacing of the stirrups.

53
Figure 1.23 – Definition of area and perimeters in Equations (1.29) - (1.31)

The torsional rigidity of a reinforced concrete beam before torsional cracking is

given by Equation (1.32) according to St. Venant’s theory.

C  Gc    b3  h (1.32)

where β is the coefficient for St. Venant’s torsional constant, obtained from Equation

(1.33):

192 b  1 (2n  1) h


  1    tanh (1.33)
 5
h n 0 (2n  1) 5
2b

The torsional rigidity expressions, adopted by Hansell and Winter (1959) (Equation 1.9),

Siev (1960) (Equation 1.10) and Sant and Bletzaker (1961) (Equation 1.16), all

correspond to the uncracked stage of a concrete beam, and they are derived from

54
Equation 1.32. Equation 1.29, on the other hand, is the torsional rigidity of a reinforced

concrete beam in the early post-cracking stage, meaning right after the formation of

diagonal tension cracks due to torsion.

The post-cracking torsional rigidity of a concrete beam is provided by the outer

thin-walled layer of concrete surrounding the corner longitudinal bars and stirrups. A

three-dimensional model called the thin-walled tube, space truss model was used by

Lampert (1973) and Hsu (1973) to develop the post-cracking torsional rigidity of

rectangular reinforced concrete beams. Later, Tavio and Teng (2004) simplified the

rigidity expression developed by Hsu (1973) and proposed a new expression. Equation

(1.29) is a modified version of the torsional rigidity expression proposed by Tavio and

Teng (2004), which is presented in Chapter V (Equation 5.27).

Equation (1.29) is the torsional rigidity of a concrete beam right after cracking.

Using the equation in critical moment calculations suggests that the whole concrete beam

is cracked diagonally at the time of buckling. In the presence of lateral supports at the

beam ends, torsional moments resulting from the out-of-plane deformations increase

from zero at midspan to a maximum value at the ends. Therefore, the middle portion of

the beam remains diagonally uncracked throughout the loading while the end portions,

under significant torsional moments, are cracked to a major degree at the time of

buckling, particularly in the case of a geometrically-imperfect beam, experiencing major

lateral deformations prior to buckling. The uncracked torsional rigidity of the beam

reasonably reflects the torsional resistance of the middle portion of the beam, yet the ends

possess torsional rigidities close to or even smaller than the post-cracking torsional

rigidity. Adopting the post-cracking torsional rigidity for the whole beam in the

55
calculations is overly-conservative while using the uncracked torsional rigidity

overestimates the resistance of a beam, leading to unsafe results. If a single torsional

rigidity expression is desired to be valid for the whole span, it should be an average of the

maximum and minimum values of the torsional rigidity of the beam along the span.

1.4.3 Summary

The analytical methods for predicting lateral torsional buckling loads of reinforced

concrete beams, presented in Section 1.4.2, considered the elastic-inelastic stress-strain

behavior of concrete and steel, the contribution of the longitudinal and shear

reinforcement to the stability and the flexural cracking of concrete. Nevertheless, the

influences of the initial geometric imperfections and the restrained shrinkage cracking of

concrete on the lateral stability of reinforced concrete beams have not been studied yet.

The analytical part of this study aimed at incorporating all the factors, including the

initial geometric imperfections and restrained shrinkage cracking of concrete, into the

formula.

For an exact analysis of the lateral stability of a reinforced concrete beam, the initial

geometric imperfections, the initial cracking condition (presence or absence of shrinkage

cracks), the experimental stress-strain curves of concrete and steel and the cross-sectional

details of the beam should be fully known. In none of the experimental studies presented

in Section 1.4.1, all of the aforementioned properties of the test specimens were reported.

Therefore, reinforced concrete beams, whose geometric and material properties are fully

known, were tested in the present study for a better evaluation of the analytical methods

presented in Section 1.4.2 and the method proposed in the present study.

56
CHAPTER II

SPECIMENS AND MATERIAL PROPERTIES

2.1 Specimens

2.1.1 Specimen Descriptions

In the experimental program, two sets of specimens were tested. The first set of

specimens was composed of six beams of four types, B36, B30, B22 and B18. The

second set of beams consisted of five beams of two different types, B44, B36L. Table 2.1

presents the specimens of the entire experimental program. Figure 2.1 and Figure 2.2

illustrate the nominal cross-sectional details of the specimens.

Each beam is denoted with the letter “B”, followed by two numbers. The first

number corresponds to the depth of the specimen in inches, while the second number is

used for the identification of the specimen. For instance, B44-1 corresponds to the first of

the identical beams having a depth of 44 inches. Additionally, specimen group B36L has

the letter “L” (representing the longer span) to distinguish it from the specimen B36.

Companion beams were identical to each other in dimensions and amount of

flexural and shear reinforcement. Furthermore, concrete from the same batch was used in

companion beams to minimize the influence of the mechanical properties of concrete on

the experimental results. Similarly, reinforcing steel of the companion beams was from

the same batch with the exception of specimen groups B22 and B18. Flexural reinforcing

bars in specimens B22-1 and B18-1 were Grade 60 while the bars in specimens B22-2

and B18-2 were Grade 40 (ASTM A615/A, 2007). Since Beams B22 and B18 buckled

57
Table 2.1 – Specimens tested in the experimental program

Specimen Number of Height, h Width, b Span Length, d/b L/b


Group Samples (in) (in) L (ft) ratio ratio
B36 1 36 2.5 20 12.45 96
B30 1 30 2.5 20 10.20 96
B22 2 22 1.5 12 12.45 96
B18 2 18 1.5 12 10.20 96
B44 3 44 3.0 39 12.45 156
B36L 2 36 3.0 39 10.20 156

before yielding of flexural reinforcement, grade of the reinforcing bars had no influence

on the buckling behavior of the beams.

Shear reinforcement was needed in the specimens to prevent shear failure. Due to

the small widths of the specimens, welded wire reinforcement (WWR) sheets were used

instead of bent reinforcing bars. Two 2x6-W2.5xW3.5 sheets, one on each side of the

flexural reinforcement, constituted the shear reinforcement of each specimen (Figures 2.1

and 2.2).

2.1.2 Experiment Design


The first set of tests was carried out to evaluate the performance of the experimental

setup. Thus, any potential shortcomings in the loading and support systems could be

discovered and corrected before the second set of experiments. Another goal of the first

set of tests was to observe the lateral-torsional buckling behavior of reinforced concrete

beams and to detect the factors affecting the lateral stability. Therefore, the first set of

specimens was designed to be quite slender so that the beams would certainly fail by

lateral-torsional buckling.

58
Figure 2.1 – First set of specimens (B36, B30, B22, and B18)

Beams B30 and B36 were tested in the first stage of the experimental program.

They had similar dimensions and cross-sectional details to the beams tested by Sant and

Bletzacker (1961), whose experimental work formed the basis of the slenderness

limitation specified in Section 10.4 of ACI 318-05 (2005) together with the experimental

study carried out by Hansell and Winter (1959). To understand scale effects, four smaller

beams of two types, B22 and B18 were tested in the first stage.

59
Figure 2.2 – Second set of specimens (B44, B36L)

In the first stage of the experimental program, several observations were made

leading to the design of the second set of specimens. The second set of beams was

constructed at a larger scale than the first set for two reasons. First, tests on B22 and B18

demonstrated that small-scale beams were extremely sensitive to experimental errors,

such as eccentricities in the applied load and deviations from vertical in the orientation of

60
the load. The smaller lateral-flexural and torsional rigidities of small-scale beams cause

the experimental results of such beams to be excessively influenced by the accidental

torsions resulting from the slight eccentricities and deviations of the applied loads. As the

scales of the specimens were increased, the slight tolerance errors became less influential.

Secondly, all specimens were constructed in the same lab environment, using the same

type of materials. Therefore, the initial geometric imperfections of the specimens of

different sizes were of the same order of magnitude. Since large-scale beams were

expected to be less affected than the small-scale beams by imperfections of the same

order of magnitude, testing beams with greater scales was preferred in the second stage of

the experimental program.

All specimens were designed to undergo elastic lateral torsional buckling, so that

the influences of the factors other than inelasticity on the lateral stability of reinforced

concrete beams could be examined. Elastic lateral torsional buckling of reinforced

concrete beams takes place when both concrete and reinforcement in the beams are

strained in the elastic portions of their respective stress-strain curves. In a simply-

supported beam, subjected to a concentrated load at midspan, the extreme compression

fibers at midspan are the most-stressed portion of the concrete beam. If the extreme

compression fiber strain at initiation of buckling is within the initial elastic portion of the

stress-strain curve of concrete, the entire compression zone of the beam behaves

elastically at initiation of buckling. Similarly, the strain in the tension reinforcement at

initiation of buckling does not reach the yield strain of steel in elastic lateral torsional

buckling of reinforced concrete beams.

61
2.2 Concrete Material Properties
The small dimensions and congested reinforcement (Figure 2.3) in the first set of beams

rendered the mechanical vibration of concrete difficult. To overcome the consolidation

problems, Self-Consolidating Concrete (SCC) was used in the first set of specimens. SCC

is a flowable type of concrete which spreads into the form and consolidates under its own

weight (Figure 2.4). The high-range water-reducing (HRWR) admixtures in SCC

decrease the viscosity of concrete and eliminate the need for mechanical vibration. The

spread of SCC was measured as 25 in. according to the slump flow test, described in

ASTM C1611 (2005). The SCC used a 3/8-in maximum size aggregate.

Mechanical properties of concrete and reinforcing steel influence the lateral

buckling behavior of reinforced concrete beams significantly. Concrete from the same

batch and reinforcing bars from the same batch of steel were used in the companion

beams to reduce differences.

For the concrete used in the first set of beams, three 6 in. x 12 in. cylinder samples

were tested on the 7th day, on the 28th day and on each test day to obtain the compressive

strength of concrete (f’c) according to ASTM C39-05 (2005). Furthermore, three more

cylinder tests were conducted on each day to determine the modulus of elasticity (Ec) and

the Poisson’s Ratio (υc) of the concrete according to ASTM C469 (2002). Different from

the first set of beams, cylinder tests were only conducted on the test days in the second

set of beams. Table 2.3 tabulates the means and the standard deviations of the test results

of each test day.

62
Figure 2.3 – Congested reinforcement in B36

Figure 2.4 – Application of self-consolidating concrete

63
Table 2.2 – Mechanical properties of concrete

Age f’c (psi) Ec (ksi) υc


Test at Test
Day Sample μ1 Sample μ Sample μ
(days) Size Size Size
B22-1 119 3 11730 3 5200 3 0.16
B22-2 129 3 11000 3 4850 3 0.17
B18-1 145 3 11460 3 5000 2 0.13
B18-2 160 3 11320 3 5000 3 0.16
B30 220 3 12220 3 5950 3 0.20
B36 249 3 12780 3 5850 3 0.17
B44-1 179 3 8470 3 4450 3 0.16
B44-2 225 3 8540 3 4450 3 0.15
B44-3 234 3 8560 3 4550 3 0.14
B36L-1 192 3 7900 3 4300 3 0.15
B36L-2 201 3 7940 3 4500 3 0.15
1
Sample Mean

In critical and ultimate bending moment calculations, the stress-strain curves of

concrete used in the specimens were needed. Therefore, compression tests were

conducted on 6 in. x 12 in. concrete cylinders to determine the experimental stress-strain

curves of concrete. Figure 2.5 to Figure 2.7 illustrate the experimental stress-strain curves

of concrete, obtained from cylinder tests. Critical moment and ultimate flexural moment

calculations of the specimens are simplified if the stress-strain curve of concrete is

expressed in a mathematical form. For this purpose, several analytical models for the

stress-strain curve of high-strength concrete were examined.

Analytical stress-strain curves from the models proposed by Carreira and Chu

(1985), Tomaszewicz (1984) and Wee et al. (1996) were included in the plots to

determine the model giving the best agreement with the experimental stress-strain curves.

64
Carreira and Chu (1985) proposed Equation (2.1) for the stress-strain relationship

high-strength concrete.

   
    

f c  f c'    o  
(2.1)
 
  1     
   
 o 

where ε and fc are the concrete strain and stress, respectively; εo is the strain at peak stress

and f’c is the compressive strength of concrete according to the cylinder tests; β is a

material parameter, given by

1
 (2.2)
1  f  o  Ec
c
'

The model proposed by Tomaszewicz (1984) adopts Equation (2.1) for the

ascending branch of the stress-strain curve. For the descending branch of the curve, on

the other hand, Equation (2.3) was developed with the introduction of a new parameter, k

to Equation (2.1).

   
    

f c  f c'    o  
(2.3)
k  
  1     
   
 o 

where k = f’c/2.90 with f’c given in ksi.

Similar to the formulation given by Tomaszewicz (1984), Wee et al. (1996)

recommended the use of Equation (2.1) for the ascending branch and a modified form of

Equation (2.1), given below, for the descending branch.

65
   
 k1      

f c  f c'    o  
(2.4)
k2  
 k   1    
 1   
 o 

where k1 = (7.26/f’c)3.0 and k2 = (7.26/f’c)1.3 with f’c given in ksi.

All three models adopt the same equation (2.1) for the ascending branch of the

stress-strain curve. Figure 2.5 to Figure 2.7 indicate that Equation (2.1) closely estimates

the ascending portions of the experimental curves. Since the stresses in all specimens

were within the initial portions of the stress-strain curves of concrete, only the ascending

portions of the curves were determined in the cylinder tests.

66
Figure 2.5 –Stress-strain curves of concrete in the first set of beams

Figure 2.6 –Stress-strain curves of concrete in B44

67
Figure 2.7 –Stress-strain curves of concrete in B36L

68
CHAPTER III

EXPERIMENTAL SET-UP, INSTRUMENTATION AND


PROCEDURE

3.1 Experimental Set-up

3.1.1 Loading Mechanism

The loading frame used for testing the beams of the present study consisted of a loading

mechanism, called the gravity load simulator, a tension jack mounted to the center pin of

the simulator, a loading cage and a ball-and-socket joint conveying the load from the

cage to the beam (Figures 3.1 and 3.2).

Gravity load simulator was first developed by Yarimci et al. (1967) and used in

sway-permitted testing of large scale frames, later in lateral torsional buckling tests of

steel I-beams by Yura and Phillips (1992) and lateral stability of polymer composite I-

shaped members by Stoddard (1997). In the present experimental program, the gravity

load simulator designed and applied by Stoddard (1997) was used.

The gravity load simulator is composed of two inclined arms and a rigid

triangular frame connected to the arms through pins. The pin connections at both ends of

the arms cause the mechanism to be unstable. The instantaneous center of the mechanism

at any configuration of the simulator is the intersection point of the extensions of the

inclined arms (Figure 3.3). The center pin (bottom pin) of the triangular frame moves in

an approximately horizontal line for certain limits of mechanism motion. Since the center

pin is directly below the instantaneous center at any configuration of the simulator, the

line of action of the load applied by a loading device connected to the center pin has a

vertical orientation passing through the instantaneous center. The applied load has

69
Figure 3.1 – Undeflected and deflected configurations of the loading frame and loading
cage

insignificant deviations from the vertical orientation in a certain range of lateral

displacement of the center pin. In the present study, a hydraulic cylinder mounted to the

center pin of the triangular frame loaded the beams vertically throughout the entire test

and did not restrain the out-of-plane translation of the loading point owing to the lateral

motion of the center pin of the simulator with the beam (Figure 3.4).

The rotational freedom of the loading point was achieved with the help of the

ball-and-socket joint (Figure 3.5), which was composed of two steel plates and a steel

70
Figure 3.2 – Loading frame

ball between them, positioned in a socket. Bottom plate of the joint, which was epoxied

to the top surface of the beam, rotated with the beam allowing the loading cage (Figure

3.6) and the hydraulic cylinder to preserve their vertical orientation. Consequently, the

applied load, transferred from the loading cage to the beam by the steel ball, continued to

have a vertical orientation even after the rotations of the beam. The socket was lightly

oiled prior to each test to diminish the friction between ball and plates.

71
Figure 3.3 – Undeflected and deflected shapes of the gravity load simulator

(a) (b)

Figure 3.4 – (a) Undeflected and (b) Deflected configurations of the gravity load
simulator

72
Figure 3.5 – Ball-and-socket joint

73
(a) (b)

Figure 3.6 –The vertical orientation of the loading cage (a) before the test; (b) after buckling

74
3.1.2 Supports

In the design of the experimental setup, the in-plane and out-of-plane supports were

selected and designed to obtain simple support conditions about the major and minor axes

of the beam. The end supports allowed rotations about the major and minor axes (Figure

3.7) while restraining rotation about the longitudinal axis of the beam (Figure 3.8).

Furthermore, the end supports restrained the in-plane (vertical) and out-of-plane (lateral)

translations (Figure 3.8), yet allowed longitudinal translation and warping deformations.

Figure 3.7 – Minor-axis rotations at the supports

75
Figure 3.8 – Lateral deformations and torsional rotations restrained at the supports

Longitudinal deformations of the beams played an important role in the design of

in-plane and out-of-plane supports. Restraining the displacements of the support sections

in longitudinal direction changes the behavior of a beam completely. When the

longitudinal displacements of a beam are prevented at the support locations, the in-plane

support conditions deviate from simple support conditions.

When analyzing beams subjected to in-plane loading, through one-dimensional

models, the supports are located at the centroids of the support sections [Figure 3.9(a)]. In

76
* C – Centroid of the Mid-pan Section of the Undeflected Beam

C’ – Centroid of the Mid-pan Section of the Deflected Beam

Figure 3.9 – In-plane support conditions: (a) in analysis models;


(b) Hinged-hinged; (c) Roller-hinged; (d) Roller-roller.

the following discussion, the centerline of the beam is assumed to be coincident with the

neutral axis. According to this assumption, the centerline continues to be unstrained as

the beam deforms in plane. In a simply-supported beam, the support sections rotate in

plane about the line of contact of the support with the beam. In a beam with support

77
conditions as in Figure 3.9(a), the end sections rotate about the horizontal centroidal axis

(major axis) of the section. Therefore, the portion above the centroidal axis displaces

longitudinally towards the mid-span of the beam while the portion below the centroidal

axis displaces outwards. In other words, the top portion of the beam shortens while the

bottom portion elongates due to the in-plane bending rotations. The centroids of the end

sections remain at their initial positions maintaining the longitudinal distance between

them. Furthermore, the centerline and the initial midspan of the beam do not undergo

translations in longitudinal direction. The in-plane flexural deformations of the beam take

place symmetrically about the mid-span of the undeflected configuration of the beam.

The in-plane support conditions are somewhat different in an experiment. The

supports are located underneath the beam. Using rollers or hinges or a combination of

them at the support locations influences the in-plane behavior and deformations of a

beam when the supports are beneath the beam. Figures 3.9(b)-(d) illustrate the in-plane

deformations of a beam with hinges at both ends, a hinge at one end and a roller at the

other end and rollers at both ends, respectively.

A beam exhibits completely different in-plane flexural behaviors under the three

different boundary conditions shown in Figures 3.9(b)-(d). The three cases are

investigated to find the boundary conditions under which a beam has an in-plane flexural

behavior in closest agreement with the case considered in the one-dimensional analysis

[Figure 3.9(a)].

In Figure 3.9(b), both ends of the beam are supported with hinges. Since the

hinges at the ends restrain the longitudinal displacement, the bottommost portions of the

end sections remain at their original positions as the beam bends in plane. The end

78
sections undergo major-axis rotations about the supports and the centroids of the end

sections displace towards mid-span. The longitudinal displacements of the centroids of

the end sections constitute a clear distinction from the in-plane flexural deformations of

the beam supported as in Figure 3.9(a). Hence, providing fixtures simulating hinges at

both ends was not adopted in the experiments.

In Figure 3.9(c), the beam is hinge-supported at one end and roller-supported at

the other end. The roller support translates in longitudinal direction while the hinge

remains in its original position when the beam bends in plane. The roller-supported end

of the beam experiences longitudinal translations as well as flexural rotations; yet the

hinged end of the beam only rotates in-plane about the support. As the beam ends

undergo bending rotations, the roller translates in longitudinal direction to assure that the

initial distance between the centroids of the end sections is preserved. The in-plane

flexural behavior of a beam with a hinge at one end and a roller at the other end matches

with the behavior in Figure 3.9 (a) in the preservation of the initial longitudinal distance

between the centroids of the end sections. Nevertheless, there are differences between the

in-plane deformations of a beam supported as in Figure 3.9(c) and a beam supported as in

Figure 3.9(a). The midspan section of the beam in Figure 3.9(c) translates in the

longitudinal direction towards the roller-supported end as the beam flexes. Hence, the

location of the midspan load also shifts longitudinally with the beam, constituting a

significant difference from the behavior of the beam supported at the centroids of the end

sections.

In Figure 3.9(d), both ends of the beam are roller-supported. The beam is

statically unstable since there is no restraining force at the support locations preventing

79
the beam from undergoing rigid-body translation in the longitudinal direction. However,

the lack of the longitudinal restraining force at the beam ends was not significant in the

present study since the longitudinal displacements and in-plane bending deformations of

the specimens were estimated to be small due to the large major-axis bending rigidities

possessed by the slender beams. Furthermore, the loading frame prevented significant

longitudinal translation of the beams. The bending rigidity of the loading frame would

provide a longitudinal force at the load point, restraining the tilting of the loading frame

due to the longitudinal translation of the beam.

Figure 3.9(d) depicts that the roller supports at both ends of the beam translate in

longitudinal direction, allowing the stretched bottom portions of the end sections to

displace outwards as the beam bends in plane. Since the centroids of the end sections do

not displace in longitudinal direction, the initial longitudinal distance between the

centroids of the end sections is maintained as the beam ends rotate in plane. Unlike a

beam with a hinge at one end and a roller support at the other end, the centerline of a

beam with roller supports at both ends does not move in longitudinal direction.

Based on the above discussion, the in-plane flexural behavior of a beam with

roller in-plane supports at both ends is closest to the behavior of a beam with the support

conditions as in a one-dimensional analysis. Therefore, each specimen was roller-

supported at both ends to have similar support conditions to a one-dimensional analysis

(Figure 3.10).

Lateral stability of a beam is also influenced by the longitudinal restraint at the

lateral supports. Figure 3.11 illustrates the two lateral support conditions in the aspect of

longitudinal restraint. Lateral support (shown with crosses in the figure) was provided at

80
Figure 3.10 – Roller supports at the beam ends in the second set of experiments

five points along the depth of the beam ends. In Figure 3.11 (a), lateral supports allowed

free translation in longitudinal direction while preventing the beam from deflecting in

lateral direction. On the other hand, lateral supports in Figure 3.11 (b) restrain the

longitudinal displacements as well as the lateral displacements at the beam ends. Hence,

lateral supports are shown as rollers in Figure 3.11 (a) and as hinges in Figure 3.11 (b).

If the lateral supports prevent the ends from rotating in plane by restraining the

longitudinal displacements, the beam ends become fixed rather than simply-supported.

Since the lateral stability of a beam is closely related to its in-plane flexural behavior, the

longitudinal restraining forces at the lateral supports should be minimized to achieve

simple support conditions in and out of plane.

In the present experimental program, out-of-plane supports were designed in a

way that the points in contact with lateral supports were allowed to translate in

longitudinal direction with insignificant levels of restraint. In that way, the beam ends

were provided with rotational freedom about the major axis.

81
Figure 3.11 – Behavior of the beams with: (a) roller supports; (b) hinged supports
in lateral direction

Two different support frames were built in the two stages of experimental

program to achieve the aforementioned lateral support conditions. In the first stage, ball

rollers were employed to support the beams laterally (Figure 3.12). A ball roller [Figure

3.12(b)] is a special type of caster, whose wheel is a steel ball capable of swiveling freely

in a socket. The rotational freedom of the ball allows free motion in any direction. The

82
use of ball rollers in the first set of experiments assured that the points on the beam in

contact with the lateral supports were not restrained from translating in longitudinal

direction. So, the lateral supports provided the support sections of the beams with in-

plane rotational freedom to achieve the simple support conditions. The ball rollers were

mounted to the support frames through threaded studs [Figure 3.12(b)].

Although the ball rollers were observed to prevent the beam ends from rotating

about the longitudinal axis of the beam and from deflecting laterally, the support forces

transferred from the beam to the ball rollers were noticed to cause the threaded rods of

the ball rollers to bend during the tests (Figure 3.13). Therefore, a new lateral support

frame (Figure 3.14) was designed prior to the second set of experiments. Rigid casters

(Figure 3.15) were used instead of the ball casters.

Figure 3.12 – (a) Support frame in the first set of experiments; (b) A ball roller in contact
with the beam

83
Figure 3.13 – Bending of the ball roller

Figure 3.14 – Lateral support frame in the second set of experiments

84
Figure 3.15 – Rigid caster in contact with the specimen

The rigid casters had a wheel rotating about an axle passing through the center of

the wheel. The casters were mounted to the lateral support frames in a way that the wheel

rotations allowed longitudinal displacements of the points of contact of the beams with

the casters (Figure 3.15). Therefore, the in-plane flexural rotations of the end sections

were not restrained to satisfy the simple support conditions about the major axis.

The casters had a mounting plate with four corner holes to bolt the caster to a

frame. Instead of bolting the casters directly to the support frame, the mounting plate of

each caster was connected edge to edge to a steel plate adjacent to the other side of the

frame (Figure 3.15) to allow the casters to move to the desired level along the height of

the frame to accommodate different beam depths. The four ½-in diameter bolts

connecting the casters to the support system provided adequate rigidity to the casters

85
against the bending moments induced by the vertical friction forces between the beams

and the caster wheels.

The support frames in the second set of experiments were mainly composed of

two HSS 3x3x1/4 structural tubes, one on each side of the beam (Figure 3.14). Each of

these tubes was supported by two diagonal knee braces. One of these braces was

extended to the top of the support member (HSS 3x3x1/4) while the other brace was

connected to the tube at one-third of the height of the tube.

In the first test of the second stage (Specimen B44-1), two casters were used on

each side of the beam to support the beam ends laterally [Figure 3.16(a)]. One of the

casters supported the topmost portion of the beam while the other caster was touching the

beam at the two-third of the height. Although two casters had sufficient capacity to

withstand the lateral forces in the tests, problems associated with deformations and

distortions at the beam ends were encountered. Since lateral support was provided at the

top halves of the beam ends only, the bottom portions of the ends displaced in the

opposite direction to the lateral deformations, after buckling (Figure 3.17). The top

portions, on the contrary, remained in their initial positions owing to the adequate lateral

support at the top. Displacement of the bottom parts of the end portions relative to the top

resulted in distortions in the cross-sectional shape of the beam. Figure 3.17 illustrates the

distortion. The distortions in the support regions did not affect the buckling moment and

the deformations in the beam prior to buckling since the bottom parts of the end sections

started moving laterally as a result of the excessive out-of-plane deformations in the post-

buckling stage. Two additional casters on each side, supporting the bottom halves of the

beam ends were used in the following experiments [Figure 3.16(b)]. The beam ends were

86
supported by four casters on each side of the beam to provide lateral translational and

rotational restraint along the depth of the beam.

Figure 3.16 – Support frame in the second set of tests: (a) B44-1; (b) B44-2

Figure 3.17 – Distortion in the cross-section at the beam end

87
3.1.3 Load, Deflection and Strain Measurements

A 50-kip load cell in the first set of experiments and a 100-kip load cell in the

second set of experiments were used. The load cells were placed in line with the jack and

loading cage in order to measure the applied load. String potentiometers were utilized to

determine the in-plane and out-of-plane deflections, the torsional rotations and distortions

at midspan.

