Sie sind auf Seite 1von 375

Advancements on the

Disturbed Stress Field Model

Claudio Esteban Osses-Henriquez

A thesis submitted in partial fulfillment


of the requirements for the degree of

Master of Science in Civil Engineering

University of Washington

2007

Program Authorized to Offer Degree: Civil & Environmental Engineering


University of Washington
Graduate School

This is to certify that I have examined this copy of a master’s thesis by

Claudio Esteban Osses-Henriquez

and have found that it is complete and satisfactory in all respects,


and that any and all revisions required by the final
examining committee have been made.

Committee Members:

Laura Lowes

Dawn Lehman

Peter Mackenzie-Helnwein

Date:
In presenting this thesis in partial fulfillment of the requirements for a master’s degree at
the University of Washington, I agree that the Library shall make its copies freely available
for inspection. I further agree that extensive copying of this thesis is allowable only for
scholarly purposes, consistent with “fair use” as prescribed in the U.S. Copyright Law. Any
other reproduction for any purpose or by any means shall not be allowed without my written
permission.

Signature

Date
University of Washington

Abstract

Advancements on the
Disturbed Stress Field Model

Claudio Esteban Osses-Henriquez

Co-Chairs of the Supervisory Committee:


Professor Laura Lowes
Department of Civil and Environmental Engineering
Professor Dawn Lehman
Department of Civil and Environmental Engineering

This thesis presents the formulation and implementation of a plane Reinforced Concrete
(RC) material model for use in Finite Element Modeling (FEM) based on the Disturbed
Stress Field Model (DSFM) (Vecchio 2000). The DSFM treats RC as a composite of concrete
and steel. Concrete behavior is simulated using a smeared rotating crack approach in which
localized deformations, such as crack slip and crack opening, are averaged over the entire
RC area. Reinforcing steel also is assumed to be smeared over the entire RC area.
The original formulation of the DSFM proposed by Vecchio uses a secant stiffness matrix
for solution of the nonlinear problem and uses the change in the secant stiffness to determine
convergence of the nonlinear problem. This approach leads to slow rates of convergence,
because of the use of the secant stiffness, and equilibrium errors, because using the change
in the secant stiffness ratio to evaluate convergence does not guarantee equilibrium of the
system. To address these issues the model was reformulated and reimplemented in Mat-
lab. This new model is referred to as the DSFM-TG model. The DSFM-TG formulation
generates a consistent tangent stiffness matrix, which could be expected to greatly increase
the rate of convergence of the nonlinear problem. Additionally, the DSFM-TG explicitly
incorporates some features which were not originally included in the DSFM in an implicit
manner such as simulation of the impact of concrete confinement, the Poisson effect, out-of-
plane reinforcement, and dowel action. Finally, the DSFM-TG model was implemented in
simple FEM algorithm in which convergence is determined on the basis of the equilibrium
error.
TABLE OF CONTENTS

List of Figures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

List of Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xiii

Chapter 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1.3 Outline of the Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

Chapter 2 DSFM-TG Reinforced Concrete Material Model . . . . . . . . . . . . . 4


2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2 Conceptual Model and Behavior Assumptions . . . . . . . . . . . . . . . . . . 4
2.3 Average Region Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3.1 Compatibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
2.3.2 Constitutive Relationships . . . . . . . . . . . . . . . . . . . . . . . . . 10
2.3.3 Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 Cracking Region Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4.1 Compatibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4.2 Constitutive Relationships . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4.3 Equilibrium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.5 1D Average Region Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.5.1 Longitudinal Steel 1D Models . . . . . . . . . . . . . . . . . . . . . . . 20
2.5.2 Tensile Concrete 1D Models . . . . . . . . . . . . . . . . . . . . . . . . 24
2.5.3 Compressive Concrete 1D Models . . . . . . . . . . . . . . . . . . . . . 32
2.6 1D Cracking Region Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.6.1 Crack Spacing Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.6.2 Crack Width Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.6.3 Aggregate Interlock Model . . . . . . . . . . . . . . . . . . . . . . . . 39
2.6.4 Dowel Action Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

i
2.7 Multiaxial Loading Effects on 1D Models Modeling . . . . . . . . . . . . . . . 42
2.7.1 Poisson Effect versus Load Induced Strains . . . . . . . . . . . . . . . 42
2.7.2 Impact of Multiaxial Concrete Stresses on Concrete Compressive Re-
sponse Parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
2.8 1D Multiaxial Loading Parameters Models . . . . . . . . . . . . . . . . . . . . 50
2.8.1 Poisson’s ratio ν 1D Models . . . . . . . . . . . . . . . . . . . . . . . . 50
2.8.2 Compression softening parameter βl 1D Models . . . . . . . . . . . . . 54
2.8.3 Confinement parameter βd 1D Models . . . . . . . . . . . . . . . . . . 59
2.9 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

Chapter 3 DSFM-TG RC Material Model Formulation for use in FEM . . . . . . 66


3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
3.2 Internal RC State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.2.1 Governing Equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
3.2.2 Nonlinear System of Equations . . . . . . . . . . . . . . . . . . . . . . 69
3.3 RC Stress Vector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.3.1 Steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
3.3.2 Concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
3.4 RC Stiffness Matrix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99
3.4.1 Steel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
3.4.2 Concrete . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
3.5 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120

Chapter 4 DSFM-TG RC Material Model Implementation for use in FEM . . . . 122


4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
4.2 RC material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
4.3 Process 1: Internal RC State . . . . . . . . . . . . . . . . . . . . . . . . . . . 125
4.3.1 Subprocess 1N LS: Solve Non-linear System of Equations . . . . . . . 127
4.3.2 Subprocess 1RJ: Residual and Jacobian . . . . . . . . . . . . . . . . 132
4.3.3 Subprocess 1CR: Crack Region Models . . . . . . . . . . . . . . . . . 141
4.3.4 Subprocess 1AR: Average Region Models . . . . . . . . . . . . . . . . 142
4.3.5 Subprocess 1M E: Multiaxial Loading Effect Models . . . . . . . . . . 144
4.4 Process 2: Concrete and Steel Material Stiffness Matrices and Stress Vectors 159
4.4.1 Subprocess 2C: Concrete material stiffness and stress vector . . . . . 159
4.4.2 Subprocess 2S: Steel material stiffness and stress vector . . . . . . . . 162
4.5 Process 3: RC material stiffness matrix and stress vector . . . . . . . . . . . . 163

ii
4.6 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

Chapter 5 Model Evaluation and Validation . . . . . . . . . . . . . . . . . . . . . 166


5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5.2 1D model tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 166
5.2.1 Crack spacing model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
5.2.2 Crack width model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 170
5.2.3 Aggregate interlock model . . . . . . . . . . . . . . . . . . . . . . . . . 172
5.2.4 Dowel action model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
5.2.5 Longitudinal Steel 1D models . . . . . . . . . . . . . . . . . . . . . . . 177
5.2.6 Compressive Concrete 1D model . . . . . . . . . . . . . . . . . . . . . 180
5.2.7 Tensile Concrete 1D model . . . . . . . . . . . . . . . . . . . . . . . . 186
5.2.8 Compression softening 1D model . . . . . . . . . . . . . . . . . . . . . 191
5.2.9 Confinement 1D model . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
5.2.10 Poisson compressive 1D model . . . . . . . . . . . . . . . . . . . . . . 199
5.2.11 Poisson tensile 1D model . . . . . . . . . . . . . . . . . . . . . . . . . 203
5.3 RC material model tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
5.3.1 Compression tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 206
5.3.2 Tension tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
5.3.3 Shear tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
5.4 RC material model in FEM analysis tests . . . . . . . . . . . . . . . . . . . . 267
5.4.1 Panel PV10 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270
5.4.2 Panel PV16 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 284
5.4.3 Panel PV17 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 299
5.5 Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 310

Chapter 6 Summary, Conclusions and Recommendations . . . . . . . . . . . . . . 313


6.1 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 313
6.2 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 314
6.3 Recommendations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 316

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 318

Appendix A Matlab code . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319


A.1 averageRegionModels.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319
A.2 bilinear.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 319

iii
A.3 confinement kupfer richart.m . . . . . . . . . . . . . . . . . . . . . . . . . . . 320
A.4 cracking spacing basic.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
A.5 crackingRegionModels.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 321
A.6 findRCstate.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 322
A.7 get steel layer strain.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 323
A.8 getRCStressStiff.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
A.9 getRJ.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 324
A.10 getSteelConcStressStiff.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
A.11 Hognestad.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
A.12 IsItConverged.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 331
A.13 isUpdateNeeded.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 332
A.14 Linear no residualSOFT Bentz 2003STIFF.m . . . . . . . . . . . . . . . . . . 332
A.15 multipleCompressionSoftening.m . . . . . . . . . . . . . . . . . . . . . . . . . 334
A.16 multipleConcrete.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 335
A.17 multipleConfinement.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 336
A.18 multipleLoadingEffectModels.m . . . . . . . . . . . . . . . . . . . . . . . . . . 336
A.19 multiplePoisson.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 338
A.20 multipleSteel.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 339
A.21 poisson kupfer comp.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
A.22 poisson V2 tension.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 340
A.23 PopovicHS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 341
A.24 principal alt.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342
A.25 Raynor.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 343
A.26 RCmaterial.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
A.27 record.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
A.28 steel dowel he kwan.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 344
A.29 transformation.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345
A.30 Trilinear.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
A.31 Vecchio 1982 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 347
A.32 Vecchio 1986 comp soft.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
A.33 Vecchio 1992 A comp soft.m . . . . . . . . . . . . . . . . . . . . . . . . . . . 349
A.34 Walraven slip resistance.m . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 350

iv
LIST OF FIGURES

Figure Number Page


2.1 Steel coordinate system,αi . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
2.2 Concrete coordinate system,12c . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.3 Average Steel Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.4 Average Concrete Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
2.5 Cracking Region Geometric Variables . . . . . . . . . . . . . . . . . . . . . . 15
2.6 Stresses at the crack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.7 Bilinear model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21
2.8 Trilinear model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.9 Raynor model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
2.10 Longitudinal Steel Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
2.11 Vecchio1982 model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
2.12 LNRBentz model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.13 Tensile Concrete Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.14 Hognestad model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
2.15 PopovicHS model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.16 Compressive Concrete Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 37
2.17 Crack Spacing Basic Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38
2.18 Aggregate Interlock Model, vc vs. wcr . . . . . . . . . . . . . . . . . . . . . . 40
2.19 Aggregate Interlock Model, vc vs. δslip . . . . . . . . . . . . . . . . . . . . . . 40
2.20 Perpendicular Concrete Principal Strains . . . . . . . . . . . . . . . . . . . . 49
2.21 Perpendicular Concrete Principal Stresses . . . . . . . . . . . . . . . . . . . . 51
2.22 Compression Softening/Confinement Interaction . . . . . . . . . . . . . . . . 51
2.23 Variable Poisson’s Ratio . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
2.24 Linear Poisson . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
2.25 Vecchio1986 model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
2.26 Vecchio1992A model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
2.27 Compression Softening Models . . . . . . . . . . . . . . . . . . . . . . . . . . 58
2.28 Confinement Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59

v
2.29 Confinement Stresses Decomposition . . . . . . . . . . . . . . . . . . . . . . . 60
2.30 Uniform(pu ) and additional normalized(pa ) confinement pressure . . . . . . . 61
2.31 Normalized confinement pressures when fcmax > 0 . . . . . . . . . . . . . . . 61
4.1 RCmaterial function and processes . . . . . . . . . . . . . . . . . . . . . . . . 124
4.2 Process 1: Internal RC Problem . . . . . . . . . . . . . . . . . . . . . . . . . . 127
4.3 Subprocess 1NLS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 154
4.4 Subprocess 1RJ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
4.5 Subprocess 1CR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 156
4.6 Subprocess 1AR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 157
4.7 Subprocess 1ME . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
4.8 Process 2: Concrete and Steel Material Stiffness Matrices and Stress Vectors 160
4.9 Process 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163
5.1 Basic Crack Spacing Model, Scr . . . . . . . . . . . . . . . . . . . . . . . . . . 168
5.2 Basic Crack Spacing Model, Scr euler and dScr . . . . . . . . . . . . . . . . . . 169
dθ12c
∂wcr ∂wcr
5.3 Basic Crack width model, wcr , , and . . . . . . . . . . . . . . . . . 171
∂εc1 ∂Scr
5.4 Walraven model, vc under variable wcr . . . . . . . . . . . . . . . . . . . . . 173
5.5 Walraven model, vceuler and derivatives under variable wcr . . . . . . . . . . . 173
5.6 Walraven model, vc under variable δslip . . . . . . . . . . . . . . . . . . . . . 174
5.7 Walraven model,vceuler and derivatives under variable δslip . . . . . . . . . . . 174
5.8 He-Kwan model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
5.9 He-Kwan model, Ed and fdeuler . . . . . . . . . . . . . . . . . . . . . . . . . . 176
5.10 Longitudinal steel models, fs . . . . . . . . . . . . . . . . . . . . . . . . . . . 178
5.11 Longitudinal steel models, Es and fseuler . . . . . . . . . . . . . . . . . . . . 179
5.12 Compressive concrete model, fc under variable εc . . . . . . . . . . . . . . . . 182
5.13 Compressive concrete model, fceuler and derivatives under variable εc . . . . . 183
5.14 Compressive concrete model, fc under variable εp . . . . . . . . . . . . . . . . 184
5.15 Compressive concrete model, fceuler and derivatives under variable εp . . . . . 184
5.16 Compressive concrete model, fc under variable fp . . . . . . . . . . . . . . . . 185
5.17 Compressive concrete model, fceuler and derivatives under variable fp . . . . . 185
5.18 Tensile concrete models, fc under variable εc . . . . . . . . . . . . . . . . . . 187
5.19 Tensile concrete models, fceuler and derivatives under variable εc . . . . . . . . 188
5.20 Tensile concrete models, fc under variable Scr . . . . . . . . . . . . . . . . . . 188
5.21 Tensile concrete models, fceuler and derivatives under variable Scr . . . . . . . 189

vi
5.22 Tensile concrete models, fc under variable φ . . . . . . . . . . . . . . . . . . . 189
5.23 Tensile concrete models, fceuler and derivatives under variable φ . . . . . . . . 190
εperp
c
5.24 Compression softening models, βd under variable − . . . . . . . . . . . . 192
εc
5.25 Compression softening models, βd under variable εc . . . . . . . . . . . . . . . 193
5.26 Compression softening models, βdeuler and derivatives under variable εc . . . . 193
5.27 Compression softening models, βd under variable εperpc . . . . . . . . . . . . . 194
5.28 Compression softening models, βd euler and derivatives under variable εperpc . . 194
f min
5.29 Confinement model, βl under variable fcmin and constant cmax . . . . . . . . 196
fc
5.30 Confinement model, βl under variable fc min . . . . . . . . . . . . . . . . . . . 197
5.31 Confinement model, βleuler and derivatives under variablefcmin . . . . . . . . . 197
5.32 Confinement model, βl under variable fcmax . . . . . . . . . . . . . . . . . . . 198
5.33 Confinement model, βleuler and derivatives under variable fcmax . . . . . . . . 198
εload
c
5.34 Poisson Compressive Model, ν under variable . . . . . . . . . . . . . . . 200
εp
5.35 Poisson Compressive Model, ν under variable εload
c . . . . . . . . . . . . . . . 200
5.36 Poisson Compressive Model, ν euler and derivatives under variable εload . . . . 201
c
5.37 Poisson Compressive Model, ν under variable εp . . . . . . . . . . . . . . . . 201
5.38 Poisson Compressive Model, ν euler and derivatives under variable εp . . . . . 202
εload
c
5.39 Poisson Tension Model, nu under variable ratio . . . . . . . . . . . . . . 204
εccr
5.40 Poisson Tension Model, nu under variable εload c . . . . . . . . . . . . . . . . . 204
5.41 Poisson Tension Model, nu euler and derivatives under variable εload . . . . . . 205
c
5.42 RC response for test cases CC(1-4)-TR-Strain . . . . . . . . . . . . . . . . . . 210
5.43 Concrete principal compressive, fc2 , response for test cases CC(1-4)-TR-Strain212
5.44 Concrete principal tensile, fc1 ,response for test cases CC(1-4)-TR-Strain . . . 213
5.45 Concrete principal out-of-plane, fcz , response for test cases CC(1-4)-TR-Strain214
5.46 Steel stress response, CC(1-4)-TR-Strain . . . . . . . . . . . . . . . . . . . . . 215
5.47 Number of Internal Iterations, CC(1-4)-TR-Strain . . . . . . . . . . . . . . . 216
5.48 Internal Strain Error, CC(1-4)-TR-Strain . . . . . . . . . . . . . . . . . . . . 217
5.49 RC response, CC1-SF-Stress, CC2-TF-Stress, and CC3-TR-Stress . . . . . . . 218
5.50 Number of External Iterations, CC1-SF-Stress, CC1-TF-Stress, and CC1-
TR-Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
5.51 External error, CC1-SF-Stress, CC1-TF-Stress, andCC1-TR-Stress . . . . . . 220
5.52 RC response, CC(1-4)-TR-Stress . . . . . . . . . . . . . . . . . . . . . . . . . 221

vii
5.53 Concrete fc2 stress response, CC(1-4)-TR-Stress . . . . . . . . . . . . . . . . 222
5.54 Concrete fc1 stress response, CC(1-4)-TR-Stress . . . . . . . . . . . . . . . . 223
5.55 Concrete fcz stress response, CC(1-4)-TR-Stress . . . . . . . . . . . . . . . . 224
5.56 Steel stress response, CC(1-4)-TR-Stress . . . . . . . . . . . . . . . . . . . . 225
5.57 Number of External Iterations, CC(1-4)-TR-Stress . . . . . . . . . . . . . . . 226
5.58 External error, CC(1-4)-TR-Stress . . . . . . . . . . . . . . . . . . . . . . . . 227
5.59 RC response, CT1-TR-Strain and CT2-TR-Strain . . . . . . . . . . . . . . . 230
5.60 Concrete fc1 response, CT1-TR-Strain and CT2-TR-Strain . . . . . . . . . . 231
5.61 Concrete fc2 response, CT1-TR-Strain and CT2-TR-Strain . . . . . . . . . . 232
5.62 In-plane steel layer 2 stress response, CT1-TR-Strain and CT2-TR-Strain . . 233
5.63 In-plane steel layer 1 stress response, CT1-TR-Strain and CT2-TR-Strain . . 234
5.64 Number of Internal Iterations, CT1-TR-Strain and CT2-TR-Strain . . . . . . 235
5.65 Internal Strain (errorε ) and stress (errorσ ) internal errors, CT1-TR-Strain
and CT2-TR-Strain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
5.66 RC response, CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-Stress . . . . . . . 237
5.67 Number of external iterations, CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-
Stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
5.68 External error, CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-Stress . . . . . . 239
5.69 RC response, CT1-TR-Stress, and CT2-TR-Stress . . . . . . . . . . . . . . . 240
5.70 Concrete fc1 response, CT1-TR-Stress, and CR2-TR-Stress . . . . . . . . . . 241
5.71 Concrete fc2 response, CT1-TR-Stress, and CR2-TR-Stress . . . . . . . . . . 242
5.72 Steel fs2 and fscr2 , CT1-TR-Stress, and CR2-TR-Stress . . . . . . . . . . . . 243
5.73 Steel fs1 and fscr1 , CT1-TR-Stress, and CT2-TR-Stress . . . . . . . . . . . . 244
5.74 Number of External Iterations, CT1-TR-Stress, and CT2-TR-Stress . . . . . 245
5.75 External Error, CT1-TR-Stress, and CT2-TR-Stress . . . . . . . . . . . . . . 246
5.76 RC response, CS1, CS2, CS3, and CS4-TR-Strain . . . . . . . . . . . . . . . . 249
5.77 Concrete response, CS1, CS2, CS3, and CS4-TR-Strain . . . . . . . . . . . . 250
5.78 Concrete shear response, CS1, CS2, CS3, and CS4-TR-Strain . . . . . . . . . 251
5.79 Steel response, fs2 and fscr2 , CS1, CS2, CS3, and CS4-TR-Strain . . . . . . . 252
5.80 Steel response, fs1 and fscr1 , CS1, CS2, CS3, and CS4-TR-Strain . . . . . . . 253
5.81 Number of Internal Iterations, CS1, CS2, CS3, and CS4-TR-Strain . . . . . . 254
5.82 Internal Error, CS1, and CS2-TR-Strain . . . . . . . . . . . . . . . . . . . . . 255
5.83 Internal Error, CS3, and CS4-TR-Strain . . . . . . . . . . . . . . . . . . . . . 256
5.84 RC response, CS1-SF-Stress, CS1-TF-Stress, CS1-TR-Stress . . . . . . . . . . 257
5.85 Number of Internal Iterations, CS1-SF-Stress, CS1-TF-Stress, CS1-TR-Stress 258