In a lateral-torsional buckling test, the in-plane deflections of a beam are

accompanied by out-of-plane deflections. Each point along the span of a slender beam

undergoes lateral displacements as well as vertical displacements. Therefore, the cable of

a potentiometer, having horizontal or vertical orientation at the beginning of the test,

deviates from its initial orientation once the beam starts deforming out of plane (Figure

3.18). Since a geometrically imperfect beam deflects out of plane and experiences

rotations even at the initial stages of loading, uncoupled lateral and vertical deflections

cannot be measured directly by using horizontally- and vertically-oriented potentiometers

even prior to the buckling of the beam.

In the first set of experiments, the coupled deflection measurements from the

potentiometers were converted into in-plane and out-of-plane deflections and rotations at

the shear center (centroid in rectangular sections), through a modified approach presented

by Zhao (1994) and extended by Stoddard (1997). This approach, presented in Appendix

E, is based on geometric relations linking the deflection measurements from three

potentiometers to deflections and rotation of the centroid.

88
Figure 3.18 – Deviation of the Initial Orientations of the Potentiometers

In the second set of experiments, the distances of the lateral and vertical

potentiometers to the beam were increased to minimize the coupling between the lateral

and vertical deflection readings. This is illustrated in Figure 3.19, which shows that the

angle of the measuring cable from horizontal (vertical in the case of a vertical string

potentiometer) in the twisted configuration of the beam decreases as the distance between

the potentiometer and the beam increases. Accordingly, the difference between the

horizontal component of the measuring cable (L.cosα in Figure 3.19) and the length of

the measuring cable (L) decreases with an increase in the distance of the potentiometer to

the beam. This means that the change in length of the measuring cable, measured by the

89
Figure 3.19 – Coupling between the in-plane and out-of-plane deflection measurements

for a lateral string potentiometer with varying distances from the specimen

potentiometer, becomes closer to the lateral deflection of the cable-attachment point on

the beam, as the distance of the potentiometer increases.

The cross-section of a concrete beam might distort when the beam buckles in a

lateral torsional mode. As shown in Figure 3.20(a), two lateral potentiometers are

adequate to determine the rotated configuration of the midspan section after buckling

when the cross-section of a beam does not distort. Nonetheless, distortions in the cross-

section of a beam cannot be detected by only measuring the out-of-plane deflections at

two different depths along the midspan section. Therefore, lateral deflections were

measured at three or more different points along the depth of each specimen at midspan

[Figure 3.20(b)] to assess the shape of the midspan section throughout the test and to

detect any possible distortion in the cross-section. Three lateral potentiometers in the first

90
Figure 3.20 – Lateral-torsional buckling (a) with; (b) without distortions in the cross-
sectional shape of the beam

stage of the experimental program (Figure 3.21) and five lateral potentiometers in the

second stage of the experimental program (Figure 3.22) were used.

During the first stage of the experimental program, Linear Variable Differential

Transducers (LVDT’s) were used for obtaining the strain distributions through the depth

of the convex and concave faces of each specimen at midspan (Figure 3.23). These

LVDT’s were replaced with electrical resistance strain gauges (Figure 3.24) during the

second stage of the experimental program.

91
Figure 3.21 – Lateral deflection potentiometers in the first set of experiments

Figure 3.22 – Lateral deflection potentiometers in the second set of experiments

92
Figure 3.23 – Strain measurement using LVDT’s in the first set of tests

3.1.3.1 LVDT Strain Measurements

Each LVDT was placed in an aluminum box glued to the side face of the specimen

(Figure 3.23). The extension rod of the LVDT core (armature) was attached to an

aluminum plate bonded to the side face of the specimen. The initial longitudinal distance

between the box and plate was the gage length, over which the strain was measured. As

93
Figure 3.24 – Strain measurement through electrical resistance strain gauges in the
second set of tests

the beam bent in and out of plane, the longitudinal distance between the box and plate

changed, causing the armature to slide inside the LVDT tube. The strain was calculated

from the slide of the armature. Nevertheless, it was found out that the slide of the

armature was not equal to the change in the longitudinal distance between the box and

plate. The out-of-plane bending deformations in the beam caused the extension rod,

connecting the armature to the plate, to bend and lose its initial straightness, which

94
caused the measurement taken by the LVDT to be different from the axial elongation or

shortening of beam at the LVDT location. Therefore, LVDT’s were not used for

measuring the strains in the second set of experiments.

3.1.3.2 Strain Measurements through Electrical Resistance Strain Gauges

In the first test of the second stage of the experimental program (Specimen B44-1),

the longitudinal strains from in-plane and out-of-plane bending and the depthwise strains

from the possible distortions in the cross-section of the beam were measured through

two-element cross strain gauges, attached to the side faces of the beam at mid-span

(Figure 3.25). The strain gage oriented in the depthwise direction was used for detecting

the possible distortions in the cross-section of the beam. Three-element rosettes were not

needed since the longitudinal and depthwise strains were estimated to be the principal

strains due to negligible shear stresses from shear forces and torsional moments at

midspan.

Strain was measured at five points along the depth of the beam (Figure 3.26) to

determine the strain distributions on the convex and concave faces of the beam. Appendix

C presents the longitudinal strain distributions along the convex and concave faces of the

second set of specimens at midspan. The depthwise strains measured at midspan of B44-1

are also given in the appendix.

The strain measurements in the first test (B44-1) indicated that the depthwise

strains did not reach significant levels prior to buckling. Therefore, in the remaining tests

individual gauges, measuring the longitudinal strains only, were used instead of cross

gauges (Figure 3.26).

95
Figure 3.25 - 2-element cross strain gauge on the side face of B44-1

In the test of B44-1, strain measurements in the tension zone were greatly

influenced by flexural cracking. Cracks which formed directly under the gauges caused

the measured strain values to be extremely high. To measure the tensile strains in the

remaining tests, the strain gauges on the tension side of the beam were installed on

aluminum strips, which were attached to the face of the beam by means of concrete drop-

in anchors and bolts (Figure 3.27) to prevent the slip of the strips during the tests. The

strain gauges installed on the aluminum strips measured the average tensile strain

between the two points, where the strip was attached to the side face of the beam.

96
Figure 3.26 – Longitudinal strain gauges along the depth of north face of specimen B44-2
at midspan

Consequently, the tensile strain measurements were not affected from the flexural

cracking in the tension zone. Full-bridge strain gauge circuits (Figure 3.27) composed of

two transverse and two longitudinal gages were installed on some of the aluminum strips

to cancel the accidental bending strains in the strips and to measure the axial strains only.

97
Figure 3.27 –Electrical resistance strain gauges on an aluminum strip for measuring the
longitudinal strain in the tension zone

3.2 Test Procedure

The beams were positioned on their sides during the construction stage. After the

concrete was set, each specimen was tilted into the vertical position and moved to the test

setup through a special lifting method which is explained in Appendix G. The sweep and

initial twisting angles of the specimens were measured prior to the tests (Appendix B).

The beams were loaded to failure. To detect the experimental cracking load of

each specimen and to explore the extension of in-plane flexural cracks, loading was

stopped at every 1-to-2 kip load increment at the initial stages of the test. Once the rate of

increase in the out-of-plane deflections and torsional rotations became large, the beams

were loaded to failure without interruption.

98
3.3 Summary of the Test Results

All specimens of the present experimental program failed in lateral torsional

buckling. Table 3.1 presents the experimental buckling loads of the specimens and the

centroidal lateral and vertical deflections and the torsional rotations at midspan at the

instant of buckling.

In all specimens, the typical crack pattern of lateral torsional buckling, which is

explained in Chapter VII, was observed. The experimental load-lateral (out-of-plane)

deflection, load-vertical (in-plane) deflection and torque-twist curves of the specimens

are presented in Appendix D. The midspan strain distributions of the beams throughout

the test are presented in Appendix C.

The tests indicated that the initial geometric imperfections, sweep (initial lateral

deflection) in particular, significantly influence the lateral stability of a reinforced

concrete beam. Section 6.3 explains the effects of the initial geometric imperfections on

the lateral stability and load-lateral deflection behavior of reinforced concrete beams in

the light of the results of the present experimental program.

99
Table 3.1 – Experimental Results of the Specimens

Buckling Lateral Vertical Angle of


Specimen Load Deflection at Deflection at Twist
(kips) Centroid (in.) Centroid (in.) (degrees)
B18-1 12.4 1.12 0.37 1.17
B18-2 12.0 1.18 0.35 0.45
B22-1 8.7 2.06 0.24 1.66
B22-2 11.0 1.44 0.22 0.90
B30 22.0 1.82 0.48 0.86
B36 39.2 0.39 0.40 0.52
B44-1 15.2 2.81 0.55 0.77
B44-2 12.0 2.12 0.48 0.55
B44-3 21.0 2.58 0.78 0.66
B36L-1 13.5 2.82 0.84 0.70
B36L-2 21.7 1.48 1.37 0.65

100
CHAPTER IV

LATERAL BENDING RIGIDITY OF RECTANGULAR


REINFORCED CONCRETE BEAMS AND INFLUENCE OF
SHRINKAGE CRACKING ON THE RIGIDITY

4.1 Introduction

Rigidity of a beam against bending moments about the minor axis is termed as the lateral

(or out-of-plane) bending rigidity. Lateral bending rigidity is the product of two

variables: (1) the second moment of area about the minor axis of the section (Iy); and (2)

the modulus of elasticity (E), reflecting the overall material resistance of the beam at the

initiation of buckling. Determination of the lateral bending rigidity of a reinforced

concrete beam is not straightforward due to the variation in Iy and E as loading

progresses. The flexural cracks in a beam render some portions of the beam ineffective in

resisting flexural moments. Therefore, sectional response of a concrete beam to lateral

bending (Iy) is not constant throughout the test. Secondly, the stress-strain behavior of

concrete is linear and elastic only up to the elastic limit, assuming that the proportionality

limit of concrete is equal to the elastic limit. If a reinforced concrete beam or some

portions of it is strained beyond the elastic limit of concrete, the material response of the

beam cannot be reflected through E and another modulus of elasticity should be used to

account for the inelastic material behavior of concrete. Accordingly, the lateral bending

rigidity expression proposed for reinforced concrete beams should take into account the

elastic-inelastic material behavior of concrete, the non-homogeneous nature of a

reinforced concrete beam, and the reduction due to cracking in the cross-sectional area of

the beam providing the bending rigidity.

101
Different lateral bending expressions for reinforced concrete beams in the

literature are summarized in Section 1.3. In the following section, only the lateral bending

rigidity expressions that are used in the analysis of the experimental results are explained

in more detail. In Section 4.3, the lateral bending rigidity expression proposed in the

present study is presented. The proposed rigidity expression is developed based on

modeling a reinforced concrete beam with a system of springs. Section 4.3 also discusses

the spring system models used by Bischoff (2007) and Bischoff and Scanlon (2007) in

explaining the differences between the effective moment of inertia expressions proposed

by Branson (1963) and Bischoff (2005). Finally, influence of restrained shrinkage

cracking on the lateral bending rigidity of reinforced concrete beams is explained in

Section 4.4, where the lateral bending rigidity expression, proposed in the present study,

is modified to account for the effect of shrinkage cracks. Furthermore, factors that

promoted the formation of the shrinkage cracks in the first set of beams and the measures

taken to prevent shrinkage cracking in the second set of beams are also discussed in

Section 4.4.

4.2 Available Lateral Bending Rigidity Expressions

In the analytical study, four different lateral flexural rigidity expressions were used in

addition to the rigidity expression proposed in the present study. The first expression is

the lateral flexural rigidity of a homogeneous and elastic beam:

b3  h
Beh  Ec  (4.1)
12

102
where Ec is the elastic modulus of concrete; b and h are the width and height of the beam,

respectively.

Equation (4.1) takes into account the contribution of the entire cross-section of a

beam to lateral bending rigidity and therefore neglects the reduction in the bending

rigidity due to flexural cracking. Furthermore, the use of Ec in the equation reveals that

the entire beam is assumed to be stressed in the elastic range of the stress-strain curve of

concrete, up to buckling, which is true in elastic lateral torsional buckling of reinforced

concrete beams only. In the case of inelastic buckling, Ec does not represent the overall

material rigidity of concrete in a beam at the time of buckling.

Although Equation (4.1) neglects the contribution of the reinforcement to the

bending rigidity, ignores the reduction in the rigidity due to the presence of flexural

cracks and considers the elastic buckling only, the equation was included in the analytical

study to determine the influence of flexural cracking, reinforcement and inelastic material

behavior of concrete on the lateral stability of concrete beams.

Equations (4.2), (4.3) and (4.4) are the lateral bending rigidity expressions

proposed by Hansell and Winter (1959), Sant and Bletzacker (1961) and Massey (1967),

respectively. The expressions were previously presented in Section 1.3 of Chapter I.

b3  c
Bhw  Esec  (4.2)
12

b3  d
Bsb  Er  (4.3)
12

b3  c
Bm  Esec   Es  I sy (4.4)
12

103
where Bhw, Bsb, Bm are the lateral flexural rigidities according to Hansell and Winter

(1959), Sant and Bletzacker (1961) and Massey (1967), respectively; c is the neutral axis

depth of the midspan section of a beam at the initiation of buckling; ; d is the effective

depth of the beam; Es is the elastic modulus of the reinforcing steel; ΣIsy is the moment of

inertia of the longitudinal reinforcing bars about the minor axis of the section; Esec and Er

are the secant and reduced modulus of elasticity of concrete corresponding to the extreme

compression fiber strain at the instant of bifurcation, respectively. Er is calculated from

Equation (4.5):

4  Ec  Etan
Er  (4.5)
 
2
Ec  Etan

where Etan is the tangent modulus of elasticity of concrete corresponding to the extreme

compression fiber strain at the instant of buckling.

Equations (4.2)-(4.4) offer different approaches to account for the possible

inelastic material behavior of concrete at the instant of buckling, by proposing the use of

different types of modulus of elasticity (Esec and Er). Equations (4.2) and (4.4) account

for the destabilizing effect of flexural cracks, by considering the minor axis moment of

inertia of the compression zone only. Finally, Equation (4.4) accounts for the contribution

of the longitudinal reinforcement to the lateral bending rigidity through the use of the

second term on the right hand side of the equation (Es.ΣIsy). Equations (4.1)-(4.4) are

included in the study to compare the results obtained from these equations to the lateral

bending rigidity values obtained from the rigidity equation proposed in the present study.

104
4.3 Proposed Lateral Bending Rigidity Expression

The proposed lateral bending rigidity expression was developed through spring models.

The idea of using springs in modeling the bending behavior of beams originated from the

works of Bischoff and Scanlon (2007) and Bischoff (2007), who used spring models to

justify the effective moment of inertia (Ie) expression developed by Bischoff (2005). For

a better understanding of the spring model used in the present study, the effective

moment of inertia expression proposed by Bischoff (2005) and the spring model

corresponding to this expression is explained in the following discussion.

Prior to the formation of flexural cracks in the tension zone of a beam, the entire

cross-section of the beam contributes to the moment of inertia, which is obtained from

Equation (4.6) by also considering the contribution of the flexural reinforcement:

2
1  h
  b  h3  b  h   y     n  1  As   d  y 
2
I ucr (4.6)
12  2

where As is the cross-sectional area of the flexural reinforcement; n is the modular ratio of

steel to concrete; y is the depth of the center of gravity of the transformed section from

the top surface of the beam. When calculating the uncracked moment of inertia, Iucr, the

flexural reinforcement is transformed into an equivalent concrete area in accordance with

the modular ratio of steel to concrete, n. The gross moment of inertia of a concrete beam

(Ig), on the other hand, is calculated from Equation (4.7):

1
Ig   b  h3 (4.7)
12

105
The contribution of the flexural reinforcement to the moment of inertia can be neglected

and Iucr can be simplified to Ig in reinforced concrete beams with low reinforcement

ratios.

When the bending moment at a cross-section reaches the cracking moment (Mcra),

flexural cracks form in the outermost layers of the tension zone. As the bending moment

increases, the flexural cracks propagate upwards, rendering a greater area in the tension

zone ineffective in resisting bending. Therefore, moment of inertia of the section

decreases as loading progresses and the moment of inertia reaches a minimum limit,

called the cracked moment of inertia (Icr) in serviceability limits. Icr is calculated from

Equation (4.8):

1
 b  c 3  n  As  d  c 
2
I cr  (4.8)
12

where c is the neutral axis depth when all fibers in the compression zone are stressed

below the elastic limit of concrete.

Bending moments exceeding Mcra result in discrete cracks along the length of a

concrete beam. The difference in the moments of inertia of the cracked parts and the

uncracked parts of a beam causes variation in the flexural rigidity along the span.

Concrete between the discrete cracks contributes to resist the tensile stresses in the beam

and increases the overall flexural rigidity. The tensile contribution of the concrete

between the cracks is called tension stiffening. Formation of discrete flexural cracks

along the span and tension stiffening raise a gradual transition of the moment of inertia of

a beam from the uncracked moment of inertia (Iucr) to the cracked moment of inertia (Icr),

as the applied moment (Ma) increases beyond Mcra. The gradual transition in the post-

106
cracking stage was taken into account by Branson (1963), who proposed an effective

moment of inertia expression, which is a weighted average of the moment of inertia of

the gross cross-section (Ig) and the moment of inertia of the fully cracked transformed

cross-section (Icr):

 M cra 
3
  M 3 
I eb     I g  1      I cr
cra
(4.9)
 M   M  
a   a

where Ieb is the effective moment of inertia according to Branson (1963); Ma is the

maximum bending moment along the span; and Mcra is the cracking moment of the beam.

Equation 4.9 is the effective moment of inertia expression recommended in ACI 318-05

(2005) Section 9.5.2 to compute the immediate vertical deflections of reinforced concrete

beams.

Branson’s (1963) expression concerning the effective moment of inertia is an

empirical equation, which is based on test results of simply-supported rectangular

reinforced concrete beams with a reinforcement ratio of 1.65 %. Later, Bischoff (2005)

found that Equation (4.9) overestimates the effective moment of inertia of concrete

beams with low steel reinforcement ratios (ρl<1%) and concrete beams reinforced with

fiber-reinforced polymer bars. Using the tension stiffening strain approach, Bischoff

(2005) was able to develop the following alternative effective moment of inertia

expression:

 M cra  1   M cra   1
2 2
1
    1     (4.10)
I ebi  M a  I g   M a   I cr

107
Equation (4.10) is different from the expression of Branson (1963), which is an

average of the rigidities of the uncracked and cracked portions of a beam. Bischoff’s

(2005) effective moment of inertia was developed through averaging the flexibilities of

the uncracked and cracked parts.

According to Bischoff and Scanlon (2007), the difference between Branson’s

(1963) effective moment of inertia (Ieb) and Bischoff’s (2005) effective moment of inertia

(Iebi) can be explained through spring models. Ieb, which is the weighted average of the

rigidities, can be obtained by modeling the uncracked and cracked parts of a beam

through springs in parallel [Figure 4.1(a)]. Iebi, nevertheless, is obtained by modeling the

uncracked and cracked portions of a beam with springs in series [Figure 4.1(b)].

Figure 4.1 (b) illustrates that springs in series carry the same load (applied

moment, Ma in this case), whereas the parallel connection of springs in Figure 4.1(a)

implies that the load is distributed to the springs in accordance with their rigidities. A

discrete crack in the span and an uncracked portion right adjacent to it are subjected to

approximately the same bending moment and therefore modeling the uncracked and

cracked parts of a beam with springs in series is more appropriate.

Later, the experiments carried out by Gilbert (2006) on simply-supported

rectangular one-way slabs revealed that the sectional resistance of reinforced concrete

flexural members with low reinforcement ratios (ρl<1%) was overestimated by Branson’s

(1963) effective moment of inertia expression, while Bischoff’s (2005) effective moment

of inertia expression produced immediate vertical deflections in close agreement with the

experimental deflections of the specimens.

108
(a)

(b)
Figure 4.1- Spring models defining (a) Branson’s (1963); (b) Bischoff’s (2005) effective
moment of inertia expression

For the present study, Figures (4.2) to (4.6) compare the experimental vertical

deflections of the second set of specimens (B44 and B36L) with the analytical values

calculated using Branson’s (1963) and Bischoff’s (2005) effective moment of inertia

expressions.

According to Bischoff and Scanlon (2007), the analytical deflection estimates

based on both Equations (4.8) and (4.9) are in close agreement with each other when the

steel reinforcement ratio of a concrete beam is above 1%. Specimens B44 and B36L had

109
Figure 4.2 – In-plane deflections of B44-1 at midspan

Figure 4.3 – In-plane deflections of B44-2 at midspan

110
Figure 4.4 – In-plane deflections of B44-3 at midspan

Figure 4.5 – In-plane deflections of B36L-1 at midspan

111
Figure 4.6 – In-plane deflections of B36L-2 at midspan

reinforcement ratios of 2.5% and 2.8%, respectively. Although in Figures 4.5 and 4.6, the

analytical deflection curves corresponding to Ieb and Iebi are only slightly different from

each other, Figures 4.2 to 4.4 reveals that Bischoff’s (2005) effective moment of inertia

expression (Iebi) produces closer estimates to the experimental values.

The experimental curves of B36L-1 and B36L-2 in Figures 4.5 and 4.6 are in

close agreement with the analytical curves corresponding to Ieb and Iebi. Nevertheless, the

experimental curves of B44-1, B44-2 and B44-3 in Figures 4.2 to 4.4 do not show a good

agreement with the analytical curves due to the significant differences between the

experimental cracking moment values of the specimens (Table 4.1) and the cracking

moment values obtained from Equation (4.11), which was used for obtaining the

analytical curves corresponding to Ieb and Iebi:

112
M cra 

I ucr  7.5  f c'  (4.11)
h y

where 7.5  f c' is the modulus of rupture of normal-weight concrete, given in ACI 318-

05 (2005) Section 9.5.2.3.

Table 4.1 – Experimental and calculated cracking moments of the second set of
specimens

Specimen Experimental Mcra Calculated Mcra


(in-kips) (in-kips)
B44-1 470 870
B44-2 530 870
B44-3 480 870
B36L-1 420 610
B36L-2 540 600

Previously, Bischoff and Scanlon (2007) and Bischoff (2007) showed that spring

models well represent the in-plane bending behavior of reinforced concrete beams.

Accordingly, the lateral bending behavior of reinforced concrete beams can also be

represented by a spring system, if the contributions of different portions of a concrete

beam can be reasonably evaluated.

The proposed spring model for the lateral bending behavior of reinforced concrete

beams makes use of the reduced modulus theory [Considère (1891) and Engesser

(1895)]. Here, the reduced modulus theory and its use in the proposed model are

explained in detail with the help of Figure 4.7.

The proposed model is based on a geometrically perfect beam, which does not

experience lateral deformations and torsional rotations prior to bifurcation. Figure 4.7 (c)

113
Figure 4.7 – Moduli of elasticity corresponding to the fibers in the compression zone of a
beam section

is the longitudinal strain distribution along the depth of a cross-section of the beam before

buckling. The longitudinal strains in the pre-buckling stage of the beam solely originate

from in-plane bending moments. The compressive strain varies linearly from zero at the

neutral axis to maximum (εco) at the extreme fibers and the strain at an arbitrary depth y

from the compression face of the beam is denoted as εc. Figure 4.7(d) is the stress-strain

curve of concrete in compression. Since the longitudinal strain is not constant along the

depth of the compression zone, compression fibers at different depths are at different

points on the stress-strain curve. For instance, Point A on the curve corresponds to the

fibers at a depth y while Point B corresponds to the outermost fibers. When bifurcation

takes place, the concave part of the section is subjected to additional compressive strains

from lateral bending while the convex part is subjected to tensile strains, as shown in

114
Figure 4.7(a). The longitudinal strain from out-of-plane bending increases from zero at

the minor axis, which is the vertical centroidal axis in symmetric sections, to maximum

(εtl and εcl) at sides. Tensile strains from lateral bending cause the fibers on the convex

side of the compression zone to be unloaded, while the additional compressive strains

result in further loading of the compression fibers on the concave side of the section.

Figure 4.7(d) illustrates that unloading of the compression fibers takes place along a line

parallel to the initial linear portion of the stress-strain curve of concrete. In other words,

the elastic modulus, Ec is valid for all unloading fibers in the compression zone,

independent of the longitudinal strain (εc) of a fiber prior to buckling. The further loading

of a compression fiber at an arbitrary depth y, on the other hand, takes place along a line

tangent to the stress-strain curve of concrete at Point A. Since the slope of the line

tangent to the curve changes along the stress-strain curve, the tangent modulus of

elasticity corresponding to the loading fibers changes along the depth of the compression

zone of the section.

Hansell and Winter (1959) analytically showed that the secant modulus of

elasticity corresponding to the extreme compression fiber strain (εco) should be used as

the material rigidity term if the entire compression zone of a section continues to be

loaded after buckling. Secant modulus of elasticity (Esec) corresponding to the extreme

compression fiber is the slope of the line connecting Point B on the stress-strain curve to

the origin O as shown in Figure 4.7 (d). The origin of the stress-strain curve corresponds

to the fibers at the neutral axis depth, which have zero longitudinal strain at the initiation

of buckling. Point B, on the other hand, corresponds to the most-stressed fibers of the

compression zone. Therefore, the line connecting Point B to the origin represents the

115
entire compression zone if all compression fibers of the section are further loaded in the

post-buckling stage.

In the present study, the compression zone of a section is divided into a loading

and an unloading portion after buckling, according to the reduced modulus theory. The

secant modulus of elasticity, Esec corresponding to the extreme compression fiber strain at

the instant of bifurcation is used as the modulus of the loading part of the compression

zone.

The spring model proposed in the present study is shown in Figure 4.8. A

reinforced concrete beam is composed of uncracked and cracked parts along the length.

Each of the uncracked and cracked portions of the beam is partitioned into a loading and

an unloading segment when buckling takes place. Applied moment is distributed to the

loading and unloading segments of a portion. Since a cracked portion along the span and

an uncracked portion adjacent to it bear approximately the same lateral bending moment,

the cracked and uncracked parts of the beam are modeled with springs in series. The

loading and unloading segments of a cracked or an uncracked portion of the beam

contribute to the resistance of the lateral bending moments in accordance with their

flexural rigidities about the minor axis of the beam section. Hence, loading and unloading

segments of a portion are modeled with springs in parallel.

In an uncracked section, the entire section contributes to resistance to the minor-

axis bending moments. Therefore, k1 and k2 contain the term h. In a cracked section, on

the other hand, concrete below the neutral axis is assumed not to contribute to the flexural

rigidity of the beam due to the flexural cracks in the tension zone of the section.

116
Figure 4.8 – Proposed spring model for the lateral bending behavior of reinforced
concrete beams

Therefore, the neutral axis depth of the section at the initiation of buckling (c) is used in

k3 and k4.

In the rigidity expressions k1, k2, k3 and k4, b/2 was adopted as the width of each

of the loading and unloading segments of a section. The widths of the loading and

unloading segments are equal if the secant modulus of concrete (Esec) corresponding to

the extreme compression fiber strain at the initiation of buckling is equal to the elastic

modulus of concrete (Ec). Having Esec equal to Ec is possible only if the entire beam is

stressed in the linear elastic range of concrete, which is known as the elastic lateral-

torsional buckling. In the case of inelastic lateral-torsional buckling, nevertheless, Esec

can be much lower than Ec, causing the widths of the loading part and the unloading part

of a section to be different. However, b/2 was used in the equations to simplify the

lateral-flexural rigidity expression.