viii
5.86 Internal Error, CS1-SF-Stress, CS1-TF-Stress, CS1-TR-Stress . . . . . . . . . 259
5.87 RC response, CS1, CS2, CS3, CS4-TR-Stress . . . . . . . . . . . . . . . . . . 260
5.88 Concrete response, CS1, CS2, CS3, CS4-TR-Stress . . . . . . . . . . . . . . . 261
5.89 Concrete shear response, CS1, CS2, CS3, CS4-TR-Stress . . . . . . . . . . . . 262
5.90 Steel response,fs2 and fscr2 , CS1, CS2, CS3, CS4-TR-Stress . . . . . . . . . . 263
5.91 Steel response,fs1 and fscr1 , CS1, CS2, CS3, CS4-TR-Stress . . . . . . . . . . 264
5.92 Number of External Iterations, CS1, CS2, CS3, CS4-TR-Stress . . . . . . . . 265
5.93 External Error, CS1, CS2, CS3, CS4-TR-Stress . . . . . . . . . . . . . . . . . 266
5.94 Panels PV10 Sand 16 boundary conditions an external loading . . . . . . . . 267
5.95 Panel PV17 boundary conditions an external loading . . . . . . . . . . . . . . 268
5.96 Panel PV10 Shear Response, First group of tests . . . . . . . . . . . . . . . . 271
5.97 Panel PV10 Shear Response, Second group of tests . . . . . . . . . . . . . . . 273
5.98 Panel PV10 Shear Response, Third group of tests . . . . . . . . . . . . . . . . 275
5.99 Panel PV10 RC Response, Third group of tests . . . . . . . . . . . . . . . . . 276
5.100Panel PV10 θ12c and θε angles, Third group of tests . . . . . . . . . . . . . . 277
5.101Panel PV10 Concrete fc1 and fc2 stress response, Third group of test . . . . . 279
5.102Panel PV10 Cracking Region Response, Third group of tests . . . . . . . . . 281
5.103Panel PV10 In-plane Steel Layer 2 Response, Third group of tests . . . . . . 282
5.104Panel PV10 In-plane Steel Layer 1 Response, Third group of tests . . . . . . 283
5.105Panel PV16 Shear Response, First group of tests . . . . . . . . . . . . . . . . 286
5.106Panel PV16 Shear Response, Second Group of Tests . . . . . . . . . . . . . . 288
5.107Panel PV16 Shear Response, Third Group of Tests . . . . . . . . . . . . . . . 290
5.108Panel PV16 RC Response, Third Group of Tests . . . . . . . . . . . . . . . . 291
5.109Panel PV16, θ12c and θε angles, Third Group of Tests . . . . . . . . . . . . . 293
5.110Panel PV16 Concrete fc1 and fc2 stress response, Third Group of Tests . . . 294
5.111Panel PV16 Cracking Region Response, Third Group of Tests . . . . . . . . . 296
5.112Panel PV16 In-plane Steel Layer 2 Response, Third Group of Tests . . . . . . 297
5.113Panel PV16 In-plane Steel Layer 1 Response, Third Group of Tests . . . . . . 298
5.114Panel PV17 Compressive Response, First Group of Tests . . . . . . . . . . . . 301
5.115Panel PV17 Compression Response, Second Group of Tests . . . . . . . . . . 303
5.116Panel PV17 Compression Response, Third Group of Tests . . . . . . . . . . . 305
5.117Panel PV17 RC Response, Third Group of Tests . . . . . . . . . . . . . . . . 306
5.118Panel PV17 θ12c and θε angles, Third Group of Tests . . . . . . . . . . . . . . 307
5.119Panel PV17 Concrete fc2 Stress Response, Third Group of Tests . . . . . . . 308
5.120Panel PV17 Concrete fc1 Stress Response, Third Group of Tests . . . . . . . 309

ix
5.121Panel PV17 Steel Stress Response, Third Group of Tests . . . . . . . . . . . . 310

x
LIST OF TABLES

Table Number Page


2.1 Longitudinal Steel Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
2.2 Tensile Concrete Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.3 Compressive Concrete Models . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
3.1 Internal RC Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
4.1 Residual Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
4.2 Jacobian Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 138
4.3 Compression Softening Strains . . . . . . . . . . . . . . . . . . . . . . . . . . 147
4.4 Confinement Stresses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 149
4.5 β ε variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
4.6 ν variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153
5.1 Basic crack spacing model test cases . . . . . . . . . . . . . . . . . . . . . . . 167
5.2 Basic crack width model test cases . . . . . . . . . . . . . . . . . . . . . . . . 170
5.3 Walraven model test cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
5.4 He-Kwan dowel model test cases . . . . . . . . . . . . . . . . . . . . . . . . . 175
5.5 Longitudinal steel model test cases . . . . . . . . . . . . . . . . . . . . . . . . 177
5.6 Compression 1D models test cases . . . . . . . . . . . . . . . . . . . . . . . . 180
5.7 Tension concrete 1D models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
5.8 Compression Softening 1D models . . . . . . . . . . . . . . . . . . . . . . . . 191
5.9 Confinement models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 195
5.10 Poisson Compressive models . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199
5.11 Poisson Tensile models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203
5.12 RC material compression tests . . . . . . . . . . . . . . . . . . . . . . . . . . 206
5.13 RC material compression test properties . . . . . . . . . . . . . . . . . . . . . 208
5.14 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
5.15 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 219
5.16 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
5.17 RC material tension tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228

xi
5.18 RC material tension test properties . . . . . . . . . . . . . . . . . . . . . . . . 229
5.19 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 235
5.20 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 238
5.21 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 245
5.22 RC material shear tests . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 247
5.23 RC material shear test properties . . . . . . . . . . . . . . . . . . . . . . . . . 248
5.24 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 254
5.25 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
5.26 Analysis Times . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
5.27 Arch Length Method Analysis Parameters . . . . . . . . . . . . . . . . . . . . 268
5.28 Internal RC Problem Analysis Parameters . . . . . . . . . . . . . . . . . . . 269
5.29 Panel PV10 Material Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 270
5.30 Panel PV10, First group of tests . . . . . . . . . . . . . . . . . . . . . . . . . 270
5.31 Panel PV10, Second group of tests . . . . . . . . . . . . . . . . . . . . . . . . 272
5.32 Panel PV10, Third group of tests . . . . . . . . . . . . . . . . . . . . . . . . 274
5.33 Panel PV16 Material Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 284
5.34 Panel PV16, First group of tests . . . . . . . . . . . . . . . . . . . . . . . . . 285
5.35 PV16 Panel, Second group of tests . . . . . . . . . . . . . . . . . . . . . . . . 287
5.36 PV16 Panel, Third group of tests . . . . . . . . . . . . . . . . . . . . . . . . . 289
5.37 Panel PV17 Material Properties . . . . . . . . . . . . . . . . . . . . . . . . . . 299
5.38 Panel PV17 First Group of Tests . . . . . . . . . . . . . . . . . . . . . . . . . 300
5.39 Panel PV17, Second Group of Tests . . . . . . . . . . . . . . . . . . . . . . . 302
5.40 Panel PV17, Third Group of Tests . . . . . . . . . . . . . . . . . . . . . . . . 304

xii
GLOSSARY

D: RC material stiffness tensor

[D]: RC stiffness matrix in the XY coordinate system

[Dc ]: concrete material stiffness in the XY coordinate system

[Ds ]: steel material stiffness in the XY coordinate system

db : bar diameter

dbi : in-plane steel layer i bar diameter

Ec : concrete tangent stiffness

Ec1 : concrete tangent stiffness in direction 1sc

Ec2 : concrete tangent stiffness in direction 2sc

Ecz : concrete tangent stiffness in direction z

Eca : concrete tangent tension softening stiffness

Ecb : concrete tangent tension stiffening stiffness

Ecto : concrete uncracked tension stiffness

Es : steel tangent stiffness

Esi : steel tangent stiffness in in-plane layer i

Escr : steel at crack tangent stiffness

Escri : steel at crack tangent stiffness in in-plane layer i

{Es }: steel tangent stiffness vector

{Escr }: steel at crack tangent stiffness vector

Esy : steel yielding stiffness

Eso : steel initial stiffness

xiii
Esh : steel hardening stiffness

fcperp : perpendicular concrete stress

fcadd : additional confinement stress

fcmax : maximum confinement stress

fcmin : minimum confinement stress

fc : concrete stress

fc1 : concrete stress in direction 1sc

fc2 : concrete stress in direction 2sc

fcz : concrete stress in direction z

fcj : concrete stress in direction j,(j = 1sc , 2sc , z)

fca : concrete tension softening stress

fcb : concrete tension stiffening stress


0
fc : concrete peak stress at peak

fp1 : modified concrete peak stress in the 1sc direction

fp2 : modified concrete peak stress in the 2sc direction

fpz : modified concrete peak stress in the zsc direction

{fp }: modified concrete peak stress vector

fsi : in-plane steel layer i longitudinal stress

fsz : out-of-plane steel layer stress

fscri : in-plane steel layer i longitudinal stress at crack

fdcri : in-plane steel layer i dowel stress at crack

fy : steel yielding stress

fsh : steel hardening stress

fu : ultimate steel stress

Gf : concrete fracture energy parameter

xiv
[J]: Jacobian matrix

[Jred ]: reduced Jacobian

Lr : concrete characteristic cracking length

nsh : hardening steel parameter (Raynor steel model)

[P ]: Poisson matrix

{p}: parameter vector

pa : normalized additional confinement stress

pu : normalized uniform confinement stress

{R}: residual vector

{R{ε} }: strain residual vector

{R{ε3P D,Load } }:
c
load concrete strain residual vector

{R{εp } }: modified concrete strain at peak residual vector

{R{fp } }: modified concrete peak stress residual vector

Rnormal : normal crack equilibrium residual

Rtangent : tangent crack equilibrium residual

Rout−of −plane : out-of-plane crack equilibrium residual

Scr : crack spacing

[Tstrain ]: strain transformation matrix

[Tstress ]: stress transformation matrix

[Tstif f ]: stiffness transformation matrix

vc : aggregate interlock shear stress at crack

wcr : crack width

β1ε : concrete strain at peak modification factor in direction 1sc

β2ε : concrete strain at peak modification factor in direction 2sc


ε
βZ : concrete strain at peak modification factor in direction z

xv
β1σ : concrete peak stress modification factor in direction 1sc

β2σ : concrete peak stress modification factor in direction 2sc


σ
βZ : concrete peak stress modification factor in direction z
ε
βd1 : compression softening concrete strain at peak modification factor in direction 1sc
ε
βd2 : compression softening concrete strain at peak modification factor in direction 2sc
ε
βdz : compression softening concrete strain at peak modification factor in direction z
ε
βl1 : confinement concrete strain at peak modification factor in direction 1sc
ε
βl2 : confinement concrete strain at peak modification factor in direction 2sc
ε
βlz : confinement concrete strain at peak modification factor in direction z
σ
βd1 : compression softening concrete peak stress modification factor in direction 1sc
σ
βd2 : compression softening concrete peak stress modification factor in direction 2sc
σ
βdz : compression softening concrete peak stress modification factor in direction z
σ
βl1 : confinement concrete peak stress modification factor in direction 1sc
σ
βl2 : confinement concrete peak stress modification factor in direction 2sc
σ
βlz : confinement concrete peak stress modification factor in direction z

βlKupf er : Kupfer confinement parameter

βlRichart : Richart confinement parameter

ε: RC strain tensor

{ε}: RC strain vector

δslip : crack slip

∆ε1cr : strain increment in in-plane steel at crack

∆dcr : steel dowel displacement at crack

{∆dcr }: steel dowel displacement at crack vector

ε: RC strain tensor

{εp }: modified concrete strain at peak vector

xvi
{ε}: RC strain vector in the XY coordinate system

ε: total average strain vector

εSI , εSN I , γSI : steel average strains in XY coordinate system

{ε3P
C
D,LOAD
}: load principal concrete strain vector

{ε3P
C
D
}: total principal concrete strain vector

{εslip }: : slip strain vector in the XY coordinate system

{εslip }12C : : slip strain vector in the 12c coordinate system

{εC }: : concrete strain vector in the XY coordinate system

{εC }12C : : concrete strain vector in the 12c coordinate system

εco : concrete strain at peak

εP 1 : modified concrete strain at peak in the 1sc direction

εP 2 : modified concrete strain at peak in the 2sc direction

εP Z : modified concrete strain at peak in the zsc direction

εc1 : concrete strain in the 1sc direction

ε2 : concrete strain in the 2sc direction

εcz : concrete strain in the z direction

εload
c1 : load concrete strain in the 1sc direction

εload
c2 : load concrete strain in the 2sc direction

εload
cz : load concrete strain in the z direction

εs : steel longitudinal strain

εy : steel yielding strain

εsh : hardening steel strain

εu : ultimate steel strain

εc : concrete strain

εccr : concrete cracking strain

xvii
εts : concrete tension softening strain

εperp
c : perpendicular concrete strain

{ε3P
c
D
}: concrete principal direction strain vector

{ε3P
c
D,Load
}: concrete principal direction Poisson strain vector

εsz : out-of-plane steel strain

εscri : in-plane steel layer i strain at crack

εsi : in-plane steel layer strain at crack

αi : in-plane steel layer i orientation with respect to the x direction

ν1 : Poisson’s ratio in direction 1sc

ν2 : Poisson’s ratio in direction 2sc

νZ : Poisson’s ratio in direction z

{ν}: Poisson vector

φI : in-plane steel layer i angle with respect to crack

{φ}: angle φ’s vector

{σ}: RC stress vector in the XY coordinate system

{σc3P D }: principal concrete stresses vector

σ: material stress tensor

σc : concrete stress tensor

{σc }12c : concrete stress vector in the 12c coordinate system

σsi : in-plane steel layer i stress tensor

{σ}: RC material stress vector

{σc }: concrete stress vector

{σs }: steel stress vector

{σsi }: in-plane steel layer i stress vector

ρI : in-plane steel layer i ratio

xviii
ρZ : out-of-plane steel layer ratio

Smx : cracking spacing in the x direction

Smy : cracking spacing in the y direction

θ12c : principal concrete stress 1sc orientation with respect to x

θε : maximum principal RC material strain direction with respect to the x-direction

xix
ACKNOWLEDGMENTS

To my wife and unborn daughter, mis papas y hermana, and all the wonderful people I
had the chance to meet at the UW.
My wife for being so patient and supporting throughout my pursuit of a Masters degree.
This has been a dream for a very long time.
My family in Chile for their financial and emotional support and their willingness to
travel all the way to Seattle to watch me graduate.
Michelle Tragesser and Bob Lee for their guidance and mentorship throughout my career
as a structural engineer.
My committee, Professors Lowes, Lehman and Mackenzie, for their guidance and sup-
port.

xx
1

Chapter 1

INTRODUCTION

1.1 Background

The research presented is based largely on the Modified Compression Field Theory
(MCFT) and the Disturbed Stress Field Model (DSFM). Both models were developed
to predict the non-linear behavior of 2D reinforced concrete (RC) elements subjected to
in-plane stresses.
The MCFT was developed by Vecchio and Collins (1986). In this theory, concrete is
assumed to be an orthotropic material and the impact of concrete cracking is represented
using a smeared, rotating crack model. Discontinuities introduced by cracking are ”smeared”
over the element to generate concrete strain and allow the use of continuum mechanics. The
orientation of a concrete crack is allowed to rotate during the analysis so that the crack
remains normal to the direction of the concrete principal tensile stress. Reinforcement steel
is assumed to act in parallel with the concrete. Interaction between concrete and steel is
considered at the crack surface where compatibility, constitutive, and equilibrium equations
are combine to determine local steel stresses and shear stresses at the crack surface. The
results of previous research including RC panel tests conducted by Vecchio and Collins were
used to develop models defining concrete and steel constitutive response.
The DSFM is an extension of the MCFT model in which slip is allowed at the crack
surface. This model is more appropriate for lightly reinforced concrete members. The
explicit inclusion of slip at the crack implies that the concrete principal strain directions
do not necessarily coincide with the RC principal strain directions, as in the case for the
MCFT model.
The MCFT and DSFM were implemented into a Nonlinear Finite Element Program
2

called Vector2 developed by Vector Analysis Group at the University of Toronto. The
program incorporates several analysis features not covered by the MCFT or DSFM such as
the modeling of the lateral expansion effect, triaxial stresses, cyclic loading, construction
sequence, buckling and bond-slip mechanisms. Vector2 uses an iterative secant stiffness
algorithm, where the convergence criteria is based on the convergence of the material secant
slopes.

1.2 Objectives

The objective of this research is to formulate and implement an RC material model


suitable for the use of the finite element method (FEM) analysis in OpenSees. OpenSees is
a software framework design to simulate seismic performance of structural and geotechnical
systems. The formulation of the compatibility, constitutive and equilibrium relationships of
the presented RC material model is based on the DSFM.
As mentioned before, an iterative secant stiffness algorithm has already been imple-
mented into the FEM program Vector2. This implementation is based on the work devel-
oped by Vecchio (2001). However, this iterative secant stiffness algorithm is not directly
suitable for use in traditional Newton-Raphson tangent based stiffness FEM analysis, which
is also the one adopted by the OpenSees FEM tools.
Using a secant stiffness rather than a tangent stiffness could have significant detrimental
effects in the convergence speed, although several authors [WHO?] have claimed that using
a secant approach leads to a more stable solution. An additional difference between the
iterative secant stiffness algorithm and a Newton-Raphson tangent-based stiffness is the
convergence criteria. While the first defines convergence as a function of the unbalanced
force residual, the second defines convergence as a function of the secant material stiffness
rate of change between iterations.
In order to implement a tangent-based stiffness material suitable for the use in OpenSees
it is necessary to conduct a complete revision of the construction of the RC material stiffness
matrix presented in the work developed by Vecchio (2001). The material stiffness matrix
is the base for the construction of the element and structure stiffness matrix. The new RC
material model presented herein will be referred to as DSFM-TG , where TG stands for
3

tangent. The DSFM-TG model includes some phenomena not originally included in the
DSFM such as Poisson effect, confinement, out of plane stresses, and dowel action. The
formulation and implementation of those phenomena are partly based on the information
presented in Vector2’s manual.

1.3 Outline of the Thesis

Chapter 2 discusses the main aspects of the DSFM material model, and the advancements
included in DSFM-TG

Chapter 3 presents the mathematical framework developed to use the DSFM-TG govern-
ing equation within a FEM analysis environment.

Chapter 4 presents the different processes developed for use in the implementation of the
DSFM-TG solution algorithms.

Chapter 5 evaluates the performance of the DSFM-TG formulation and implementation


in physical and numerical terms. A

Chapter 6 presents the main conclusions obtained from this research, along with recom-
mendations for future research.

Appendix A lists the Matlab code used to implement the DSFM-TG .


4

Chapter 2

DSFM-TG REINFORCED CONCRETE MATERIAL MODEL

2.1 Introduction

The basis of the DSFM-TG model and its governing equations are presented in this
Chapter. Section 2.2 introduces the conceptual model and assumptions used. Sections
2.3 and 2.4 present the compatiblity, constitutive, and equilbrium relationships for the RC
average and cracking regions. Sections 2.5 and 2.6 present the 1D compatibility and consti-
tutive models used in the average and cracking region. Section 2.7 explains the modeling of
multiaxial loading effects on the 1D models used in the average region. Section 2.8 describes
the different models used to determine the multiaxial parameters described in Section 2.7.