117
The equivalent rigidity (keq) of the spring system in Figure 4.8 is obtained from

Equation (4.12) by also using the weight factors for the cracked and uncracked parts of a

beam, previously employed by Branson (1963) and Bischoff (2005):

1  M cra 
m
1   M m  1
    1   cra    (4.12)
keq  M cr  k1  k2   M cr   k3  k 4
 

where Mcra and Mcr are the cracking moment and the critical moment of a beam,

respectively. Using the expressions for k1, k2, k3 and k4, given in Figure 4.8

1  M cra 
m
1   M m  1
   3 3
 1   cra    3 3
(4.13)
keq  M cr  h b hb   M cr   E  c  b  E  c  b
Esec   Ec  sec c
24 24 24 24

After simplifications, the lateral flexural rigidity of a reinforced concrete beam, keq is

obtained from Equation (4.14):

b3  c Esec  Ec
keq   m
(4.14)
24 M  c 
1   cra     1 
 M cr   h 

ACI 318-05 (2005) suggests the use of a value of 3 for the power m in Equation

4.9 to obtain an average rigidity for the entire span of a reinforced concrete beam with

discrete cracks along the span. Bischoff (2005), on the other hand, stated that a value of

m=2 in his effective moment of inertia expression (Equation 4.10) correlates well with

Branson’s original equation. In the present study, the spring system (Figure 4.8) models

the cracked and uncracked portions of a concrete beam with springs in series, similar to

118
the spring model used by Bischoff (2005). Therefore, using a value of m=2 was assumed

to be more appropriate.

The lateral bending rigidity in Equation (4.14) can be formulated as the product of

the modulus of elasticity of concrete and the effective moment of inertia of the beam

about the minor axis, leading to Equation (4.15):

 
 3 
  b c  1    Esec  Ec 
keq    (4.15)
 12  M   c   
m 
2


 1   cra    h  1
  M cr    

The expression in the square brackets in Equation (4.15) is the effective moment of

inertia of the beam about the minor axis. The remaining part of the equation, on the other

hand, is the overall modulus of elasticity of concrete (Eo) of the beam, calculated from

Equation (4.16):

 E  Ec 
Eo   sec  (4.16)
 2 

Considering the contribution of the longitudinal reinforcement to the lateral

bending rigidity is meaningful if two criteria are satisfied. First, the longitudinal rebars in

a beam should remain unyielded till the buckling moment to contribute to the lateral

bending resistance. Secondly, longitudinal rebars should be located close to the sides of

the beam to increase the lateral distance from the minor axis, which constitutes the

moment arm of the bars in lateral bending. Previously, the contribution of the

longitudinal rebars to the lateral bending rigidity was taken into account by Massey

(1967) and by Revathi and Mennon (2006). The second term in Equation (4.4), proposed

119
by Massey (1967), and the term ψ.((Es/Ec).ΣIsy in Equation (1.29), proposed by Revathi

and Mennon (2006) correspond to the longitudinal reinforcement. In the spring model

employed in the present study, rigidity contributions of the longitudinal rebars to the

lateral bending rigidities of the uncracked and cracked sections of a beam can be

represented by a spring connected in parallel to the other two springs of each of the

cracked and uncracked parts of the beam. In other words, the number of parallel springs

in each of the cracked and uncracked portions should be increased to three if the

contribution of the longitudinal reinforcement is desired to be included in the rigidity

expression. Accordingly, the lateral bending rigidity expression is modified, giving

Equation (4.17):

keq  
   
 Esec  Ec   h  b3 24  Es  I sy    Esec  Ec   c  b3 24  Es  I sy 
   (4.17)
   
 Esec  Ec   b3 24  c   M cra M cr   1   M cra M cr   h   Es  I sy

m m

where Es is the elastic modulus of the reinforcing steel; ΣIsy is the total moment of inertia

of the longitudinal reinforcing bars about the minor axis of the section. When the

longitudinal reinforcement yields prior to buckling, Es becomes zero and Equation (4.17)

reduces to Equation (4.15). Similarly, the contribution of the longitudinal reinforcement

(EsΣIsy) vanishes and the equation simplifies to Equation (4.15) if the longitudinal

reinforcing bars are located along the minor axis of the beam section (ΣIsy=0). In the

specimens of the present experimental program, for example, the longitudinal rebars were

located along the vertical centroidal axis of the beam, which coincides with the minor

axis in the case of elastic lateral-torsional buckling. Therefore, Equation (4.15) was used

in the critical moment calculations of the beams, buckling elastically.

120
4.4 Influence of Shrinkage Cracking on the Lateral Bending Rigidity

Shrinkage is defined as the volume change in a concrete member due to the loss

of water arising from the difference in relative humidity between concrete and the

surrounding environment. If a concrete member is allowed to shrink freely, it will

experience longitudinal deformations. If the shrinkage deformations of a beam are

restrained, on the other hand, tensile stresses develop in the beam, resulting in cracking of

concrete. Restrained shrinkage cracks may reduce the lateral bending resistance of a

concrete beam considerably.

Influence of shrinkage restraint stresses on the flexural rigidity of concrete beams

was studied by Scanlon and Bischoff (2008), who stated that shrinkage restraint stresses

in a beam reduce the cracking moment. Scanlon and Bischoff (2008) proposed the use of

a reduced effective cracking moment in Equation (4.10) in the presence of restrained

shrinkage cracks in a beam. The reduced cracking moment value proposed by Scanlon

and Bischoff (2008) is equal to 2/3 of Mcr. Scanlon and Bischoff (2008) also stated that

the influence of shrinkage cracks on the flexural rigidity becomes more pronounced when

a beam has a low longitudinal reinforcement ratio (ρl<0.8%).

Considering the influence of the shrinkage cracks on the cracking moment,

Equation (4.14) can be modified, leading to Equation (4.18):

 
 3 
  b c  1    Esec  Ec 
keq     (4.18)
 12   
2
 M cra   c  
  2


 1      1 
  M cr   h  

121
where ω is equal to 1 in the absence of restrained shrinkage cracks in a beam and ω is

equal to 2/3 in the presence of restrained shrinkage cracks.

Shrinkage restraint comes from several sources. For instance, free shrinkage of a

beam can be prevented by the structure (slab, beams) surrounding the beam. Longitudinal

reinforcement in a beam and the formwork in the construction stage also have restraining

effects on the shrinkage deformations of a beam.

In the first set of specimens of the present experimental program, shrinkage

cracking of concrete was observed (Figure 4.9). To overcome the cracking problem in the

second set of beams, the reasons for the formation of the shrinkage cracks were

investigated. According to the investigation, the use of self-compacting concrete (SCC)

instead of the conventionally vibrated ordinary concrete (OC) in the first set of beams

might have enhanced the degree of shrinkage cracking of the beams.

Previously, various researchers investigated the vulnerability of SCC to shrinkage

cracking. Loser and Leemann (2008) stated that shrinkage of a concrete mixture is

primarily related to the volume of the paste in the mixture. Owing to the higher paste

volume and lower aggregate content, SCC has greater total shrinkage and a higher

shrinkage rate, and therefore, an earlier age of cracking than OC with comparable

compressive strength if rapid drying of concrete takes place. Loser and Leemann (2008)

also recommended the use of shrinkage reducing admixtures (SRA) in SCC to reduce

shrinkage and increase the age of cracking of concrete. Turcry et al. (2006) conducted an

experimental study, through which they concluded that an SCC mixture cracks earlier

than the OC mixture with the same compressive strength due to the higher shrinkage rate.

Turcry and Loukili (2006) explained the higher shrinkage rate of SCC with its lower

122
Figure 4.9 – Shrinkage cracking in B30 prior to the test

bleeding capacity than OC as a result of the higher binder content in SCC. Similarly,

Leemann and Hoffmann (2005) found out that SCC has a shrinkage rate 30% higher than

OC with the same compressive strength.

The high shrinkage rate of SCC might have reduced the age of shrinkage cracking

of concrete and caused shrinkage cracks to form before the removal of the first set of

beams from the forms. Surfaces of the beams exposed to air were kept moist using wet

burlaps. However, the lower bleeding capacity and the higher shrinkage rate of SCC

might have resulted in rapid drying of the surface and induced tensile stresses to concrete

before the subsequent rewetting of the burlaps on the beams.

123
Another potential stimulus for the shrinkage cracking of the first set of beams was

the late removal of the beams from their forms. The first set of beams was kept in the

forms for approximately two weeks. Formwork of a concrete beam constitutes a restraint

for the free shrinkage deformations. Although the open surfaces of the beams were

maintained wet till the dismantling of the forms, the forms might have caused restrained

shrinkage cracks to form due to higher shrinkage rate of SCC.

The early cracking age of the first set of beams in the forms was also related to

the specimen geometry. The beams were cast on their sides to facilitate the concrete pour.

Position of the beams in the formwork caused one of the lateral faces of each beam to be

uncovered, providing a large surface for the evaporation of the bleeding water.

Furthermore, the small widths of the specimens facilitated the drying to reach the internal

regions and affect the entire beam rapidly. Weiss and Shah (2001) carried out an

experimental study in which they observed that thinner concrete sections are less resistant

to shrinkage cracking and the age of cracking decreases as the specimen thickness

decreases.

In the second set of specimens, some measures were taken to prevent shrinkage

cracking of concrete. First, conventionally vibrated ordinary concrete (OC) was used

rather than SCC to increase the age of shrinkage cracking of concrete through the lower

shrinkage rate. Secondly, the beams were removed from the forms in less than a week to

eliminate the shrinkage restraint for concrete as early as possible.

Addition of the Eclipse Shrinkage Reducing Admixture (SRA), produced by

Grace Construction Products, was another protective measure against restrained

shrinkage cracking of concrete. Studies done by several researchers indicated the

124
favorable influence of SRA on the reduction of the total shrinkage and the shrinkage rate

of concrete. The experimental study conducted by Shah et al. (1992) indicated that

addition of SRA’s to concrete greatly reduced the free shrinkage deformations and the

widths of the shrinkage cracks in the case of restrained shrinkage. Lura et al. (2007)

experimentally showed that the addition of SRA’s to mortar reduces the evaporation of

water from the surface of the mortar and causes smaller tensile stresses to develop at the

surface. Therefore, mortar mixtures with SRA have fewer and narrower shrinkage cracks

than the mixtures without SRA under the same environmental conditions.

Efficiency of the measures taken to avoid shrinkage cracking of concrete was

examined through some methods. First, shrinkage cracks could not be detected in any of

the beams constructed in the second phase of experimental program. Nevertheless, the

presence of micro-cracks in concrete cannot be observed through visual inspection.

Hence, two more methods were used to measure the shrinkage strains in the beams to

investigate shrinkage cracking of concrete at the micro level. First, prismatic specimens

with and without SRA were prepared from the concrete mixtures used in the beams.

Sampling of concrete was done according to ASTM C192 (2007). Length changes of

specimens were measured according to the test method described in ASTM C157 (2006).

Six specimens were prepared from each of the concrete mixtures used in Beams B44 and

B36L. The SRA contents and curing conditions of the specimens are tabulated in Table

4.2.

In Figure 4.10, the percent length changes of Specimens 3, 4 and 5 are compared

to illustrate the influence of SRA on the volume change of concrete. According to the

plot, the length change of the specimen without SRA (Specimen 5) was measured to be

125
Table 4.2 – Descriptions of the shrinkage specimens from the concrete mixtures used in
B44 and B36L

Addition Curing
Specimen
of SRA Conditions
Same Conditions
1 Yes
as the Beams
Same Conditions
2 Yes
as the Beams
In the moist room
3 Yes
for 28 days
In the moist room
4 Yes
for 28 days
In the moist room
5 No
for 28 days
In the moist room
6 No
for 28 days

Figure 4.10 – Length changes of specimens with and without SRA from the concrete
mixture of B44

126
close to the length changes of the specimens with SRA (Specimens 3, 4) in the first 120

days after the concrete pour. Later, Specimen 5 experienced greater length changes than

the other two specimens.

Secondly, strains on the lateral faces of the specimens were continuously

measured through DEMEC (Demountable Mechanical) gages to determine the restrained

shrinkage stresses in the beams for monitoring for the formation of shrinkage cracks in

concrete. Directions of the principal stresses originating from restrained shrinkage are not

known. Hence, three independent strain measurements in different directions are needed

to determine the principal strains and stresses at a certain point. To measure the stresses

at the surfaces of the specimens, delta strain rosettes were formed at two different

locations on the side face of each specimen in specimen group B36L.

A DEMEC gage is a mechanical device which measures the distance between two

points. The gage has two conical points, one at each end of an invar bar. One of the

conical points is fixed and the other conical point can move in a certain range. To

measure the distance between two fixed points on a surface, the conical points of the gage

are inserted into the holes drilled at the fixed point. The initial distance between the two

fixed points on the beam is the gage length over which the strain is measured.

In the present study, four screw anchors were embedded into the fresh concrete at

each strain measurement location according to the pattern shown in Figure 4.11. Four

screws positioned in this pattern form a delta strain rosette. Strain in each direction is

obtained by dividing the change in the distance between two points to the initial distance

between the points, measured on the concrete pour day.

127
Figure 4.11 – Delta rosette for principal strain measurement at a point

Variation of the two principal strains in time is illustrated in Figures 4.12 and 4.13

for Specimens B36L-2 and B36L-3, respectively. Cracking strain of concrete (εcr) in

uniaxial tension is also shown in each plot. εcr is calculated from Equation (4.19):

ft
 cr  (4.19)
Ec

where ft is the splitting tensile strength of concrete, which is obtained from Equation

(4.20):

ft  6.4  f c' (4.20)

128
Figure 4.12 – Principal strains on the side face of B36L-2

Figure 4.13 – Principal strains on the side face of B36L-3

129
Equation (4.20) is the tensile strength of concrete in uniaxial tension according to

Mirza et al. (1979). Although the tensile strength of concrete reduces in the presence of a

compressive stress in the perpendicular direction the reduction is ignorable considering

the small values of the compressive principal strains in Figures 4.12 and 4.13 on the day

when the tensile principal strains reached the peak values.

The plots indicate that tensile strains developed at the surfaces of the beams till

the removal of the beams from the formwork, which constituted a restraint for the free

shrinkage deformations. After the removal of the beams from the forms, on the other

hand, the beams were subjected to compressive strains originating from the free

shrinkage deformations of concrete. The tensile principal strains in the beams prior to the

dismantling of the forms only slightly exceeded the cracking strain of concrete for a short

period of time. Therefore, the potential shrinkage cracks in the beams are expected to be

narrow and small in number.

Since the specimens of the present study did not experience significant restrained

shrinkage cracking according to the aforementioned measurements, the multiplier ω in

Equation (4.18) was taken 1 in the evaluation of the lateral bending rigidities of the

specimens. ω can be taken 2/3 as a conservative assumption when the restrained cracking

condition of concrete in a beam is not known.

130
CHAPTER V

TORSIONAL RIGIDITY OF RECTANGULAR REINFORCED


CONCRETE BEAMS

Resistance of a beam to lateral torsional buckling is determined by the lateral bending

rigidity and the torsional rigidity of the beam. The present chapter briefly introduces the

torsional behavior of reinforced concrete beams and explains the evaluation of the

torsional rigidity of a concrete beam in the light of the experimental torque-twist curves

of the test specimens.

5.1 Torsional Behavior of Reinforced Concrete Beams

The torsional behavior of reinforced concrete beams is explained with the help of Figure

5.1, which is the typical torque-twist curve of a reinforced concrete beam with shear

reinforcement. The torque-twist curve in the figure can be divided into three distinct

segments: OA, AB and BC. The initial linear segment (OA) ends at point A, which

corresponds to the initiation of the diagonal cracking in the beam. The slope of the line

OA is termed as the uncracked torsional rigidity of the beam, (GC)u. Prior to the

formation of the diagonal tension cracks, the torsional rigidity is related to the shear

strains around the perimeter of the cross-section of a beam. The entire beam behaves as a

solid and homogeneous body and the contribution of the flexural and shear reinforcement

to the torsional rigidity can be neglected. The uncracked torsional rigidity expressions

existing in the literature are presented in Section 5.2.1.

131
Figure 5.1 – Typical torque-twist curve of a reinforced concrete beam with shear
reinforcement

The second segment (AB) starts when the applied torque reaches the cracking

torque, Tcr. According to ACI 318R-05 Section 11.6.1, Tcr can be determined from

Equation (5.1):

 Acp
2 
Tcr  4  fc'   (5.1)
 pcp 
 

where f’c is the cylinder compressive strength of concrete in psi; Acp is the gross cross-

sectional area of the beam and pcp is the perimeter of the cross-section.

132
Cracking torque of a reinforced concrete beam is the torsional strength of the

plain concrete beam with the same dimensions. Equation (5.1) was developed based on

the assumption that a plain concrete beam fails in torsion when the principal tensile stress

in the beam becomes equal to the tensile strength of concrete (f’t), which can be obtained

from Equation (5.2):

ft '  4  f c' (5.2)

Equation (5.2) is the tensile strength of concrete under biaxial tension and compression. It

was used instead of Equation (4.20), which is the tensile strength of concrete under

uniaxial tension, to account for the compressive and tensile principal stresses in a beam

under pure torsion.

Hsu (1984) developed a criterion for the torsional failure of plain concrete

members, which is based on the skew-bending theory, developed by Hsu (1968) to

explain the torsional behavior of concrete beams. According to the skew-bending theory,

the failure plane of a concrete beam, loaded in pure torsion, makes a 45-degree angle

with the longitudinal axis of the beam. The applied torque can be decomposed into two

components: a component parallel to the failure surface (bending component, Tb in

Figure 5.2) and a component perpendicular to the surface (twisting component, Tt in

Figure 5.2). According to Hsu (1984), torsional failure of a plain concrete beam takes

place when the tensile stress on the lateral face of the beam (σt in Figure 5.2) induced by

the bending component of the applied torque reaches the modulus of rupture of concrete.

Accordingly, the torsional strength of a plain concrete member is obtained from Equation

(5.3).

133
Figure 5.2 – Components of the axial torque on the failure surface of a concrete beam
according to the skew-bending theory

b2  h
Tnp    0.85  f r  (5.3)
3

where fr is the modulus of rupture of concrete, expressed in terms of f’c according to

Equation (5.4):

 10 
f r  21  1  2  3 f c' (5.4)
 b 

The multiplier 0.85 in Equation (5.3) is the reduction in the modulus of rupture of

concrete resulting from the stresses induced by the twisting component of the applied

torque. The compressive stresses (σc in Figure 5.2) from the twisting component are in

perpendicular direction to the tensile stresses from the bending component on the lateral

face of the beam. Interaction of the tensile and compressive stresses induces a 15-percent

reduction in the modulus of rupture.

134
Based on previous experimental results [Hsu (1968)], Hsu (1984) established that

torsional reinforcement in a beam (including both longitudinal and shear reinforcement)

increases the cracking torque of the beam, although it does not influence the uncracked

torsional rigidity. The cracking torque of a concrete beam reinforced with longitudinal

reinforcement and closed stirrups is obtained from the torsional strength of the plain

concrete beam with the same dimensions, Tnp (Equation 5.3), using Equation (5.5):

Tcr  (1  4  t )  Tnp (5.5)

where ρt is the total volumetric reinforcement ratio of the beam, calculated from Equation

(5.6):

t  l   s (5.6)

Volumetric ratio of the longitudinal reinforcement, ρl, and volumetric ratio of the shear

reinforcement, ρs, are determined according to Equations (5.7) and (5.8), respectively.

Al
l  (5.7)
Acp

At  p1
s  (5.8)
Acp  s

where Al is the total cross-sectional area of the longitudinal reinforcement; At is the cross-

sectional area of one leg of a stirrup; p1 is the perimeter of the area bounded by the

centerline of a stirrup; s is the spacing of the stirrups.

135
In Section 5.3 of the present chapter, the maximum torsional moments in the

specimens at the initiation of buckling will be compared to the cracking torques, obtained

from Equation (5.1) and (5.5), to determine the cracking conditions of the beams at the

instant of buckling.

In the uncracked stage, the entire solid section is effective in resisting the

torsional moments. Upon diagonal cracking, the concrete core of the section becomes

ineffective. In the post-cracking stage, therefore, torsional rigidity of a reinforced

concrete beam is provided by the outer skin of the section enclosing the closed stirrups

and the longitudinal corner bars. The horizontal plateau (AB) in Figure 5.1 corresponds

to the redistribution of the shear forces in the beam as the transition from the uncracked

condition to the post-cracking condition takes place. Segment AB becomes less

pronounced as the torsional reinforcement ratio of a concrete beam increases.

Initial portion of the post-cracking segment (BC) of the torque-twist curve is

linear. The slope of the initial linear portion is denoted as the post-cracking torsional

rigidity of the beam (GcrCcr in Figure 5.1). The post-cracking torsional rigidity

expressions in the literature are introduced in Section 5.2.2.

Beyond the linear portion of BC, the torque-twist curve curves gradually to

horizontal until the applied torque reaches the torsional strength of the beam, Tn. In the

curved portion of BC, torsional rigidity decreases with an increase in the torque. At the

torsional strength level, the beam does not possess torsional rigidity, and thus, torsional

failure takes place after a short time.

136
5.2 Torsional Rigidity of Rectangular Reinforced Concrete Beams

Torsional cracking changes the behavior of a reinforced concrete beam completely.

Different distributions of the strains from torsion in the pre- and post-cracking stages of

loading create different equilibrium conditions. Owing to the differences in the torsional

behavior of a concrete beam before and after diagonal cracking, the following discussion

classifies the torsional rigidity expressions, existing in the literature, into two separate

groups: the uncracked torsional rigidity expressions and the post-cracking torsional

rigidity expressions for reinforced concrete beams.

5.2.1 Uncracked Torsional Rigidity Expressions

A reinforced concrete beam is considered as an elastic and homogeneous body prior to

the formation of the diagonal tension cracks. The torsion of elastic and homogeneous

beams was studied by St. Venant (1856), who developed a semi-inverse method to solve

the equations from the theory of elasticity, defining the torsion of noncircular sections.

Using Fourier series, St. Venant (1856) reached the torsional rigidity expression for the

rectangular sections:

(GC )u  c  b3  h  Gc (5.9)

where Gc is the modulus of rigidity of concrete, calculated from Equation (5.10) and βc is

the coefficient for St. Venant’s torsional constant, obtained from Equation (5.11):

Ec
Gc  (5.10)
2  (1  )

where υ is the Poisson’s ratio of concrete and Ec is the elastic modulus of concrete.

137
1  192 b  1 (2n  1) h 
 c   1  5
  5
 tanh  (5.11)
3   h n0 (2n  1) 2b 

Equation (5.11) indicates that St. Venant’s torsional constant depends on the height-to-

width (h/b) ratio of a cross-section.

In discussing the torsional rigidity of rectangular sections, Wang (1953) stated

that the first term of the infinite series in Equation (5.11) gives the value of the sum to

within 0.5 percent. Therefore, for practical purposes, βc can be approximated to a simpler

form, considering the first term of the series only:

1  192 b h 
c   1  5
  tanh  (5.12)
3   h 2b 

According to Timoshenko and Goodier (1970), for narrow rectangular cross-sections

h
tanh 1 (5.13)
2b

Accordingly, Equation (5.12) can be simplified to Equation (5.14), if the beam has a

narrow rectangular cross-section:

1  b
 c   1  0.63   (5.14)
3  h 

The torsional rigidity expression adopted by Siev (1960) uses the above form of βc.

Assuming that the modulus of rigidity of concrete (Gc) obtained from Equation (5.10) is

valid at the time of buckling, Siev (1960) proposed Equation (5.15):

138
 b3  h  b 
(GC ) s  Gc     1  0.63    (5.15)
 3  h 

Another approximate form of Equation (5.11) was presented by Yen (1975), based on the

studies of Kollbrunner and Bassler (1969):

1  b b5 
 c   1  0.630   0.052 5  (5.16)
3  h h 

Hansell and Winter (1959) simplified Equation (5.11) to the following form:

2
1  b
 c   1  0.35   (5.17)
3  d 

Using the above from of βc, Hansell and Winter (1959) proposed a torsional rigidity

expression for rectangular reinforced concrete beams:

 b3  c  b 
2
(GC ) hw  Gc'   1  0.35    (5.18)
 3  d  

where G’c is the reduced modulus of rigidity of concrete according to Hansell and Winter

(1959), calculated from Equation (5.19):

Esec
Gc'  (5.19)
2  1  

where Esec is the secant modulus of elasticity of concrete corresponding to the extreme

compression fiber strain at the initiation of buckling.

139
Equation (5.19) takes into account both elastic and inelastic lateral torsional

buckling of reinforced concrete beams. Elastic modulus of concrete (Ec) and the modulus

of rigidity calculated from Ec (Equation 5.10) do not reflect the true material rigidity of a

beam if some fibers of the beam are stressed beyond the elastic limit of the stress-strain

curve of concrete, as in the case of inelastic lateral torsional buckling. Hansell and Winter

(1959) suggested to use the reduced shear modulus (G’c) to account for the reduction in

the overall material rigidity of the beam when the beam buckles inelastically. In the case

of elastic lateral torsional buckling, on the other hand, G’c becomes equal to Gc since all

fibers throughout the beam are stressed within the elastic range of the stress-strain curve

of concrete and Esec is equal to Ec

Sant and Bletzacker (1961) approximated the parameter βc to 1/3, which is

commonly used in thin-walled sections and proposed the following torsional rigidity

expression for narrow rectangular reinforced concrete beams:

b3  d
(GC ) sb  Gr  (5.20)
3

where Gr is the reduced modulus of rigidity of concrete according to Sant and Bletzacker

(1961), calculated from Equation (5.21):

Er
Gr  (5.21)
2  1  

where Er is the double modulus of concrete corresponding to the extreme compression

fiber strain at the instant of buckling, calculated from Equation (5.22):

140
4  Ec  Etan
Er  (5.22)
 
2
Ec  Etan

where Etan is the tangent modulus of concrete corresponding to the extreme compression

fiber strain at the instant of buckling.

Equation (5.20) depicts that Sant and Bletzacker (1961) preferred to relate the

shear modulus of concrete to the double modulus of elasticity, Er to account for the

possible inelastic material behavior at the initiation of buckling.

In the above discussion, simplified versions of St. Venant’s torsional constant

were presented. Equations (5.12), (5.14), (5.16) and (5.17) are the simplified forms of

Equation (5.11). Figure 5.3 compares the values obtained from the simplified forms of

Equation (5.11) to the actual values of βc obtained from Equation (5.11).

Figure 5.3 shows that Equations (5.12), (5.14), (5.16) and (5.17) are in good

agreement with Equation (5.11) for b/h smaller than unity. However, the use of a constant

value of 1/3 for βc, proposed by Sant and Bletzacker (1961), is meaningful only when the

section is quite narrow. For b/h>0.1, assuming βc=1/3 will introduce significant errors to

the calculations. To conclude, the aforementioned simplified versions of Equation (5.11)

can be used instead of Equation (5.11) to facilitate the uncracked torsional rigidity

calculations.

5.2.2 Post-Cracking Torsional Rigidity

In the post-cracking stage of loading, the torsional rigidity of a reinforced concrete beam

is provided by the outer skin concrete, since the concrete core is rendered ineffective by

141
Figure 5.3 – Comparison of the coefficients βc calculated from different equations

the diagonal tension cracks. The outer skin concrete is assumed to form a thin-walled

tube including the closed stirrups and the longitudinal corner bars.

The post-cracking torsional rigidity expressions developed by previous

researchers are based on a 3-D model, denoted as the thin-walled tube space truss model,

which is based on Rausch’s (1929) space truss analogy. According to the model, a solid

beam turns into a thin-walled tube after formation of the diagonal cracks. The thin-walled

tube, providing the post-cracking torsional rigidity, is a space truss, which is composed of

three different types of members (Figure 5.4). Helical concrete strips between the

diagonal tension cracks form the compression struts which are assumed to be connected

to the closed stirrups and the longitudinal reinforcing bars at the joints through hinges.