2.2 Conceptual Model and Behavior Assumptions

RC is modeled as a composite material where the contribution of steel and concrete are
added to find the total RC response. A set of compatibility, constitutive, and equilibrium
equations link the global behavior of the RC composite with that of the steel and concrete
constituents.
RC is divided into an average and a cracking region. The average region stress state
is similar to that of the region located between cracks, and its response is representative
of the global RC behavior. The RC stresses in this region are considered average. The
integration of the average region stresses over a given volume of RC results in the internal
forces associated to that volume. The cracking region is located in the vicinity of the crack,
and its response is representative of local RC behavior.
The average strains over an RC region are called average RC strains, ε. In an FEM
formulation these strains are typically determined from the nodal displacements and dis-
placement shape functions. The average steel strains, εs , (steel strains in the average
region) are assumed to be equal to the average RC strains. The average concrete strains,
5

εc , (concrete strains in the average region) are equal to the average RC strains, ε, minus
the smeared slip strain,εslip , due to the slip, δslip at the crack.
Reinforcing steel is normally placed in multiple layers. DSFM-TG accounts for multiple
in-plane layers arbitrarily oriented, and a single out-of-plane layer normal to the plane.
Each in-plane steel layer is modeled as a smeared fixed orthotropic material. The material is
smeared in the sense that the strength and stiffness of discrete reinforcing bars are smeared
over a unit area to simulate 2D continuum. The material is fixed in the sense that the
orientation of the materials physical properties defined by the orientation of the steel does
not change during the analysis. The material is orthotropic in the sense that it provides
longitudinal and dowel stress response. The dowel response is only considered in the cracking
region where the presence of localized displacements significantly the dowel stresses and their
importance for the behavior of that region. The out-of-plane steel layer is modeled as a
uniaxial material acting normal to the plane; this steel layer may carry tension in the out-
of-plane direction due to the expansion of the concrete. Tension in this steel layer generates
an out-of-plane compressive stress in the concrete that acts to confine the concrete.
Concrete is modeled as a smeared, rotating orthotropic material. The concrete model is
a smeared model in the sense that localized deformations due to concrete crack opening is
smeared over a unit area of concrete. The model is an orthotropic model in the sense that
the physical properties rotate with the orientation of the principal concrete strains. The
rotating orthotropic material is different from a fixed orthotropic material, such as wood or
the in-plane layers used in this model, for which the orientation of the materials physical
properties are fixed during the analysis. Concrete contributes to the RC material response
by carrying compression and tension stresses in the concrete principal strain directions.
The material state of the continuum is determined by enforcing equilibrium, compati-
bility, and constitutive for both the local cracking region and the global average region. At
the crack, the concrete tensile stress normal to the crack surface is assumed to be zero. To
satisfy equilibrium, steel stresses increase with respect to that of the average region, and
concrete shear stress develops parallel to the crack (due to the aggregate interlock at the
crack). As mentioned before, steel dowel action is considered in the cracking region, but
not in the average region.
6

Phenomena such as buckling, bar slip, and yielding due to combined longitudinal and
dowel stress interaction could be, but have not been included in DSFM-TG model.
The concrete constitutive models are derived from 1D testing. In general, concrete sub-
jected to in-plane deformation will experience biaxial loading, and even triaxial loading if
out-of-plane reinforcement is included. The presence of lateral stresses can significantly
change the concrete response from what a 1D constitutive model would predict. The ap-
proach taken in the DSFM-TG model is to capture the effect of multiaxial loading by
modifying the 1D constitutive model parameters.
Different constitutive models are used for concrete in tension and compression. The
1D compressive constitutive models are modified to account for lateral tensile loading and
confinement. Compression softening refers to the lost of compression strength and stiffness
in concrete due to the presence of lateral tensile strains. Confinement refers to the increase
in concrete strength due to compressive lateral stresses. No modifications are applied to the
basic concrete tension constitutive models due to multi-axial loading effect.
Lateral expansion of concrete due to Poisson effect is considered in the DSFM-TG model.
The introduction of lateral expansion complicates the use of the constitutive models because
it introduces expansion, or Poisson, strains, εpoisson
c , that are not directly related to stresses.
A simple example of this would be the lateral expansion strains observed in a concrete
cylinder loaded uniaxially in compression. These lateral strains do not generate lateral
stresses, and therefore cannot be directly used with a 1D constitutive model to compute the
stress in the radial direction. It is necessary then to introduce the concept of load strains,
εload
c , which are the component of the material strains that can be used with 1D constitutive
models to compute stresses.
Out-of-plane concrete stress arises when out-of-plane reinforcement is introduced, as out-
of-plane reinforcement constrains the out-of-plane expansion of concrete. It is necessary to
capture this effect in the model because restrain of out-of-plane concrete expansion can
produce significant out-of-plane concrete confinement stresses, which increases the in-plane
compressive strength capacity of the concrete.
The primary assumptions used in the DSFM-TG model are
7

• Steel reinforcement is uniformly distributed within the concrete volume.

• The orientation of the cracks rotates and is defined by the orientation of the average
concrete principal strains.

• Concrete crack spacing and width is uniform

• Reinforcing ”dowel” effect is only present in the cracking region

• Perfect bond exists between reinforcing steel and concrete in the average region

• The material response of reinforcing steel and concrete is defined using 1D constitutive
models.

• Concrete response in tension is independent of action in the perpendicular direction.

2.3 Average Region Modeling

This section describes the compatibility, constitutive, and equilibrium relationships em-
ployed to simulate the behavior of the average region.

2.3.1 Compatibility

The average RC strain state vector representation, {ε}, in the global XY coordinate
system is  
 εx 
 
 
 
{ε} = εy (2.1)
 
 
γ 
 
xy

In a typical FEM implementation these strains are computed from element nodal displace-
ments using element shape functions. The average steel strain state, εs , is assumed to be
equal to that of the RC continuum:
εs = ε (2.2)

The average concrete strain state, εc , is equal to that of the RC if the concrete is uncracked
or there is no slip, δslip , on the crack surface. Slip at the crack results in a shift in the
concrete average strains with respect to the average RC strains:

εc = ε − εslip (2.3)
8

The average slip strain, εslip , is determined by smearing the local slip displacement at the
crack over the average region.
The vector representations of the average concrete and in-plane steel strain state, {εc }
and {εs }, in the global XY coordinate system are
 

 ε 


 cx 

{εc } = εcy (2.4)

 


γ  
cxy
 

 ε 


 sx  
{εs } = εsy (2.5)

 


γ  
sxy

To use the 1D steel constitutive models it is necessary to transform the average steel strain
vector from the XY global coordinate system to the coordinate systems aligned with the bar
reinforcing axes. The steel coordinate system associated with the ith in-plane steel layer is
identified as αi , and it is shown in Figure 2.1. To use the 1D concrete constitutive models it
is necessary to transform the average concrete strain vectors from the XY global coordinate
system to the coordinate systems aligned with the concrete principal strain directions. The
concrete coordinate system is identified as 12c , and it is shown in Figure 2.2.

Figure 2.1: Steel coordinate system,αi


9

Figure 2.2: Concrete coordinate system,12c

The average steel strain vector in the αi coordinate system is


 

 ε 


 si 

{εs }αi = [Tstrain (αi )] · {εs } = ε (2.6)
 sni 
 
 γsi 
 

where αi is the angle between the bar axis and the X axis as shown in Figure 2.1. The
average concrete strain vector written in the 12c coordinate system is
 

 ε 

 c1 

{εc }12c = [Tstrain (θ12c )] · {εc } = εc2 (2.7)

 


0 

where θ12c is the angle between the maximum principal concrete stress direction and the
X axis as shown in Figure 2.2. It is calculated from the concrete strains in the XY global
system as shown on Equation 2.8
µ ¶
1 γcxy
θ12c = arctan (2.8)
2 εcx − εcy

The transformation from the XY coordinate system to a new coordinate system, as per-
10

formed in equations 2.6 and 2.7, is accomplished using the strain transformation matrix,
[Tstrain ]:
 
cos2 (ω) sin2 (ω) cos(ω) sin(ω)
 
 2 2 
[Tstrain (ω)] =  sin (ω) cos (ω) − cos(ω) sin(ω)  (2.9)
 
2 2
−2 cos(ω) sin(ω) 2 cos(ω) sin(ω) cos (ω) − sin (ω)

where ω is measure counterclockwise from the X axis.


Even though DSFM-TG is a 2D plane stress RC material model, the out-of-plane con-
crete compression stress can have an impact in the in-plane response of concrete. However,
while the total stress in the out-of-plane direction is zero, in the presence of out-of-plane
steel, concrete out-of-plane stress may not be zero. Typically in a RC concrete continuum
out-of-plane compression stress is induced when concrete expansion is restrained by out-of-
plane steel. To simulate this effect, the out-of-plane RC strain, εz is computed, and the
average out-of-plane concrete, εcz , and steel strains, εsz , are assumed to be equal to that of
the RC continuum.

εcz = εz (2.10)

εsz = εz (2.11)

2.3.2 Constitutive Relationships

The DSFM-TG uses 1D constitutive models to find the stresses in the concrete and steel
stresses in the average region. The stresses are calculated in the material coordinate systems,
αi and 12c , as shown in Figures 2.3 and 2.4, respectively. The steel and concrete constitutive
relationship are presented below.

Steel

The steel in the average region is assumed to carry stress only in the longitudinal direction
of the steel bars. The steel will likely provide some resistance due to dowel action; however,
11

Figure 2.3: Average Steel Stresses

Figure 2.4: Average Concrete Stresses


12

in the average region, dowel action will be dominated by steel action in the longitudinal
direction, and, thus, is ignored. In the cracking region, slip on the crack surface can result
in significant deformation perpendicular to the bar axis and thus significant dowel action.
Dowel action is explained in Section 2.4

Longitudinal Steel Response: The longitudinal stress, fsi , in the in-plane layer i is a
function of the longitudinal steel strain component εsi of the of the {εs }αi strain vector.

fsi = fs (εsi , material model parameters) (2.12)

Any 1D steel constitutive model could be included in the model. In the current DSFM-
TG implementation the following models are implemented:

• Bilinear

• Trilinear

• Raynor [Raynor (2000)]

The Bilinear model is made of linear elastic portion and a flat yielding plateau. The
Trilinear model is similar to the Bilinear model, but it also has a hardening portion. The
Raynor model is similar to the trilinear model but differs from it in that it has a sloped
(non-flat) yielding plateau and the hardening portion is curved instead of linear. A detailed
description of these models can be found in Section 2.5. The same constitutive model used
to calculate fsi is used to calculate the out-of-plane steel stress fsz .

fsz = fs (εsz , material model parameters) (2.13)

Concrete

The 2D concrete response is defined by 1D constitutive models acting in the principal


concrete stress directions. The concrete 1D constitutive models are categorized as tensile
and compressive response models. Whether a tensile or compressive constitutive model is
13

used for the calculation of fcj (j = 1, 2, z) depends on the load component of the strain in
that direction εload
cj .



tensile εload
cj ≥ 0,
concrete response = (2.14)

compressive εload < 0
cj

The load strain, εload


cj , is defined as

εload
cj = εcj − εpoisson
cj (2.15)

where εpoisson
cj is the strain due to Poisson effect. A more detailed explanation is presented
in Section 2.7.

Tensile Concrete Response: The tensile concrete stress fcj (j = 1, 2, z) is a function


of εload
cj .

fcj = fctension (εload load


cj , material model parameters), εcj ≥0 (2.16)

Any constitutive models could be included in the DSFM-TG model. In the current
implementation, the following models are implemented:

• Vecchio1982 [Vecchio and Collins (1986)]

• LNRBentz [Vecchio (2000)]

The Vecchio1982 model was the original model included in the MCFT, and the LNR-
Bentz is the model used in the DSFM. Both models consists of a linear precracking portion.
After cracking the Vecchio1982 uses a tension softening stress model, while the LNRBentz
uses a combination of tension softening and stiffening to determine the stresses. A detailed
description of these models can be found in Section 2.5.

Compressive Concrete Response: The compressive concrete stress fcj (j = 1, 2, z) is


a function of εload
cj and the modified peak stress and strain at peak, fpj and εpj .
14

fccompression (εload load


cj , εpj , fpj , material model parameters), εcj <0 (2.17)

The parameters εpj and fpj are the modified peak stress and strain at peak in the
direction of fcj due to compression softening and confinement. A detailed explanation of
how εpj and fpj are determined is presented in Section 2.7. The concrete constitutive models
in DSFM-TG are

• Hognestad [Hognestad (?)]

• PopovicHS [Popovic (1973)]

The Hognestad model consists of a parabolic curve symmetric about the peak concrete
stress. The PopovicHS model accounts for the variations in shape of the stress-strain curve
for different concrete strengths. A detailed description of these models can be found in
Section 2.5.

2.3.3 Equilibrium

Equilibrium needs to be achieved between the external and internal RC material stresses.
Since the DSFM-TG model is a 2D plane stress material model the RC stress in the out-
of-plane direction must be equal to zero:

ρz fsz + fcz = 0 (2.18)

where ρz is defined as the ratio of out-of-plane steel in the out-of-plane direction z.


Consideration of equilibrium in the in-plane direction results in nodal forces that equilibrate
average RC stresses.

2.4 Cracking Region Modeling

Once the principal tensile stress in the concrete exceeds the cracking strength, cracks
develop in the concrete continuum and new steel and concrete stresses arise in the vicinity of
the crack. These stresses are in equilibrium with the average RC region stresses. However,
15

it is assumed that concrete tensile stress on the crack surface is zero, thus, equilibrium of
cracking and average stresses requires additional stresses to develop through other mech-
anisms. These mechanisms include increases steel stress in the longitudinal, dowel action,
and shear on the concrete crack surface, which requires slip at the concrete crack surface.

2.4.1 Compatibility

The geometry of cracked concrete is described in Figure 2.5. It can be observed in


Figure 2.5 that slip on the crack surface does not result in deformation in the average
concrete region because concrete between cracks is assumed to move as a rigid body in
presence of slip. The slip does introduce a localized displacement that when smeared over
the average region, results in an average slip strain, εslip . Equation 2.19 shows the average
slip strain vector representation in the 12c coordinate system.

Figure 2.5: Cracking Region Geometric Variables

 

 0 


 

 
{εslip }12c = 0 (2.19)

 

 δ  
 slip 

Scr
16

where δslip and Scr are as defined in Figure 2.5. The sum of the average concrete strain and
the average slip strain must equal the average RC strain to maintain geometric compatibility.
This requirement, written in the XY coordinate system, is

[Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ) = {ε} (2.20)

where εc1 and εc2 are the concrete strains in the average region in the 1sc and 2sc
directions respectively. The average concrete strain can be seen as the average RC strains
minus the average slip strain for the average region. Equation 2.20 can be reduced to
   

 εc1 
 

 
  εx 

    
[Tstrain (−θ12c )] · εc2 = εy (2.21)

   

 δ     
 γxy 
 slip 
Scr
The longitudinal steel stresses increase in the cracking region because steel must carry
the tensile load dropped by the concrete. The strain in the ith in-plane steel layer at the
crack , εscri , is
εscri = εsi + ∆ε1cr cos2 (φi ) (2.22)

where
φi = θ12c − αi (2.23)

and ∆ε1cr is the strain increase in a steel layer that is oriented perpendicularly to the
crack. The cos2 (φi ) reduces the strain increase for a steel layer oriented at an oblique angle
to the crack. The dowel displacement (perpendicular to bar axis) of the ith in-plane steel
layer at the crack, ∆dcri , is
∆dcri = δslip cos(φi ) (2.24)

This displacement defines the dowel response of the steel.


The crack width, wcr , and spacing, Scr , are estimated using the crack width and spacing
17

models.

Scr = Scr (Smx , Smy , θ12c ) (2.25)

wcr = wcr (εc1 , Scr ) (2.26)

The crack spacing and width models included in DSFM-TG are

• Basic crack spacing model [Vecchio (2000)]

• Basic crack width model [Vecchio (2000)]

These two models were part of the original MCFT formulation and are also part of the
DSFM. The Basic crack spacing model calculates the crack spacing by using set cracking
spacing Smx and Smy values for the x and y direction, and the angle that defines the
orientation of the crack θ12c . The Basic crack model estimates the crack width by assuming
that all the RC elongation in the 1sc direction occurs at the crack. A detailed description
description of these models is presented in Section 2.6.

2.4.2 Constitutive Relationships

The DSFM-TG recognizes three types of stresses at the crack: longitudinal and dowel steel
stresses and shear stress on the crack surface due to aggregate interlock.

Steel:

Longitudinal Response: The longitudinal steel stress at the crack, fscri , in the in-plane
layer i is a function of the longitudinal steel strain at the crack ,εscri , defined in Equation
2.22:

fscri = fs (εscri , material model parameters) (2.27)

The specific constitutive model used in the cracking regions is the same as that in average
region.
18

Dowel Action Response: The dowel action stresses at the crack are caused by the
bending of the bars due to the dowel slip, ∆dcri , displacement at the crack (Equation 2.24).
The dowel stress for the in-plane layer i is defined:

fdcri = fd (∆dcri , material model parameters) (2.28)

Relatively few dowel response models have been proposed by previous researchers. The
model implemented in the DSFM-TG constitutive model is proposed by He-Kwan (2001).
This model is described in detail in Section 2.6.

Concrete:

Aggregate Interlock Response: The shear stress on the concrete crack surface, vc ,
arises due to the aggregate interlock resistance against the slip δslip . It is a function of the
slip on the crack surface δslip , and the crack width, wcr :

vc = vc (δslip , wcr , material model parameters) (2.29)

The model proposed by Walreven (1981) is one the very few aggregate interlock models
in the literature. This model is implemented in the DSFM-TG and it is disucrssed in detail
in Section 2.6.

2.4.3 Equilibrium

The stresses in the average and cracking region equilibrate to maintain the internal equilib-
rium in the RC continuum. Figure 2.6 shows the stresses acting on a piece of the continuum
concrete. Stresses acting on the left and bottom and top surface belong to the average
region. The stresses acting on the right surface belong to the cracking region. The following
equation defines equilibrium normal to the crack orientation:

np
X np
X np
X
fc1 + ρi fsi cos2 (φi ) = ρi fscri cos2 (φi ) + ρi fdcri cos(φi ) sin(φi ) (2.30)
i=1 i=1 i=1
19

where np is the number of in-plane steel layers and all the other variables are as previously
defined. The left side of the equation represents the stresses normal to the crack in the
average region and the right represents the stresses in the cracking region. Equation 2.30can
be simplified to yield:

np
X np
X
2
fc1 − ρi (fscri − fsi ) cos (φi ) − ρi fdcri cos(φi ) sin(φi ) = 0 (2.31)
i=1 i=1

Enforcing equilibrium in the direction parallel to the crack surface results in the following
equation:

np
X np
X np
X
ρi fsi sin(φi ) cos(φi ) = −vc + ρi fscri sin(φi ) cos(φi ) − ρi fdcri cos2 (φi ) (2.32)
i=1 i=1 i=1

The left side of the equation represents the stresses tangent to the crack in the average region
and the right represents the stresses in the cracking region. Simplification of Equation 2.32
results in the following:

np
X np
X
vc − ρi (fscri − fsi ) sin(φi ) cos(φi ) + ρi fdcri cos2 (φi ) = 0 (2.33)
i=1 i=1

From Equation 2.33 it can be seem that a symmetric steel layer configuration about a
line normal to the crack would not generate shear stress vc .

2.5 1D Average Region Models

This section presents the 1D constitutive models used to predict the steel and concrete
stresses. These models are intended for use in simulating the response of systems so mono-
tonically increasing loads. A natural extension of the DSFM-TG would be to modify the
presented constitutive models to simulate response under cyclical loading. Since these mod-
els are used in the context of non-linear problem, expressions for their derivatives have been
also derived.
20

Figure 2.6: Stresses at the crack

2.5.1 Longitudinal Steel 1D Models

Three longitudinal response models have been included in the model DSFM-TG . The three
longitudinal models are

• Bilinear

• Trilinear

• Raynor

The Bilinear model is made of linear elastic portion and a flat yielding plateau. The
Trilinear model is similar to the Bilinear model, but it also has a hardening portion. The
Raynor model is similar to the trilinear model but differs from it in that it has a sloped (non-
flat) yielding plateau and the hardening portion is curved instead of linear. Expressions for
the tangent stiffness are presented for each model:
21

Figure 2.7: Bilinear model

Bilinear

This constitutive model consists of an initial linear-elastic response and a zero stress gain
in the post-yield regime. The stress is calculated as:


Eso εs |εs | ≤ εy
fs = (2.34)

f sign(ε ) |ε | > ε
y s s y

where Eso is the initial stiffness, εy is the yielding strain, and fy is the yielding stress. The
model and its parameters are shown in the Figure 2.7.
The tangent stiffness Es is defined as



dfs Eso |εs | ≤ εy
Es = (2.35)
dεs 
0 |εs | > εy

Trilinear

This constitutive model consists of an initial linear portion, a flat yielding plateau, a linear
hardening portion, and a zero-strength portion where the the steel bar fails. The stress is
22

Figure 2.8: Trilinear model

calculated as: 



 Eso εs |εs | ≤ εy





fy sign(εs ) εy < |εs | ≤ εsh
fs = (2.36)



 (fy + Esh (|εs | − εsh ))sign(εs ) εsh < |εs | ≤ εu





0 |εs | > εu

where Esh and εsh are the hardening stiffness and strain, and εu and fu are the ultimate
strain and stress at failure. The model and its parameters are shown in Figure 2.8.
The tangent stiffness, Es is defined as:




 Eso |εs | ≤ εy





dfs 0 εy < |εs | ≤ εsh
Es = (2.37)
dεs 


 Esh εsh < |εs | ≤ εu





0 |εs | > εu
23

Figure 2.9: Raynor model

Raynor

This constitutive model was developed by Raynor (2000) to provide a better representation
of the stress-strain response of steel as observed in lab tests. It differs from the Trilinear
model in that some hardening occurs in the yield plateau, and the yield plateau is followed
by a nonlinear hardening zone. Stress is defined as follows:




Eso εs |εs | ≤ εy





(fy + (|εs | − εy )Esy ) sign(εs ) εy < |εs | ≤ εsh
fs = µ µ ¶ ¶
(εu − |εs |) nsh (2.38)


 u
 f − (fu − fsh ) sign(εs ) εsh < |εs | ≤ εu

 (εu − εsh )



0 |εs | > εu

where Esy is the initial yielding stiffness, fsh is the hardening stress, and the parameter nsh
defines the shape of the hardening portion of the strain-stress curve (nsh = 1 results in a
trilinear response). The model and its parameters are shown in Figure 2.9.
24

Table 2.1: Longitudinal Steel Models

Case Model εy fy εsh fsh εsu fu nsh


[mm/mm] [MPa] [mm/mm] [MPa] [mm/mm] [MPa]
1 Bilinear 2.00E-03 200 - - - - -
2 Trilinear 2.00E-03 200 1.00E-02 - 1.50E-02 220 -
3 Raynor 2.00E-03 200 1.00E-02 205 1.50E-02 220 1
4 Raynor 2.00E-03 200 1.00E-02 205 1.50E-02 220 3
5 Raynor 2.00E-03 200 1.00E-02 205 1.50E-02 220 6

The tangent stiffness Es is defined:






Eso |εs | ≤ εy





dfs Esy µ ¶
εy < |εs | ≤ εsh
Es = (ε − |ε |)nsh −1 (2.39)
dεs 
nsh (fu − fsh )
u s
εsh < |εs | ≤ εu



 (εu − εsh )nsh



0 |εs | > εu

Figure 2.10 shows a plot of the different analysis cases contained in Table 2.1.