142
Figure 5.4 –Thin-walled tube space truss model

Compressive stresses in the tube are carried by the compression struts while the tensile

stresses are carried by the shear and longitudinal reinforcement. Using the equilibrium of

forces and compatibility of strains in the space truss, Lampert (1973) was able to develop

a post-cracking torsional rigidity expression (Equation 5.23) for rectangular reinforced

concrete beams.

4  Es  A23
Gcr  Ccr  (5.23)
 4  n    A2 1 1 
p22     
 p2  ti l  s 

where A2 is the area bounded by the lines connecting the centers of the longitudinal

corner bars; p2 is the perimeter of the area bounded by the lines connecting the centers of

the corner bars; λ is a multiplier for the concrete strain [λ=3 according to Lampert

143
(1973)]; n is the modular ratio of steel to concrete; ti is the wall thickness of the tube. ti is

the smaller of b/6 and b2/5; b is the width of the beam and b2 is the smaller dimension of

the rectangle formed by the lines connecting the centers of the longitudinal corner bars.

The three terms in the denominator of Equation (5.23) correspond to the

contributions of the concrete struts, the longitudinal reinforcing bars and the closed

stirrups, respectively.

Hsu (1973) proposed a similar equation using the thin-walled tube space truss

model:

4  Es  Ae2  Acp
Gcr  Ccr  (5.24)
 4  n  Acp 1 1 
pe2     
 pe  te l  s 

where Ae is the area bounded by the centerline of the effective wall; pe is the perimeter of

the area bounded by the centerline of the effective wall; Acp is the gross area of the

section; te is the effective wall thickness. Based on the previous experimental results, Hsu

(1973) proposed an empirical equation to obtain te:

te  1.4   l   s   b (5.25)

Later, Hsu (1990) introduced the concept of shear flow zone. According to the

concept, thickness of the thin-walled tube, providing the post-cracking torsional rigidity,

is the thickness of the shear flow zone (td) in the post-cracking stage. Thickness of the

shear flow zone depends on the applied torque according to the following equation:

4  Ta
td  (5.26)
Acp  f c'

144
where Ta is the applied torque. Equation (5.26) was obtained by Hsu (1990) using the

softened truss model.

Finally, Tavio and Teng (2004) developed an equation for the torsional rigidity of

a reinforced concrete beam at cracking, using the shear flow zone concept:

4    Es  Ao2  Acp
 GC cr  (5.27)
 1 1 
p   
2

 l  s 
o

where Ao is the area bounded by the centerline of the shear flow zone; po is the perimeter

of the area bounded by the centerline of the shear flow zone and μ is a multiplier. Tavio

and Teng (2004) stated that a value of μ= 1.5 matches well with the experimental data in

the literature.

Equations (5.23) and (5.24) correspond to the post-cracking torsional rigidity

(GcrCcr in Figure 5.1), which is the slope of the initial linear portion of the post-cracking

segment of the torque-twist curve. Equation (5.27), on the other hand, corresponds to

torsional cracking at rigidity [(GC)cr in Figure 5.1], which is the slope of the secant line

connecting the end point of the horizontal plateau of the torque-twist curve (Point B in

Figure 5.1) to the origin. The lack of the first term in the denominator of Equation (5.27)

depicts that Tavio and Teng (2004) neglected the contribution of the concrete

compression struts to the torsional rigidity at cracking in order to simplify the expression.

5.3 Experimental Torsional Rigidities of the Test Beams

Figures (5.5) & (5.6) illustrate the experimental torque-twist curves of B44-1 and B36L-1

to explain the torsional behavior of a reinforced concrete beam in a lateral-torsional

145
buckling test. The experimental torque-twist curves of the remaining specimens are

presented in Appendix D.

Figure 5.5 –Experimental torque-twist curve of Specimen B44-1

Figure 5.6 – Experimental torque-twist curve of Specimen B36L-1

146
Application of a single concentrated load at mid-span and simple support

conditions in and out of plane at the beam ends resulted in the non-uniform distribution

of the torsional moment along the span of each specimen, as previously shown in Figure

3.34(b). Due to the non-uniform moment distribution, the torque-twist curve of a test

beam is somewhat different from the typical torque-twist curve of a reinforced concrete

beam under uniform torque throughout the span. In Figure 5.1, the horizontal plateau

(AB) corresponds to the diagonal cracking throughout the entire span of a beam when the

applied torque reaches the cracking torque. In other words, the entire beam is subject to

diagonal cracking at the same stage of loading and the redistribution of the strains

throughout the whole span creates a noticeable softening in the beam beyond the initial

linear portion of the torque-twist curve. According to the experimental results obtained

by Hsu (1968), the horizontal plateau is distinct in reinforced concrete beams with closed

stirrups up to a total volumetric reinforcement ratio (ρt) of 0.04-0.05.

In Figures (5.5) & (5.6), on the other hand, the torque-twist curve does not have a

pronounced horizontal plateau due to the progressive reduction in the torsional rigidity of

the beam. In Figure 5.7, the torque-twist curve of B44-2 is approximated with a series of

linear segments with decreasing slopes to illustrate that the overall torsional rigidity of

the beam reduces gradually as the diagonal tension cracks, existing in the support zones

earlier in the test, spread towards the inner portions of the span in the further stages of

loading.

The maximum torsional moment in the beam at the initiation of buckling (Tb) is

shown with a heavy solid line on Figures (5.5) & (5.6). In the torque-twist curve of each

specimen, Tb falls into the first linear segment of the curve, which has the greatest slope

147
Figure 5.7 – Approximation of the torque-twist curve of B44-2 into a series of line
segments

among all segments. The fact that progressive reduction in the slope of the curve starts

beyond Tb manifests the absence of the diagonal tension cracks in the entire beam at the

time of buckling. In other words, all of the specimens were diagonally uncracked at the

initiation of buckling. The same conclusion can be drawn from Table 5.1, which tabulates

cracking torques of the specimens, according to Equations (5.1) and (5.5), together with

the maximum torsional moments in the beams at the start of buckling (Tb). Tb of each of

the specimens, except B18-2 and B30, is smaller than the cracking torques obtained from

both equations. Tb values of B18-2 and B30, on the other hand, are slightly larger than

Tcr, obtained from Equation (5.1). However, B18-2 and B30 are accepted as completely

uncracked at the initiation of buckling, since Tb of each specimen is smaller than Tcr

148
according to Equation (5.5), which is developed based on the results of several tests,

carried out by Hsu (1968).

Table 5.1 – Maximum torsional moments at the initiation of buckling and the cracking
torques of the specimens

Tcr (in-kips)
Specimen Tb* (in-kips)
Equation (5.1) Equation (5.5)
B18-1 12.6 8.5 34.9
B30 48.5 41.1 78.3
B36 27.6 50.8 95.9
B44-1 45.9 70.4 114.1
B44-2 36.2 70.6 114.5
B44-3 30.7 71.1 114.9
B36L-1 53.6 60.0 96.7
B36L-2 49.8 60.0 97.0
*
Maximum Measured Torsional Moment in the Beam at the Initiation of
Buckling

Table 5.2 tabulates the torsional rigidities of the beams, calculated from Equations

(5.9), (5.18), (5.20), (5.24) and (5.27). The table also includes the slopes of the initial

linear segments of the experimental torque-twist curves of the specimens, for the sake of

comparison. The slope of the experimental torque-twist curve of a beam under the

loading and support conditions of the present study cannot be directly compared to the

torsional rigidity values calculated from the aforementioned equations, due to the non-

uniform torsional moment distribution along the beam span. The experimental torque-

twist curves were obtained by plotting twist per unit length of the beam (θ) against the

maximum torsional moment along the span (Tmax). The torsional moment in the beam

decreases from maximum at the ends to minimum at mid-span. Therefore, Tmax

corresponds to the laterally-supported beam ends only. The slope of a torque-twist curve

is the torsional rigidity of a beam if the torsional moment is constant through the entire

149
Table 5.2 – Torsional rigidities of the specimens

Torsional Rigidity (x106 in2-kips)


Specimen (GC)m (GC)u (CG)hw (GC)sb GcrCcr (GC)cr
Fig. 5.4 Eq. 5.9 Eq. 5.18 Eq. 5.20 Eq. 5.24 Eq. 5.27
B44-1 1.067 0.751 0.284 0.669 0.045 0.115
B44-2 0.890 0.768 0.289 0.685 0.046 0.115
B44-3 1.166 0.798 0.297 0.712 0.047 0.116
B36L-1 1.059 0.685 0.273 0.614 0.043 0.101
B36L-2 0.988 0.720 0.282 0.646 0.044 0.101
B36 0.859 0.500 0.192 0.453 0.032 0.069
B30 0.612 0.410 0.151 0.369 0.026 0.056
B22-1 0.178 0.059 0.023 0.051 0.005 0.010
B22-2 0.198 0.048 0.019 0.042 0.004 0.008
B18-2 0.181 0.041 - 0.034 0.003 0.007

span. In the case of the test specimens, the slope of the experimental torque-twist curve,

plotting Tmax vs. θ, is greater than the actual torsional rigidity of the beam, since only the

support zones resist large torsional moments in the order of Tmax. However, the slopes of

the experimental curves are shown in Figures (5.5) & (5.6) and tabulated in Table 5.2 to

compare the order of magnitude of the experimental torsional rigidities of the beams with

the analytical values calculated from the uncracked and post-cracking rigidity

expressions.

To conclude, the experimental torque-twist curves of the specimens reveal that the

torsional behavior of a reinforced concrete beam, buckling elastically, is closely predicted

by St. Venant’s theory, and thus, the torsional rigidity of a concrete beam at the buckling

instant can be obtained from Equation (5.9). The profound difference between the slopes

of the initial linear portions of the experimental torque-twist curves [(GC)m] and the post-

150
cracking torsional rigidities of the beams (last two columns in Table 5.2) clearly indicates

that concrete beams, buckling elastically, do not undergo diagonal cracking. In the next

section, some modifications to Equation (5.9) are proposed to account for the case of

inelastic lateral-torsional buckling in reinforced concrete beams.

5.4 Proposed Torsional Rigidity Expression

As explained in the previous section, the torsional constant (C) of a reinforced concrete

beam is closely estimated by St. Venant’s theory prior to the formation of diagonal

tension cracks. The material rigidity term in Equation (5.9) is the shear modulus of

rigidity of concrete (Gc), which is obtained from the elastic modulus (Ec) through

Equation (5.10).

The use of Ec in the critical moment calculations is appropriate only if the beam

buckles elastically. To account for both elastic and inelastic lateral torsional buckling of

concrete beams, another modulus of elasticity, termed as the overall modulus of elasticity

[Equation (4.16)], was proposed. Using a simplified form of St. Venant’s torsional

constant, presented in Section 5.2.1, and accounting for the possible inelastic material

behavior of concrete at the instant of buckling, the following torsional rigidity expression

is proposed for the rectangular reinforced concrete beams:

 b3  h  b 
(GC )o  Go    1  0.63    (5.28)
 3  h 

where Go is the overall modulus of rigidity of concrete, calculated from Equation (5.29):

151
Eo
Go  (5.29)
2  1  

where Eo is the overall modulus of elasticity of concrete, obtained from Equation (5.30):

 E  Ec 
Eo   sec  (5.30)
 2 

152
CHAPTER VI

CRITICAL MOMENT CALCULATIONS AND INFLUENCES OF


THE INITIAL GEOMETRIC IMPERFECTIONS ON THE LATERAL
STABILITY OF REINFORCED CONCRETE BEAMS

6.1 Introduction

A geometrically perfect beam buckles when the applied moment reaches a critical value,

denoted as the critical moment (Mcr). In the presence of initial geometric imperfections,

on the other hand, the ultimate moment-carrying capacity, also termed as the limit

moment (ML), of a beam is smaller than the critical moment (Mcr) corresponding to the

perfect initial configuration of the beam.

In Section 6.2, the critical moment calculations of beams are presented.

Determination of the critical moment of a beam includes the evaluation of its torsional

and lateral bending rigidities. Therefore, Section 6.2 is linked to Chapters IV and V.

In Section 6.3, the influences of the initial geometric imperfections on the lateral

stability of reinforced concrete beams are explained. Section 6.3 also presents an equation

to calculate the limit moment (ML) of a concrete beam with initial lateral imperfections

(sweep) and initial twisting angle from the critical moment (Mcr) corresponding to the

initially perfect configuration of the beam.

6.2 Critical Moment Calculations

Timoshenko and Gere (1963) developed critical moment expressions for beams with

different cross-sectional shapes, loading and support conditions. Equation (6.1) is a very

general form of the critical moment expression, developed by Vacharajittiphan et al.

153
(1974) considering the influence of the in-plane (vertical) deformations of a beam prior to

buckling on the lateral stability:

  2 ECw 
EI y GJ  1  
C1  GJL2 
M cr   (6.1)
C2  L  EI   GJ   2 ECw  
y
1    1   1  
 EI x    EI x  GJL2  

where C1 is a constant corresponding to the loading conditions of a beam; C2 is a constant

corresponding to the support conditions; Mcr is the critical moment; L is the unbraced

length; EIx, EIy ,GJ, ECw are the in-plane, out-of-plane, torsional and warping rigidities of

a beam, respectively.

Smitses and Hodges (2006) stated that the effect of warping rigidity (ECw) is

considerable in thin-walled open cross-sections only. According to Timoshenko and Gere

(1963), Cw can be taken zero in a beam with narrow rectangular cross-section. Hence,

Equation (6.1) simplifies to Equation (6.2):

C1 EI y GJ
M cr   (6.2)
C2  L  EI y   GJ 
1    1  
 EI x   EI x 

The expression in the square root in the denominator of Equation (6.2)

corresponds to the in-plane (vertical) deformations of a beam prior to buckling. In deep

beams, the in-plane flexural rigidity (EIx) is significantly greater than the out-of-plane

flexural rigidity (EIy) and the torsional rigidity (GJ). Therefore, the square root term in

the denominator is very close to unity in deep beams. Ignoring this term does not change

154
the calculated values to a major extent. For instance, Beams B44 of the present study had

a EIy/ EIx ratio of 0.0048 and a GJ / EIx ratio of 0.0080. Using these values, the square

root term in the denominator becomes 0.994, which corresponds to a 0.6% change in the

critical moment. When the square root term in the denominator is ignored, Equation (6.2)

reduces to Equation (6.3):

C1
M cr   EI y GJ (6.3)
C2  L

According to Allen and Bulson (1980), the constant C1 has a value of 4.23 for a

single concentrated load at midspan. The effective length ratio C2 has a value of 1.00

when a beam is simply-supported in and out of plane.

The applied load has an additional destabilizing effect on the beam when it is

applied above the centroid of the section (Figure 1.22 of Chapter I). On the contrary, the

load has a stabilizing effect on the beam when it is applied below the centroid. Equations

(1.21) – (1.23), proposed by Stiglat (1991), account for the influence of the location of

the load application point with respect to the centroid of the section. Timoshenko and

Gere (1963) developed a critical load expression considering the influence of the location

of the point of application of load with respect to the centroid of the midspan cross-

section. Accordingly, Equation (6.3) can be modified to Equation (6.4) to account for this

effect:

4.23  e EI y 
M cr  1  1.74     EI y GJ (6.4)
L  L GJ 

155
where e is the initial vertical distance of the load from the shear center of the beam

section.

EIy and GJ are the rigidities of a homogeneous and elastic beam. The lateral

bending rigidity (Bo) and the torsional rigidity [(GC)o] of a reinforced concrete beam are

different from EIy and GJ due to the differences in behavior between a reinforced

concrete beam and a homogeneous and elastic beam, such as cracking of concrete,

elastic-inelastic material behaviors of concrete and reinforcing steel, etc. In the present

study, Equation (4.17) and Equation (5.28) are proposed for calculating the lateral

bending rigidity and the torsional rigidity of a reinforced concrete beam, respectively.

Considering all the aforementioned changes, Equation (6.5) is proposed:

4.23  e Bo 
M cr   1  1.74     B   GC o (6.5)
L 

L  GC o  o

6.3 Influences of Sweep and Initial Twisting Angle on the Lateral Stability of
Reinforced Concrete Beams

Initial geometric imperfections play a crucial role in the stability of beams. Concrete

girders possess three different types of geometric imperfections: camber (initial in-plane

deformation), sweep (initial out-of-plane deformation) and initial twisting angle.

Influence of sweep on the lateral stability of reinforced concrete beams is two-

fold. First, the out-of-plane deformations of a beam are affected by the sweep. A

geometrically perfect beam does not experience lateral deformations and twisting

rotations until bifurcation buckling takes place. When the buckling moment is reached, a

156
beam free from sweep undergoes very large lateral deformations and rotations at a

constant moment level. Nevertheless, the load-deflection behavior of a beam is different

in the presence of sweep, which causes the beam to undergo lateral deformations in the

pre-buckling stage of loading. Lateral deformations start with the initiation of loading and

grow at a relatively low rate in the pre-buckling stage. Once the beam buckles, the lateral

deformations and twisting rotations grow at much higher rates while the moment carried

by the beam is maintained at an approximately constant level.

Similarly, the initial twisting angle in a beam causes the beam to experience

twisting rotations even prior to buckling. The twisting rotations, growing slowly in the

pre-buckling stage, become very large after buckling takes place.

The second effect of sweep is the reduction in the ultimate load carried by a

concrete beam. A geometrically perfect beam buckles when the maximum moment

carried by the beam reaches the critical moment. Nonetheless, the moment carrying

capacity of a beam with sweep is smaller than the critical moment (Mcr). The maximum

moment on the load-deflection curve of an imperfect beam is termed as the limit moment

(ML) of the beam, which should be distinguished from the critical moment.

To clarify the above discussion, the experimental load-lateral deflection curves of

the beams in specimen groups B44 and B36L are illustrated in Figures 6.1 and 6.2,

respectively. The load-deflection curves do not start from the origin. The sweep of each

beam at the centroid of midspan section was taken as the initial lateral deflection

(deflection at zero load). The load-lateral deflection curves of the other specimens are

presented in Appendix D.

157
Figure 6.1 – Lateral top deflections of B44-1 and B44-2 at midspan

Figure 6.2 – Lateral top deflections of B36L-1 and B36L-2 at midspan

158
According to Figures 6.1 and 6.2, the load-lateral deflection curve of a reinforced

concrete beam has an initial linear portion and a curved portion. The curved portion turns

into an approximately horizontal line beyond the limit load, meaning that the lateral

deflections in the beam increase excessively at a constant load level when the beam

buckles. The two-fold influence of sweep on the lateral stability of reinforced concrete

beams can be observed in Figures 6.1 and 6.2. The beam with the greatest sweep

experiences larger out-of-plane deformations than its companion before reaching the

ultimate moment. Furthermore, the ultimate moment carried by the beam having the

largest sweep is smaller than the ultimate moment carried by its companion.

The influences of sweep on the limit moment and the load-deflection behavior of

a concrete beam can be understood by considering the differences between the load-

lateral deflection curves corresponding to the identical beams. First, the load-deflection

curves of the companion beams differ in the slope of the initial linear portion of the

curve, which increases as the girder sweep decreases. Secondly, sweep affects the

sharpness of the curved portion of the load-deflection curve. When the girder sweep

increases, the linear portion of the curve ends at lower load levels and the slope of the

curve decreases from a maximum to zero along a greater portion of the curve, creating a

smoother curved portion.

The differences between the load-lateral deflection curves of identical beams with

different sweeps can be clearly observed in Figure 6.2. B36L-1 and B36L-2 were

identical in nominal dimensions, cross-sectional details and nominal material strengths,

and they only differed in initial geometric imperfections (Table B.4 in Appendix B). The

initial lateral deformations of B36L-1 and B36L-2 were measured as 15/16 and 3/8

159
inches, respectively, at the top of the beams at midspan. The initial linear portion of the

load-deflection curve of Specimen B36L-2 is steeper than the linear portion of the curve

of Specimen B36L-1. Furthermore, the load-deflection behavior of the beam with greater

sweep (B36L-1) ceased to be linear at earlier stages of loading than the beam with

smaller sweep (B36L-2). B36L-1, which buckled at a load of 13.5 kips, has a linear load-

deflection behavior up to 5 kips. On the other hand, the load-deflection curve of B36L-2,

with a buckling load of 21.6 kips, remains linear up to 15 kips. Since the linear portion of

the curve of B36L-1 ends at smaller loads and the slope of the curve gradually decreases

up to the buckling load, the load-deflection curve of B36L-1 has a smoother curve

beyond the linear portion. B36L-2, on the contrary, has a sharp curve beyond the linear

portion due to the rapid decrease in the slope of the curve beyond the longer linear

portion.

The reduction in the limit moment (ML) of a reinforced concrete beam due to

sweep has not been studied extensively in the literature. Burgoyne and Stratford (2001)

stated that the additional stresses associated with the initial minor-axis curvature created

by sweep are responsible for the reduction in the ultimate moment of a concrete beam.

Longitudinal strains in a beam with an initial lateral curvature originate from in-plane and

out-of-plane bending moments. Figure 6.3 illustrates the longitudinal strain distributions

in the cross-section of a beam caused by the major-axis and minor-axis bending

moments. In a geometrically perfect beam, strains from minor-axis bending [Figure

6.3(a)] are not present up to buckling. In a beam with initial lateral deformations,

nevertheless, the minor-axis curvature created by the sweep produces longitudinal strains

even prior to the application of load. Since the beam undergoes lateral deformations as

160
Figure 6.3 – Longitudinal strain distributions in a cross-section from major-axis and
minor-axis bending moments

loading progresses, the additional stresses associated with the minor-axis curvature

increase. According to Burgoyne and Stratford (2001), cracking of concrete led by the

longitudinal stresses from the minor-axis bending moments results in the reduction of the

lateral bending rigidity of the beam, which decreases the limit moment (ML). As the

sweep of a concrete beam increases, the initial longitudinal strains related to the minor-

axis curvature increase and the reduction in the buckling resistance of a beam due to

cracking takes place earlier in the loading history. Hence, the limit moment of a concrete

beam decreases with the increasing sweep.

161
To inspect the correctness of the above statements, the strain data obtained in the

second set of experiments was examined. As previously explained in Section 3.1.3,

longitudinal strains were measured on the lateral faces of each beam at midspan.

Assuming that minor axis of the cross-section is coincident with the vertical centroidal

axis, the compressive strain on the concave face of a beam (εcl in Figure 6.3) resulting

solely from the lateral bending moment is equal to the tensile strain on the convex face

(εtl) originating from lateral bending. Therefore, the longitudinal strain from major-axis

bending at a particular depth can be obtained by averaging the two strains measured on

the convex and concave faces of the beam at that depth. The difference between the strain

measured on the concave face and the average of the two strains is the compressive strain

(εcl) created by the minor-axis bending only while the difference between the strain

measured on the convex face and the average of the two strains is the tensile strain (εtl)

from lateral bending.

Figures 6.4 and 6.5 illustrate the extreme compression fiber strains of Beams B44

and B36L at midspan resulting from the in-plane bending moments only. The extreme

compression fibers at midspan are the most stressed compression fibers of a beam. The

load-strain curves in Figures 6.4 and 6.5 are linear up to the limit load (PL). The linear

relationships in the figures imply that the in-plane bending moments created elastic

material response in the beams. The load-strain curves in Figures 6.4 and 6.5 show a

different character from the load-lateral deflection curves of the beams, shown in Figures

6.1 and 6.2. The load-deflection curves do not remain linear up to the limit load. Beyond

a certain limit, the slope of the load-deflection curve starts decreasing until vanishing at

162
Figure 6.4 – Extreme compression fiber strains of B44-1 and B44-2 from major-axis
bending

Figure 6.5 – Extreme compression fiber strains of B36L-1 and B36L-2 from major-axis
bending

163
the ultimate load. The continuous decrease in the slope of the load-deflection curve refers

to the reduction in the lateral bending rigidity of the beam. This reduction is not related to

the increase in major-axis bending strains since the linear load-strain relationship of each

beam is preserved up to buckling, unlike the load-lateral deflection relationship.

Figures 6.6 and 6.7 depict the minor-axis bending strains on the convex, tension,

faces of the specimens at the top. The load-strain curves in the figures do not start from

the origin. The strain at zero load corresponds to the initial strain created by the minor-

axis curvature associated with the sweep. Each curve ends at the point corresponding to

the limit load carried by the specimen prior to buckling. Both figures indicate that the

minor axis bending strains of the companion beams at the limit load were approximately

equal to each other. Furthermore, a comparison of Figures 6.6 and 6.7 with Figures 6.1

and 6.2 indicates that the load-minor axis bending strain curve of each beam has the same

character as the load-lateral deflection curve of the beam. The same character of the two

curves implies that the reduction in the lateral bending rigidity of a beam, which leads to

instability, is related to the increase in the strains from the minor-axis bending moments.

Moreover, the approximately equal values of the in-plane bending strains of the

companion beams at limit load imply that a concrete beam loses its stability when the

minor-axis bending strains in the beam reach certain levels. Therefore, the experimental

results of the present study agree with the statements of Burgoyne and Stratford (2001),

who associated the instability failure of an imperfect concrete beam with the reduction in

its lateral bending rigidity due to the cracking of concrete caused by the increase in the

strains from minor-axis curvature as the load increases.

164
Figure 6.6 –Extreme top strains on the convex faces of B44 caused by minor-axis
bending

Figure 6.7 –Extreme top strains on the convex faces of B36L caused by minor-axis
bending

165
In the present study, an equation was developed to calculate the ultimate load-

carrying capacity of a reinforced concrete beam with initial out-of-plane deformations.

The limit load of an imperfect beam is obtained by reducing the critical load of a perfect

beam in an amount equal to the influence of sweep to the load-carrying capacity of the

beam. A similar approach was previously used by Burgoyne and Stratford (2001) to

obtain the minor-axis curvature of a beam under a certain load.

To understand the following discussion, twisting angles in a beam accompanying

the lateral deformations should be taken into consideration. A geometrically imperfect

beam experiences twisting rotations and lateral deformations prior to buckling when

loaded. Owing to the twisting rotations, the major and minor axes of a beam rotate about

the longitudinal axis passing through the centroid of the cross-section (Figure 6.8). ν,

which will be used in the following equations, is the lateral deflection of the centroid of

the midspan section in the direction of the major axis of the twisted configuration of the

section (x”x” in Figure 6.8). ν is measured from a longitudinal axis passing though the

centroids of the end sections of a beam. In other words, ν measured with respect to the

perfect configuration of a beam.

Burgoyne and Stratford (2001) expressed the load-deflection behavior of a beam

under its self-weight as follows, when the initial lateral imperfection and the critical self-

weight of the beam are known:

o
 (6.6)
w
1
wcr

166
Figure 6.8 – Rotation of the major and minor axes of a section due to twist

where w is the self-weight per unit length of the beam; wcr is the critical self-weight per

unit length, which causes buckling of the beam; νo is the initial imperfection at the center

of the beam in the direction of the major axis of the initial configuration of midspan

section (Figure 6.8).

167
Equation (6.6) is modified to account for the concentrated midspan loading used in the

present study:

o
 (6.7)
P
1
Pcr

where P is the applied load and Pcr is the critical load corresponding to the beam free

from initial geometric imperfections.

Equations (6.6) and (6.7) are based on a method developed by Southwell (1932).

Southwell’s (1932) method is used for obtaining the buckling load and the initial

imperfections of a beam by examining the experimental load and deflection

measurements of the beam under loads much lower than the critical load. Figure 6.9(a) is

the Southwell plot for the lateral deflection data of Specimen B44-1. The inverse slope of

the plot gives the buckling load of the specimen, while the absolute value of the x-

intercept of the plot is the initial lateral centroidal deflection of the beam at midspan. In

Appendix H, Southwell’s (1932) method and the modified versions of the method

proposed by Meck (1977) and Massey (1963) are discussed in more detail and the

application of the three methods to the data obtained in the present experimental program

are described.