2.5.2 Tensile Concrete 1D Models

The tensile behavior of concrete before cracking is typically considered to be linear elastic
with a material stiffness similar to that observed under small compressive loads. After
cracking, concrete strength decreases with increasing crack width. Post-cracking tension
response is simulated using two models: tension softening and tension stiffening [Vecchio,
(2000)].
The tension softening model is intended to simulate the reduction in tensile strength
observed immediately after cracking has occurred as the concrete crack opens. The tension
softening model predicts a rapid loss in strength that is linear with the crack opening. The
tension stiffening model is intended to simulate the tensile stiffness provided by uncracked
concrete between cracks due bonding between concrete and steel.
25

250

200

150
Case 1: Bilinear
Case 2: Trilinear
Case 3: Raynor, nsh = 1
Case 4: Raynor, n = 3
sh
100 Case 5: Raynor, nsh = 6

50

0
0 0.005 0.01 0.015
εs [mm/mm]

Figure 2.10: Longitudinal Steel Models


26

The tensile concrete constitutive models included in DSFM-TG are

• Vecchio1982 [Vecchio (1986)]

• LNRBentz [Vecchio (2000)]

The Vecchio1982 model was the original model included in the MCFT, and the LNR-
Bentz is the model used in the DSFM. Both models consists of a linear precracking portion.
After cracking the Vecchio1982 uses a tension softening stress model, while the LNRBentz
uses a combination of tension softening and stiffening to determine the stresses. Expressions
for the following derivatives have been derived for each model:

∂fc
• Ec =
∂εc
∂fc

∂{φ}
∂fc

∂Scr

where {φ} is a column vector containing the angles φi at which the steel layers cross the
crack.  
 
 φ1 

 


 

 φ2 
 

 

{φ} = .. (2.40)

 

 .. 
 

 


 

φ 
 
np

∂fc ∂fc
The derivatives and are added to these models to account for the presence of
∂{φ} ∂Scr
the {φ} and Scr input variables. The derivatives are used to solve the internal RC problem,
and also to construct the RC stiffness matrices as described in detailed in Chapter 3.

Vecchio1982

This model was originally included in the MCFT developed by Vecchio and Collins (1986),
and it is based on data from panel tests performed at the University of Toronto by Vec-
chio and Collins (1982). This model is appropriate for RC material with well distributed
27

reinforcement. The stress is calculated as:





Ecto εc εc ≤ εccr
fc = 0 (2.41)

 f
 √t εc > εccr
1 + 200εc

0
where Ecto is the initial stiffness and εccr and ft are the cracking strain and stress. After
cracking, concrete strength decreases with increasing crack width. Post-cracking tension re-
sponse is simulated using a tension stiffening model. The tension stiffening model simulates
the tensile stiffness provided by uncracked concrete between cracks due to the bonding con-
crete and steel between cracks. occurs due to the bonding due of the steel and the concrete
between cracks. Tension stiffening is a reflect of the interaction between steel and concrete.
The model and its parameters are shown in Figure 2.11.

Figure 2.11: Vecchio1982 model

The following derivatives are defined:





∂fc Ecto εc ≤ εccr
Ec = = 0 (2.42)
∂εc 
 100ft
− √ √ εc > εccr
200εc (1 + 200εc )2
∂fc
= {0, 0, ..0} (2.43)
∂{φ}
∂fc
= 0 (2.44)
∂Scr
28

LNRBentz

The LNRBentz (Linear tension softening with no residual and Bentz tension stiffening)
was originally included in the DSFM model by Vecchio (2000)]. It is a combination of
linear tensions softening with no residual and the Bentz (?) tension stiffening model. The
tension softening model simulates the reduction in tensile strength observed immediately
after cracking has occurred as the concrete crack opens. The LNRBentz model uses tension
softening to capture the behavior immediately after cracking. As the cracks opens the
tension stiffening mechanism more becomes more prominent than the tension softening.
Once the stress predicted by the tension softening model is lower than that of tension
stiffening, the LNRBentz switches to the tension stiffening curve as shown in Figure 2.12.
The tensile stress is calculated as:


Ecto εc εc ≤ εccr
fc = (2.45)

max(f , f ) ε > ε
ca cb c ccr

Where fca is the softening stress and fcb is the stiffening stress. The softening stress fca is
calculated as: µ ¶
0 εc − εccr
fca = ft 1 − (2.46)
εts − εccr

where εts is the terminal strain or strain at which the tension stress fca becomes zero. The
εts strain is calculated as:

2Gf
εts = 0, 1.1 εccr 6 εts 6 10 εccr (2.47)
Lr ft

where Gf is a fracture energy parameter and Lr is the characteristic length. The parameter
Gf is a measurement of the energy required to form a crack a unit area of crack. It is
considered to be a fundamental material property and maybe determined from experimental
test data. The parameter Lr is the characteristic length defined as the length over which
the crack damage is assumed to be distributed. The values of Gf and Lr are assigned the
29

following values:

N
Gf = 75 (2.48)
m
Scr
Lr = (2.49)
2

The above values were selected from the DSFM [Vecchio (2000)] and the Vector2 manual
[Wong and Vecchio (2002)].
The stiffening stress fcb is based on the Bentz 2003 model [Vecchio (2000)] and is calcu-
lated as:
0
f
fcb = √t (2.50)
1 + Ct εc

where Ct is a parameter that estimates the degree of bonding between concrete and steel
and it is calculated as:

Ct = 2.2 m (2.51)
1
m = (2.52)
Pnp 4ρi | cos(φi )|
i=1
dbi

where the parameter m reflects the ratio between the bonding area and the tributary RC
area. The bar diameter of the in-plane layer i bars dbi is in [mm]. The model and its
parameters are shown in Figure 2.12. The following derivatives are defined:




Ecto εc ≤ εccr


∂fc
Ec = = Eca εc > εccr and fca > fcb (2.53)
∂εc 




Ecb εc > εccr and fca 6 fcb




{0, 0, ..0} εc ≤ εccr


∂fc
= {0, 0, ..0} εc > εccr and fca > fcb (2.54)
∂{φ} 




 −m2 {t} εc > εccr and fca 6 fcb

(2.55)
30

Figure 2.12: LNRBentz model

where {t} and {0,0,..0} are row vectors of the same length as the number of in-plane
steel layers np.




0
 εc ≤ εccr


∂fc (εc − εccr ) Gf
= − εc > εccr and fca > fcb (2.56)
∂Scr 
 (ε 2 2
ts − εccr ) Lr



0 εc > εccr and fca 6 fcb

where

0
−ft
Eca = (2.57)
εts − εccr
0
Ct ft
Ecb = − √ √ (2.58)
2 Ct εc (1 + Ct εc )2
½ ¾
ρ1 sin(φ1 )sign(cos(φ1 )) ρnp sin(φnp )sign(cos(φnp ))
{t} = −4 , ..., (2.59)
db1 dbnp

In order to show the differences between the tensile models and the impact of the input
parameters on the stress, a set of strain-stress curves case have been developed. They are
listed in Table 2.2. Figure 2.13 shows the plots associated to the cases shown in Table
2.2. Case 1 is the basic case against which the other cases are compared. Case 2 shows
how an increase in the cracking spacing Scr reduces the tensile softening stress, but it does
31

Table 2.2: Tensile Concrete Models


0
Case Model Scr φ ρ db fc
[mm] [MPa] % [mm] [MPa]
1 LNRBentz 40 0 1 5 -20
2 LNRBentz 60 0 1 5 -20
3 LNRBentz 40 10 1 5 -20
4 LNRBentz 40 0 5 5 -20
5 LNRBentz 40 0 1 3 -20
8 Vecchio1986 40 0 1 5 -20

1.5

0.5
Case 1
Case 2
Case 3
Case 4
Case 5
Case 8

0
0 0.5 1 1.5 2 2.5 3 3.5
εc [mm/mm] −3
x 10

Figure 2.13: Tensile Concrete Models


32

not affect the tensile stiffening stress. Case 3 shows how using an angle oblique rather than
normal to the crack also reduces the stiffening response, but it does not affect the tensile
softening. Case 4 shows how increasing the amount of reinforcement increases the tensile
stiffening response, but it does not affect the tensile softening. Case 5 shows how using
bars with smaller diameters increases the tensile stiffening stress as a result of the better
bonding, but it does not affect the tensile softening. Case 8 shows the Vechio1982 model,
which does not explicitly includes the influence of the specific values of parameters such as
Scr , db , or ρ.

2.5.3 Compressive Concrete 1D Models

Compression constitutive models define the response of concrete under compression comprise
of multiple linear or curvilinear segments scaled to predict a maximum compression stress of
0
fc at a strain εco . The current implementation modifies the peak stress and strain at peak
to account for the compression softening and confinement effect. The parameters fp and εp
are introduced to represent the adjusted maximum strength, fp , and strain at maximum,
εp . The two compression models incorporated in DSFM-TG are

• Hognestad [Hognestad (?)]

• PopovicHS [Popopovic (1973)]

The Hognestad model consists of a parabolic curve symmetric about the peak concrete
stress. The PopovicHS model accounts for the variations in shape of the stress-strain curve
for different concrete strengths.
Expressions for the following derivatives have been derived for each model:

∂fc
• Ec =
∂εc
∂fc

∂εp
∂fc

∂fp
33

Figure 2.14: Hognestad model

∂fc ∂fc
The derivatives and are added to these models to account for the presence of
∂εp ∂fp
the εp and fp input variables. The derivatives are used to solve the internal RC problem,
and also to construct the RC stiffness matrices as described in detailed in Chapter 3.

Hognestad

The Hognestad (?) model is a simple parabola symmetric about the peak strength point
defined by εp and fp . This is the original model used the MCFT model [Vecchio (1986)].
This model is more appropriate for normal strength concrete. The stress is calculated as:

 Ã µ ¶ µ ¶2 !

 εc εc
fp 2 − εc ≥ 2εp
fc = ε p ε p (2.60)


0 εc < 2εp

Figure 2.14 shows the model and its parameters. The following derivatives are defined:
34

 µ ¶

 2 εc
fp −2 2 εc ≥ 2εp
Ec = εp εp (2.61)


0 εc < 2εp
 µ ¶

 ε ε2
∂fc fp −2 2c + 2 3c εc ≥ 2εp
= εp εp (2.62)
∂εp 

0 εc < 2εp
 µ ¶ µ ¶2

 εc εc
∂fc 2 − εc ≥ 2εp
= ε p ε p (2.63)
∂fp 

0 εc < 2εp

PopovivcHS

The PopovicHS or Popovic High Strength model was developed by Popovic (1973). This
is the model used by Vecchio (2000) in the DSFM formulation. PopovicHS is appropriate
for high strength for which stress-strain response is approximated linear up to stresses ap-
proaching the maximum strength and the stress decays rapidly once the maximum strength
is reached. Using the PopovicHS model, concrete stress is calculated as:
µ ¶
εc
n
εp
fc = fp µ ¶nk (2.64)
εc
n−1+
εp

where the parameter n increases the linearity (decreases the nonlinear effects) of the pre-
peak response.
fp
n = 0.8 − (2.65)
17

and the k parameter increases the rate at which the the post-peak decays. The parameter
k is computed as: 

1 εc ≥ εp
k= (2.66)
 max(0.67 − fp , 1)

εc < εp
62
Notice that the decay parameter k increases for higher strength concretes. Figure 2.15 shows
the model and its parameters.
35

Figure 2.15: PopovicHS model

The following derivatives are defined:


 
µ ¶nk
 n n2 k εc 
 εp εp εp 
 
Ec = fp  µ ¶nk − Ã µ ¶ !2  (2.67)
 
 n − 1 + εc
nk
εc 
εp n−1+
εp
 
2
µ ¶nk+1
 εc n k εc 
 n 2 
∂fc  εp εp εp 
= fp − µ ¶nk + Ã µ ¶ !2 (2.68)
∂εp  
 n − 1 + εc
nk
εc 
εp n−1+
εp
µ ¶ µ ¶µ ¶
εc dn εc dn
n + fp fp n +a
∂fc εp dfp εp dfp
= µ ¶nk − Ã µ ¶nk !2 (2.69)
∂fp εc εc
n−1+ n−1+
εp εp
36

Table 2.3: Compressive Concrete Models

Case Model εp fp
[mm/mm] [MPa]
1 PopovicHS -0.003 -10
2 PopovicHS -0.003 -30
3 PopovicHS -0.003 -50
6 Hognestad -0.003 -10
7 Hognestad -0.003 -30
8 Hognestad -0.003 -50

where
µ ¶ µ ¶µ ¶
εc nk εc dn dk
a = log k+n (2.70)
εp εp dfp dfp
dn 1
= − (2.71)
dfp 17


0
dk k=1
= (2.72)
dfp  1 k 6= 1

62

In order to show the differences between the compression models and the impact of the
input parameters on the stress, a set of strain-stress curves case have been developed. They
are listed in Table 2.3.
Figure 2.16 shows a plot of the different cases contained in Table 2.3. The PopovicHS
cases 1, 2, and 3 stress-strain curves shape varies significantly as the fp changes. Case 3
shows a more triangular shape about the peak, while case 1 (weaker concrete) is practically
flat after peak, which suggests that this model does not accommodate well the behavior
of low strength concretes. The Hognestad cases 4, 5, and 6 have stress-strain curves that
maintain the same parabolic shape independent of the concrete strength.
37

−5

−10

−15

−20

−25

−30

−35 Case 1:PopovicHS


Case 2:PopovicHS
−40 Case 3:PopovicHS
Case 6:Hognestad
Case 7:Hognestad
−45
Case 8:Hognestad

−50
−7 −6 −5 −4 −3 −2 −1 0
εc [mm/mm] −3
x 10

Figure 2.16: Compressive Concrete Models

2.6 1D Cracking Region Models

This section presents the models that estimate crack spacing,Scr , width,wcr , and slip stress
along the crack, vc . These quantities are important to determine the internal state of the RC
material. Since these models are used in the context of a non-linear problem, expressions
for their derivatives have been also develop.

2.6.1 Crack Spacing Model

The crack spacing model used in DSFM-TG is that used by Vecchio (2000) in the original
DSFM formulation, and it is refered to as the basic crack spacing model. The average crack
spacing perpendicular to the crack, Scr , is calculated as:

1
Scr = (2.73)
| cos(θ12c )| | sin(θ12c )|
+
Smx Smy
38

where Smx and Smy are the average crack spacings in the X and Y direction, respec-
tively. Figure 2.17 shows how Scr varies with the inclination of the principal concrete
strain, θ12c . The Scr matches Smx when the principal concrete stress direction is hori-

Case 1: S = 50mm, S == 50 mm
mx my
Case 2: S = 40mm, S == 60 mm
mx my

−315 −270 −225 −180 −135 −90 −45 0 45 90 135 180 225 270 315 360
θ [Deg]
12c

Figure 2.17: Crack Spacing Basic Model

zontal (θ12c = −360◦ , −180◦ , 0◦ , 180◦ , 360◦ ) and Smy when the direction is vertical (θ12c =
−270◦ , −90◦ , 90◦ , 270◦ ). The following derivative is also defined:
µ ¶
dScr 2 − sin(θ12c )sign(cos(θ12c )) cos(θ12c )sign(sin(θ12c ))
= −Scr + (2.74)
dθ12c Smx Smy

2.6.2 Crack Width Model

The crack width model used in DSFM-TG comes from from Vecchio (2000) DSFM formu-
lation, and it is named basic crack width model herein. The average crack width, wcr , is
calculated as:
wcr = Scr εc1 (2.75)

This model estimates the crack width by assuming that all the concrete elongation in the
maximum principal direction is concentrated at the crack. In reality this expression is an
upper bound for the crack width since some elongation occurs between cracks. The two
39

following derivatives are defined:

∂wcr
= εc1 (2.76)
∂Scr
∂wcr
= Scr (2.77)
∂εc1

2.6.3 Aggregate Interlock Model

The aggregate interlock model used in DSFM-TG is based on the Walraven (1981) model
which is also used in the DSFM. This model is named Walraven herein, and it estimates
the shear stress vc at the crack:
vc = δslip fwcr (2.78)

where 

1.8 wcr
−0.8 + max((0.234 w −0.707 − 0.2), 0) · |f | w 6= 0
cr cc cr
fwcr = (2.79)

0 wcr = 0

where fcc is the concrete cube strength in [MPa]. In this mode, fwcr is intended to simulate
the impact of the concrete strength on the shear stress vc . Thus, fwcr becomes zero for wcr
greater than about 1.25 mm. Also fwcr is zero when wcr is zero, as there cannot be shear
on the crack surface if the crack is not open. And vc tends to infinity as wcr tends to zero.
In order to illustrate the Walraven model’s behavior, a set of cases have been plotted
in Figures 2.18 and 2.19. Figure 2.18 shows how vc rapidly decreases as wcr increases, and
that the fcc value does not influence vc for cracking widths over about 1.25 mm. Figure
2.19 shows the linear relationship between vc and wcr for a constant wcr . Cases In Figure
2.19, cases 6 and 8 lay on top of each other because the value of fcc does not influence vc
wcr greater than 1.25 mm.
For this model, the following two derivatives are defined:

∂vc dfwcr
= δslip (2.80)
∂wcr dwcr
∂vc
= fwcr (2.81)
∂δslip
40

Case 1:δ =1 mm,f =20 MPa


slip cc

Case 2:δslip=2 mm,fcc=20 MPa

Case 3:δslip=1 mm,fcc=40 MPa

Case 4:δslip=2 mm,fcc=40 MPa

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2


wcr [mm]

Figure 2.18: Aggregate Interlock Model, vc vs. wcr

Case 5:wcr=1 mm,fcc=20MPa

Case 6:w =2 mm,f =20MPa


cr cc

Case 7:wcr=1 mm,fcc=40MPa

Case 8:wcr=2 mm,fcc=40MPa

0.5 1 1.5 2
δslip [mm]

Figure 2.19: Aggregate Interlock Model, vc vs. δslip


41

where 

−1.44wcr
−1.8 − 0.1654w −1.707 |f | w ≤ 1.25 mm
dfwcr cr cc cr
= (2.82)
dwcr 
−1.44w−1.8
cr wcr > 1.25 mm

2.6.4 Dowel Action Model

The dowel action model included in DSFM-TG is based on the work by He and Kwan (?)
in which the dowel action is modeled assuming that the rebar acts as a beam on an elastic
foundation. The dowel force for a single bar of the in-plane layer i is computed as:

Vdi = Esi Isi λ3i ∆dcri < Vdu (2.83)


q
2
Vdui = 1.27dbi |fc0 |fy (2.84)
πd4bi
Isi = (2.85)
64
r
4 kci dbi
λi = (2.86)
4Esi Isi
p
127c fc0
kci = 2/3
(2.87)
dbi
(2.88)

where ∆dcri is the dowel displacement, λi is a parameter representing the relative stiffness
of the bar to the surrounding concrete, Vdui is the ultimate shear force associated with the
flexural hinging of the bar and crushing of the surrounding concrete, kci is a parameter
associated with the concrete stiffness, and c is a parameter associated with the bar spacing,
which has been set to be 0.8 in this model. The variables dbi and Isi are the diameter and
the moment of inertia of the in-plane layer i bars. The shear stress at the crack due to
dowel action in a single bar of the in-plane layer i, fdcri , is calculated as:

Vdi
fdcri = (2.89)
Asi

where Asi is the area of the bar and Vdi is as calculated in Equation 2.83.
42

2.7 Multiaxial Loading Effects on 1D Models Modeling

The 1D constitutive models were developed using experimental data from 1D tests. To
use these models for multiaxial load conditions, it is necessary to account for the effects of
multiaxial loading on those 1D constitutive models. The multiaxial loading effects consid-
ered in the DSFM-TG are

• Strain due to Poisson effect versus loading

• The impact of multiaxial concrete stresses on concrete compressive response parame-


ters

The first effect refers to the distinction that must be made between the load-driven com-
ponent of the material strain and the strain due to Poisson effect. The second refers to the
impact of the multi-axial stress state on the 1D concrete compression response. Ideally, the
model would simulate strength and ductility gain due to confinement as well as strength
loss due to tension in perpendicular direction. However, the current model accounts only
0
for modifications on the peak strength, fc , and strain at peak, εco , and not on the ductility.
The DSFM-TG does not account for other multi-axial effects such as the effect of lateral
load on the concrete tensile response or dowel-longitudinal steel stress interaction. A series
of multi-axial loading parameters were defined to quantify the multiaxial load effects listed
above:

• Poisson’s ratio, ν

• Confinement modification factor, βl

• Compression softening modification factor, βd

2.7.1 Poisson Effect versus Load Induced Strains

Poisson strains arise due to loading in the plane perpendicular to that under consideration.
Both concrete and steel are subjected to strain due to the Poisson effect. The Poisson
43

induced concrete strains are defined as:

εpoisson
c1 = −ν2 εload load
c2 − νZ εcz , (2.90)

εpoisson
c2 = −ν1 εload load
c1 − νZ εcz , (2.91)

εpoisson
cz = −ν1 εload load
c1 − ν1 εc2 , (2.92)

where νj , εload
cj (j = 1, 2, z) are the Poisson’s ratios and the load strain in the direction of fcj .

The load strain εload


cj is the component of the strain due to the fcj stress. The load concrete
strain εload
cz in the out-of plane direction is zero if there is no out-of-plane reinforcement.
The average concrete strains are the sum of the load and Poisson induced concrete strains:

poisson
εc1 = εload
c1 + εc1 = εload load load
c1 − ν2 εc2 − νZ εcz , (2.93)
poisson
εc2 = εload
c2 + εc2 = εload load load
c2 − ν1 εc1 − νZ εcz , (2.94)

εcz = εload poisson


cz + εcz = εload load load
cz − ν1 εc1 − ν2 εc2 , (2.95)

The equations above can be written as

{ε3P
c
D
} = [Po ] · {ε3P
c
D,Load
}, (2.96)
44

where
 

 ε 
 c1 
 
{ε3P
c
D
} = εc2 (2.97)

 

ε 
 
cz
 

ε load 


 c1  
{ε3P
c
D,Load
} = εload
c2 
(2.98)

 

εload 
cz
 
1 −ν2 −νZ
 
 
[Po ] = −ν1 1 −νZ  (2.99)
 
−ν1 −ν2 1

The principal concrete load strain vector {ε3P


c
D,Load
} can be written as a function of the
principal concrete strain vector {ε3P
c
D } as

{ε3P
c
D,Load
} = [Po ]−1 · {ε3P
c
D
} = [P ] · {ε3P
c
D
} (2.100)

where

[P ] = [Po ]−1 (2.101)


 
1 − νZ ν2 ν2 (1 + νZ ) νZ (ν2 + 1)
1


= ν1 (1 + νZ ) 1 − ν1 νZ νZ (1 + ν1 ) (2.102)
a 
ν1 (ν2 + 1) ν2 (1 + ν1 ) 1 − ν1 ν2
a = 1 − νZ ν2 − ν1 ν2 − 2ν1 νZ − ν1 νZ (2.103)

If no out-of-plane reinforcement is present then εload


cz is zero and Equation 2.100 simplifies

to:
45

εcz = −ν1 εload


c1 − ν2 εc2
load
(2.104)
1
εload
c1 = (εc1 + ν2 εc2 ) (2.105)
1 − ν1 ν2
1
εload
c2 = (ν1 εc1 + εc2 ) (2.106)
1 − ν1 ν2
εload
cz = 0 (2.107)

The load concrete strains, εload load load


c1 , εc2 , and εcz , are appropriate for use in calculating the

concrete average stresses, fc1 , fc2 , and fcz . The νj Poisson’s ratios are functions of εload
cj ,

and εpj :
νj = ν(εload
cj , εpj , material model properties) (2.108)

where εpj is the modified strain at peak due to compression softening and confinement. The
Poisson modulus models included in DSFM-TG are

• Variable Poisson’s Ratio [Kupfer (1973)]

• Linear Poisson [Wong and Vecchio (2002)]

The Variable Poisson’s Ratio model is used for concrete in compression, and it varies the
value of ν based on the ratio between the compressive strain and strain at peak. The Linear
Poisson model is used for concrete in tension, and it varies the value of ν based on the ratio
between the tensile strain and the cracking strain. These models are described in detail in
Section 2.8.
The longitudinal and normal average steel strains, εsi and εsni , can be defined in terms
of load and Poisson induced steel strains:

poisson
εsi = εload
si + εsi = εload load
si − νsteel εsni (2.109)
poisson
εsni = εload
sni + εsni = εload load
sni − νsteel εsi (2.110)

However, the impact of Poisson induced strain is ignored for reinforcing steel. This simpli-
fication is justified by the mechanism of concrete-steel load transfer. The ribs existing in
46

deformed rebars provide strain compatibility between concrete and steel in the direction of
the bar axis, but do not do so as effectively in the radial direction. Thus, radial steel may
not be consistent with the average RC strains. For the current implementation it is assumed
that the radial strains due to concrete stresses surrounding the bar are small enough to be
ignored and:
εload
sni = 0 (2.111)

is an acceptable simplification. Thus,

εsi = εload
si (2.112)

εsni = −νsteel εload


si (2.113)

The longitudinal steel strain, εsi , is used as an approximation of the longitudinal load
steel strain, εload load
si , to calculate the longitudinal steel stress fsi . The calculation of εsni may

become necessary if a bond-slip model is added to this model in the future.

2.7.2 Impact of Multiaxial Concrete Stresses on Concrete Compressive Response Parame-


ters

The 1D concrete compression constitutive models used in DSFM-TG are derived from
1D tests. However, testing under multi-dimensional loading provides data for modifying 1D
models to account for various multiaxial load scenarios.
The results of 1D multi-dimensional testing indicate that the 1D concrete stress-strain
curve can be determined by two parameters. The results of experimental tests indicate also
0
that 1D response with multi-dimensional loading can be determined by modifying fc and
εco . Thus, a modified peak stress, fp , and strain, εp , are computed as follows for use in the
47

basic 1D constitutive model:

0
fp = β σ fc (2.114)

εp = β ε εco (2.115)

β ε = β ε (βdε , βlε , εc , εco ) (2.116)

β σ = βdσ βlσ (2.117)

where βd (< 1) and βl (> 1) are the modification factors due to compression softening and
confinement, respectively. The superscript σ and ε stand for the specific parameter that
it is being modified. In general compression softening and confinement models could be
strength-only, or strength-strain. Strength-only:

βdσ = βd (2.118)

βlσ = βl (2.119)

βdε = 1 (2.120)

βlε = 1 (2.121)

Strain-strength:

βdσ = βd (2.122)

βlσ = βl (2.123)

βdε = βd (2.124)

βlε = βl (2.125)

Compression softening is the reduction in the concrete peak response due to transverse
cracking. Confinement is the increase in the concrete peak response due to compressive
stresses in the plane normal to the direction in which the compressive response is being
48

predicted. The factor β ε is calculated as:






 βdε βlε βdε βlε ≥ 1 or εc ≥ βdε βlε εco


εc
βε = εc < βdε βlε εco (2.126)

 ε co



ε
co εc < εco

This ensures that the compression concrete response descends only after εc < εco [Wong and
Vecchio, (2002).The compression softening factor βd is a function of the total compressive
concrete strain εc in the direction under consideration (i.e. not εload
c ) and the perpendicular
tensile strain εperp
c , and it is calculated as

βd = βd (εc , εperp
c , material model properties) , for εc ≤ 0, εperp
c > 0, wcr > 0(cracked)
(2.127)
If any of the conditions shown on the right of Equation 2.127 are not met, there is no
reduction in the compressive response due to compression softening, and βd = 1. Since
there are two principal strains normal to the direction under consideration, εperp
c is defined
as the maximum of these strains

εperp
c = max(εperp perp
c1 , εc1 ) (2.128)

where the εperp


c1 and εperp
c1 strains are total concrete strains. Figure 2.20 shows the perpendic-
ular concrete principal strains εperp
c1 and εperp
c1 . The parameters βd2 , and βdz are calculated

as:

βd2 = βd (εc2 , εc1 ) (2.129)

βdz = βd (εcz , εc1 ) (2.130)

The compression softening parameter βl1 is always 1, because the crack orientation is never
transverse to the direction 1sc (it is always normal). The two models available in DSFM-
TG are
49

Figure 2.20: Perpendicular Concrete Principal Strains


50

• Vecchio1986 [Vecchio and Collins (1986)]

• Vecchio1992A [Vecchio (2000)]

The Vecchio1986 model is a strength-only model and was originally included in the MCFT
model. The Vecchio1992A model is a stress-strain model, and it is included in the DSFM.
These models are described in detail in Section 2.8.
The confinement parameter βl is a function of the stresses acting in the plane normal
that under consideration

βl = βl (fcmin , fcmax , material model properties) , for fcmin < 0 (2.131)

where the perpendicular principal stresses fcmin and fcmax are shown in Figure 2.21. If the
condition shown in Equation 2.131 is not met (i.e. there is not compression in at least
one perpendicular direction) then there is no confinement, and βl = 1. The confinement
parameters βl1 , βl2 , and βlz are calculated as:

βl1 = βl (min(fc2 , fcz ), max(fc2 , fcz ), material model properties) (2.132)

βl2 = βl (min(fc1 , fcz ), max(fc1 , fcz ), material model properties) (2.133)

βlz = βl (min(fc1 , fc2 ), max(fc1 , fc2 ), material model properties) (2.134)

The simultaneous incorporation of the effect of compression softening and confinement


allows DSFM-TG to estimate the cases when both effects are present, such as in Figure
2.22.

2.8 1D Multiaxial Loading Parameters Models

2.8.1 Poisson’s ratio ν 1D Models

Experimental data indicate that the Poisson’s ratio for concrete is a function of the
concrete strain, and that it increases under compression and diminishes under tension.
To simulate this behavior two Poisson’s models are included in the DSFM-TG , one for
compression and one for tension:
51

Figure 2.21: Perpendicular Concrete Principal Stresses

Figure 2.22: Compression Softening/Confinement Interaction


52

Figure 2.23: Variable Poisson’s Ratio

• Variable Poisson’s Ratio [Kupfer (1973)]

• Linear Poisson [Wong (2002)]

where the first and second models address the tension and compression cases, respectively.
These models are described below:

Variable Poisson’s Ratio

Under uniaxial compression stress, internal micro-cracks develop that increase the lateral
expansion ratio or Poisson modulus. This behavior is modeled by the Variable Poisson’s
ratio model [Kupfer, (1973)]:


 εp

 ν0 ≤ εload
c <0
à µ ¶2 ! 2
ν= 2εload εp (2.135)

 c
≤ min(2.5ν0 , 0.5) εload
 ν0 1 + 1.5 −1 c <
εp 2

where ν0 is the initial isotropic Poisson’s ratio, εp is the modified strain at peak stress, and
εload
c is the load concrete strain. The model and its parameters are shown in Figure 2.23.
The following derivatives are required for solution of the material state and computation
53

of a consistent material tangent:


 µ load ¶

 εc 2
dν 3ν0 2 −1 , εload
c < εp /2 and ν ≤ min(2.5ν0 , 0.5)
= ε p ε p (2.136)
dεload
c 

0, otherwise
 µ load ¶ load

 ε ε
∂ν −3ν0 2 c − 1 c 2 , εload c < εp /2 and ν ≤ min(2.5ν0 , 0.5)
= ε p εp (2.137)
∂εp 

0, otherwise

Linear Poisson

Under uniaxial tension, the Poisson’s ratio remains constant up to cracking. After cracking
has occurred the Poisson’s ratio rapidly decays to zero since most of the deformation occurs
in the vicinity of the crack, which does not generate contraction in the lateral directions.
Using the linear model, the Poisson’s ratio, ν, is calculated as:


ν0 , 0 < εload ≤ εccr ,

 c

 µ load

ε
ν = ν0 2 − c , εccr < εload
c ≤ 2εccr (2.138)

 εccr



0, εload
c > 2εccr

where εccr is the cracking strain. The model and its parameters are shown in Figure 2.24.

Figure 2.24: Linear Poisson


54

The required derivatives are defined:


 ν
dν − 0 εc < εp /2 and ν ≤ min(2.5ν0 , 0.5)
= εccr (2.139)
dεload
c 
0 otherwise
∂ν
= 0 (2.140)
∂εp

2.8.2 Compression softening parameter βl 1D Models

Compression softening refers to the reduction in the compressive concrete response due to
0
the presence of transverse cracking. The factor βl is used to modified the εco , and fc . The
two compression models included in DSFM-TG are:

• Vecchio1986 [Vecchio and Collins (1986)]

• Vechio1992A [Vecchio (2000)]

The first model is strength-only (it modifies only εco ) and the second is a strain-strength
0
model (it modifies both εco and fc ). Expressions for the following derivatives have been
derived for each model:

∂βd

∂εc
∂βd

∂εperp
c

Vecchio1986

This model was originally included in the MCFT model [Vecchio (1986)]. It was developed
to be used with the Hognestad (?) concrete compressive response model. The factor βd is
calculated as:
µ ¶
1
βd = min 1, (2.141)
0.8 + 0.34r
55

Figure 2.25: Vecchio1986 model

where

εperp
c
r = − (2.142)
εco
(2.143)

where εperp
c is the tensile train in the perpendicular direction. Figure 2.25 shows the model
and its parameters. The following derivatives are defined:

∂βd
= 0 (2.144)
∂εc

 0.34βd2
∂βd  βd < 1
= εco (2.145)
∂εperp
c 
0 βd ≥ 1
56

This is a stress-only model, which means that the factor βd only affects the peak stress and
not the strain at peak stress:

βdε = 1 (2.146)
∂βd ε
= 0 (2.147)
∂εc
∂βd ε
= 0 (2.148)
∂εperp
c
βdσ = βd (2.149)
∂βd σ ∂βd
= (2.150)
∂εc ∂εc
∂βd σ ∂βd
= (2.151)
∂εperp
c ∂εperp
c

Vecchio1992A

This model was developed from a panel test conducted at the University of Toronto [Vecchio
and Collins (1982)]. It was developed to be used with the PopovicHS [Popovic (1973)]
compressive concrete model. The factor βd is calculated as:
µ ¶
1
βd = min 1, (2.152)
1 + Cs Cd
Cs = 0.55 (2.153)


0.35(r − 0.28)0.8 r ≥ 0.28
Cd = (2.154)

0 r < 0.28
µ perp ¶
εc
r = min − , 400 (2.155)
εc

Figure 2.26 shows the model and its parameters. The following derivatives are defined:
 µ perp ¶

 εc
∂βd −0.28βd Cs (r − 0.28)−0.2 βd < 1 and r < 400
= ε2c (2.156)
∂εc 

0 otherwise
 µ ¶

 1
∂βd 0.28βd Cs (r − 0.28)−0.2 βd < 1 and r < 400
= εc (2.157)
∂εperp
c 

0 otherwise
57

Figure 2.26: Vecchio1992A model

This is a strain-strength model, which means that the factor βd affects both the peak
stress and the strain at peak.

βdε = βd (2.158)
∂βd ε ∂βd
= (2.159)
∂εc ∂εc
∂βd ε ∂βd
= (2.160)
∂εperp
c ∂εperp
c
βdσ = βd (2.161)
∂βd σ ∂βd
= (2.162)
∂εc ∂εc
∂βd σ ∂βd
= (2.163)
∂εperp
c ∂εperp
c

Figure 2.27 compares the the two presented models using a εperp
c = εco .
58

Case 1: Vecchio1992A
Case 5: Vecchio1986

5 10 15 20 25 30 35 40 45 50
perp
−εc /εc

Figure 2.27: Compression Softening Models


59

2.8.3 Confinement parameter βd 1D Models

The only confinement model included in DSFM-TG is the Kupfer-Richart model. This is
derived by combining the results of a set of biaxial tests performed by Kupfer (1969) and
uniform pressure triaxial tests performed by Richart (1928). The factor βl is calculated as:

βl = 1 + βlKupf er + βlRichart , (2.164)

where βlKupf er and βlRichart are the Kupfer and the Richart confinement factors. Figure 2.28
shows the stresses confining the dashed line direction. The confining stresses fcmin can be

Figure 2.28: Confinement Stresses

decomposed into a fcmax and fcadd as shown in Figure 2.29. Notice that the stresses fcmax
and fcmin are negative; therefore, fcmin is a larger compressive stress than fcmax . The stress
fcmax represents a uniform confining stress while fcadd represents an additional confining
stress in one of the directions.
The uniform component of the confinement effect is captured by the βlRichart , while the
additional component of the stress is captured by βlKupf er . The decomposed confinement
60

Figure 2.29: Confinement Stresses Decomposition

0
stresses are normalized by fc , leading to two new variables:

fcadd
pa = (2.165)
fc0
fcmax
pu = (2.166)
fc0

The first term is called the additional normalized pressure term, while the second is the uni-
form normalized pressure, and they are shown on Figure 2.30. The confinement parameters
βlKupf er and βlRichart are calculated using the normalized pressures as:

βlKupf er = 4.1pu (2.167)

βlRichart = 0.92pa − 0.76p2a (2.168)

If fcmax > 0 then pa and pu become:

fcmin
pa = (2.169)
fc0
pu = 0 (2.170)

where the effect of fcmax > 0 is captured by the compression softening factor βd .
61

Figure 2.30: Uniform(pu ) and additional normalized(pa ) confinement pressure

Figure 2.31: Normalized confinement pressures when fcmax > 0


62

2.9 Summary and Conclusions

The DSFM compatibility, constitutive, and equilibrium equations are an extension of the
Modified Compression Field Theory (MCFT), which was developed using data from panel
tests conducted by Vecchio and Collins (1982) at the University of Toronto. The DSFM-
TG incorporates additional features with respect to the DSFM such as the dowel effect
DSFM-TG , concrete confinement, Poisson effect (lateral expansion of concrete), and out-
of-plane steel. Information provided in the Vector2 user’s manual (2002) was used to incor-
porate these effects.
The main conclusions that were obtained from this Chapter are

• The concrete constitutive, Poisson effect and confinement equations are the results
of the independent application of experimentally derived 1D expressions in the three
principal concrete stress directions.

• The crack opening, wcr , and the steel strains at the crack, εscr , are not accounted
when calculating the average RC strains.

• The stresses in the average region are not average.

• The perfect steel concrete bonding is not compatible with the DSFM governing equa-
tions when the concrete is cracked.

• The variables ∆ε1cr and wcr are not directly related.

• The interaction between the dowel an longitudinal steel stresses at the crack is not
accounted.