Figure 6.9(b) compares the experimental load-lateral deflection curve of

Specimen B44-1 to the analytical curve obtained by using Equation (6.7). Due to the

close agreement between the experimental and analytical curves, Equation (6.7) was used

for expressing the load-deflection behavior of a reinforced concrete beam with initial

geometric imperfections in closed form when developing the limit load (PL) formula.

168
Figure 6.9 - (a) Southwell (1932) Plot; (b) Load-Deflection Plot for Specimen B44-1

169
It is to be noted that P is applied at the top of the beam. The deflection of the

point of application of load (midwidth of the top face) in the direction of the major axis

of the twisted configuration (νt) is related to the deflection of the centroid (ν) according to

Equation (6.8):

h
 t     tan( ) (6.8)
2

where φ is the angle of twist of the beam at midspan corresponding to the load P.

Similarly, the initial deflection of the load application point in the major-axis direction

(νto) is related to the initial centroidal deflection according to Equation (6.9):

h
 to   o   tan(o ) (6.9)
2

where φo is the initial angle of twist of the beam.

Using Southwell’s (1932) method, a relation between φ and φo similar to Equation (6.7)

is obtained:

o
 (6.10)
P
1
Pcr

Secondly, the relation between the vertical applied load and the lateral deflection should

be assessed. The additional lateral deflection of the point of application of load (νt - νto)

is created by the component of the vertical applied load (P) in the direction of the major

axis of the twisted section (P.sinφ):

170
 h  h  P  sin( )  L3
  tan( )   
  o  tan(o )   (6.11)
 2  2 48  Ec  I y

where L is the unbraced length of the beam; Ec is the elastic modulus of concrete and Iy is

the second moment of area about the vertical centroidal axis of the beam section.

Although the minor-axis bending rigidity of a concrete beam can be much lower than EcIy

right before buckling, the use of EcIy in Equation (6.11) was found out to agree much

better with the experimental results of the present study.

Combining Equations (6.7), (6.10) and (6.11) and using the small angle

assumption, the limit load PL is related to the critical load of a beam according to

Equation (6.12):

 o  o  h 2    48  Ec  I y 
PL  Pcr  (6.12)
ult  L3

where φult is the angle of twist of the beam at midspan corresponding to PL. The sweep of

the centroid (uo in Figure 6.8) is related to νo according to Equation (6.13):

uo   o  cos(o ) (6.13)

Using Equation (6.13) and the small angle assumption, Equation (6.12) is modified to

Equation (6.14):

 uo  o  h 2    48  Ec  I y 
PL  Pcr  (6.14)
ult  L3

171
In Chapter VII, analytical load estimates from Equation (6.14) are compared to

the experimental results and the analytical estimates from other methods in the literature.

Furthermore, simplifications to Equation (6.14) for design codes are presented in Chapter

VIII.

172
CHAPTER VII

EXPERIMENTAL RESULTS AND OBSERVATIONS AND


CORRELATION OF THE ANALYTICAL AND EXPERIMENTAL
RESULTS

Section 7.1 presents the crack patterns of the specimens after buckling and some of the

experimental results. Section 7.2 compares the analytical estimates from different

formulations to the experimental limit loads of the specimens measured in the tests.

7.1 Experimental Results and Observations

7.1.1 Cracks Patterns of the Specimens

Restrained shrinkage cracks of the specimens were marked prior to the tests to

distinguish the initial cracks from the cracks formed after the application of load.

Vertical flexural cracks extending through the entire depth of the beam at midspan and

diagonal tension cracks outside the midspan region constituted the typical crack pattern

on the convex faces of the specimens (Figure 7.1) after buckling. Few diagonal cracks in

the vicinity of the end supports and vertical flexural cracks only in the bottom portion of

the beam at and around midspan were observed on the concave faces of the beams after

lateral torsional buckling (Figure 7.2).

All specimens of the experimental program were failed by lateral torsional

buckling. Therefore, similar crack patterns were observed in all specimens. At the initial

stages of loading, flexural cracks initiated and propagated in the tension zone of each

beam around midspan. These vertical cracks were visible both on convex and concave

faces of the beam (Figure 7.3). As the applied load was increased, approaching to the

173
Figure 7.1 –Typical crack pattern on the convex faces of the specimens after buckling

Figure 7.2 –Typical crack pattern on the concave faces of the specimens after buckling

174
Figure 7.3 – Flexural cracks on the concave face of B44-3 at midspan before buckling

critical load value, lateral deflections and rotations in the beam increased and lateral

bending and torsion became more dominant on the crack patterns of the beams. The

flexural tension cracks in the outermost fibers of the tension zone, initiated by the in-

plane flexural moments, extended upwards on the convex side of the beam (Figure 7.4)

due to the tensile strains introduced by lateral bending. On the contrary, the flexural

cracks stopped propagating and closed up to a certain extent on the concave side as a

result of the compressive strains originating from the lateral bending moments (Figure

7.3). Since the lateral bending moment reached its maximum value at midspan, the

vertical flexural cracks extending through the entire depth of the beam were encountered

around midspan on the convex faces of the beams.

175
Figure 7.4 –Vertical cracks on the convex face of B44-2 at midspan after buckling

Cracking outside the midspan region in the test beams was in the form of diagonal

tension cracks resulting from the torsional moments and shear forces due to the large

lateral displacements after buckling. Figures 7.5 and 7.6 illustrate the directions of the

shear and principal stresses in the beams due to the shear forces and torsional moments,

respectively. Direct shear stresses (shear stresses due to the shear forces) coincided with

the shear stresses from torsion on the convex sides of the beams while the direct shear

stresses opposed the shear stresses from torsion on the concave sides. Since the shear

stresses from both sources added up on the convex side, the diagonal tension cracks were

pronounced on the convex faces of the beams (Figure 7.7). Nevertheless, few or no

diagonal cracks could be spotted on the concaves face of the beams (Figure 7.8).

176
Figure 7.6(b) shows the typical torsional moment diagram of the beams tested in

the present study. Ignoring the location of the point of application of load with respect to

the shear center of the midspan section, the torque induced by the lateral deflections in

each beam increased from zero at midspan to a maximum value at the beam ends.

Accordingly, the torsional moments were greater in the vicinity of the laterally-supported

beam ends. The greater shear stresses from torsion around the supports overcame the

direct shear stresses and few diagonal tension cracks became visible on the concave faces

of the beam at and around the supports (Figure 7.8). The diagonal tension cracks on the

concave side in the vicinity of the end supports were in perpendicular direction (reversed)

to the diagonal tension cracks on the convex side due to torsional restraint at the beam

ends. The diagonal cracks on the convex and concave sides were connected to each other

at the top of the beam, where the extensions of the diagonal tension cracks were visible

(Figure 7.9).

177
Figure 7.5 –Directions of the shear and principal stresses due to the shear forces

178
Figure 7.6 –Directions of the shear and principal stresses due to the torsional moments

179
Figure 7.7 –Diagonal tension cracks on the convex face of B36L-1 after buckling

Figure 7.8 –Diagonal tension cracks on the concave face of B18-2 after buckling

180
Figure 7.9 –Diagonal tension cracks continuing on the top surface of B44-2 after

buckling

7.1.2 Experimental Results

The experimental load-lateral (out-of-plane) deflection, load-vertical (in-plane) deflection

and torque-twist curves of the specimens are presented in Appendix D. The midspan

strain distributions of the beams throughout the test are presented in Appendix C.

All specimens of the present study experienced elastic lateral-torsional buckling.

Figures 7.10 to 7.12 illustrate the measured greatest compressive strains of the specimens

at the initiation of buckling on the experimental stress-strain curve of concrete. The point

corresponding to each of the specimens is on the initial portion of the stress-strain curve

of concrete, implying that concrete in each specimen behaved elastically at initiation of

181
buckling. Similarly, the reinforcing bars in the specimens were measured to be strained

within the elastic range of the stress-strain curve of steel (Appendix C).

Figure 7.10 – Maximum compressive strains in the first set of beams, illustrated on the
stress-strain curve of concrete

Figure 7.11 – Maximum compressive strains in B44 at the time of buckling

182
Figure 7.12 – Maximum compressive strains in B36L at the time of buckling

7.2 Correlation of the Analytical and Experimental Results

In Table 7.1, analytical estimates according to four different formulae in the literature and

the formula proposed in the present study are tabulated together with the experimental

limit loads of the specimens. Table 7.2 presents the experimental to analytical load ratios

of the specimens according to the formulae used in Table 7.1. The experimental to

analytical load ratios corresponding to different analytical formulae are also compared in

Figure 7.13 for each specimen. Specimen B44-3 is not included in Tables 7.1 and 7.2 and

Figure 7.13, since the experimental data of this beam, which buckled in the opposite

direction to its sweep, was not considered reliable to be compared to the analytical

estimates from different solutions.

The equations used for obtaining the analytical estimates in Tables 7.1 and 7.2

and Figure 7.13 are presented here. The limit load estimates according to the method

183
proposed in the present study are obtained from Equation (7.1), which was previously

given in Chapter VI [Equation (6.14)]:

16.92  Bo   GC o  e Bo  uto   48  Ec  I y 
PL   1  1.74    (7.1)
L2

 L  GC o 
 sin(ult )  L3

where L is the unbraced length of the beam; Bo is the lateral bending rigidity, obtained

from Equation (7.2); (GC)o is the torsional rigidity, calculated from Equation (7.3); e is

the vertical distance of the load application point from the centroid of the midspan cross

section; uto is the sweep at the top of the beam at midspan; Ec is the elastic modulus of

concrete; Iy is the second moment of area of the beam section about the minor axis; φult is

the angle of twist of the beam at midspan corresponding to the limit load.

 
 3 
  b c  1    Esec  Ec 
Bo     (7.2)
 12   
2
 M cra   c  
  2


 1      1 
  M cr   h  

where b and h are the width and height of the beam, respectively; c is the depth of the

neutral axis from the compression face; Mcra is the cracking moment; Mcr is the critical

moment; Esec is the secant modulus of elasticity of concrete corresponding to the extreme

compression fiber strain at midspan at the instant when Pcr is reached; ω is a constant,

which has a value of 1 in the absence of restrained shrinkage cracks in concrete and a

value of 2/3 in the presence of restrained shrinkage cracks.

Esec  Ec  b3  h  b 
(GC )o    1  0.63    (7.3)
4  1     3  h 

184
where υ is Poisson’s ratio of concrete.

First, the torsional and lateral bending rigidities of a beam should be calculated

from Equations (7.3) and (7.2), respectively. Next, the limit load (Pult) can be calculated

from Equation (7.1), using the calculated values of Bo and (GC)o. Equations (7.2) and

(7.3) correspond to the bending and torsional rigidities of a rectangular reinforced

concrete beam when the beam is subjected to an applied load of Pcr. For example, Esec in

the equations is the secant modulus of elasticity of concrete corresponding to extreme

compression fiber strain at midspan when the maximum in-plane bending moment in the

beam is equal to Mcr. Furthermore, c in Equation (7.2) is the neutral axis depth of the

beam section resisting a moment of Mcr. Considering the terms Esec, c and Mcr in the

rigidity expressions, Equations (7.2) and (7.3) also depend on Equation (7.1). Due to this

interdependence, an iterative procedure is needed to calculate the limit load of a beam. In

Appendix F, the critical moment calculations of one of the specimens are shown as an

example.

The limit load of a beam depends on the girder sweep and the angle of twist at the

instant when the limit load is reached. Table 7.3 tabulates uto and φult values, which were

used in Equation (7.1) to calculate the limit loads of the specimens. uto values in the table

are the sweep values measured at the top of each specimen at midspan. The initial lateral

imperfections of the specimens measured at different points along the length of each

beam are presented in Appendix B. φult values in Table 7.3 on the other hand, are the

twisting angles calculated according to the method described in Section 3.1.3 of Chapter

III from the lateral deflection measurements taken in the tests.

185
Table 7.1 – Experimental and analytical critical load values of the specimens
Analytical Load, Pan (kips)
Experimental Sant &
Specimen Load, Elastic & Homog. Hansell& Present
Material, Pel, Winter (1959), Bletzacker (1961), Massey (1967), Study,
Pex (kips) Psb, Eq. (7.10) Pm, Eq. (7.14)
Eq. (7.4) Phw, Eq. (7.7) Pult, Eq. (7.1)
B18-1 12.4 26.7 10.2 19.8 17.7 9.0
B18-2 12.0 23.7 9.4 17.6 15.6 8.9
B22-1 8.7 33.2 12.8 25.4 21.4 8.0
B22-2 11.0 27.5 10.7 21.0 18.7 *
B30 22.0 86.6 31.9 69.1 51.9 17.0
B36 39.2 101.5 39.1 80.6 67.3 40.6
B44-1 15.2 38.8 13.2 31.2 24.1 15.6
B44-2** 12.0 40.1 13.7 32.6 24.2 7.2
B36L-1 13.5 36.7 13.4 30.0 22.9 11.4
B36L-2 21.7 38.7 14.0 31.7 24.5 18.2
* Sweep of Specimen B22-2 is not known.

** Specimen B44-3 buckled in opposite direction to its sweep. Due to this unusual situation, the experimental data of B44-3
was not considered reliable to compare to the analytical estimates.

186
Table 7.2 – Experimental–to-analytical critical load ratios of the specimens

Elastic & Homog. Hansell& Winter Sant & Bletzacker Massey Present Study,
Specimen
Material, Pex/Pel (1959), Pex/Phw (1961), Pex/Psb (1967), Pex/Pm Pex/Pult
B18-1 0.46 1.22 0.63 0.70 1.38
B18-2 0.51 1.28 0.68 0.77 1.35
B22-1 0.26 0.68 0.34 0.41 1.09
B22-2 0.40 1.03 0.52 0.59 *
B30 0.25 0.69 0.32 0.42 1.29
B36 0.39 1.00 0.49 0.58 0.97
B44-1 0.39 1.15 0.49 0.64 0.97
B44-2** 0.30 0.88 0.37 0.50 1.67
B36L-1 0.37 1.01 0.45 0.59 1.18
B36L-2 0.56 1.54 0.68 0.88 1.19
Mean 0.39 1.05 0.50 0.61 1.23
S.D. 0.10 0.26 0.13 0.15 0.22
COV% 26 25 27 24 18
* Sweep of Specimen B22-2 is not known.

** Specimen B44-3 buckled in opposite direction to its sweep. Due to this unusual situation, the experimental data of
B44-3 was not considered reliable to compare to the analytical estimates.

187
Figure 7.13 – Experimental-to-analytical critical load ratios of the specimens according to different formulae

188
Table 7.3 – Measured sweeps and angles of twist at limit load of the specimens at
midspan

Angle of Twist
Specimen Sweep, uto (in) at Limit Load
φult (deg)
B18-1 7/16 1.17
B18-2 1/8 0.45
B22-1 11/16 1.66
B30 5/8 0.86
B36 7/32 0.52
B44-1 9/16 0.77
B44-2 25/32 0.55
B36L-1 3/4 0.70
B36L-2 11/32 0.65

The column denoted as the elastic and homogeneous material in Table 7.1 gives

the analytical values calculated by assuming that a reinforced concrete beam is a

homogeneous and elastic body, not subjected to cracking under loads. Regarding these

assumptions, the ultimate load (Pel) of a concrete beam can be calculated from Equation

(7.4):

16.92  Bel   GC u  e Bel 


Pel    1  1.74    (7.4)
L2 

L  GC u 

where Bel is the lateral flexural rigidity of an elastic and homogeneous beam with a solid

rectangular cross section; (GC)u is the uncracked torsional rigidity of a reinforced

concrete beam (Equation 5.9 in Chapter V). Bel and (GC)u are calculated from Equations

(7.5) and (7.6), respectively:

189
b3  h
Bel  Ec  (7.5)
12

Ec
(GC )u    c  b3  h (7.6)
2  1  

where βc is the coefficient for St. Venant’s torsional constant, obtained from Equation

(5.11).

Although Equations (7.4) - (7.6) do not reflect the true behavior of a reinforced

concrete beam, the results obtained from Equation (7.4) are included in Table 7.1 to

compare the limit load of each specimen to the limit load of a homogeneous and elastic

beam with the same dimensions as the specimen.

The analytical ultimate load estimates according to Hansell and Winter (1959)

were obtained from Equations (7.7):

16.92  Bhw   GC hw  e Bhw 


Phw   1  1.74    (7.7)
L2 

L  GC hw 

where Bhw and (GC)hw are the lateral flexural and torsional rigidities according to the

expressions proposed by Hansell and Winter (1959), given as

b3  c
Bhw  Esec  (7.8)
12

Esec  b3  c  b 
2
 GC hw    1  0.35    (7.9)
2  1    3  d  

where d is the effective depth of the beam from the compression face.

190
The ultimate load values according to Sant and Bletzacker (1961) are obtained

from the following equation:

16.92  Bsb   GC  sb  e Bsb 


Psb   1  3.48    (7.10)
L2 

L  GC sb 

where Bsb and (GC)sb are the lateral flexural and torsional rigidities according to the

expressions proposed by Sant and Bletzacker (1961), given as

b3  d
Bsb  Er  (7.11)
12

Er  b3  d 
 GC sb     (7.12)
2  1    3 

where Er is the reduced modulus of elasticity of concrete corresponding to the extreme

compression fiber strain at midspan at the instant when the applied load is equal to Psb. Er

is calculated from

4  Ec  Etan
Er  (7.13)
 
2
Ec  Etan

where Etan is the tangent modulus of elasticity of concrete corresponding to the extreme

compression fiber strain at midspan at the instant when the applied load is equal to Psb.

The expression in parenthesis in Equation (7.10), which corresponds to the

destabilizing or stabilizing effect of the load applied above or below the centroid of the

midspan section, is different from the respective expression in Equations (7.1) and (7.7).

In their study, Sant and Bletzacker (1961) considered the influence of the distance of the

191
load application point from the centroid of the section and included the expression in

parenthesis in Equation (7.10) to account for this influence. In the present study, the

original critical moment expression [Equation (7.10)] proposed by Sant and Bletzacker

(1961) was used together with the torsional and lateral bending rigidity expressions

developed by Sant and Bletzacker (1961).

Finally, the ultimate load values according to Massey’s (1967) formulation were

obtained from the following equation:

16.92  Bm   GC m  e Bm 
Pm    1  1.74    (7.14)
L2 

L  GC m 

where Bm and (GC)m are the lateral flexural and torsional rigidities according the

expressions proposed by Massey (1967), given as

b3  c
Bm  Esec   Es  I sy (7.15)
12

 Gc'  c  b3  h  1   Gs  Gc'   bs3  ts    b1  d1  Ao  Es


2
 GC m (7.16)
3 2 2  s

where ΣIsy is the moment of inertia of the longitudinal steel about the minor axis of the

section; bs and ts are the width and thickness of the longitudinal reinforcement layer,

respectively (illustrated in Figure 1.18 in Chapter I); γ is a constant defined by Cowan

(1953); b1 and d1 are the breadth and depth of the cross-sectional area enclosed by the

closed stirrup, respectively (Figure 1.18); s is the spacing of the stirrups; Ao is the cross-

sectional area of one leg of the stirrup; Es and Gs are the modulus of elasticity and

192
modulus of rigidity of steel, respectively; G’c is the reduced modulus of rigidity of

concrete, calculated according to the following equation:

Esec
Gc'  (7.17)
2  (1   )

Tables 7.1 and 7.2, and Figure 7.13 indicate that the analytical estimates produced

by the proposed method are in good agreement with the experimental results.

Furthermore, the analytical estimates from the proposed method never exceeded the

experimental ultimate loads, with the exception of Specimens B36 and B44-1, for which

the analytical estimates are only 3-4% greater than the experimental values. The

experimental to analytical load ratios corresponding to the proposed method were in the

range of 0.97-1.67 with a mean value of 1.23 and coefficient of variation of 0.18 (Table

7.2).

Among the formulae proposed by the previous researchers, the equation given by

Hansell and Winter (1959) closely estimated the limit loads of the specimens with the

experimental to analytical load ratios in the range of 0.68-1.54 (a mean value of 1.05 and

coefficient of variation of 0.25). The formulae proposed by Sant and Bletzacker (1961)

and Massey (1967) constantly overestimated the limit loads of the specimens. The

analytical estimates from Sant and Bletzacker’s (1961) formula sometimes reached 2.5-3

times the limit loads measured in the tests.

As mentioned before, uo and φult are needed to calculate the limit load of a beam

according to Equation (7.1). φult, in particular, is a quantity which is determined by

testing a beam to failure. Since testing a beam to failure is not always possible,

particularly in a real construction, a value for φult should be assumed in the limit load

193
calculations. In Chapter VIII, Equation (7.1) is modified by assuming constant values for

φult to simplify the equation.

194
CHAPTER VIII

SUMMARY AND CONCLUSIONS

8.1 Summary

The present study investigated the lateral stability of rectangular reinforced concrete

beams both analytically and experimentally. The experimental part of the study provided

the experimental results of reinforced concrete beams, whose initial geometric

imperfections, shrinkage cracking conditions and material properties are completely

known. In the analytical part of the study, the following formula was developed for

estimating the limit loads (PL) of simply-supported rectangular reinforced concrete beams

with initial geometric imperfections, subjected to a concentrated load at midspan:

4  M cr uto   48  Ec  I y 
PL   (8.1)
L sin(ult )  L3

where PL is the limit load; L is the unbraced length of the beam; uto is the sweep at the top

of the beam at midspan; Ec is the elastic modulus of concrete; Iy is the second moment of

area of the beam section about the minor axis; φult is the angle of twist of the beam at

midspan corresponding to the limit load (PL). Mcr is the critical moment corresponding to

the geometrically perfect configuration of the beam, obtained from Equation (8.2):

4.23  e Bo 
M cr   1  1.74     B   GC o (8.2)
L 

L  GC o  o

195
where Bo is the lateral bending rigidity, obtained from Equation (8.3); (GC)o is the

torsional rigidity, calculated from Equation (8.4); e is the vertical distance of the load

application point from the centroid of the midspan cross section.

 
 3 
  b c  1    Esec  Ec 
Bo    (8.3)
 12   
2
 M cra   c  
  2


 1      1 
  M cr   h  

Esec  Ec  b3  h  b 
(GC )o    1  0.63    (8.4)
4  1     3  h 

where b and h are the width and height of the beam, respectively; c is the depth of the

neutral axis from the compression face; Mcra is the cracking moment, obtained from

Equation (8.5); ω is a constant, which has a value of 1 in the absence of restrained

shrinkage cracks in concrete and a value of 2/3 in the presence of restrained shrinkage

cracks and υ is Poisson’s ratio of concrete. Esec is the secant modulus of elasticity of

concrete corresponding to the extreme compression fiber strain at midspan at the instant

when Mcr is reached. Esec is calculated from Equation (8.6).

M cra 

I ucr  7.5  f c'  (8.5)
h y

where 7.5  f c' is the modulus of rupture of normal-weight concrete, given in ACI 318-

05 (2005) Section 9.5.2.3; y is the depth of the center of gravity of the transformed

196
section from the top surface of the beam; Iucr is the moment of inertia of the transformed

section about the major axis, obtained from Equation (8.7).

fc
Esec  (8.6)
c

where fc and εc are the extreme compression fiber stress and extreme compression fiber

strain at midspan corresponding to the critical moment Mcr.

2
1  h
  b  h3  b  h   y     n  1  As   d  y 
2
I ucr (8.7)
12  2

where As is the cross-sectional area of the flexural reinforcement; n is the modular ratio of

steel to concrete; d is the effective depth of the centroid of tension reinforcement from

compression face. When calculating the uncracked moment of inertia, Iucr, the flexural

reinforcement is transformed into an equivalent concrete area in accordance with the

modular ratio of steel to concrete, n.

Due to the interdependence of Mcr, Bo and (GC)o, an iterative process is needed

when calculating the limit load of a beam. This interdependence is also present in the

methods proposed by Hansell and Winter (1959), Sant and Bletzacker (1961) and Massey

(1967). The primary reason for the interdependence of the critical moment and the lateral

bending and torsional rigidities is that the modulus of elasticity and modulus of rigidity

terms in the rigidity expressions depend on the maximum compressive strain in the beam

corresponding to the critical moment, as explained in Appendix F in more detail.

The experimental stage of the study consisted of testing eleven reinforced

concrete beams with d/b ratios between 10.20 and 12.45 and L/b ratios between 96 and

197
156. Beam thickness, depth and unbraced length were 1.5 to 3.0 in., 18 to 44 in., and 12

to 39.75 ft, respectively. The test beams were simply-supported in and out of plane and

subjected to a single concentrated load at midspan. The end supports allowed warping

deformations in the beams while restraining the torsional rotations at the beam ends. The

loading mechanism used in the tests provided lateral translational and rotational freedom

at the load application point and ensured that the vertical orientation of the applied load

was maintained throughout the entire test. The in-plane (vertical) and out-of-plane

(lateral) deformations, the torsional rotations and the strain distributions in the beams

were measured at midspan, continuously along the tests.

There are several factors influencing the lateral stability of reinforced concrete

beams. Initial geometric imperfections, shrinkage cracking, contribution of the

longitudinal and shear reinforcement to the torsional and lateral bending rigidities, creep,

inelastic stress-strain properties of concrete and reinforcing steel, loading and support

conditions are some of the major factors affecting the lateral stability. Investigating the

influences of several factors in the same experiments renders the analysis and

interpretation of the experimental data cumbersome. In the present experimental program,

the main factor whose effects were investigated is the initial lateral imperfections (sweep)

of a beam. To detect the effects of sweep, the influences of some of the other factors were

minimized or eliminated through the following ways:

 The specimens of the present experimental program were designed in such a way that

both concrete and reinforcing steel in the beams remained in the elastic ranges of their

stress-strain curves throughout the loading process. By eliminating the inelastic

stress-strain properties of concrete and steel, pure elastic material behavior was

198
attained and the effects of inelasticity on the lateral stability of the specimens were

eliminated. Nevertheless, the lateral and torsional rigidity expressions proposed in the

present study account for both elastic and inelastic material behaviors of concrete and

reinforcing steel to be applicable for all reinforced concrete beams with different

dimensions, reinforcement details and material properties.

 Restrained shrinkage cracking of concrete was tried to be prevented through the

measures described in Section 4.4. Addition of Shrinkage Reducing Admixtures

(SRA) to concrete and early removal of the beams from the forms minimized the

amount and extent of shrinkage cracking in the second set of specimens.

 The specimens were designed in a way that the contributions of the longitudinal and

shear reinforcement to the lateral stability of the beams were negligible. The

longitudinal reinforcing bars were located along the vertical centroidal axis of the

beam section. Presuming that the minor axis of the beam is coincident with the

vertical centroidal axis throughout the test, the longitudinal reinforcement did not

contribute to the lateral bending rigidities of the beams since the second moment of

area of the longitudinal reinforcement (ΣIsy) is equal to zero in this design.

Furthermore, the specimens did not experience diagonal tension cracking up to

buckling. Prior to the formation of diagonal tension cracks, a reinforced concrete

beam behaves as a solid and homogeneous body, whose torsional rigidity is provided

by the entire cross-section. In the pre-cracking stage, contributions of the flexural and

shear reinforcement to the torsional rigidity are negligible. Due to the absence of

torsional cracks in the beams up to buckling, contribution of the reinforcement to the

resistance of torsional moments was disregarded.

199
 All specimens of the experimental program were tested under identical loading and

support conditions. Consequently, the effects of the loading and support conditions on

the test results of companion beams (beams with identical nominal dimensions,

reinforcement details and material properties) were eliminated.

 Concrete from the same batch and reinforcing steel from the same batch were used in

the companion beams to keep the material properties constant so that the

experimental results of the companion beams were not affected from the differences

between the mechanical properties of the beams.

 To eliminate the influence of creep on the lateral stability, the test beams were loaded

continuously up to the ultimate load without interruption.

Although the experimental program aimed at eliminating the influences of the

aforementioned factors on the lateral stability of the specimens, all factors affecting the

lateral stability of a reinforced concrete beam were taken into consideration in the

development of the proposed analytical method. For instance, Equation (4.18) in Chapter

IV gives the lateral bending rigidity of a reinforced concrete beam, considering the

influence restrained shrinkage cracking. For the possible contribution of the longitudinal

reinforcing bars, on the other hand, Equation (4.16) was developed. Furthermore, both

torsional and lateral bending rigidity expressions developed in the present study are

applicable to the cases of elastic and inelastic lateral-torsional buckling.

8.2 Conclusions

Some of the conclusions drawn from the present study for the lateral stability of

rectangular reinforced concrete beams follow:

200
 The load predictions from the proposed analytical method showed good correlation

with the experimental results. The analytical to experimental limit load ratios of the

specimens ranged from 0.97 to 1.67 for the proposed analytical method. The

analytical method developed in the present study is superior to the analytical methods

proposed by Hansell and Winter (1959), Sant and Bletzacker (1961) and Massey

(1967) in incorporating the effects of sweep, shrinkage cracking and inelastic stress-

strain properties of concrete into the buckling formula.

 Among the former methods considered in the present study, the analytical method

proposed by Hansell and Winter (1959) produced better correlation with the

experimental results than the methods proposed by Sant and Bletzacker (1961) and

Massey (1967), which constantly overestimated the ultimate loads of the beams.

 In contrast to the methods in the literature, the load estimates produced by the

proposed method were smaller or slightly larger than the experimental buckling loads

of the specimens (Table 7.2), which makes the proposed method more conservative

than the methods proposed by Hansell and Winter (1959), Sant and Bletzacker (1961)

and Massey (1967). Even some of the load estimates from the method proposed by

Hansell and Winter (1959), which was in closer agreement with the experimental

results than the methods proposed by Sant and Bletzacker (1961) and Massey (1967),

were greater than the experimental buckling loads of the specimens.

 In the case of elastic lateral torsional buckling, reinforced concrete beams do not

undergo diagonal tension cracking before buckling. Therefore, the uncracked

torsional rigidity closely reflects the torsional resistance of a reinforced concrete

beam at the instant of buckling, if the beam buckles elastically. Moreover,

201
contributions of the longitudinal and shear reinforcement to the torsional rigidity of a

reinforced concrete beam is negligible in elastic buckling since the reinforcement of a

concrete beam has a major influence on the torsional behavior of a concrete beam

only after diagonal cracking.

 The influence of the longitudinal reinforcement on the lateral stability of a concrete

beam originates from its contribution to the lateral bending rigidity. Similarly,

restrained shrinkage cracking of concrete affects the lateral stability of a beam by

reducing its lateral bending rigidity. Lateral bending rigidity expression given in

Equation (4.17) considers the increase in the lateral bending rigidity of a reinforced

concrete beam due to the contribution of the flexural reinforcement while the rigidity

expression given in Equation (4.16) was developed regarding the negative influence

of shrinkage cracking of concrete on the lateral bending rigidity.

 A geometrically imperfect concrete beam does not reach the critical moment

corresponding to the geometrically perfect configuration of the beam. The additional

stresses originating from the initial minor-axis curvature from sweep cause the

imperfect beam to crack earlier in the loading process (at smaller load levels).

Cracking on the convex side of the beam greatly reduces the lateral bending rigidity

and the load-carrying capacity of the beam starts decreasing before reaching the

critical load. The maximum load on the load-lateral deflection curve of an imperfect

beam is denoted as the limit load, which should be distinguished from the critical

load. The limit load formula proposed in the present study is based on reducing the

critical load by an amount equal to the destabilizing effect of the sweep on the beam.

202
 Results of the present experimental program indicated that an increase in the sweep

can greatly reduce the ultimate load resisted by a reinforced concrete beam before

losing its stability. The significant differences between the ultimate loads of the

companion beams mainly originated from the different initial lateral imperfections of

the beams. For instance, the experimental ultimate load of Specimen B36L-2 (11/32

in sweep) was 38% greater than the experimental ultimate load of Specimen B36L-1

(3/4 in sweep).

 To calculate the limit load of an imperfect reinforced concrete beam from Equation

(7.1), the angle of twist (φult) of the beam needs to be known. φult is a parameter

indicating the torsional rotations in a beam till buckling and it is determined by

testing a beam to failure. In the design of a concrete beam, a value for φult should be

assumed. The φult values of the test specimens ranged from 0.45 degrees to 1.66

degrees. However, most of the specimens had a midspan angle of twist between 0.55

degrees to 0.75 degrees at the limit load level. As φult decreases, the reduction in the

limit load of a beam increases. Therefore, assuming a value between 0.55 and 0.60

degrees for φult seems reasonable and safe according to the available experimental

data. However, more experimental data is needed to make more accurate

assumptions.

8.3 Future Research

The analytical method, proposed in the present study, incorporates several factors

influencing the lateral stability of reinforced concrete beams. The factors with

considerable influence were restated in Section 8.1. The experimental stage of the study,

nonetheless, aimed at investigating the effects of the initial lateral imperfections on the

203
buckling behavior of reinforced concrete beams. To achieve this goal, the effects of the

other influential factors were tried to be minimized, if not eliminated. Hence, further

experiments are needed to investigate the effects of the factors, which were disregarded

in the present experimental program. For example, all specimens of the experimental

program were failed by elastic lateral torsional buckling. Further experiments on

reinforced concrete beams subject to inelastic lateral torsional buckling are needed to

investigate the degree of agreement between the experimental ultimate loads of the beams

and the estimates produced by the proposed analytical method in the case of inelastic

lateral torsional buckling. Similarly, the accuracy of the analytical estimates needs to be

explored experimentally when the longitudinal reinforcement of a concrete beam

contributes to the lateral stability to a major extent. As stated before, the contribution of a

longitudinal rebar to the lateral bending rigidity of a beam increases as the lateral distance

of the bar from the minor axis increases. Further experiments on reinforced concrete

beams with longitudinal reinforcing bars distributed along the sides of the beams can be

useful to examine the accuracy of the estimates from the proposed analytical method

when the reinforcement plays an important role in resisting the lateral bending moments.

Finally, more experimental data is needed to determine the common values of the

twisting angle at limit load (φult) of reinforced concrete beams, which is used in Equation

(8.1) for calculating the limit load of a reinforced concrete beam with initial geometric

imperfections. Based on a statistical analysis on the results of a large number of lateral

torsional buckling tests of reinforced concrete beams, the appropriate values of φult can be

determined and recommended in the structural concrete codes for using in Equation (8.1).

204
APPENDIX A

NOMINAL AND MEASURED DIMENSIONS OF THE SPECIMENS

The actual dimensions of a concrete beam can be significantly different than the nominal

dimensions. For a more precise analytical study, the actual dimensions of a beam,

determined from several measurements throughout the beam, should be used when

calculating the analytical critical loads. In this concern, the dimensions of the specimens

were measured at several locations throughout each beam. The tables of the present

appendix tabulate the nominal and measured dimensions of the specimens together with

the means (μ), standard deviations (σ) and the percent coefficients of variation (cv %) of

the measurements. Standard deviations and coefficients of variation of the measurements

are included in the tables to reflect the degree of variation in the measurements of a

dimension.

Table A.1 and A.2 present the heights of the specimens, measured at several

locations along the lengths of the beams. The locations of the measurements are shown in

Figure A.1.

Tables A.3 and A.5, on the other hand, present the widths of the specimens,

measured at several locations along the depth and length of each beam. The locations of

the measurement points are shown in Figure A.2.

Each of the first set of beams was cut at mid-length after the test to determine the

actual locations of the longitudinal reinforcing bars in the beams. The widths of the

beams measured along the cuts are tabulated in Table A.4.

205
Table A.6 presents the unbraced lengths of the first set of beams. Table A.7, on

the other hand, presents the total lengths of the second set of beams, measured at five

different depths, as shown in Figure A.3. The unbraced span lengths of B44 and B36L

were measured as 39.33 ft.

Table A.1 – Nominal and measured heights of the first set of specimens

Specimen B 36 B 30 B 22-1 B 22-2 B 18-1 B 18-2


Nominal Height (in) 36.000 30.000 22.000 22.000 18.000 18.000
Measurement
Point
(Fig.A.1a&b)
1 35.906 30.000 22.000 22.063 18.000 18.000
2 35.969 30.063 22.031 22.125 18.094 18.094
Measured Height (in)

3 36.063 30.063 22.031 22.094 18.156 18.125


4 36.094 30.031 22.000 22.063 18.125 18.094
5 36.063 29.969 21.969 22.063 18.125 18.063
6 36.031 29.875 22.000 22.094 18.063 18.094
7 36.031 29.938 22.000 22.094 18.125 18.094
8 36.000 29.875 22.031 22.000 18.063 18.000
9 36.031 29.938 22.000 22.000 18.125 18.000
10 36.094 30.000 22.000 22.000 18.156 18.000
11 35.875 29.969 21.938 22.188 18.000 18.188
μ 36.014 29.975 22.000 22.071 18.094 18.068
σ 0.068 0.063 0.026 0.055 0.053 0.059
cv % 0.19 0.21 0.12 0.25 0.29 0.33

206
Table A.2 – Nominal and measured heights of the second set of beams

Specimen B44-1 B44-2 B44-3 B36L-1 B36L-2


Nominal Height (in) 44 44 44 36 36
Measurement Point
(Figure A.1c)
1 44 44 44 1/16 36 5/32 36 1/16
2 44 43 31/32 44 1/16 36 36
3 44 1/8 44 43 15/16 36 3/32 36 1/8
4 44 1/16 44 1/32 44 3/32 36 1/32 36 1/32
5 44 3/32 44 44 36 36
6 44 1/8 44 1/32 44 1/16 36 1/32 36
7 44 3/16 44 44 1/8 36 36 1/16
Measured Height (in)

8 44 1/16 44 44 5/32 35 7/8 36


9 44 1/8 43 29/32 44 3/32 36 35 7/8
10 44 1/8 43 29/32 44 3/32 36 36 1/16
11 44 43 31/32 44 3/32 36 1/16 36
12 43 15/16 44 1/32 44 1/8 36 36
13 43 7/8 44 1/32 44 3/32 36 1/8 36 1/16
14 43 27/32 44 1/32 44 3/16 36 36 1/32
15 43 13/16 44 1/16 43 7/8 36 1/8 36 1/16
16 43 13/16 44 44 36 1/8 36 1/16
17 43 13/16 43 29/32 44 3/32 36 3/32 36
18 43 25/32 44 1/64 44 1/16 36 1/8 36 1/32
19 43 13/16 44 3/32 44 36 3/32 36
20 43 27/32 44 1/8 44 1/32 36 1/16 36 1/16
21 43 15/16 44 1/8 44 1/32 36 1/32 36
μ 43 31/32 44 1/64 44 1/16 36 3/64 36 1/32
σ 1/8 1/16 5/64 1/16 3/64
% cv 0.30 0.17 0.16 0.18 0.13

207
Figure A.1–Height measurement points along the lengths of the beams
(All dimensions are in feet)

208
Table A.3 – Nominal and measured widths of the first set of specimens along the span

Specimen B 36 B 30 B 22-1 B 22-2 B 18-1 B 18-2


Nominal Width (in) 2.500 2.500 1.500 1.500 1.500 1.500
Meas. Point
(Fig. A.2a&b)
1 2.438 2.563 1.578 1.563 1.563 1.484
2 2.500 2.500 1.578 1.563 1.578 1.500
3 2.438 2.438 1.563 1.563 1.578 1.453
Measured Width (in)

4 2.438 2.500 1.594 1.547 1.516 1.500


5 2.469 2.531 1.625 1.563 1.547 1.547
6 2.469 2.438 1.531 1.469 1.547 1.609
7 2.531 2.469 1.578 1.500 1.594 1.594
8 2.469 2.500 1.547 1.500 1.516 1.516
9 2.438 2.500 1.609 1.531 1.516 1.516
10 2.438 2.531 1.500 1.531 1.516 1.516
11 2.469 2.500 1.453 1.516 1.516 1.594
12 2.406 2.500 1.547 1.484 1.516 1.578
μ 2.458 2.498 1.559 1.527 1.542 1.534
σ 0.032 0.035 0.046 0.032 0.029 0.045
cv % 1.301 1.401 2.951 2.096 1.881 2.934

Table A.4 –Widths of the first set of specimens along the depth of midspan section

Specimen B 36 B 30 B 22-1 B 22-2 B 18-1 B 18-2


Nominal Width 2.500 2.500 1.500 1.500 1.500 1.500
1 2.575 2.586 1.576 1.447 1.567 1.632
2 2.595 2.540 1.630 1.469 1.557 1.647
3 2.617 2.537 1.590 1.474 1.550 1.643
4 2.655 2.557 1.583 1.495 1.529 1.592
5 2.652 2.583 1.588 1.497 1.539 1.611
6 2.617 2.567 1.565 1.512 1.535 1.605
Measured Height (in)

7 2.643 2.614 1.549 1.524 1.547 1.612


8 2.616 2.622 1.540 1.494 1.587
9 2.641 2.639 1.517
10 2.643 2.660 1.527
11 2.613 2.637
12 2.605 2.619
13 2.602
14 2.552
15 2.550
16 2.496
17 2.496
18 2.498
19 2.490
μ 2.587 2.597 1.578 1.496 1.546 1.616
σ 0.056 0.039 0.026 0.025 0.012 0.016
cv % 2.147 1.508 1.652 1.646 0.788 0.990

209
Figure A.2–Width measurement points along the lengths of the specimens
(All dimensions are in feet)

210
Table A.5 – Nominal and measured widths of the second set of beams

Specimen B44-1 B44-2 B44-3 B36L-1 B36L-2


Nominal Width (in) 3.000 3.000 3.000 3.000 3.000
Measurement Point
(Figure A.2c)
1 3.015 3.000 3.028 3.111 3.030
2 3.038 3.000 3.049 3.080 3.037
3 3.070 3.063 3.100 3.116 3.068
4 3.028 3.031 3.077 3.102 3.162
5 3.094 3.000 3.070 3.070 3.060
6 3.043 3.012 3.074 3.287 3.150
7 3.000 3.003 3.042 3.283 3.280
8 3.051 2.999 3.105 3.159 3.189
9 3.063 3.043 3.017 3.125 3.346
10 3.038 3.002 3.011 3.206 3.336
11 3.067 2.996 3.084 3.325 3.165
12 3.032 3.031 3.144 3.265 3.268
13 2.997 3.054 3.005 3.150 3.303
14 2.963 3.039 3.009 3.172 3.297
15 3.033 2.960 2.909 3.159 3.238
16 3.012 3.055 3.054 3.170 3.244
17 3.031 3.028 3.086 3.130 3.333
18 3.103 3.037 3.122 3.144 3.179
19 3.149 3.009 3.042 3.143 3.218
Measured Width (in)

20 3.019 3.015 3.016 3.270 3.164


21 3.042 2.982 3.027 3.163 3.214
22 3.063 2.992 3.127 3.093 3.298
23 3.055 3.072 3.034 3.245 3.273
24 3.067 3.036 3.040 3.219 3.337
25 3.107 3.122 3.061 3.160 3.147
26 3.109 3.122 2.973 3.283 3.266
27 3.064 3.042 2.859 3.284 3.296
28 3.051 3.050 2.892 3.205 3.186
29 3.036 3.120 2.982 3.281 3.103
30 3.106 3.094 3.031 3.258 3.365
31 3.070 3.074 3.096 3.142 3.125
32 3.022 3.121 3.115 3.121 3.111
33 3.033 3.022 3.090 3.245 3.066
34 3.053 3.056 3.100 3.216 3.195
35 3.095 3.026 3.101 3.239 3.105
36 3.076 3.100 3.137 3.171 3.065
37 3.065 3.006 3.111 3.280 3.070
38 3.085 3.012 3.020 3.123 3.107
39 3.025 3.084 3.089 3.207 3.062
40 3.073 2.982 2.993 3.161 3.162
41 3.082 3.096 3.048 3.061 3.166
42 2.998 3.083 3.115 3.068 3.197
43 3.054 3.037 2.988 3.192 3.109
44 3.128 3.031 3.001 3.310 3.243
45 2.944 3.063 3.072 3.190 3.145
46 3.047 3.156 3.128 3.149 3.208
47 3.046 3.156 3.180 3.160 3.226
48 3.017 3.188 3.147 3.143 3.185
μ 3.051 3.048 3.054 3.184 3.190
σ 0.039 0.050 0.068 0.070 0.091
% cv 1.3 1.6 2.2 2.2 2.8

211
Table A.6 – Nominal and measured span lengths of first set of specimens

Specimen B 36 B 30 B 22-1 B 22-2 B 18-1 B 18-2


Beam Length (in) 252.000 252.000 156.000 156.000 156.000 156.000
Span (in) 240.000 240.000 144.000 144.000 144.000 144.000
Measurement
Depth
(Figure A.3)
1 239.500 239.875 143.750 143.750 143.750 143.719
239.375 239.813 143.750 143.781 143.781 143.688
Measured

2
Span (in)

3 239.375 239.750 143.750 143.750 143.750 143.688


4 239.375 239.688 143.750 143.813 143.781 143.688
5 239.500 239.688 143.750 143.813 143.750 143.688
μ 239.425 239.763 143.750 143.781 143.763 143.694
σ 0.061 0.073 0.000 0.028 0.015 0.012
cv % 0.025 0.030 0.000 0.019 0.010 0.008

Table A.7 – Nominal and measured total lengths of the second set of specimens

Specimen B44-1 B44-2 B44-3 B36L-1 B36L-2


Nominal Length (ft) 39.75 39.75 39.75 39.75 39.75
Measurement
Depth
(Figure A.3)
1 39.76 39.75 39.75 39.78 39.78
Length (ft)
Measured

2 39.77 39.74 39.76 39.79 39.77


3 39.78 39.74 39.76 39.79 39.77
4 39.77 39.75 39.77 39.80 39.78
5 39.77 39.75 39.77 39.79 39.78
μ 39.77 39.75 39.76 39.79 39.78
σ 0.01 0.01 0.01 0.01 0.01
% cv 0.02 0.01 0.02 0.02 0.01

Figure A.3–Length measurement depths of the specimens

212
APPENDIX B

MEASURED INITIAL GEOMETRIC IMPERFECTIONS AND


PERMANENT DEFORMATIONS OF THE SPECIMENS

The initial lateral deformations, sweeps, of the test beams are tabulated in Tables B.1-

B.4. The tables present the lateral deformations at the extreme top, mid-height and

extreme bottom of each beam. Sweep was measured at three different levels along the

depth of each beam to obtain the initial angles of twist of the specimens. Furthermore, the

initial lateral deflections of the beams were measured at several points along the length of

each beam to determine the exact curved shapes of the specimens out of plane.

Knowledge of the lateral bows of the test beams prior to loading was essential in the

analytical stage of the present study. The number and locations of the measurement

points along the lengths of the beams are illustrated in Figure B.1. Southward deflections

in Tables B.1-B.4 are positive while the northward ones are negative according to the

sign convention shown in Figure B.1.

213
Table B.1 – Measured sweeps of Specimens B18

Measurement Sweep (in)


Point along
B18-1 B18-2
the Span
(Figure B.1a) Top Bottom Top Midheight Bottom
1 -1/8 -3/16 -1/32 -1/32 -1/32
2 -5/16 -5/16 -1/16 -1/16 -1/16
3 -7/16 -7/16 -1/8 -1/8 -1/8
4 -7/16 -7/16 -1/8 -1/8 -1/8
5 -7/16 -7/16 -1/8 -1/8 -1/8
6 -7/16 -7/16 -3/32 -3/32 -3/32
7 -1/4 -5/16 -1/16 -1/16 -1/16

Table B.2 – Measured sweeps of Specimens B30 and B36

Measurement Sweep (in)


Point along
B30 B36
the Span
(Figure B.1b) Top Midheight Bottom Top Midheight Bottom
1 -1/4 -3/16 -3/16 1/32 1/64 1/32
2 -7/16 -3/8 -1/4 1/8 5/32 1/8
3 -5/8 -9/16 -1/2 1/8 5/32 1/8
4 -5/8 -9/16 -1/2 7/32 3/16 7/32
5 -5/8 -9/16 -1/2 7/32 3/16 7/32
6 -1/2 -7/16 -3/8 7/32 3/16 7/32
7 -1/4 -3/16 -1/8 1/8 3/32 1/8

214
(a)

(b)

(c)

Figure B.1 – Imperfection measurement points on beams (a) B18; (b) B30 and B36;
(c) B44 and B36L (All dimensions are in feet.)

215
Table B.3 – Measured sweeps of Specimens B44

Measurement Sweep (in)


Point along
B44-1 B44-2 B44-3
the Span
(Figure B.1c) Top Midheight Bottom Top Midheight Bottom Top Midheight Bottom
1 1/16 1/8 1/8 1/32 1/8 5/32 -11/32 -3/16 -1/4
2 3/16 1/8 1/8 1/16 5/32 1/8 -9/16 -3/8 -5/16
3 3/16 3/16 3/16 3/16 3/16 3/16 -5/8 -9/16 -1/2
4 1/8 1/8 1/8 9/32 7/32 7/32 -29/32 -13/16 -3/4
5 3/16 3/16 3/16 13/32 11/32 3/8 -1 1/8 -7/8 -1
6 1/4 1/4 3/16 17/32 1/2 9/16 -1 7/32 -1 5/32 -1 3/16
7 3/8 3/16 3/16 21/32 11/16 5/8 -1 3/8 -1 5/32 -1 1/4
8 1/4 3/16 1/4 23/32 5/8 19/32 -1 1/2 -1 9/32 -1 7/16
9 3/8 3/16 1/8 3/4 11/16 5/8 -1 9/16 -1 3/8 -1 1/2
10 9/16 3/8 1/4 25/32 23/32 13/16 -1 3/8 -1 3/8 -1 1/2
11 5/16 5/16 5/16 3/4 13/16 27/32 -1 5/16 -1 5/16 -1 1/2
12 5/16 1/4 5/16 25/32 27/32 7/8 -1 3/8 -1 5/16 -1 1/2
13 1/4 1/4 5/16 27/32 27/32 29/32 -1 9/32 -1 1/4 -1 7/16
14 1/4 3/16 1/4 7/8 13/16 27/32 -1 1/4 -1 5/32 -1 3/8
15 1/4 3/16 1/4 3/4 11/16 3/4 -1 1/8 -1 1/8 -1 1/4
16 3/16 3/16 3/16 19/32 9/16 5/8 -7/8 -13/16 -1 1/16
17 3/16 1/8 3/16 17/32 17/32 17/32 -5/8 -9/16 -3/4
18 1/4 1/8 1/4 3/8 7/16 3/8 -7/16 -3/8 -3/8
19 1/4 1/8 1/4 1/4 1/4 1/4 -1/4 -3/16 -1/4

216
Table B.4 – Measured sweeps of Specimens B36L

Measurement Sweep (in)


Point along
B36L-1 B36L-2
the Span
(Figure B.1c) Top Midheight Bottom Top Midheight Bottom
1 -3/16 -3/16 -3/16 -1/8 -1/4 -1/16
2 -3/8 -7/16 -3/16 -3/16 -1/8 -1/4
3 -9/16 -5/8 -5/16 -1/4 -1/8 -9/32
4 -5/8 -5/8 -1/4 -1/8 -3/16 -5/16
5 -11/16 -7/8 -5/16 -5/32 -3/16 -11/32
6 -7/8 -3/4 -5/16 -3/8 -3/8 -7/16
7 -7/8 -5/8 -7/16 -9/32 -1/2 -1/2
8 -3/4 -7/8 -9/16 -11/32 -1/2 -3/8
9 -15/16 -3/4 -19/32 -11/32 -9/16 -3/8
10 -15/16 -7/8 -11/16 -3/8 -9/16 -5/16
11 -1 -13/16 -5/8 -1/4 -3/8 -3/8
12 -3/4 -1 -11/16 -13/32 -1/2 -3/8
13 -3/4 -3/4 -1/2 -15/32 -1/2 -11/32
14 -13/16 -5/8 -9/16 -1/2 -1/2 -11/32
15 -11/16 -5/8 -1/2 -1/2 -1/2 -3/8
16 -3/4 -9/16 -3/8 -5/16 -1/4 -3/16
17 -5/8 -9/16 -5/16 -1/4 -5/16 -3/16
18 -1/4 -1/2 -1/4 -3/16 -5/16 -1/8
19 -1/8 -1/4 -1/8 -5/32 -1/4 -1/16

217
Figures B.2 to B.10 illustrate the sweeps of the specimens at mid-height. Figures

B.11-B.15 depict the permanent lateral deformations of the specimens at midheight,

measured after the complete removal of the applied load while Figures B16-B.20 show

the permanent angles of twist of the specimens along the beam length. The sinusoidal

curves, obtained from Equations B.1 and B.2, are included in Figures B.2-B.20 to

compare the initial lateral bow and the permanent deformed shape of the centerline of

each beam to the sinusoidal curve.

 z 
u  z   uo  sin   (B.1)
 L 

 z 
  z   o  sin   (B.2)
 L 

where z = the longitudinal distance from the end of a beam, L = the total length of the

beam; u(z) and uo = sweep of the beam at mid-height, at a distance z from the end and at

mid-span respectively; φ(z) and φo = the angles of twist at a distance z and at mid-span,

respectively. Figures B.11-B.20 reveal that the permanent deformations of a buckled

beam are in close agreement with the sinusoidal curve which implies that the buckled

shape of a beam can be best approximated by the sinus function.

218
Figure B.2 - Sweep at midheight of B18-1

Figure B.3 - Sweep at midheight of B18-2

219
Figure B.4 – Sweep at midheight of B30

Figure B.5 - Sweep at midheight of B36

220
Figure B.6 - Sweep at midheight of B44-1

Figure B.7 - Sweep at midheight of B44-2

221
Figure B.8 - Sweep at midheight of B44-3

Figure B.9 - Sweep at midheight of B36L-1

222
Figure B.10 - Sweep at midheight of B36L-2

Figure B.11 – Permanent lateral deformation at midheight of B30

223
Figure B.12 – Permanent lateral deformation at midheight of B44-1

Figure B.13 – Permanent lateral deformation at midheight of B44-2

224
Figure B.14 - Permanent lateral deformation at midheight of B44-3

Figure B.15 - Permanent lateral deformation at midheight of B36L-1

225
Figure B.16 – Permanent torsional rotations of B30

Figure B.17 – Permanent torsional rotations of B44-1

226
Figure B.18 – Permanent torsional rotations of B44-2

Figure B.19 – Permanent torsional rotations of B44-3

227
Figure B.20 – Permanent torsional rotations of B36L-1

228
APPENDIX C

MIDSPAN STRAIN DISTRIBUTIONS OF THE BEAMS AT


DIFFERENT LOAD LEVELS

The present appendix presents the longitudinal strain distributions along the depth of the

midspan section of each of the second set of specimens. As explained in Chapter III, in

the second set of tests longitudinal strains were measured continuously throughout the

tests, through strain gages attached on the convex and concave faces of the beams at mid-

span. Convex and concave faces of a beam are at equal distances from the mid-width of

the beam. The minor axis of the cross section of a beam coincides with the vertical

centroidal axis, if the beam section is symmetric about the midwidth. Longitudinal strains

originating from out-of-plane bending increase from zero at the minor axis to a maximum

at the outermost fibers of the section. In other words, the minor axis of a section is only

strained by major-axis bending because the lateral bending stresses vanish at minor axis.