The 1D constitutive and multiaxial models associated with the concrete are used in the
three principal concrete directions independently. This ignores the relationship between
them, particularly between ν1 , ν2 , and νZ , which are known to be linked to each other.
Again, only a full 3D model for the concrete could capture effects of this type. However,
DSFM-TG is used within the context of a 2D RC material model, where in general the
stresses are going to be dominated by the compression in one or two directions at the most.
The localized crack displacements are the crack opening or width wcr , and the slip along
the crack δslip . The DSFM treats these localized displacements by averaging them over the
63

average RC region. The idea behind averaging them is to create an average RC strain that
can be later used to calculate average stresses from the constitutive models. The average
region is defined as the region of the RC concrete where the strain and the stresses are
average. Using also the DSFM assumption that there is perfect bonding between concrete
and steel, it can be concluded that concrete and steel strain are equal and average in the
average region.
The concept of smearing or averaging localized displacements over the average region
is conducted in a simplified or partial manner. The average concrete strains {εc } or the
concrete strains in the average region are calculated as the average RC strain {ε} minus the
average slip strains {εslip }, which comes from the δslip at the crack. The crack opening wcr
is not included in the calculation of {εc }, which leads to an overestimation of the concrete
strains in the average region. This problem is not that significant for the concrete stress
fc1 calculation in the average region since the tensile concrete models were developed using
the average total strain ε1 and not average concrete strain εc1 . Therefore, using a εc 1 = ε1
is still a good approximation to calculate the concrete stress fc1 , even though it presents
some incompatibility problems.
The situation of the steel is similar to that of the concrete. The DSFM compatibility
equation used to calculate average steel strains {εs } states that they are equal to the average
RC strains ({εs } = {ε}). This equation contradicts the perfect bonding between concrete
and steel, since the {εs } =
6 {εc } in the average RC region. Additionally the steel strains
at the crack εscri are greater than those in the average region. In order to maintain strain
compatibility the steel strains in the average region {εs } need to be lower than {ε} which
contradicts the DSFM equation {εs } = {ε}. The effect of the slip δslip is also not included
in the steel strains in the average region. The impact of not properly calculating the steel
strains in the average region leads to an overestimation of the steel stresses in the average
region, since no unloading occurs due to cracking.
An additional shortcoming of the DSFM is the definition of average and cracking regions.
The average region intends to be a region of average strain and stresses located between
cracks. This is not possible. The strains in the average region (defined as between cracks)
are not average since they are are affected by the cracking region localized displacements.
64

Defining the concept of average stresses could lead to the idea that stresses are averaged for
a given average strain state {ε}. In reality, the stresses in the average region and cracking
region are equal due to equilibrium, and they are not the results of an averaging.
Another important aspect that could be improved in the DSFM assumption is the per-
fect bonding between steel and concrete. In reality, this occurs only when the concrete is
uncracked. When the concrete is cracked the steel strain at the crack increases, which is
only possible if there is a relative displacement between the steel bars and the concrete sur-
rounding them. The increase in the steel strain at the crack does not result in a reduction
of the average steel strain, which is similar to the problem the way the crack opening or
width wcr is handle for the concrete. Additionally if there is δslip at the crack the steel
strains in the average region are not modified to account for the offset introduced by the
δslip , like the concrete strains are. Therefore, under the presence of slip δslip , the steel and
concrete strains in the average region are not the same, which is incompatible with the
perfect bonding assumption. Not including the effect of the δslip in the calculation of the
steel strains in the average region is another simplification of the model.
The DSFM model does not directly relates the ∆ε1cr and wcr , and neither of them
have an impact in the calculation of the average concrete and steel strains. The practical
implication of not including ∆ε1cr as part of the internal RC strain calculation is that
yielding of the steel at the crack, which occurs due to large ∆ε1cr , does not result in
softening of the system, since the deformation is calculated based on the strains in the
average region. Experimental results suggest that a large portion of an element occurs
due to localized elongation of concrete at the crack. This behavior was observed in the
experimental tests performed on Panels PV16 and PV17.
In order to link ∆ε1cr and wcr and include them into the calculation of strains in the
average zone, it is necessary to develop an expression for an equivalent unbonded length
near the crack. The unbonded length would define an area for the cracking region, which
could be used in the integratation the strains in this region and not only in the average
region as it is now in the DSFM.
Slip at the crack δslip occurs only when the steel crossing the crack is not symmetric
along the 1sc axis (line normal to the crack). In a way, this deviates from the common idea
65

of slip δslip due to external shearing stress parallel to the orientation of the crack. Since
the crack orientation rotates with the principal concrete stress direction, the shear stresses
do not generate slip δslip by themselves. The asymmetry of the reinforcement crossing the
crack is the real responsible for slip δslip .
The confinement effect was included to capture the additional ductility observed in
confined concrete. The confinement is greatly enhanced when the concrete is confined in
two directions rather than one. The only way to obtain confinement in two directions is
by using out-of-plane reinforcement. The out-of-plane reinforcement gets activated due to
outward expansion of concrete. This out-of-plane expansion of concrete is due to the Poisson
effect caused by in-plane concrete compressive strains.
The model incorporates the confinement effect by increasing the peak strength and
strain, but it does not directly affect the post-peak response (i.e. make it more ductile).
Actually, the PopovicHS model increases the decay after peak which is counterproductive
with the experimental behavior observed. In order to achieve this ductile behavior it is
necessary to incorporate new models for the post peak response that decay slower due to
the confinement effect.
The dowel effect is only included at the crack since this is where the localized bending of
the bars occurs. No interaction between the longitudinal and dowel stresses is included. This
may overestimate the stresses that the bars can handle at the crack. When the longitudinal
stress of the bar exceeds the yielding stress, the bending stiffness is greatly reduced. This
effect could be captured by developing interaction curves between bending and axial stresses.
66

Chapter 3

DSFM-TG RC MATERIAL MODEL FORMULATION FOR USE IN


FEM

3.1 Introduction

Nonlinear displacement-based FEM software typically use an incrementally advancing


solution algorithm, in which load (or displacement) demands on a structure are modified
using small increments to achieve a desired load history. A solution is found at each load
stage to satisfy the incremented demand. Within a load increment, the structural problem
is linearized using the structure tangent stiffness, which is the derivative of the internal
structure forces with respect to the associated displacements. This structure stiffness is
calculated by assembling the tangent stiffnesses from each individual finite element that
composes the structure. The element tangent stiffness is computed by integrating the tan-
gent to the material stiffness over the volume of the element. Thus, to expedite the solution
process, it is desirable that the material constitutive models provide not only the stress

state, σ, for a particular strain state, ε, but also the tangent material stiffness, D = .

The DSFM-TG 1D constitutive models do not account for cyclical loading; therefore,
no history variables are included in the material model formulation for use in FEM. A
natural extension of DSFM-TG would be to modify the 1D consitutitve models to account
for cyclic loading, which would also involve incorporating additional internal RC variables.
The definition of the internal RC problem and a solution strategy are described in Section
3.2.
For any nonlinear material it is typically necessary to solve a multivariable nonlinear
system of equations to determine the stress state, given the strain state. The solution of the
material state includes the solution for the complete set of material variables that define
the state.
Section 3.3 presents the derivation of the expressions for the calculation of the RC stress
67

vector {σ}. Section 3.4 presents the derivation of expressions for the calculation of the
RC stiffness matrix [D]. Alternative non-consistent material stiffness matrix expressions
such as secant fixed, and tangent fixed (concrete only) are also presented in this section for
comparison purposes.

3.2 Internal RC State

3.2.1 Governing Equations

The internal RC state for a given strain state, ε, is governed by the set of compatibility, con-
stitutive, equilibrium, and multiaxial effect equations defined in Chapter 2. The governing
equations are listed below:

Compatibility:
{ε} = [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ) (3.1)

Equation 3.1 is the same as Equation 2.20 presented in Section 2.4. Equation 3.1 links
the average RC strains with the average concrete strain and slip strain.

Poisson effect:
{ε3P
c
D,Load
} = [P ] · {ε3P
c
D
} (3.2)

Equation 3.2 is the same as Equation 3.15 presented in Section 2.7. Equation 3.2 links the
load concrete average strains with the total concrete average strains.

Modified strain at peak:

εp1 = β1ε εco (3.3)

εp2 = β2ε εco (3.4)

εpz = βzε εco (3.5)

Equations 3.3, 3.4, and 3.5 are the same as Equation 2.115 (presented in Section 2.7) applied
in the three concrete principal direction 1sc , 2sc , and z. Equations 3.3, 3.4, and 3.5 link the
68

strain εco with the modified peak strains εp1 , εp2 and εpz .

Modified peak stress:

0
fp1 = β1σ fc (3.6)
0
fp2 = β2σ fc (3.7)
0
fpz = βzσ fc (3.8)

Equations 3.6, 3.7, and 3.8 are the same as Equation 2.114 (presented in Section 2.7) applied
in the three concrete principal direction 1sc , 2sc , and z. Equations 3.6, 3.7, and 3.8 link the
0
stress fc with the modified peak stresses fp1 , fp2 and fpz .

Normal crack equilibrium:

np
X np
X
2
fc1 − ρi (fscri − fsi ) cos (φi ) − ρi fdcri cos(φi ) sin(φi ) = 0 (3.9)
i=1 i=1

Equation 3.9 is the same as Equation 2.31 presented in Section 2.4. Equation 3.9 defines
the equilibrium normal to crack.

Tangent crack equilibrium:

np
X np
X
vc − ρi (fscri − fsi ) sin(φi ) cos(φi ) + ρi fdcri cos2 (φi ) = 0 (3.10)
i=1 i=1

Equation 3.10 is the same as Equation 2.33 presented in Section 2.4. Equation 3.10 defines
the equilibrium tangent to crack.

Out-of-plane equilibrium:
ρz fsz + fcz = 0 (3.11)

Equation 3.11 is the same as Equation 2.18 presented in Section 2.3. Equation 3.11 defines
the equilibrium in the out-of-plane direction.
69

3.2.2 Nonlinear System of Equations

Equations 3.1 to 3.11 represent a total of 15 scalar Equations (equations 3.1 and 3.2 have
three scalar equations each). The only known variables in the above list of equations are
0
material response parameters εco , fc , and ρi the strain defining the current state, {ε}. The
remaining variables shown in equations 3.1 to 3.11 are identified as Group 1 variables an
are listed in Table 3.1. The variables on which the Group 1 variables depend directly are
identified in Table 3.1 as Group 2 variables. Similarly, Group 3 and Group 4 variables are
the variables on which the Group 2 and Group 3 variables directly depend.
Note in Table 3.1 the concrete stresses fc1 and fcz are shown to directly depend on
εload
cj , εpj , fpj (j=1,z), and θ12c . This covers the generic compression-tension concrete stress

case (compression depends on εload load


cj , εpj , and fpj , and tension on εcj , and θ12c ). A similar

approach was taken for the Poisson’s ratios νj (j = 1, 2, z) which is shown as directly
dependant on εload load and ε , and tension model
cj , εpj (compression model depends on εcj pj

only on εload
cj ).

The selected basic internal variables or the variables that completely define the internal
RC state are shown in bold in Table 3.1, and they are: εc1 , εc2 , εcz , εload load load
c1 , εc2 , εcz , εp1 ,

εp2 , εpz , fp1 , fp2 , fpz , θ12c , δslip , and ∆ε1cr . They represent a total of 15 scalar unknowns,
which matches the number of scalar governing equations. Thus, the general or cracked
internal RC problem becomes a 15 × 15 Non-linear system of equations problem.
The crack equilibrium Equations 3.9 and 3.10 are only valid if the concrete status is
cracked. If the concrete status is uncracked, these two equations are not used and the
number of equations reduces to 13. The number of unknowns variables is also reduced to
13 since the values for δslip and ∆ε1cr are set to zero when the concrete is uncracked.
70

Table 3.1: Internal RC Variables


Group 1 Group 2 Group 3 Group 4
[Tstrain (−θ12c )] θ12c - -
{εslip }12c δslip - -
Scr θ12c -
{εc }12c εc1 , εc2 - -
3P D,Load
{εc } εload load load
c1 , εc2 , εcz - -
[P ] ν1 εload
c1 , ε p1 -
ν2 εload
c2 , εp2 -
νZ εload
cz , εpz -
{ε3P
c
D
} εc1 , εc2 , εcz -
εp1 - - -
εp2 - - -
εpz - - -
β1ε ε
βd1 εc1 , εc2 , εcz -
ε
βl1 fc2 εload
c2 , εp2 , fp2 , θ12c
fcz εload
cz , εpz , fpz , θ12c
εload
c1 - -
β2ε ε
βd2 εc2 , εc1 , εcz -
ε
βl2 fc1 εload
c1 , ε p1 , fp1 , θ12c
fcz εload
cz , ε pz , fpz , θ12c
εload
c2 - -
βzε ε
βdz εcz , εc1 , εc2 -
ε
βlz fc1 εload
c1 , ε p1 , fp1 , θ12c
fc2 εload
c2 , ε p2 , fp2 , θ12c
εload
cz - -
fp1 - - -
fp2 - - -
fpz - - -
β1σ σ
βd1 εc1 , εc2 , εcz -
σ
βl1 fc2 εload
c2 , ε p2 , fp2 , θ12c
fcz εload
cz , ε pz , fpz , θ12c
β2σ σ
βd2 εc2 , εc1 , εcz -
σ
βl2 fc1 εload
c1 , ε p1 , fp1 , θ12c
fcz εload
cz , ε pz , fpz , θ12c
βzσ σ
βdz εcz , εc1 , εc2 -
σ load
βlz fc1 εc1 , εp1 , fp1 , θ12c
fc2 εload
c2 , εp2 , fp2 , θ12c
fc1 εload
c1 , εp1 , fp1 , θ12c - -
fscri εscri εsi {ε}
∆ε1cr -
φi θ12c
fsi εsi {ε} -
fdcri δslip - -
φi θ12c -
vc δslip - -
wcr Scr θ12c
εc1 -
fsz εsz εcz -
fcz εload
cz , ε pz , f pz , θ 12c - -
71

Unknown vector {x} and residual vector {R}

The unknown vector {x} includes the 15 basic scalar variables and is defined as:

{x}T = { {ε3P
c
D T
} , {ε3P
c
D,Load T
} , {εp }T , {fp }T , {p}T } (3.12)

where
 

 ε 

 c1 

{ε3P
c
D
} = εc2 , (3.13)

 


ε  
cz
 

 ε 
load 
 c1 
 
{ε3P
c
D,Load
} = εload
c2 
, (3.14)

 
εload 
 
cz
 

 
εp1 
 

{εp } = εp2 , (3.15)

 

ε 
 
pz
 

 f  
 p1 
 
{fp } = fp2 , (3.16)

 

f 
 
pz
 

 θ 

 12c 
 
{p} = δslip , (3.17)

 

∆ε 
 
1cr

where 3P D stands for the 3 principal concrete stress directions. A residual vector, {R}, is
defined from the governing internal RC for use in solving the 15 basic scalar variables:

{R}T = { {R{ε} }T , {R{ε3P D,Load } }T , {R{εp } }T , {R{fp } }T , {Req }T }, (3.18)


c
72

where

{R{ε} } = {ε} − [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ), (3.19)

{R{ε3P D,Load } } = {ε3P


c
D,Load
} − [P ] · {ε3P
c
D
}, (3.20)
c

{R{εp } } = {εp } − {β ε }εco , (3.21)


0
{R{fp } } = {fp } − {β σ }fc , (3.22)

{Req } = { Rnormal , Rtangent , Rout−of −plane }T , (3.23)

where the vectors {β ε } and {β σ } defined as:


 

βε

 1
 
{β ε } = β2ε , (3.24)

 

β ε 
 
z
 

βσ 
 1
 
{β σ } = β2σ (3.25)

 

β σ 
 
z

and the scalar equilibrium equations Rnormal = 0, Rtangent = 0, and Rout−of −plane = 0
are

np
X np
X
2
Rnormal = fc1 − ρi (fscri − fsi ) cos (φi ) − ρi fdcri cos(φi ) sin(φi ) (3.26)
i=1 i=1
np
X np
X
Rtangent = vc − ρi (fscri − fsi ) sin(φi ) cos(φi ) + ρi fdcri cos2 (φi ) (3.27)
i=1 i=1
Rout−of −plane = ρz fsz + fcz (3.28)

A perfectly converged internal RC state is defined by a residual vector {R} equal to a 15x1
zero vector.
73

Jacobian matrix [J]

The Nonlinear system of equations is linearized within an infinitesimal differential by the


Jacobian matrix, [J]:

∂R({x}, {ε})
[J] = (3.29)
∂{x}

where {R} is a function of {x} and {ε} as defined in Equation 3.18. The Jacobian is a
15 × 15 matrix composed of the following 3x3 submatrices:
2 ∂{R{ε} } ∂{R{ε} } ∂{R{ε} } ∂{R{ε} } ∂{R{ε} } 3
, , , ,
6 ∂{ε 3P D }
c ∂{ε3P D,Load
} ∂{εp } ∂{fp } ∂{p} 7
6 c 7
6 ∂{R{ε3P D,Load } } ∂{R{ε3P D,Load } } ∂{R{ε3P D,Load } } ∂{R{ε3P D,Load } } ∂{R{ε3P D,Load } } 7
6 c
, c
, c
, c
, c 7
6 7
6 ∂{ε3P
c
D}
∂{ε3P
c
D,Load
} ∂{εp } ∂{fp } ∂{p} 7
6 ∂{R{εp } } ∂{R{εp } } ∂{R{εp } } ∂{R{εp } } ∂{R{εp } } 7
6 7
[J] = 6 , , , , 7
6 ∂{ε3P
c
D}
∂{ε3P D,Load
} ∂{εp } ∂{fp } ∂{p} 7
6 c 7
6 ∂{R{fp } } ∂{R{fp } } ∂{R{fp } } ∂{R{fp } } ∂{R{fp } } 7
6 , , , , 7
6 ∂{ε3P D}
∂{ε3P D,Load
} ∂{εp } ∂{fp } ∂{p} 7
6 c c 7
4 ∂{Req } ∂{Req } ∂{Req } ∂{Req } ∂{Req } 5
, , , ,
∂{ε3P
c
D}
∂{ε3P
c
D,Load
} ∂{εp } ∂{fp } ∂{p}
(3.30)

The 3 × 3 submatrices are found by applying a differential operator to the 3 × 1 vectors


{R{ε} }, {R{ε3P D,Load } }, {R{εp } }, {R{fp } }, and {Req }.
c

Differential d{R{ε} } Applying a differential operator to Equation 3.19

d{R{ε} } = d({ε} − [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c )) (3.31)

= d{ε} − d[Tstrain (−θ12c )] · ({εslip }12c + {εc }12c )

− [Tstrain (−θ12c )] · d({εslip }12c + {εc }12c ) (3.32)


0
= [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c )dθ12c

− [Tstrain (−θ12c )] · (d{εslip }12c + d{εc }12c ) (3.33)

where
0 d[Tstrain (−θ12c )]
Tstrain (−θ12c ) = (3.34)
d(θ12c )
74

The differential d{ε} is a 3 × 1 zero vector because {ε} is constant within the internal RC
problem. The derivatives dθ12c and d{εc }12c can be expressed as:

dθ12c = {e1 }T · d{p} (3.35)

d{εc }12c = [R2D ] · d{ε3P


c
D
} (3.36)

and
 
1 0 0
 
 
[R2D ] = 0 1 0 (3.37)
 
0 0 0
n oT
{e1 } = 1, 0, 0 (3.38)

The strain vector {εslip }12c can be written as:

δslip
{εslip }12c = {e3 } (3.39)
Scr

where

n oT
{e3 } = 0, 0, 1 (3.40)

Applying a differential operator to Equation 3.39 yields to


µ ¶ µ ¶
12c δslip 1 δslip
d{εslip } =d {e3 } = dδslip − 2 dScr {e3 } (3.41)
Scr Scr Scr

The differential dδslip can be written using the definition of vector {p} as:

dδslip = {e2 }T · d{p} (3.42)

where
n oT
{e2 } = 0, 1, 0 (3.43)
75

and

dScr
dScr = dθ12c (3.44)
dθ12c
dScr
= {e1 }T · d{p} (3.45)
dθ12c

It should be noted that Scr is a function of Smx , Smy , as well as θ12c . However since Smx
and Smy are considered to be material parameters that do not change during the analysis
the derivatives with respect to these parameters are zero. In general during the presentation
of the Jacobian [J] derivation only the arguments that are not constant (e.g. θ12c ) will be
shown and not the constant arguments (e.g. Smx and Smy ), that do not affect the differential
of the variable being studied (e.g. Scr ). Using equations 3.45 and 3.42 in 3.41:
µ ¶
1 δslip dScr
d{εslip }12c = {e2 }T · d{p} − 2 {e1 }T · d{p} {e3 } (3.46)
Scr Scr dθ12c
µ ¶
1 T δslip dScr T
= {e3 } · {e2 } − 2 {e3 } · {e1 } · d{p} (3.47)
Scr Scr dθ12c
= [H1 ] · d{p} (3.48)

where [H1 ] is
1 δslip dScr
[H1 ] = {e3 } · {e2 }T − 2 {e3 } · {e1 }T (3.49)
Scr Scr dθ12c

Notice that the differential d{p} was extracted from the parenthesis in Equation 3.46 by
using the following identity:

({a}T · {b}){c} = ({c} · {a}T ) · {b} (3.50)

where {a}, {b} and {c} are 3x1 vectors. Substituting equations 3.35, 3.36, and 3.48 in 4.38
76

yields:

0
d{R{ε} } = [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ) · {e1 }T · d{p}

− [Tstrain (−θ12c )] · ([H1 ] · d{p} + [R2D ] · d{ε3P


c
D
}) (3.51)

= −[Tstrain (−θ12c )] · [R2D ] · d{ε3P


c
D
}
³ 0 ´
+ [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ) · {e1 }T − [Tstrain (−θ12c )] · [H1 ] · d{p}

(3.52)

From Equation the following submatrices of the Jacobian, [J], can be found:

∂{R{ε} }
= −[Tstrain (−θ12c )] · [R2D ] (3.53)
∂{ε3P
c
D}

∂{R{ε} }
= [0] (3.54)
∂{ε3P
c
D,Load
}
∂{R{ε} }
= [0]
∂{εp }
∂{R{ε} }
= [0] (3.55)
∂{fp }
∂{R{ε} } 0
= [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ) · {e1 }T
∂{p}
− [Tstrain (−θ12c )] · [H1 ] (3.56)

where [0] is a 3x3 zero matrix.