The specimens of the present study were designed symmetrically about the

midwidth. Therefore, the minor axis of each specimen was coincident with the vertical

centroidal axis, and thus, the concave and convex faces of the beam were at equal

distances from the minor axis, assuming that the beams were perfectly symmetric about

the vertical centroidal axes, as designed. Accordingly, the compressive strains from

lateral bending on the concave faces were equal to the tensile strains from lateral bending

on the convex faces of the beams at mid-span. The longitudinal strain distributions along

the minor axes of the specimens, which originate solely from in-plane bending, were

obtained by averaging the longitudinal strains measured on the convex and concave faces

229
Figure C.1 – Loads and lateral deflections corresponding to the strain distributions in
Figures C.2 to C.4

of the beams. The longitudinal strains distributions on the convex and concave faces and

along the minor axes of the beams were illustrated in the following figures for different

load levels along the tests. The applied loads and lateral centroidal deflections

corresponding to the strain distributions in the figures are shown on the load-lateral

deflection curves of the specimens (Figures C.1, C.5, C.8, C.12 and C.15).

230
Figure C.2 – Midspan strain distributions of B44-1 at the initial stages of loading

231
Figure C.3 – Midspan strain distributions of B44-1 close to buckling

232
Figure C.4 – Midspan strain distributions of B44-1 after buckling

Figure C.5 – The Loads and lateral deflections corresponding to the strain distributions in
Figures C.6 and C.7

233
Figure C.6 – Midspan strain distributions of B44-2 at the initial stages of loading

234
Figure C.7 – Midspan strain distributions of B44-2 at the initiation of buckling

Figure C.8 – The Loads and lateral deflections corresponding to the strain distributions in
Figures C.9 to C.11

235
Figure C.9 – Midspan strain distributions of B44-3 at the initial stages of loading

236
Figure C.10 – Midspan strain distributions of B44-3 at different load levels

237
Figure C.11 – Midspan strain distributions of B44-3 close to buckling

238
Figure C.12 – The Loads and lateral deflections corresponding to the strain distributions
in Figures C.13 and C.14

239
Figure C.13 – Midspan strain distributions of B36L-1 at the initial stages of loading

240
Figure C.14 – Midspan strain distributions of B36L-1 close to buckling

241
Figure C.15 – The Loads and lateral deflections corresponding to the strain distributions
in Figures C.16 to C.18

242
Figure C.16 – Midspan strain distributions of B36L-2 at the initial stages of loading

243
Figure C.17 – Midspan strain distributions of B36L-2 at different load levels

Figure C.19-C.21 illustrate the depthwise strains on the convex and concave faces

of Specimen B44-1 at different load levels. The depthwise strains at mid-span of B44-1

were measured to determine the exact state of stress at mid-span and to detect any

distortions in the beam during the test. The measured strain values indicate that the

depthwise strains did not reach significant levels prior to buckling. The relatively higher

values of depthwise strains in the post-buckling stage (Figure C.21) are related to the

excessive out-of-plane deformations after buckling.

244
Figure C.18 – Midspan strain distributions of B36L-2 prior to and after buckling

245
Figure C.19 – Depthwise strains along the midspan section of B44-1 at the initial stages of loading

246
Figure C.20 – Depthwise strains along the midspan section of B44-1 at the final stages of loading

247
Figure C.21 – Depthwise strains along the midspan section of B44-1 during unloading

248
APPENDIX D

EXPERIMENTAL LOAD-DEFLECTION PLOTS OF THE


SPECIMENS

This appendix is a collection of the experimental load-deflection plots of the specimens.

The load-lateral (out-of-plane) deflection, load-vertical (in-plane) deflection and the

torque-twist plots of the specimens are presented in the appendix.

Each of Figures D.1 to D.10 shows the lateral deflections of a specimen,

measured at different points along the depth of the beam at mid-span. The depths of the

measurement points are also shown on the figures for the sake of comparison. As

previously explained in Section 3.1.3 of this dissertation, the goal of measuring the lateral

deflections of the beams at different depths was to evaluate the torsional rotations and

distortions in the beams.

In Figures D.11 and D.12, the load-lateral centroidal deflection curves of the

companion beams are compared for the Specimen Groups B44 and B36L, respectively.

The load-lateral deflection behavior of a reinforced concrete beam is greatly influenced

by its initial lateral imperfections. Due to the significant influence of sweep on the

stability, the centroidal sweep of each beam at midspan is illustrated on the respective

load-lateral deflection curve.

Each of the load-deflection curves in Figures D.1 to D.12 does not start from the

origin. The initial deflection value of each curve (the deflection at zero load) corresponds

to the initial lateral deformation of the beam at the particular depth. Including the initial

lateral imperfections in the plots was important particularly in Figures D.11 and D.12 to

illustrate the influence of sweep on the buckling behaviors of the companion beams. In

249
Section 6.3 of this dissertation, the effects of sweep on the load-lateral deflection

behaviors of the companion beams in Specimen Groups B44 and B36L were explained.

Figure D.1 – Out-of-plane deflections of B18-2 at midspan

Figure D.2 – Out-of-plane deflections of B22-1 at midspan

250
Figure D.3 – Out-of-plane deflections of B22-2 at midspan

Figure D.4 – Out-of-plane deflections of B30 at midspan

251
Figure D.5 – Out-of-plane deflections of B36 at midspan

Figure D.6 – Out-of-plane deflections of B44-1 at midspan

252
Figure D.7 – Out-of-plane deflections of B44-2 at midspan

Figure D.8 – Out-of-plane deflections of B44-3 at midspan

253
Figure D.9 – Out-of-plane deflections of B36L-1 at midspan

Figure D.10 – Out-of-plane deflections of B36L-2 at midspan

254
Figure D.11 – Lateral centroidal deflections of Beams B44 at midspan

Figure D.12 – Lateral centroidal deflections of Beams B36L at midspan

255
Figures D.13 to D.21 illustrate the load-vertical deflection curves of the

specimens. The plots also include analytical load-deflection curves, obtained by using the

cracked moment of inertia and two different effective moment of inertia expressions in

Equation (D.1), which gives the in-plane deflections at midspan of a simply-supported

beam, subjected to a concentrated load at midspan.

P  L3
vc  (D.1)
48  EI x

where P is the applied load; L is the span length; EIx is the in-plane flexural rigidity.

Two of the analytical curves in each figure correspond to the effective moment of

inertia expressions proposed by Branson (1963) and Bischoff (2005). The third analytical

curve, on the other hand, corresponds to the cracked moment of inertia, which is the

moment of inertia of a beam section when the entire tension zone of the section is

rendered ineffective in resisting bending moments due to flexural cracking. The effective

moments of inertia according to Branson (1963) and Bischoff (2005) and the cracked

moment of inertia are calculated from Equations (D.2), (D.3) and (D.4), respectively.

 M cra 
3
  M 3 
I eb     I g  1      I cr
cra
(D.2)
M  M
 a    a  

 M cra  1   M cra   1
2 2
1
    1     (D.3)
I ebi  M a  I g   M a   I cr

1
 b  c 3  n  As  d  c 
2
I cr  (D.4)
12

where Ig is the gross moment of inertia (Equation D.5); Ma is the maximum in-plane

256
bending moment in the beam; Mcra is the cracking moment of the beam; c is the neutral

axis depth from the compression face when all fibers in the compression zone are stressed

below the elastic limit of concrete; b is the width of the beam; As is the total cross-

sectional area of the longitudinal reinforcement; n is modular ratio of steel to concrete.

1
Ig   b  h3 (D.5)
12

where h is the height of the beam.

In Specimens B22 and B30, restrained shrinkage cracking was detected. Based on

the studies of Scanlon and Bischoff (2008), the term Mcra in Equations (D.2) and (D.3)

was replaced with 2Mcra/3 to account for the reduction in the effective moments of inertia

of B22 and B30 due to the presence of shrinkage cracks in concrete.

The experimental load-deflection curves of Beams B18-2, B36, B44 and B36L

are in close agreement with the analytical curves corresponding to the effective moments

of inertia proposed by Branson (1963) and Bischoff (2005). The load-deflection

behaviors of Beams B22 and B30, nonetheless, are not closely estimated by Equation

(D.1) when effective moment of inertia is used in the equation. Figures D.14 and D.15

indicate that the initial linear portions of the experimental load-vertical deflection curves

of Specimens B22 and B30 are coincident with the analytical line corresponding to the

cracked moment of inertia (Icr), most probably due to shrinkage cracking of concrete.

Figures D.22 to D.26 illustrate the torque-twist curves of the test specimens. The

ordinate axes represent the maximum torsional moment in a beam (Tmax), meaning the

torque at the beam ends, while the x-axes represent the twist at mid-span (θ), calculated

by dividing the angle of twist of the beam at midspan to the longitudinal distance from

257
support to mid-span. The red line in each figure indicates the maximum torsional moment

(Tb) in the beam at the instant of buckling.

Figure D.13 - In-plane deflections of Beam B18-2 at midspan

Figure D.14 - In-plane deflections of Beams B22 at midspan

258
Figure D.15 - In-plane deflections of Beam B30 at midspan

Figure D.16 - In-plane deflections of Beam B36 at midspan

259
Figure D.17 – In-plane deflections of B44-1 at midspan

Figure D.18 – In-plane deflections of B44-2 at midspan

260
Figure D.19 – In-plane deflections of B44-3 at midspan

Figure D.20 – In-plane deflections of B36L-1 at midspan

261
Figure D.21 – In-plane deflections of B36L-2 at midspan

Figure D.22 –Experimental torque-twist curve of Specimen B44-1

262
Figure D.23 –Experimental torque-twist curve of Specimen B44-2

Figure D.24 –Experimental torque-twist curve of Specimen B44-3

263
Figure D.25– Experimental torque-twist Curve of Specimen B36L-1

Figure D.26– Experimental torque-twist curve of Specimen B36L-2

264
APPENDIX E

METHOD FOR THE EVALUATION OF THE CENTROIDAL


DEFLECTIONS AND ROTATION OF A BEAM

This appendix presents the approach presented by Zhao (1994) and extended by Stoddard

(1997), which was used in the present study to convert the coupled deflection

measurements from the potentiometers into in-plane and out-of-plane deflections and

rotations at the shear center (centroid in rectangular sections).

The direction at which the beam buckles changes the geometric relations.

Therefore, the equations given by Stoddard (1997) were modified to account for the

direction of buckling. In the present study, a beam experiencing out-of-plane

deformations towards the lateral potentiometers is assumed to buckle in positive direction

(Figure E.1). Conversely, buckling away from the lateral potentiometers is defined as

buckling in negative direction. Equations corresponding to the both directions of buckling

are presented here.

At the beginning of the test, two potentiometers, T and B (denoting the top and

bottom potentiometers, respectively) were positioned horizontally while a third

potentiometer, V (denoting the vertical potentiometer) was positioned vertically as shown

in Figure E.1.

uc, vc and φc are the out-of-plane and in-plane deflections and the angle of twist at

shear center, respectively. To, Bo and Vo are the initial string lengths while Tf, Bf and Vf

are the final string lengths of potentiometers T, B and V, respectively. The lateral and

vertical deflections of the point Bp are denoted as Bx and By, respectively. Using the

265
Figure E.1 – Potentiometer configuration in the test

Pythagorean theorem for triangles 1 and 2 (Figure E.1), Equations (E.1) and (E.2) are

obtained.

 Bo  Bx 2  By2  B 2f (E.1)

V 
2
O  By  Bx2  V f2 (E.2)

When the beam buckles in negative direction, Equation (E.3) should be used instead of

Equation (E.1) while Equation (E.2) remains unchanged.

 Bo  Bx 2  By2  B 2f (E.3)

266
The solution of Equations (E.1) and (E.2) yields two sets of solution, (Bx1, By1) and (Bx2,

By2).

Bo  A1  Vo  A2
Bx1  (E.4)
A3

Vo  A4  Bo  A2
B y1  (E.5)
A3

Bo  A1  Vo  A2
Bx 2  (E.6)
A3

Vo  A4  Bo  A2
By 2  (E.7)
A3

where,

A1  Bo2  Vo2  B 2f  V f2 (E.8)

A2 
  
 Bo4  B 4f  Vo4  V f4  2  Bo2  B 2f  Vo2  V f2  2  B 2f  Vo2  V f2 
(E.9)
2  Vo2  V f2


A3  2  Bo2  Vo2  (E.10)

A4  Bo2  Vo2  B 2f  V f2 (E.11)

Similarly, the solution of Equations (E.2) and (E.3) yields,

 Bo  A1  Vo  A2
Bx 3  (E.12)
A3

Vo  A4  Bo  A2
By3  (E.13)
A3

267
 Bo  A1  Vo  A2
Bx 4  (E.14)
A3

Vo  A4  Bo  A2
By 4  (E.15)
A3

After calculating Bx and By, the angle of twist in the beam can be obtained from

the unbuckled and buckled configurations of the beam. In Figure E.2, edges of triangle 1

are determined from geometry. Using the Pythagorean theorem for the triangle, Equation

(E.16) is developed.

2
x   
2 c  
 B  B  b  cos   h  sin   b  cos  
 o c 2 c 
   
2 (E.16)
     
  By  b 2  sin c  h  1  cos c   b 2  sin c 
 
 T f2

Equation (E.16) can be simplified to become

2
 Bo  Bx  h  sin c 2   B y  h  1  cos c    T f2 (E.17)

When the beam buckles in the negative direction, Equation (E.17) changes to Equation

(E.18).

2
 Bo  Bx  h  sin c 2   B y  h  1  cos c    T f2 (E.18)

The solution of Equation (E.17) yields

c1  a tan 2   a a a
, 

 1 C  D  E 1 Ga  h  B y  Da  Ea  

(E.19)
 2 Fa 2 Fa  Da 

268
Figure E.2 – Angle of twist calculations

c 2

 a tan 2  1 
Ca  Da  Ea 1 Ga
, 
  h  By  
Da  Ea 
 (E.20)
 2 Fa 2 Fa  Da 

where,

Ca  Bo2  B y  2  h  Bo  Bx  2  h3  4  By  h 2  h  Bo2  h  Bx2


(E.21)
 By3  3  h  By2  By  T f2  2  Bo  Bx  By  h  T f2  Bx2  By

2
Da    Bo  Bx  (E.22)

269
  B 4  4  B  B3  6  B 2  B 2  2  B 2  B 2  2  B 2  T 2 
 x o x o x x y x f 
 
 4  h  Bx  By  4  Bo  Bx  By  4  Bo  Bx  T f
2 2 2

 
Ea   8  h  Bo  Bx  By  4  Bo3  Bx  4  h  By  T f2  (E.23)
 
 4  h  Bo2  By  2  Bo2  By2  4  h  B y3  2  B y2  T f2 
 
 4  h 2  T 2  4  h 2  B 2  2  B 2  T 2  B 4  B 4  T 4 
 f y o f o y f 
 


Fa  h 2  By2  2  By  h  Bo2  Bx2  2  Bo  Bx  h  (E.24)

 4  h 2  B  B  2  h 2  B 2  2  h 2  B 2  B 4 
 o x o x o

  B 4  4  B3  B  6  B 2  B 2  B 2  B2 
 x o x o x o y 
 2 2 2 3 2 2
Ga   2  h  Bo  B y  Bo  T f  4  Bo  Bx  Bx  B y  (E.25)
 2 2 2 2

 2  h  Bx  B y  Bx  T f  2  Bo  Bx  B y 
 
 4  h  Bo  Bx  B y  2  Bo  Bx  T f2 
 

It is to be noted that the roots of Equation (E.18) can be obtained by changing the signs of

the terms containing Bx in Equations (E.21) to (E.25).

Finally, the lateral and vertical displacements of the shear center (uc and vc) are

determined from the following geometric relations in terms of the angle of twist at the

shear center (φc) and the lateral and vertical displacements of point Bp (Bx and By):

h b
uc  Bx   sin c  1  cos c  (E.26)
2 2

h b
vc  By   1  cos c   sin c (E.27)
2 2

270
h b
uc  Bx   sin c  1  cos c  (E.28)
2 2

h b
vc  By   1  cos c   sin c (E.29)
2 2

Equations (E.26) and (E.27) are valid when the beam buckles in the positive direction

(Figure E.3) while Equations (E.28) and (E.29) are used when the beam buckles in the

negative direction.

For each set of (Bx, By), two different twist angles are obtained according to

Equations (E.19) and (E.20). Similarly, two different sets of out-of-plane and in-plane

centroidal deflections are calculated using Equations (E.26) to (E.29) for each set of (Bx,

By). Since two different sets of roots are obtained by solving Equations (E.1) and (E.2),

there are four different sets of deflection and rotation values of the centroid. To choose

the correct solution set, each set was compared to the deflection measurements taken by

the string potentiometers. The solution set in closest agreement with the experimental

data was chosen.

271
Figure E.3 – Centroidal deflection calculations

272
APPENDIX F

CRITICAL MOMENT CALCULATIONS OF THE SPECIMENS

This appendix presents the procedures used in the critical moment calculations of the

specimens. Each section presents the critical load calculations according to one of the

methods described in Chapter VII. The equations used in the calculations are shown on

the left halves of the following pages. On the right halves of the pages, on the other hand,

the equations and the meaning of the terms used in the equations are explained.

The critical and ultimate load calculations presented in Sections F.2 to F.5 require

an iterative procedure because of the interdependence of the variables. The iterative

procedure was carried out through the programming tools of Mathcad 14.0 (2005). The

programs used in the iterative procedure are given in Section F.2.

In this appendix, εc denotes the extreme compression fiber strain of a beam at

midspan.

F.1 Critical Load Calculations Assuming that a Reinforced Concrete Beam is an


Elastic and Homogeneous Body, Free from Cracking

Lateral flexural rigidity of the


1 
Beh  Ec    b3  h  entire beam section with elastic
 12  material behavior
Elastic modulus of concrete
Ec
obtained from cylinder tests

Width, depth and span length of the


b, h, L
beam, respectively

1  192 b  1 (2n  1) h  Coefficient for St. Venant’s


 c   1     tanh 
3  5
 h n0 (2n  1) 5
2b  torsional constant

273
Ec
Gc  Modulus of rigidity of concrete
2  (1   )
Poisson’s ratio of concrete from

material tests
Uncracked torsional rigidity of the
3
(GC )u  c  b  h  Gc beam according to St. Venant’s
theory
C1  e Beh 
M el   1  1.74    B   GC u Critical moment of a beam
C2  L 

L  GC u  eh
The loading factor for a beam
C1  4.23 loaded with a single concentrated
load at midspan
The effective length factor
C2  1.00 accounting for the simple support
conditions in and out of plane
The expression accounting for the
 e Beh  destabilizing effect of the load,
1  1.74  


L  GC u  applied above the centroid of the
beam section

4 The critical load of a beam by also


Pel    M el  M s  taking into account the self-weight
L of the beam

Bending Moment at Midspan


Ms Originating from the Self-Weight
of the Beam

F.2 Critical Load Calculations according to Hansell and Winter’s (1959)


Formulation

The critical moment calculations according to Hansell and Winter (1959) require an

iterative procedure. Using an analogy with the tangent modulus theory in inelastic

buckling of columns, Hansell and Winter (1959) stated that the secant modulus of

elasticity of concrete corresponding to the strain at the extreme compression fibers is the

modulus of the compression zone of a beam section in bending. Since the secant modulus

274
(Esec) depends on the extreme compression fiber strain (εc), Esec was denoted as a function

of εc in the calculations, Esec (εc).

According to Hansell and Winter (1959), the lateral bending rigidity (Bhw) and the

torsional rigidity (Chw) of a reinforced concrete beam is provided by the compression

zone only. Ignoring the rotations in the neutral axis of a section due to the twisting

rotations in the beam, the compression zone is a rectangular area (bxc).

The neutral axis depth (c) of the beam section and the strain at the extreme

compression fibers (εc) depend on the critical moment (Mhw). To calculate Mhw, the lateral

bending rigidity (Bhw) and the torsional rigidity (Chw) of a beam are needed. Since Bhw and

Chw are functions of c and εc, there is interdependence between Mhw and c, εc. To calculate

Mhw, c and εc, programming tools of Mathcad 13.0 (2005) were used. c and εc were

obtained from the following programs:

c  for  c  0.0001  0.0002  0.0035

c c  0.002in


As Es  s c   c
while c 
 
0.5 fc  c b
 10
5
 
in  ( c  100in) if  s c   c   y

c  c  0.0001in
As fy
otherwise
 
0.5 fc  c b

 c
   
break if Mhw  c  c  As Es  s c   c  d    10lbin if  s c   c   y
 3
 
 c
 
break if Mhw  c  c  As fy d    10lbin if  s c   c   y
 3
 
return c

275
 c  for  c  0.0001  0.0002  0.0035

c c  0.002in

 
As Es  s c   c
while c 
 
0.5 fc  c b
 10
5
 
in  ( c  100in) if  s c   c   y

c  c  0.0001in
As fy
otherwise
 
0.5 fc  c b

 c
   
break if Mhw  c  c  As Es  s c   c  d    10lbin if  s c   c   y
 3
 
 c
 
break if Mhw  c  c  As fy d    10lbin if  s c   c   y
 3
 
return  c

The expressions used in the programs are explained as follows:

Strain in the centroid of the tension


 s  c,  c  reinforcement, which is a function
of c and εc
 
f c'     c 
fc  c    o  Stress at the extreme compression
k 
 c  fibers
 1   
 o 

1 0.00689  f c' The factor ‘0.00689’ is used to


 and k  convert the compressive strength in
f c' 20
1 psi to MPa, since the stress-strain
 o  Ec equation is given in terms of MPa.

The above stress-strain relationship is the stress-strain model proposed by Carreira and

Chu (1985) for high-strength concrete. In Section 2.2 of this dissertation, Carreira and

Chu’s (1985) model was shown to be in perfect agreement with the experimental stress-

strain curves of concrete used in the specimens. Therefore, the above equation was used

in the critical moment calculations to link the strains in concrete to the stresses.

276
Yield strain and yield stress of the
εy, fy reinforcing steel, determined from
material tests
Modulus of elasticity of the
Es reinforcing steel, determined from
material tests
 4.23 
 L  Bhw   c , c    GC hw   c , c  
M hw   c , c      Critical moment, which is a
e Bhw   c , c    function of c and εc
 1  0.74    


L  GC  hw
  c , c   
 
The lateral bending rigidity
b3  c
Bhw   c , c   Esec   c   expression proposed by Hansell
12 and Winter (1959)

Esec   c   b3  c  b   The torsional rigidity expression


2
 GC hw   c , c     1  0.35    according to Hansell and Winter
2  1    3  d  
(1959)
The secant modulus is the slope of
fc  c  the line connecting the point (εc, fc)
Esec   c  
c on the stress-strain curve to the
origin
4
Phw    M hw  M s  The critical load of the beam
L

The programs run until there is a negligible difference between the bending

moment obtained from the critical moment expression [Mhw (εc, c)] and the bending

moment obtained from the stress distribution in the cross-section. The strain

measurements in the experiments indicated that all compression fibers in the specimens

were stressed within the elastic limit of concrete (elastic lateral torsional buckling).

Therefore, bending moment resistance of the beam section was calculated based on a

triangular stress distribution in the compression zone of the section.

277
F.3 Critical Load Calculations according to Sant and Bletzacker’s (1961)
Formulation

When calculating the critical load of a beam from the formula proposed by Sant and

Bletzacker (1961), an iterative procedure is needed. Therefore, the programs for c and εc,

shown in Section F.2, are used in the critical moment calculations based on Sant and

Bletzacker’s (1961) formulation.

Sant and Bletzacker (1961) suggested that the reduced modulus of elasticity (Er)

corresponding to the extreme compression fiber strain (εc) is the modulus of a beam

section at the instant of buckling. The reduced modulus theory assumes that a portion of

the beam (the convex side) undergoes unloading while the remaining portion of the beam

(the concave side) is further loaded when the beam buckles. The lateral bending rigidity

(Bsb) and the torsional rigidity (Csb) expressions proposed by Sant and Bletzacker (1961)

include the reduced modulus of elasticity (Er) as the material term.

The following discussion presents the equations proposed by Sant and Bletzacker

(1961) and important details from the calculation procedure:

 4.23 
 L  Bsb   c , c    GC  sb   c , c   Critical moment expression
M sb   c , c     
proposed by Sant and Bletzacker
e Bsb   c , c   
 1  3.48     (1961)
 L  GC  sb   c , c   
 
The expression accounting for the
 e Bsb   c , c   location of the load application point
1  3.48    with respect to the centroid of the

 L  GC sb  c , c   beam section, given by Sant and
Bletzacker (1961)
The lateral bending rigidity
b3  d
Bsb   c   Er   c   expression proposed by Sant and
12 Bletzacker (1961)

278
Er   c  b3  d The torsional rigidity expression
 GC sb   c    proposed by Sant and Bletzacker
2  1    3 (1961)
Reduced modulus of elasticity of
4  Ec  Etan   c  concrete which is a geometric
Er   c   average of the elastic modulus (Ec)
 
2
Ec  Etan   c  and the tangent modulus of elasticity
[Etan(εc)] corresponding to the
extreme compression fiber strain
Tangent modulus of elasticity is the
d slope of the line tangent to the stress-
Etan   c    f c   c   strain curve at the point
d c 
corresponding to the extreme
compression fibers.

F.4 Critical Load Calculations according to Massey’s (1967) Formulation

The iterative procedure explained in Section F.2 is used in the critical moment

calculations based on the formula proposed by Massey (1967). Similar to Hansell and

Winter (1959), Massey (1967) used the secant modulus theory. However, Massey (1967)

also included the contributions of the longitudinal and shear reinforcement of a beam to

the lateral bending and torsional rigidity expressions. The following discussion presents

the equations proposed by Massey (1967) and important details from the calculation

procedure:

 4.23 
 L  Bm   c , c    GC m   c , c  
M m   c , c      Critical moment, which is a function
m c    of c and εc
e B  , c 
 1  1.74   

 L  GC m   c , c   
 
The lateral bending rigidity
b3  c
Bm   c , c   Esec   c    Es  I sy expression proposed by Massey
12 (1967)

279
The contribution of the longitudinal
Es  I sy reinforcement to the lateral bending
rigidity. When steel yields, Es = 0.

 
   b3  h  G '     
 c c c

1  The torsional rigidity expression
 GC m  c     Gs  Gc'   c    bs3  ts   proposed by Massey (1967)
3 
   b12  d1  Ao  Es 
 
 2 2 s 
Esec   c  The reduced modulus of rigidity of
Gc'   c  
2  1    concrete

1  Contribution of the Longitudinal


 Gs  Gc'   c    bs3  ts Reinforcement to the Torsional
3  Rigidity
  b12  d1  Ao  Es Contribution of the Shear
Reinforcement to the Torsional
2 2  s Rigidity
Width and thickness of the
bs , ts longitudinal reinforcement layer,
respectively, as illustrated in Figure
1.18
Width and depth of the cross-sectional
b1 , t1 area enclosed by a closed stirrup,
respectively (Figure 1.18)
Cross-sectional area of one leg of a
Ao, s stirrup and spacing of the stirrups,
respectively
γ a constant defined by Cowan (1953)

F.5 Ultimate Load Calculations according to the Proposed Method

Different from the other methods, the method proposed in the present study accounts for

the reduction in the ultimate load of a beam due to sweep. First, the critical moment of a

reinforced concrete beam is calculated using the lateral bending and torsional rigidity

expressions proposed in the present study. The critical load corresponds to the

280
geometrically perfect configuration of the beam. Next, the limit load of the imperfect

beam is calculated by reducing the critical load an amount equal to the influence of the

sweep on the load-carrying capacity. The equations used in the proposed method are as

follows:

 4.23 
 L  Bo   c , c    GC o   c  
  Critical moment, which
M p  c , c    
p c   
e B  , c  is a function of c and εc
 1  1.74   
 L  GC o   c   
  

 
 
 b3  c 1   Esec   c   Ec  The lateral bending
Bo   c , c     2   rigidity expression
 12    M cra   c    2  proposed in the present
 1      1 
 Mp  h  study
   

 E     Ec   b3  h  b  The torsional rigidity


(GC )o   c    sec c   1  0.63    expression proposed in
 4  1     3  h  the present study

M cra Cracking moment of


the beam
The factor accounting
2 / 3 in the presence of shrinkage cracks for the reduction in the
 lateral bending rigidity
1 in the absence of shrinkage cracks in the presence of
shrinkage cracks
Critical load
4 corresponding to the

Pp  M p  M s   L
geometrically perfect
configuration of the
beam
Limit load of the beam
Pult  Pp 

uto  48  Ec  I y  with a sweep of uto at
3
sin(ult )  L the top of the beam at
midspan

281

uto  48  Ec  I y  Reduction in the
buckling load due to
sin(ult )  L3 sweep

1 Second moment of area


I y   b3  h of the beam section
12 about the minor axis
The angle of twist of
the beam at midspan at
the instant when Pult is
ult reached. A value of
0.60 degrees was found
to be appropriate for
most of the beams.

The above equations indicate that the rigidity expressions [Bp and (GC)p] and the

critical moment (Mp) are interdependent. Therefore, the iterative approach, summarized

in Section F.2, is used to calculate the neutral axis depth (c) and the extreme compression

fiber strain (εc) at midspan at the instant when buckling initiates. Then, the critical

moment (Mp) and the limit load (Pult) are obtained.

282
APPENDIX G

CONSTRUCTION DETAILS OF THE SPECIMENS

This appendix presents some specific details about the specimens of the

experimental program.

G.1 Compression Reinforcement

Lateral torsional buckling arises from the differential behaviors of the tension and

compression sides of a beam. The compression side of a beam is subjected to

compressive stresses from in-plane bending. When the compressive stresses reach critical

levels, the compression side of the beam buckles out of plane. The tension side of the

beam, on the other hand, tends to remain stable. The out-of-plane deformations of the

compression side cause the tension side to deform out of plane due to the integrity of the

beam. However, the out-of-plane deformations of the tension side are much smaller than

the deformations of the compression side as a result of the stabilizing effect of the tensile

stresses from in-plane bending. The differential out-of-plane deformations along the

depth of the beam result in the rotation of the beam about its longitudinal axis. Hence,

lateral torsional buckling creates out-of-plane deformations and torsional rotations in a

beam.

The stresses in the compression side of a beam are the main cause for lateral

torsional buckling. Increasing the out-of-plane bending rigidity of the compression side

can restrain the excessive lateral deformations of the compression side, which can indeed

prevent lateral torsional buckling. Compression reinforcement contributes to the lateral

283
bending rigidity of the compression side. Konig and Pauli (1990) indicated

experimentally that the compression reinforcement significantly increases the buckling

load of a reinforced concrete beam and decreases the out-of-plane deflections of the

compression side at buckling. Considering the stabilizing effect of the compression

reinforcement, the specimens of the present experimental program did not contain

compression reinforcement to ensure that the beams failed in lateral torsional buckling.

The shear reinforcement of the specimens was composed of two layers of welded

wire reinforcement (WWR), separated by the longitudinal reinforcing bars (Figure G.1).

Due to the lack of compression reinforcement in the beams, spacers were needed to

maintain the distance between the WWR sheets in the compression side. For this purpose,

spacers cut from reinforcing bars were placed between the WWR sheets (Figure G.1).

The lengths of the spacers were smaller than the development lengths of the reinforcing

bars, from which the spacers were cut.

G.2 Tilt-up and Lifting Mechanisms of the Specimens

The concrete beams of the present experimental program were cast on their sides to

facilitate the mechanical vibration of concrete and to ensure the spread of concrete into

the entire form, flowing around the congested reinforcement. A mechanism was needed

to tilt up the beams, leaning on their sides, and move them to the test setup using the

crane.

Two different lifting systems were used in the two stages of the experimental

program. Each of the lifting points in the first set of beams consisted of a headed cast-in-

place anchor embedded 7 inches into concrete (Figure G.2). Two lifting points in Beams

284
Figure G.1 – Reinforcement in Specimen B36

Figure G.2 – The Lifting mechanism in the first set of beams

285
B18 and B22 and four lifting points in Beams B30 and B36 provided adequate shear

capacity to tilt up the beams and adequate tensile capacity to lift the beams.

The beams were lifted through cables attached to the beams at the lifting points.

An important consideration for a beam hanging from cables is the angle of inclination of

the cables lifting the beam. Stratford and Burgoyne (1999) found out that the buckling

load of a concrete beam increases as the angle of inclination of the cables increase and

the cables approach to the vertical orientation. Accordingly, the test beams were lifted

with vertically-oriented cables when moving to the test setup. As shown in Figure G.3,

the specimen was connected to a steel spreader beam with vertical ropes and the spreader

beam was connected to the hook of the crane with inclined ropes.

Beams in Specimen Groups B44 and B36L were heavier than the first set of

beams (B22, B18, B30 and B36). The lifting points used in the first set of beams were not

able to provide adequate shear and tension capacity to tilt up and lift the second set of

beams. Therefore, a new lifting point was designed and used in the second stage of the

experiments.

The lifting points in the second set of beams were composed of a steel channel

and two reinforcing bars welded to the channel. The channel section was included in the

mechanism to resist the shear forces at the lifting points during the tilt-up process. The

reinforcing bars, on the other hand, provided adequate tensile capacity for the lifting

mechanisms while lifting the beams in the vertical position. A nut was welded to the

inside of the channel, so that a bolt can be fixed to the lifting mechanism when

connecting the lifting points to the crane.

286
Figure G.3 – Use of spreader beams for lifting the beams

To tilt up and lift the specimens, swift lifting eyes (also known as hoist rings)

were attached to the steel channels of the lifting mechanisms by means of high-strength

steel bolts (Figure G.4). The bail of a swift lifting eye can pivot about the base of the eye

in order to compensate for the direction of lifting. The ability of the bail to pivot about

the base made it possible to tilt up and lift the beams continuously without the need for

rearranging the lifting system between the tilt-up and lifting processes. The lifting

mechanisms used in the specimens were designed according to ACI 318-05 (2005)

Appendix D. Different failure mechanisms in the appendix were considered in the design

of the lifting points to prevent any possible damage to the beams during the tilt-up and

lifting processes.

287
Figure G.4 – The Lifting point in the second set of beams connected to the spreader beam

288
APPENDIX H

DETERMINATION OF THE EXPERIMENTAL BUCKLING LOADS


OF THE SPECIMENS

This appendix introduces the techniques developed by Southwell (1932), Meck

(1977) and Massey (1963) for determining the critical loads (limit loads in the case of

geometrically imperfect beams) of beams by analyzing their experimental load-deflection

data.

In elastic flexural buckling, Southwell plot is a common technique used to obtain

the buckling load of a member from its experimental data. In an axially loaded column,

for example, there is a linear relationship between uc/P and uc, where P is the axial load

on the column and uc is the lateral deflection at midlength of the column. The slope of the

uc/P vs. uc plot is equal to 1/Pcr, where Pcr is the critical load of the column.

Lateral torsional buckling, nevertheless, is more complicated than flexural

buckling of columns. A beam subjected to lateral torsional buckling undergoes out-of-

plane deformations and torsional rotations at the same time. Cheng and Yura (1988) used

two different types of Southwell (1932) plots to analyze the data of their lateral buckling

experiments on coped steel beams. Accordingly, uc/P was plotted against uc and φc/P was

plotted against φc. uc and φc are the lateral centroidal deflection and the angle of twist at

midspan, respectively and P is the concentrated load applied at midspan of the beam.

Cheng and Yura (1988) found out that the critical loads obtained from both plots were

almost the same for each specimen. However, the critical loads obtained from the uc/P vs.

uc plots were used, since Cheng and Yura (1988) considered the lateral deflection data in

289
the tests more reliable than the twist data due to the localized distortions in the test

beams.

Meck (1977) proposed the use of a “skewed” version of Southwell plot for lateral

torsional buckling of beams. Accordingly, uc/P should be plotted against φc and φc/P

should be plotted against uc. The geometric mean of the inverse slopes of the two plots

gives the critical load (Pcr).

Massey (1963) proposed a modification to the original Southwell (1932) plot to

be applicable to lateral torsional buckling experiments. According to Massey (1963), the

term P in the ordinates of the original Southwell (1932) plots should be replaced with P2

for the case of lateral torsional buckling. Similarly, Stratford and Burgoyne (1999) stated

that a deflection/(load)2 vs. deflection plot is more appropriate for a beam subject to

lateral torsional buckling, based on the studies of Allen and Bulson (1980).

Mandal and Calladine (2002) investigated the use of classical Southwell (1932)

plot and the modified versions of Southwell (1932) plot proposed by Meck (1977) and

Massey (1963) in lateral torsional buckling experiments and reached several important

conclusions. In their study, Mandal and Calladine (2002) analytically showed that the

lateral deflection (u) and the twist (φ) of a beam are proportional to each other after the

initial stages of loading in a lateral torsional buckling experiment. The direct coupling

between u and φ becomes more pronounced as the load is increased. Consequently, the

critical loads obtained from original Southwell (1932) plot and Meck’s (1977) “skewed”

version of the Southwell (1932) plot should not be different to a major extent.

In a Southwell (1932) plot, the data points corresponding to the initial stages of

loading do not lie on the straight line, which is formed by the majority of the data points.

290
According to Cheng and Yura (1988), the deviation of the initial points from the ultimate

straight line is caused by the initial restraints in the test setup and other experimental

errors which are more influential at the initial stages of loading when the applied load is

small. Based on the analysis of the experimental data obtained by Cheng and Yura

(1988), Mandal and Calladine (2002) found out that the deviation of the initial data points

from the eventual straight line is greater in the Massey’s (1963) version of the Southwell

(1932) plot. This is most probably due to the use of P2 instead of P in Massey’s (1963)

plots.

For the sake of illustration, Figures H.1 to H.3 illustrate the standard Southwell

(1932) plots and Meck’s (1977) and Massey’s (1963) versions of the Southwell plots,

respectively, for Specimen B44-2.

Figures H.1 and H.2 agree with the observations of Cheng and Yura (1988), who

considered the lateral deflection data in their tests more reliable. Almost all the data

points in the first plot of Figure H.1 lie on a straight line. In the second plot of Figure

H.1, nonetheless, the data points are too scattered, causing the determination of a straight

line to be more complicated. In the Meck’s (1977) version of the plots, the data points in

both plots are scattered since the twist data is used in both of the plots. The experimental

data of the other specimens showed the same characteristic. The large dispersion of the

data points makes the determination of the experimental buckling load more difficult

when the twist data is used in any version of the Southwell (1932) plot. Therefore, the use

of lateral deflection data in the original Southwell (1932) plot is considered easier and

more reliable in the determination of the experimental buckling load of a reinforced

291
concrete beam. The dispersion in the twist data might have been caused by the distortions

in the beams, particularly in the midspan region close to the point of application of load.

Figure H.1 – Southwell (1932) plots for Specimen B44-2

292
Pcr  0.274  42.26  11.6kips

Figure H.2 – Meck’s (1977) version of the Southwell (1932) plots for Specimen B44-2

293
Figure H.3 – Massey’s (1963) version of the Southwell (1932) plots for Specimen B44-2

294
The plots in Figures H.1 to H.3 agree also with the conclusions drawn by Mandal

and Calladine (2002). The critical loads obtained from the classical Southwell (1932)

plots (Figure H.1) are close to the critical load value obtained from Meck’s (1977)

version of the plots (Figure H.2). In Table H.1, the critical loads of the second set of

specimens, obtained from the classical and Meck’s (1977) version of the Southwell

(1932) plots are tabulated together with the ultimate loads measured during the tests. The

last column in Table H.1 corresponds to the critical load values obtained from the

classical Southwell (1932) plot using the lateral deflection data. The tabulated values

show that the critical loads according to the classical and Meck’s version of the

Southwell (1932) plots are in close agreement for all specimens.

Table H.1 – Critical loads from the classical and Meck’s (1977) version of the Southwell
(1932) plots for the second set of beams

Experimental Critical Load (kips)


Specimen Limit Load Meck Southwell
(kips) (1977) (1932)
B44-1 15.2 18.7 17.3
B44-2 12.1 11.6 13.6
B44-3 21.0 21.5 20.8
B36L-1 13.5 18.0 17.3
B36L-2 21.6 26.9 26.8

Figure H.3 indicates that the data points in Massey’s (1963) version of the plots

are more scattered than the classical and Meck’s (1977) versions of the plots.

Furthermore, the data points corresponding to the initial stages of loading lie further from

the eventual straight line in Massey’s (1963) version of the plots, as previously

established by Mandal and Calladine (2002).

295
Finally, Table H.1 indicates that both the classical version and the Meck’s (1977)

version of Southwell (1932) plot overpredict the limit loads of the second set of

specimens. Therefore, the limit loads of the specimens measured in the tests were used in

the analytical study to be more conservative.

296
REFERENCES

American Concrete Institute (ACI) (2005), “Building Code Requirements for Structural
Concrete and Commentary”, ACI 318-05 and ACI R318-05, Farmington Hills,
Michigan.

Allen, H. G. and Bulson, P. S. (1980), Background to Buckling, McGraw Hill,


Maidenhead UK, p. 582

ASTM A 615/A 615M (2008), “Standard Specification for Deformed and Plain Carbon-
Steel Bars for Concrete Reinforcement”, ASTM International, West
Conshohocken, Pennsylvania.

ASTM C 39/C 39M (2005), “Standard Test Method for Compressive Strength of
Cylindrical Concrete Specimens”, ASTM International, West Conshohocken,
Pennsylvania.

ASTM C 157/C 157M (2006), “Standard Test Method for Length Change of Hardened
Hydraulic-Cement Mortar and Concrete”, ASTM International, West
Conshohocken, Pennsylvania.

ASTM C 192/C 192M (2007), “Standard Practice for Making and Curing Concrete
Test Specimens in the Laboratory”, ASTM International, West Conshohocken,
Pennsylvania.

ASTM C 469 (2002), “Standard Test Method for Static Modulus of Elasticity and
Poisson's Ratio of Concrete in Compression”, ASTM International, West
Conshohocken, Pennsylvania.

ASTM C 1611/ C 1611M (2005), “Standard Test Method for Slump Flow of Self-
Consolidating Concrete”, ASTM International, West Conshohocken,
Pennsylvania.

Bischoff, P. H. (2005), “Reevaluation of Deflection Prediction for Concrete Beams


Reinforced with Steel and Fiber Reinforced Polymer Bars”, Journal of Structural
Engineering, ASCE, Vol. 131, No. 5, pp. 752-762.

297
Bischoff, P. H. (2007), “Rational Model for Calculating Deflection of Reinforced
Concrete Beams and Slabs”, Canadian Journal of Civil Engineering, Vol. 34,
No. 8, pp. 992-1002.

Bischoff, P. H. and Scanlon A. (2007), “Effective Moment of Inertia for Calculating


Deflections of Concrete Members Containing Steel Reinforcement and
Fiber-Reinforced Polymer Reinforcement”, ACI Structural Journal, Vol. 104,
No. 1, pp. 68-75.

Branson, D. E. (1963), “Instantaneous and Time-Dependent Deflections of Simple and


Continuous Reinforced Concrete Beams”, HPR Publication 7, Part 1, pp. 1-78,
Alabama Highway Department, Bureau of Public Roads.

Burgoyne, C. J. and Stratford, T. J. (2001), “Lateral Instability of Long-Span Prestressed


Concrete Beams on Flexible Bearings”, The Structural Engineer, Vol. 79, No. 6,
pp.23-26.

Carreira, D. J. and Chu K. (1985), “Stress-Strain Relationship for Plain Concrete in


Compression”, ACI Journal, Vol. 82, No. 6, pp. 797-804.

Cheng, J. J. R. and Yura, A. Y. (1988), “Lateral Buckling Tests on Coped Steel Beams”,
Journal of Structural Engineering, ASCE, Vol. 114, No. 1, pp. 16-30.

Considère, A. (1891), “Resistance des Pièces Comprimées”, Congrès International de


Procèdes de Construction, Paris, 3:371.

Cowan, H. J. (1953), “The Theory of Torsion Applied to Reinforced Concrete Design –


Part 2”, Civil Engineering and Public Works Review (London), Vol. 48, No. 568,
pp. 455-480.

Engesser, F. (1895), “Über Knickfragen”, Schweizerische Bauzeitung, Vol. 26, pp. 24-26.

Gilbert, R. I. (2006), “Discussion of "Reevaluation of Deflection Prediction for Concrete


Beams Reinforced with Steel and Fiber Reinforced Polymer Bars" by Peter H.
Bischoff”, Journal of Structural Engineering, ASCE, Vol. 132, No. 8,
pp. 1328-1330.

Hansell, W. and Winter G. (1959), “Lateral Stability of Reinforced Concrete Beams”,


ACI Journal, Proceedings, Vol. 56, No. 3, pp. 193-214.

298
Hsu, T. T. C. (1968), “Plain Concrete Rectangular Sections ”, Torsion of Structural
Concrete, SP 18, pp. 203-238, American Concrete Institute, Detroit.

Hsu, T. T. C. (1973), “Post-Cracking Torsional Rigidity of Reinforced Concrete


Sections”, ACI Journal, Proceedings, Vol. 70, No. 5, pp. 352-360.

Hsu, T. T. C. (1984), Torsion of Reinforced Concrete, Van Nostrand Reinhold Company


Inc., New York.

Hsu, T. T. C. (1990), “Shear Flow Zone in Torsion of Reinforced Concrete”, Journal of


Structural Engineering, ASCE, Vol. 116, No. 11, pp. 3206-3226.

Kollbrunner, C. F. and Bassler, K. (1969), Torsion in Structures, Springer-Verlag.,


New York.

König, G. and Pauli, W. (1990), “Ergebnisse von Kippversuchen an Schlanken


Fertigteilträgern aus Stahlbeton und Spannbeton”, Beton- und Stahlbetonbau, Vol.
85, No. 10, pp. 253-258.

Lampert, P. (1973), “Postcracking Rigidity of Reinforced Concrete Beams in Torsion and


Bending ”, Analysis of Structural Systems for Torsion, SP 35, pp. 385-433,
American Concrete Institute, Detroit.

Leemann, A. and Hoffmann, C. (2005), “Properties of Self-Compacting and


Conventional Concrete - Differences and Similarities”, Magazine of Concrete
Research, Vol. 57, No. 6, pp. 315-319.

Loser, R. and Leemann, A. (2009), “Shrinkage and Restrained Shrinkage Cracking of


Self-Compacting Concrete Compared to Conventionally Vibrated Concrete”,
Materials and Structures, Vol. 42, No. 1, pp. 71-82.

Lura, P., Pease, B., Mazzotta, G. B., Rajabipour, F., and Weiss, J. (2007), “Influence of
Shrinkage-Reducing Admixtures on Development of Plastic Shrinkage Cracks”,
ACI Materials Journal, Vol. 104, No.2, pp. 187-194.

Mandal, P. and Calladine, C. R. (2002), “Lateral-Torsional Buckling of Beams and the


Southwell Plot”, International Journal of Mechanical Sciences, Vol. 44, No. 12,
pp. 2557-2571.

299
Marshall, W. T. (1948), “The Lateral Stability of Reinforced Concrete Beams”, Journal,
Institution of Civil Engineers (London), Vol. 30, No. 6, pp. 194-196.

Massey, C. (1963), “Elastic and Inelastic Lateral Instability of I-Beams”, The Engineer,
Vol. 216, No. 5622, pp. 672-674.

Massey, C. (1967), “Lateral Instability of Reinforced Concrete Beams under Uniform


Bending Moments”, ACI Journal, Proceedings, Vol. 64, No. 3, pp. 164-172.

Massey, C. and Walter, K. R. (1969), “The Lateral Stability of a Reinforced Concrete


Beam Supporting a Concentrated Load”, Building Science, Vol. 3, No. 1,
pp. 183-187.

Mathcad Student Edition, Version 14.0, Mathsoft, Inc., 2007.

Mathematica Version 6.0 for Windows, Wolfram Research, Inc., 2007.

Meck, H. R. (1977), “Experimental Evaluation of Lateral Buckling Loads”, ASCE


Journal of Engineering Mechanics Division, Proceedings, Vol. 103, No. 2, pp.
331-337.

Mirza, S. A., Hatzinikolas, M., and MacGregor, J. G. (1979), “Statistical Descriptions of


Strength of Concrete”, ASCE Journal of Structural Division, Proceedings, Vol.
105, No. 6, pp. 1021-1037.

Nawy, E. D. (2005), Reinforced Concrete: a fundamental approach, Pearson Prentice


Hall, Upper Saddle River, New Jersey, pp. 52-53.

Revathi, P. and Menon, D. (2006), “Estimation of Critical Buckling Moments in Slender


Reinforced Concrete Beams”, ACI Structural Journal, Vol. 103, No. 2,
pp. 296-303.

Rausch, E. (1929), “Berechnung des Eisenbetons gegen Verdrehung”, Ph.D. thesis,


Technische Hochschule, Berlin.

300
Saint-Venant, B. de (1856). “Mémoire sur la Torsion des Prismes (lu à l’Académie le 13
juin 1853)”, Mémoires des Savants Etrangers, Mémoires Présentés par Divers
Savants à l’Académie des Sciences, de l’Institut Impérial de France et Imprimé
par son Ordre, V. 14, p. 233-560.

Sant, J. K. and Bletzaker, R. W. (1961), “Experimental Study of Lateral Stability of


Reinforced Concrete Beams”, ACI Journal, Proceedings Vol. 58, No. 6,
pp. 713-736.

Scanlon, A. and Bischoff, P. H. (2008), “Shrinkage Restraint and Loading History Effects
on Deflections of Flexural Members”, ACI Structural Journal, Vol. 105, No.4, pp.
498-506.

Shah, S. P., Karaguler, M. E., and Sarigaphuti, M. (1992), “Effects of Shrinkage


Reducing Admixtures on Restrained Shrinkage Cracking of Concrete”, ACI
Materials Journal, Vol. 89, No.3, pp. 289-295.

Siev, A. (1960), “The Lateral Buckling of Slender Reinforced Concrete Beams”,


Magazine of Concrete Research (London), Vol. 12, No. 36, pp. 155-164.

Simitses, G. J., and Hodges, D. H. (2006), Fundamentals of Structural Stability, Elseviser


Inc., Jordan Hill, Oxford, pp. 251-256.

Southwell, E. V. (1932), “On the Analysis of Experimental Observations in Problems of


Elastic Stability”, Proceedings of Royal Society of London, Vol. 135, pp. 601-616.

Stiglat, K. (1971), “Näherungsberechnung der Kritischen Kipplasten von


Stahlbetonbalken”, Die Bautechnik, Vol. 48, No. 3, pp. 98-100.

Stiglat, K. (1991), “Zur Näherungsberechnung der Kipplasten von Stahlbeton- und


Spannbetonträgern über Vergleichsschlankheiten”, Beton- und Stahlbetonbau,
Vol. 86, No. 10, pp. 237-240.

Stoddard, W. P. (1997), “Lateral-Torsional Buckling Behavior of Polymer Composite


I-Shaped Members”, Ph.D. thesis, Georgia Institute of Technology, Atlanta,
Georgia.

301
Stratford, T. J. and Burgoyne, C. J. (1999), “Lateral Stability of Long Precast Concrete
Beams”, Proceedings of the Institution of Civil Engineers: Structures and
Buildings, Vol. 134, No. 2, pp.169-180.

Tavio, and Teng, S. (2004), “Effective Torsional Rigidity of Reinforced Concrete


Members”, ACI Structural Journal, Vol. 101, No. 2, pp. 252-260.

Timoshenko, S. P. and Gere, J. M. (1963), Theory of Elastic Stability, International


Edition, McGraw-Hill Book Co., New York, pp. 251-277.

Timoshenko, S. P. and Goodier, J. N. (1970), Theory of Elasticity, International Edition,


McGraw-Hill Book Co., New York, pp. 309-313.

Tomaszewicz, A. (1984), “Betongens Arbeidsoliagram”, FCB/SINTEF Rapport, STF65


A84065.

Turcry, P. and Loukili, A. (2006), “Evaluation of Plastic Shrinkage Cracking of Self-


Consolidating Concrete”, ACI Materials Journal, Vol. 103, No. 4, pp. 272-279.

Turcry, P., Loukili, A., Haidar, K., Pijaudier-Cabot, G., and Belarbi A. (2006), “Cracking
Tendency of Self-Compacting Concrete Subjected to Restrained Shrinkage:
Experimental Study and Modeling”, Journal of Materials in Civil Engineering,
ASCE, Vol. 18, No. 1, pp. 46-54.

Vacharajittiphan, P., Woolcock, S. T., and Trahair, N. S. (1974), “Effect of In-plane


Deformation on Lateral Buckling.” Journal of Structural Mechanics, Vol. 3, No.
1, pp. 29-60.

Von Kármán, T. (1910), Encyklopädie der Matematischen Wissenschaften, Vol. IV4,


p. 349.

Wang, C. (1953), Applied elasticity, McGraw-Hill Book Co. Inc., New York, pp. 85-89.

Wee, T. H., Chin, M. S., and Mansur, M. A. (1996), “Stress-Strain Relationship of High-
Strength Concrete in Compression”, Journal of Materials in Civil Engineering,
ASCE, Vol. 8, No. 2, pp. 70-76.

302
Weiss, W. J. and Shah, S. P. (2002), “Restrained Shrinkage Cracking: The Role of
Shrinkage Reducing Admixtures and Specimen Geometry”, Materials and
Structures, Vol. 34, No. 246, pp. 85-91.

Yarimci, E., Yura, J. A., and Lu, L. W. (1967), “Techniques for Testing Structures
Permitted to sway”, Experimental Mechanics, Vol. 7, No. 8, pp. 321-331.

Yen, B. T. (1974), “Beams”, Structural steel design, editor Lambert Tall, 2nd Edition,
The Ronald Press Company, New York, p. 196.

Yura, J. A. and Phillips, B. A. (1992), “Bracing Requirements for Elastic Steel Beams”,
Research Report 1239-1, Center for Transportation Research, The University of
Texas at Austin, Texas.

Zhao, X. L., Hancock, G. J., and Trahair, N. S. (1994), “Lateral Buckling Tests of
Cold-Formed RHS Beams”, Research Report R699, School of Civil and Mining
Engineering, The University of Sydney, Australia.

303
VITA

İLKER KALKAN

İlker Kalkan was born on July 22nd, 1981 in Mardin, Turkey. He attended Middle

East Technical University in Ankara, Turkey and received a B.S. degree in Civil

Engineering with emphasis on Construction Management and Engineering in 2004. He

earned an M.S. degree in 2006 and a Ph.D. degree in 2009 in Structural Engineering,

Mechanics and Materials from Georgia Institute of Technology, Atlanta, Georgia. He

took classes from the Daniel Guggenheim School of Aerospace Engineering at Georgia

Institute of Technology as his minor field of study in Ph.D. In his doctoral studies, he

investigated the lateral stability of rectangular reinforced concrete beams analytically and

experimentally.

304

View publication stats

You might also like