Differential d{R{ε3P D,Load } } Applying a differential operator to Equation 4.39:


c

d{R{ε3P D,Load } } = d({ε3P


c
D,Load
} − [P ] · {ε3P
c
D
}) (3.57)
c

= d{ε3P
c
D,Load
} − d[P ] · {ε3P
c
D
} − [P ] · d{ε3P
c
D
} (3.58)

The Poisson matrix [P ] is a function of the Poisson modulus ν1 , ν2 , and νZ :

[P ] = [P (ν1 , ν2 , νZ )] (3.59)
77

Applying a differential operator to matrix [P ]

X ∂[P ]
d[P ] = dνj (3.60)
∂νj
j=1,2,z

Defining the Poisson modulus vector {ν} as:

{ν} = {ν1 , ν2 , νZ }T (3.61)

where the last Equation indicates that the {ν} is a function of {ε3P
c
D,Load
} and {εp }. The
differential dνj (j = 1, 2, z) can be written as a function of d{ν}

νj = {ej }T · {ν} (3.62)

dνj = {ej }T · d{ν} (3.63)

However, the Poisson’s ratios are a function of {ε3P


c
D,Load
} and {εp }. Thus, the differential
d{ν} is

∂{ν} ∂{ν}
d{ν} = · d{ε3P
c
D,Load
}+ · d{εp } (3.64)
∂{ε3P
c
D,Load
} ∂{εp }

Using Equations 3.63 and 3.64 in Equation 3.60:

X ∂[P ] ¡ ¢
d[P ] = {ej }T · d{ν} (3.65)
∂νj
j=1,2,z
à à !!
X ∂[P ] ∂{ν} ∂{ν}
= {ej }T · 3P D,Load
· d{ε3P
c
D,Load
}+ · d{εp } (3.66)
∂νj ∂{ε c } ∂{εp }
j=1,2,z
ÃÃ ! µ ¶ !
X ∂[P ] ∂{ν} ∂{ν}
T 3P D,Load T
= {ej } · · d{εc } + {ej } · · d{εp }
∂νj
j=1,2,z ∂{ε3P
c
D,Load
} ∂{εp }
(3.67)
78

Substituting Equation 3.67 into the second term of Equation 3.58 yields:
 (Ã !
X ∂[P ] ∂{ν}
d[P ] · {ε3P
c
D
} =  T
{ej } · 3P D,Load
· d{ε3P
c
D,Load
}
∂νj ∂{εc }
j=1,2,z
µ ¶ ¾¸
T ∂{ν}
+ {ej } · · d{εp } · {ε3Pc
D
} (3.68)
∂{εp }

The differentials d{ε3P


c
D,Load
} and d{εp } can be extracted from Equation 3.68.
 
X ∂[P ] ∂{ν}
d[P ] · {ε3P
c
D
} =  · {ε3P
c
D
} · {ej }T ·  · d{ε3P
c
D,Load
}
j=1,2,z
∂νj ∂{ε3P
c
D,Load
}
 
X ∂[P ] ∂{ν} 
+  · {ε3P
c
D
} · {ej }T · · d{εp } (3.69)
∂νj ∂{εp }
j=1,2,z

= [Q1 ] · d{ε3P
c
D,Load
} + [Q2 ] · d{εp }, (3.70)

where [Q1 ] and [Q2 ] are

X ∂[P ] ∂{ν}
[Q1 ] = · {ε3P
c
D
} · {ej }T · (3.71)
j=1,2,z
∂νj ∂{εc D,Load }
3P

X ∂[P ] ∂{ν}
[Q2 ] = · {ε3P
c
D
} · {ej }T · (3.72)
∂νj ∂{εp }
j=1,2,z

Substituting Equation 3.70 into Equation 3.58:

d{R{ε3P D,Load } } = d{ε3P


c
D,Load
} − ([Q1 ] · d{ε3P
c
D,Load
}
c

+ [Q2 ] · d{εp }) − [P ] · d{ε3P


c
D
} (3.73)

= ([I] − [Q1 ]) · d{ε3P


c
D,Load
} − [Q2 ] · d{εp } − [P ] · d{ε3P
c
D
} (3.74)
79

where [I] is a 3 × 3 identity matrix. From Equation 3.74 the following submatrices of the
Jacobian, [J], are found:

∂{R{ε3P D,Load } }
c
= −[P ] (3.75)
∂{ε3P
c
D}

∂{R{ε3P D,Load } }
c
= [I] − [Q1 ] (3.76)
∂{ε3P
c
D,Load
}
∂{R{ε3P D,Load } }
c
= −[Q2 ] (3.77)
∂{εp }
∂{R{ε3P D,Load } }
c
= [0] (3.78)
∂{fp }
∂{R{ε3P D,Load } }
c
= [0] (3.79)
∂{p}

Differential d{R{εp } } Applying a differential operator to Equation 4.40:

d{R{εp } } = d({εp } − {β ε }εco ) (3.80)

= d{εp } − d{β ε }εco (3.81)

Vector {β ε } is a function of {βdε },{βlε }, and {ε3P


c
D,Load
}; thus

{β ε } = {{β ε }( {βdε }, {βlε } {ε3P


c
D,Load
})} (3.82)

where

{βdε } = {βd1
ε ε
, βd2 ε T
, βdz } (3.83)

{βlε } = {βl1
ε ε
, βl2 ε T
, βlz } (3.84)

Applying a differential operator to {β ε }:

∂{β ε } ∂{β ε } ∂{β ε }


d{β ε } = · d{βd
ε
} + · d{βl
ε
} + · d{ε3P
c
D,Load
} (3.85)
∂{βdε } ∂{βlε } ∂{ε3P
c
D,Load
}
= [S1 ] · d{βdε } + [S2 ] · d{βlε } + [S3 ] · d{ε3P
c
D,Load
} (3.86)
80

Where [S1 ], [S2 ] and [S3 ] are


 
∂β1ε
 ∂β ε 0 0 
 d1 
∂{β ε }  ∂β2ε 
[S1 ] = =
 0 ε 0   (3.87)
∂{βdε }  ∂βd2 
 ∂βzε 
0 0 ε
∂βdz
 ε 
∂β1
 ∂β ε 0 0 
 l1 
∂{β ε }  ∂β2ε 
[S2 ] = =
 0 ε 0   (3.88)
∂{βlε }  ∂βl2 
 ε
∂βz 
0 0 ε
∂βlz
 
∂β1ε
 ∂εload 0 0 
 c1 
ε
∂{β }  ∂β2ε 
[S3 ] = =
 0 0 
 (3.89)
3P D,Load
∂{εc }  ∂εload
c2 
 ∂βzε 
0 0
∂εload
cz

The vectors {βdε } and {βlε } are a function of {ε3P


c
D } and {σ 3P D }, respectively:
c

{βdε } = {βdε }({ε3P


c
D
}) (3.90)

{βlε } = {βlε }({σc3P D }) (3.91)

where {σc3P D } is defined as:

{σc3P D } = {fc1 , fc2 , fcz }T (3.92)

Applying a differential operator to {βdε } and {βlε }:

∂{βdε }
d{βdε } = · d{ε3Pc
D
} (3.93)
∂{ε3Pc
D}

∂{βlε }
d{βlε } = · d{σc3P D } (3.94)
∂{σc3P D }
d{βdε } = [S4 ] · d{ε3P
c
D
} (3.95)

d{βlε } = [S5 ] · d{σc3P D } (3.96)


81

where [S4 ] and [S5 ] are defined as:


 ε ε 
∂βd1 ∂βd1
 ∂εc1 0 
 ∂β ε ∂εc2 
∂{βdε }  d2 ε
∂βd1 
[S4 ] = 3P D
=  0  (3.97)
∂{εc }  ∂εc1 ∂εc2 
 ∂β ε ε 
∂βdz
dz
0
∂εc1 ∂εcz
 ε ε 
∂βl1 ∂βl1
0
 ∂fc2 ∂fcz 
ε
∂{βl }  ε ε 
 ∂βl2 ∂βl2 
[S5 ] = =  0  (3.98)
∂{σc3P D }  ∂fc1 ∂fcz 
 ∂β ε ε
∂βlz 
lz
0
∂fc1 ∂fc2

The vector {σc3P D } is a function of {ε3P


c
D,Load
}, {εp }, {fp }, θ12c :

{σc3P D } = {σ 3P D ({ε3P
c
D,Load
}, {εp }, {fp }, θ12c )} (3.99)

Thus, applying a differential operator to {σc3P D } results in:

∂{σc3P D } ∂{σc3P D } ∂{σc3P D }


d{σc3P D } = · d{ε3P
c
D,Load
}+ · d{εp } + · d{fp }
∂{ε3P
c
D,Load
} ∂{εp } ∂{fp }
3P
∂{σc } D
+ · dθ12c (3.100)
∂θ12c
ε f
= [Dc3P D,Load ] · d{ε3P
c
D,Load
} + [Dc p ] · d{εp } + [Dc p ] · d{fp }

+ {bθ12c } · dθ12c (3.101)


ε f
= [Dc3P D,Load ] · d{ε3P
c
D,Load
} + [Dc p ] · d{εp } + [Dc p ] · d{fp }

+ {bθ12c }{e1 }T · d{p} (3.102)


82

ε f
where [Dc3P D,Load ], [Dc p ], [Dc p ], and {bθ12c } are
 
∂fc1
 ∂εload 0 0 
 c1 
∂{σc3P D }  ∂fc2 
[Dc3P D,Load ] = =
 0 0 
 (3.103)
3P D,Load
∂{εc }  ∂εload
c2 
 ∂fcz 
0 0
∂εload
cz
 ∂f 
c1
0 0
 ∂εp1 
∂{σc3P D } 
 ∂fc2 
ε
[Dc p ] = = 0 0   (3.104)
∂{εp }  ∂εp2 
 ∂fcz 
0 0
∂εpz
 ∂f 
c1
0 0
 ∂fp1 
∂{σc3P D }  ∂fc2 
f
[Dc p ] = = 0 0   (3.105)
∂{fp }  ∂f p2 
 ∂fcz 
0 0
∂fpz
 

 ∂fc1 

 
3P D

 ∂θ12c 

∂{σc } ∂f c2
{bθ12c } = = (3.106)
∂θ12c 
 ∂θ12c 

 


 0  

Substituting Equations 3.95, 3.96, and 3.102 into Equation 3.86:

¡ ¢ ¡ ¢
d{β ε } = [S1 ] · [S4 ] · d{ε3P
c
D
} + [S2 ] · [S5 ] · d{σc3P D } + [S3 ] · d{ε3P
c
D,Load
}(3.107)

= [S1 ] · [S4 ] · d{ε3P


c
D
} + ([S2 ] · [S5 ] · [Dc3P D,Load ] + [S3 ]) · d{ε3P
c
D,Load
}
ε f
+ [S2 ] · [S5 ] · [Dc p ] · d{εp } + [S2 ] · [S5 ] · [Dc p ] · d{fp }

+ [S2 ] · [S5 ] · {bθ12c } · {e1 }T · d{p} (3.108)


83

And substituting Equation 3.108 into Equation 3.81 results in:

d{R{εp } } = d{εp } − ([S1 ] · [S4 ]εco ) · d{ε3P


c
D
}

− (([S2 ] · [S5 ] · [Dc3P D,Load ] + [S3 ])εco ) · d{ε3P


c
D,Load
}
ε f
− ([S2 ] · [S5 ] · [Dc p ]εco ) · d{εp } − ([S2 ] · [S5 ] · [Dc p ]εco ) · d{fp }

− ([S2 ] · [S5 ] · {bθ12c } · {e1 }T εco ) · d{p} (3.109)

which simplifies to

d{R{εp } } = −([S1 ] · [S4 ]εco ) · d{ε3P


c
D
}

− (([S2 ] · [S5 ] · [Dc3P D,Load ] + [S3 ])εco ) · d{ε3P


c
D,Load
}
ε f
+ ([I] − [S2 ] · [S5 ] · [Dc p ]εco ) · d{εp } − ([S2 ] · [S5 ] · [Dc p ]εco ) · d{fp }

− ([S2 ] · [S5 ] · {bθ12c } · {e1 }T εco ) · d{p} (3.110)

From Equation 3.110 the following submatrices can be found:

∂{R{εp } }
= −[S1 ] · [S4 ]εco (3.111)
∂{ε3P
c
D}

∂{R{εp } }
= −([S2 ] · [S5 ] · [Dc3P D,Load ] + [S3 ])εco (3.112)
∂{ε3P
c
D,Load
}
∂{R{εp } } ε
= [I] − [S2 ] · [S5 ] · [Dc p ]εco (3.113)
∂{εp }
∂{R{εp } } f
= −[S2 ] · [S5 ] · [Dc p ]εco (3.114)
∂{fp }
∂{R{εp } }
= −[S2 ] · [S5 ] · {bθ12c } · {e1 }T εco (3.115)
∂{p}

Differential d{R{fp } } Applying a differential operator to Equation 4.41:

0
d{R{fp } } = d({fp } − {β σ }fc ) (3.116)
0
= d{fp } − d{β σ }fc (3.117)
84

Vector {β σ } is a function of {βdσ } and {βlσ }:

{β σ } = {β σ ( {βdσ }, {βlσ })} (3.118)

Applying a differential operator to the last equation:

∂{β σ } ∂{β σ }
d{β σ } = · d{β σ
d } + · d{βlσ } (3.119)
∂{βdσ } ∂{βlσ }
= [S6 ] · d{βdσ } + [S7 ] · d{βlσ } (3.120)

where [S6 ] and [S7 ] are


 
∂β1σ
 ∂β σ 0 0 
 d1 
∂{β σ }  ∂β2σ 
[S6 ] = =
 0 σ 0   (3.121)
∂{βdσ }  ∂βd2 
 σ
∂βz 
0 0 σ
∂βdz
 σ 
∂β1
 ∂β σ 0 0 
 l1 
∂{β σ }  ∂β2σ 
[S7 ] = =
 0 σ 0   (3.122)
∂{βlσ }  ∂βl2 
 ∂βzσ 
0 0 σ
∂βlz

Vectors {βdσ } and {βlσ } are a function of {ε3P


c
D } and {σ 3P D }, respectively:
c

{βdσ } = {βdσ ({ε3P


c
D
})} (3.123)

{βlσ } = {βlσ ({σc3P D })} (3.124)

(3.125)
85

Applying a differential operator {βdσ } and {βlσ }:

{βdσ }
d{βdσ } = · d{ε3Pc
D
} (3.126)
∂{ε3Pc
D}

= [S8 ] · d{ε3P
c
D
} (3.127)
{βl }σ
d{βlσ } = · d{σc3P D } (3.128)
∂{σc3P D }
= [S9 ] · d{σc3P D } (3.129)

where [S8 ] and [S9 ] are


σ σ 
∂βd1 ∂βd1
 ∂εc1 0 
 ∂β σ ∂εc2 
{βdσ }  d2 σ
∂βd1 
[S8 ] = 3P D
=  0  (3.130)
∂{εc }  ∂εc1 ∂εc2 
 ∂β σ σ 
∂βdz
dz
0
∂εc1 ∂εcz
 σ σ 
∂βl1 ∂βl1
0
 ∂fc2 ∂fcz 
σ
{βl }  σ σ 
 ∂βl2 ∂βl2 
[S9 ] = =  0  (3.131)
∂{σc3P D }  ∂fc1 ∂fcz 
 ∂β σ σ
∂βlz 
lz
0
∂fc1 ∂fc2

Substituting Equations 3.127, 3.129, and the definition of d{σc3P D } from Equation 3.102
into Equation 3.120:

d{β σ } = [S6 ] · [S8 ] · d{ε3P


c
D
} + [S7 ] · [S9 ] · d{σc3P D } (3.132)

= [S6 ] · [S8 ] · d{ε3P


c
D
}
ε
+ [S7 ] · [S9 ] · [Dc3P D,Load ] · d{ε3P
c
D,Load
} + [S7 ] · [S9 ] · [Dc p ] · d{εp }
f
+ [S7 ] · [S9 ] · [Dc p ] · d{fp } + [S7 ] · [S9 ] · {bθ12c } · {e1 }T · d{p} (3.133)
86

Substituting Equation 3.133 into Equation 3.117:

d{R{fp } } = d{fp }
¡
− [S6 ] · [S8 ] · d{ε3P
c
D
}
ε
+ [S7 ] · [S9 ] · [Dc3P D,Load ] · d{ε3P
c
D,Load
} + [S7 ] · [S9 ] · [Dc p ] · d{εp }
´ 0
f
+ [S7 ] · [S9 ] · [Dc p ] · d{fp } + [S7 ] · [S9 ] · {bθ12c } · {e1 }T · d{p} fc (3.134)
0
= −fc [S6 ] · [S8 ] · d{ε3P
c
D
}
0
− fc [S7 ] · [S9 ] · [Dc3P D,Load ] · d{ε3P
c
D,Load
}
0 ε
− fc [S7 ] · [S9 ] · [Dc p ] · d{εp }
0 f
+ ([I] − fc [S7 ] · [S9 ] · [Dc p ]) · d{fp }
0
− fc [S7 ] · [S9 ] · {bθ12c } · {e1 }T · d{p}

(3.135)

From the last equations the following submatrices can be found:

∂{R{fp } } 0
= (−[S6 ] · [S8 ]) fc (3.136)
∂{ε3P
c
D}

∂{R{fp } } ³ ´ 0
= −[S7 ] · [S9 ] · [Dc3P D,Load ] fc (3.137)
∂{ε3P
c
D,Load
}
∂{R{fp } } ¡ ε ¢ 0
= −[S7 ] · [S9 ] · [Dc p ] fc (3.138)
∂{εp }
∂{R{fp } } ³ ´ 0
f
= [I] − [S7 ] · [S9 ] · [Dc p ] fc (3.139)
∂{fp }
∂{R{fp } } ¡ ¢ 0
= −[S7 ] · [S9 ] · {bθ12c } · {e1 }T fc (3.140)
∂{p}

Differential d{Req } This differential is made of three scalar equilibrium equations:


Rnormal , Rtangent , and Rout−of −plane . Applying a differential to the normal component
87

of the {Req }:
( np np
)
X X
dRnormal = d fc1 − ρi (fscri − fsi ) cos2 (φi ) − ρi fdcri cos(φi ) sin(φi ) (3.141)
i=1 i=1
np
X ³ ¡ ¢ ´
= dfc1 − ρi d (fscri − fsi ) cos2 (φi ) + ρi d (fdcri cos(φi ) sin(φi ))
i=1
(3.142)
np
X ³
= dfc1 − ρi (dfscri − dfsi ) cos2 (φi ) + (fscri − fsi )d cos2 (φi )
i=1
´
+ dfdcri cos(φi ) sin(φi ) + fdcri d(cos(φi ) sin(φi )) (3.143)

The differential dfc1 is

dfc1 = d({e1 }T · {σc3P D }) (3.144)

= {e1 }T · d{σc3P D } (3.145)


ε
= {e1 }T · ([Dc3P D,Load ] · d{ε3P
c
D,Load
} + [Dc p ] · d{εp }
f
+ [Dc p ] · d{fp } + {bθ12c } · {e1 }T · d{p}) (3.146)

The variable fscri is a function of εscri ; thus, the differential dfscri is determined as follows:

fscri = fscri (εscri ) (3.147)

dfscri = dfscri (εscri ) (3.148)


dfscri
= dεscri (3.149)
dεscri
= Escri dεscri (3.150)

where

εscri = εsi + ∆ε1cr cos2 (φi ) (3.151)


88

Thus the differential dεscri is defined:

∂εscri ∂εscri ∂εscri


dεscri = dεsi + d∆ε1cr + dφi (3.152)
∂εsi ∂∆ε1cr ∂φi

where dεsi is

dεsi = 0 (3.153)

since εsi is computed directly from {ε} which means that is constant within the nonlinear
∂εscri ∂εscri
system of equation context. And d∆ε1cr , , and are
∂∆ε1cr ∂φi

d∆ε1cr = {e3 }T · d{p} (3.154)

dφi = dθ12c = {e1 }T · d{p} (3.155)


∂εscri
= cos2 (φi ) (3.156)
∂∆ε1cr
∂εscri
= −2∆ε1cr cos(φi ) sin(φi ) (3.157)
∂φi

The differential dφi is equal to dθ12c because φi = θ12c − αi , and αi is a constant angle that
defines the orientation of the in-plane steel layer ith . Substituting the Equations 3.153 to
3.157 into Equation 3.152:

dεscri = cos2 (φi ){e3 }T · d{p} + cos2 (φi ){e1 }T · d{p} (3.158)

= (cos2 (φi ){e3 }T − 2∆ε1cr cos(φi ) sin(φi ){e1 }T ) · d{p} (3.159)

= {c1i } · d{p} (3.160)

where

{c1i } = cos2 (φi ){e3 }T − 2∆ε1cr cos(φi ) sin(φi ){e1 }T (3.161)

Substituting Equation 3.161 into 3.150 yields

fscri = Escri {c1i } · d{p}. (3.162)


89

Definition of dRnormal from Equation 3.143 requires also definition of the differential dfdcri :

fdcri = fdcri (∆dcri ) (3.163)

Thus,

fdcri = fdcri (∆dcri ) (3.164)

dfdcri = dfdcri (∆dcri ) (3.165)


dfdcri
= d∆dcri (3.166)
d∆dcri
= Edcri d∆dcri . (3.167)

The differential d∆dcri is

∆dcri = δslip cos(φi ) (3.168)

d∆dcri = d(δslip cos(φi )) (3.169)

= dδslip cos(φi ) + δslip d cos(φi ) (3.170)

= cos(φi ){e2 }T · d{p} − δslip sin(φi )dφi (3.171)

= cos(φi ){e2 }T · d{p} − δslip sin(φi )dθ12c (3.172)

= cos(φi ){e2 }T · d{p} − δslip sin(φi ){e1 }T · d{p} (3.173)

= (cos(φi ){e2 }T − δslip sin(φi ){e1 }T ) · d{p} (3.174)

, = {c2i } · {p} (3.175)

where {c2i } is
{c2i } = cos(φi ){e2 }T − δslip sin(φi ){e1 }T (3.176)

Introducing equation 3.176into Equation 3.167:

fdcri = Edcri {c2i } · d{p} (3.177)

The differentials d cos2 (φi ) and d(sin(φi ) cos(φi )) are defined as


90

d cos2 (φi ) = −2 cos(φi ) sin(φi )dφi (3.178)

= −2 cos(φi ) sin(φi )dθ12c (3.179)

= −2 cos(φi ) sin(φi ){e1 }T · d{p} (3.180)

= {c3i } · d{p}, (3.181)

where {c3i } is

{c3i } = −2 cos(φi ) sin(φi ){e1 }T , (3.182)

and

d(sin(φi ) cos(φi )) = (cos2 (φi ) − sin2 (φi ))dφi (3.183)

= (cos2 (φi ) − sin2 (φi ))dθ12c (3.184)

= (cos2 (φi ) − sin2 (φi )){e1 }T · d{p} (3.185)

= {c4i } · d{p}, (3.186)

where {c4i } is

{c4i } = (cos2 (φi ) − sin2 (φi )){e1 }T (3.187)

Substituting the differential defined in Equations 3.162, 3.177, 3.181, and 3.186 into
3.143 results in:

np
X ¡
dRnormal = dfc1 − ρi (Escri {c1i }) cos2 (φi ) + (fscri − fsi ){c3i }
i=1
+ Edcri {c2i } cos(φi ) sin(φi ) + fdcri {c4i } ) · d{p} (3.188)

= dfc1 + {gnormal } · d{p} (3.189)

where {gnormal } is a 1 × 3 vector associated with the impact of the steel stress variations
91

on Rnormal . The steel stress variations are due to variations in θ12c , ∆1cr , and θ12c (terms
of {p}).

np
X ³
{gnormal } = − ρi Escri cos2 (φi ){c1i } + (fscri − fsi ){c3i }
i=1
´
+ Edcri cos(φi ) sin(φi ){c2i } + fdcri {c4i } (3.190)

Using Equation 3.146 in 3.189:

ε f
dRnormal = {e1 }T · ([Dc3P D,Load ] · d{ε3P
c
D,Load
} + [Dc p ] · d{εp } + [Dc p ] · d{fp }

+ ({bθ12c } · {e1 }T + {gnormal }) · d{p} (3.191)

The next scalar Equation in {Req } is Rtangent defined in Equation 3.27. Applying a differ-
ential operator to Equation 3.27:

à np np
!
X X
2
dRtangent = d vc − ρi (fscri − fsi ) sin(φi ) cos(φi ) + ρi fdcri cos (φi ) (3.192)
i=1 i=1
np
X ¡ ¢
= dvc − ρi d((fscri − fsi ) sin(φi ) cos(φi )) − d(fdcri cos2 (φi )) (3.193)
i=1
np
X
= dvc − ρi ((dfscri − dfsi ) sin(φi ) cos(φi ) (3.194)
i=1
¢
+ (fscri − fsi )d(sin(φi ) cos(φi )) − dfdcri cos2 (φi ) − fdcri d cos2 (φi ) (3.195)

Complete definition of dRtangent requires the definition of the differentials dvc , dfscri , dfsi ,
dfdcri , d(sin(φi ) cos(φi ), and d cos2 . The differential dfsi was already determined to be zero,
and the remaining differential, except dvc , are already defined in Equations 3.162, 3.177,
3.181, and 3.186 into 3.195. Substituting these Equations in Equation 3.195:

np
X
dRtangent = dvc − ρi (Escri sin(φi ) cos(φi ){c1i } (3.196)
i=1
¢
+ (fscri − fsi ){c4i } − Edcri cos2 (φi ){c2i } − fdcri {c3i } · d{p} (3.197)

= dvc + {gtangent } · d{p} (3.198)


92

where {gtangent } is a 1 × 3 vector associated with the impact of the steel stress variations
on Rtangent . The steel stress variations are due to variations in θ12c , ∆1cr , and θ12c (terms
of {p}).

np
X ³
{gtangent } = − ρi Escri sin(φi ) cos(φi ){c1i } + (fscri − fsi ){c4i }
i=1
´
− Edcri cos2 (φi ){c2i } − fdcri {c3i } (3.199)

In order to complete an expression for dRtangent and expression for dvc is found using the
definition of vc :
vc = vc (δslip , wcr ) (3.200)

Thus,

dvc = d(vc (δslip , wcr )) (3.201)


∂vc ∂vc
= dδslip + dwcr (3.202)
∂δslip ∂wcr
∂vc ∂vc
= {e2 }T · d{p} + dwcr (3.203)
∂δslip ∂wcr

Where the differential dwcr is computed using the definition of wcr

wcr = wcr (εc1 , Scr ) (3.204)

Thus,

dwcr = d(wcr (εc1 , Scr )) (3.205)


∂wcr ∂wcr
= dεc1 + dScr (3.206)
∂εc1 ∂Scr
∂wcr ∂wcr dScr
= {e1 }T · d{ε3P c
D
}+ {e1 }T · d{p} (3.207)
∂εc1 ∂Scr dθ12c
93

Substituting Equation 3.207 into Equation 3.203:

∂vc
dvc = {e2 }T · d{p}
∂δslip
µ ¶
∂vc ∂wcr T 3P D ∂wcr dScr T
+ {e1 } · d{εc } + {e1 } · d{p} (3.208)
∂wcr ∂εc1 ∂Scr dθ12c
∂vc ∂wcr
= {e1 }T · d{ε3P
c
D
}
∂wcr ∂εc1
µ ¶
∂vc T ∂vc ∂wcr dScr T
+ {e2 } + {e1 } · d{p} (3.209)
∂δslip ∂wcr ∂Scr dθ12c
= {a1 } · d{ε3P
c
D
} + {a2 } · d{p} (3.210)

where {a1 } and {a2 } are 1 × 3 vectors defined as:

∂vc ∂wcr
{a1 } = {e1 }T (3.211)
∂wcr ∂εc1
∂vc ∂vc ∂wcr dScr
{a2 } = {e2 }T + {e1 }T (3.212)
∂δslip ∂wcr ∂Scr dθ12c

Substituting 3.210 into 3.198 provides a final definition for dRtangent :

dRtangent = {a1 } · d{ε3P


c
D
} + ({a2 } + {gtangent }) · d{p} (3.213)

The final definition in d{req } to be computed is dRout−of −plane :

dRout−of −plane = d(ρz fsz + fcz ) (3.214)

= ρz dfsz + dfcz (3.215)

The first differential dfsz is

dfsz
dfsz = dεsz (3.216)
dεsz
= Esz dεsz (3.217)
94

where given the definition εsz = εcz , the differential dεsz can be written as

dεsz = dεcz (3.218)

= {e3 }T · d{ε3P
c
D
} (3.219)

Thus, substituting Equation 3.219 into Equation 3.217.

dfsz = Esz {e3 }T · d{ε3P


c
D
} (3.220)

The variables fcz can be written as {e3 }T · {σc3P D }. Thus, the differential dfcz is

dfcz = d({e3 }T · {σc3P D }) (3.221)

= {e3 }T · d{σc3P D } (3.222)

Substituting Equation 3.102 into Equation 3.222:

ε f
dfcz = {e3 }T · ([Dc3P D,Load ] · d{ε3P
c
D,Load
} + [Dc p ] · d{εp } + [Dc p ] · d{fp }

+ ({bθ12c } · {e1 }T )) · d{p} (3.223)

Introducing Equations 3.223 and 3.220 in Equation 3.215:

dRout−of −plane = ρz Esz {e3 }T · d{ε3P


c
D
} + {e3 }T · d{σc3P D } (3.224)

= ρz Esz {e3 }T · d{ε3P


c
D
}
³
T 3P D,Load ε
+ {e3 } · [Dc ] · d{ε3P
c
D,Load
} + [Dc p ] · d{εp }
´
f
+ [Dc p ] · d{fp }({bθ12c } · {e1 }T ) · d{p} (3.225)

The differentials dRnormal , dRtangent , and dRout−of −plane can be combined to define d{Req }:

d{Req } = dRnormal {e1 } + dRtangent {e2 } + dRout−of −plane {e3 } (3.226)


95

The first term is found using Equation 3.191:

³
ε f
dRnormal {e1 } = {e1 }T · [Dc3P D,Load ] · d{ε3P
c
D,Load
} + [Dc p ] · d{εp } + [Dc p ] · d{fp }
¢
+ ({bθ12c } · {e1 }T + {gnormal }) · d{p} {e1 } (3.227)

Extracting the differentials d{ε3P


c
D,Load
}, d{εp }, d{fp }, and d{p}:

ε
dRnormal {e1 } = {e1 } · {e1 }T · [Dc3P D,Load ] · d{ε3P
c
D,Load
} + {e1 } · {e1 }T · [Dc p ] · d{εp }
f
+ {e1 } · {e1 }T · [Dc p ] · d{fp }

+ ({e1 } · {e1 }T · {bθ12c } · {e1 }T + {e1 }{gnormal }) · d{p} (3.228)


ε
= [E1 ] · [Dc3P D,Load ] · d{ε3P
c
D,Load
} + [E1 ] · [Dc p ] · d{εp }
f
+ [E1 ] · [Dc p ] · d{fp }

+ ([E1 ] · {bθ12c } · {e1 }T + {e1 } · {gnormal }) · d{p} (3.229)

where [E1 ] is

[E1 ] = {e1 } · {e1 }T (3.230)

Similarly, the second term is found using Equation 3.213:

¡ ¢
dRtangent {e2 } = {a1 } · d{ε3P
c
D
} + ({a2 } + {gtangent }) · d{p} {e2 } (3.231)

= {e2 } · {a1 } · d{ε3P


c
D
} + {e2 } · ({a2 } + {gtangent }) · d{p}

(3.232)
96

Finally, the third term is found using Equation 3.225:

¡
dRout−of −plane {e3 } = ρz Esz {e3 }T · d{ε3P
c
D
}
³
ε
+ {e3 }T · [Dc3P D,Load ] · d{ε3P c
D,Load
} + [Dc p ] · d{εp }
´ ´
f
+ [Dc p ] · d{fp } + ({bθ12c } · {e1 }T ) · d{p} {e3 } (3.233)

= ρz Esz {e3 } · {e3 }T · d{ε3P


c
D
}

+ {e3 } · {e3 }T · [Dc3P D,Load ] · d{ε3P


c
D,Load
}
ε
+ {e3 } · {e3 }T · [Dc p ] · d{εp }
f
+ {e3 } · {e3 }T · [Dc p ] · d{fp }

+ {e3 } · {e3 }T · ({bθ12c } · {e1 }T ) · d{p} (3.234)

= ρz Esz [E3 ] · d{ε3P


c
D
}
ε
+ [E3 ] · [Dc3P D,Load ] · d{ε3P
c
D,Load
} + [E3 ] · [Dc p ] · d{εp }
f
+ [E3 ] · [Dc p ] · d{fp } + [E3 ] · ({bθ12c } · {e1 }T ) · d{p}

(3.235)

where [E3 ] is

[E3 ] = {e3 } · {e3 }T (3.236)

Introducing the first, second, and third term from Equations 3.229, 3.232, and 3.235 into
97

Equation 3.226:

d{Req } = dRnormal {e1 } + d{Req }{e2 } + d{Req }{e3 } (3.237)

= (ρz Esz [E3 ] + {e2 }{a1 }) · d{ε3P


c
D
} (3.238)

+ ([E1 ] + [E3 ]) · [Dc3P D,Load ] · d{ε3P


c
D
}
ε
+ ([E1 ] + [E3 ]) · [Dc p ] · d{εp }
f
+ ([E1 ] + [E3 ]) · [Dc p ] · d{fp }
³
+ ([E1 ] + [E3 ]){bθ12c } · {e1 }T
´
+ {e1 } · {gnormal } + {e2 }({a2 } + {gtangent }) · d{p} (3.239)

(3.240)

From Equation 3.240 the following submatrices of the Jacobian, [J] can be found:

∂{Req }
= ρz Esz [E3 ] + {e2 }{a1 } (3.241)
∂{ε3P
c
D}

∂{Req }
= ([E1 ] + [E3 ]) · [Dc3P D,Load ] (3.242)
∂{ε3P
c
D,Load
}
∂{Req } ε
= ([E1 ] + [E3 ]) · [Dc p ] (3.243)
∂{εp }
∂{Req } f
= ([E1 ] + [E3 ]) · [Dc p ] (3.244)
∂{fp }
∂{Req }
= ([E1 ] + [E3 ]){bθ12c } · {e1 }T + {e1 } · {gnormal } + {e2 }({a2 } + {gtangent })
∂{p}
(3.245)

This completes the derivation of all the Jacobian [J] submatrices.


98

3.3 RC Stress Vector

The RC stress state is calculated as the sum of the contribution of the concrete and the
in-plane steel:

σ = σc + σs , (3.246)
np
X
σs = σsi , (3.247)
i=1

where σ, σc , σs , and σsi are the RC, concrete, steel, and ith in-plane steel layer average
stress tensors, and np is the number of in-plane layers. The RC stress vector or average RC
stress vector, {σ}, is shown in Equation 3.248, which is the same as Equation 3.246, but
written in a vector representation in the XY coordinate system:

{σ} = {σc } + {σs }, (3.248)


np
X
{σs } = {σsi }, (3.249)
i=1

where {σ} is the average RC stress vector, which has been discusses previously in this
document, and {σc } and {σsi } are the vector representation in the XY coordinate system
of the average concrete and ith in-plane layer steel stress states, respectively.

3.3.1 Steel

The vector representation of σsi in the αi coordinate system is

 

 ρ f 

 i si 
 
{σsi }αi = 0 , (3.250)

 


 0 

where ρi is the steel ratio in the direction of ith the in-plane layer. The use of ρi has the
effect of smearing the steel stresses over the RC material. The vector representation of σsi
99

in the XY coordinate system is

{σsi } = [Tstress (−αi )] · {σsi }αi (3.251)

where [Tstress ] is the stress transformation matrix as in Equation 3.279, and αi is the angle
between the ith in-plane steel layer and the X-axis measured counterclockwise.

3.3.2 Concrete

The vector representation of σc in the 12c (concrete principal direction, 12c coordinate
system is  

 
fc1 
 

{σc }12c = fc2 (3.252)

 

0
 

which can be transformed to the XY coordinate system:

{σc } = [Tstress (−θ12c )] · {σc }12c (3.253)

where θ12c is the angle of the maximum principal concrete strain direction with respect to
the X-axis, measured counterclockwise.

3.4 RC Stiffness Matrix

The RC material stiffness tensor, D, linearizes the relationship between the stress and
strain tensors for an infinitesimal differential:

dσ = D : dε (3.254)

dσc = Dc : dε (3.255)

dσsi = Dsi : dε (3.256)

where the Dc and Dsi are defined as the concrete, and the ith in-plane steel layer stiffness
tensors. An expression for D can be found by applying the differential operator to Equation
100

3.246:

dσ = d(σc + σs ) (3.257)
Xnp
= dσc + dσsi (3.258)
i=1

introducing Equations 3.254, 3.255, and 3.256 into Equation 3.258 yields:

np
X
D : dε = Dc : dε + Dsi : dε (3.259)
i=1

defining the the steel stiffness tensor Ds as

np
X
Ds = Dsi (3.260)
i=1

and substituting 3.260 into 3.259 yields to an expression for [D]:

D : dε = (Dc + Ds ) : dε (3.261)

D = Dc + Ds (3.262)

Equations 3.262, 3.255, 3.256, and 3.260 written in matrix and vector notation yield:

[D] = [Dc ] + [Ds ] (3.263)

d{σc } = [Dc ] · d{ε} (3.264)

d{σsi } = [Dsi ] · d{ε} (3.265)


Xnp
[Ds ] = [Dsi ] (3.266)
i=1

where [D], [Dc ], [Ds ], and [Dsi ] are the RC, concrete, steel and ith in-plane steel layer
material stiffness matrices in the XY coordinate system, respectively.
101

3.4.1 Steel

Two different expressions for the steel material stiffness matrix are presented: secant fixed
and tangent fixed. Here the term ”fixed” refers to the fact the steel layers have a fixed
orientation. The secant fixed stiffness matrix uses the traditional secant stiffness for the
material, and is the one used originally in the DSFM model. The tangent stiffness uses the
traditional tangent stiffness for the material and is the one used in DSFM-TG model.

Secant Fixed

The DSFM uses a secant fixed stiffness matrix for the ith in-plane steel layer, which is
defined as:
[Dsi ] = [Tstif f (−αi )]T · [Dsi ]αi · [Tstif f (−αi )] (3.267)

where
 
ρE 0 0
 i si 
αi  
[Dsi ] =  0 0 0 (3.268)
 
0 0 0
fsi
E si = (3.269)
εsi

The [Tstif f ] matrix is the stiffness transformation, which is defined in Equation 3.281.

Tangent Fixed

The original DSFM steel stiffness matrix, [Dsi ] is not tangent, which unable its used in
DSFM-TG. A new consistent tangent is developed for use in inDSFM-TG . It is called
fixed, because it uses the fact the the orientation of the steel is fixed. The tangent fixed
stiffness matrix associated with the ith in-plane steel layer material stiffness, [Dsi ], is found
102

by applying a differential operator to Equation 3.251:

d{σsi } = d([Tstress (−αi )] · {σsi }αi ) (3.270)


0
= −[Tstress (−αi )] · {σsi }αi dαi + [Tstress (−αi )] · d{σsi }αi (3.271)

= [Tstress (−αi )] · d{σsi }αi (3.272)

As dαi is zero because αi , the angle defining the orientation of the steel layer i with respect
to the x-axis, is constant. As noted previously the fixed angle αi is where the denomination
fixed material comes from. The differential d{σsi }αi can be expressed as:

d{σsi }αi = [Dsi ]αi · d{εs }αi (3.273)

where the [Dsi ]αi is the ith in-plane steel layer material stiffness in the αi coordinate system.
The differential d{εs }αi is found by applying a differential operator to Equation 2.6 from
Section 2.3:

d{εs }αi = d([Tstrain (αi )] · {εs }) (3.274)

= [Tstrain (αi )] · d{εs } (3.275)

given that the average steel strain {εs } equal the average strain {ε}

d{εs }αi = [Tstrain (αi )] · d{ε} (3.276)

where [Tstrain ] is the strain transformation matrix as defined in Equation 3.280. Substituting
Equations 3.276 into 3.273 leads to:

d{σsi }αi = [Dsi ]αi · [Tstrain (αi )] · d{ε} (3.277)

substituting this into 3.272 results in

d{σsi } = [Tstress (−αi )] · [Dsi ]αi · [Tstrain (αi )] · d{ε} (3.278)


103

The stress, strain, and stiffness transformation matrices, [Tstress ], [Tstrain ], and [Tstif f ] are
defined as:
 
cos2 (ω) sin2 (ω) 2 cos(ω) sin(ω)
 
 
[Tstress (ω)] =  sin2 (ω) cos2 (ω) −2 cos(ω) sin(ω)  (3.279)
 
− cos(ω) sin(ω) cos(ω) sin(ω) cos2 (ω) − sin2 (ω)
 
cos2 (ω) sin2 (ω) cos(ω) sin(ω)
 
 
[Tstrain (ω)] =  sin2 (ω) cos2 (ω) − cos(ω) sin(ω)  (3.280)
 
2 2
−2 cos(ω) sin(ω) 2 cos(ω) sin(ω) cos (ω) − sin (ω)
 
2 2
cos (ω) sin (ω) − cos(ω) sin(ω)
 
 2 2 
[Tstif f (ω)] =  sin (ω) cos (ω) cos(ω) sin(ω)  (3.281)
 
2 2
2 cos(ω) sin(ω) −2 cos(ω) sin(ω) cos (ω) − sin (ω)

where ω is the rotation angle, measured positive when rotated counterclockwise. The trans-
formation matrices are related by the following identities:

[Tstress (ω)] = [Tstif f (ω)]T (3.282)

[Tstrain (ω)] = [Tstif f (−ω)] (3.283)

Substituting Equations 3.282 and 3.283 into Equation 3.278:

d{σsi } = [Tstif f (−αi )]T · [Dsi ]αi · [Tstif f (−αi )] · d{ε} (3.284)

The matrix [Dsi ] is found by comparing Equation 3.284 with Equation 3.265:

[Dsi ] = [Tstif f (−αi )]T · [Dsi ]αi · [Tstif f (−αi )] (3.285)

This is the standard stiffness transformation used for fixed materials, and it is identical to
the transformation used in the steel secant fixed material stiffness matrix shown in Equation
3.267.
104

The matrix [Dsi ]αi can be found by applying a differential operator to Equation 3.250:
 

 
ρi fsi 
 

d{σsi }αi = d 0 (3.286)

 


 0 

= d(ρi fsi {e1 }) (3.287)


dfsi
= ρi dεsi {e1 } (3.288)
dεsi
= ρi Esi dεsi {e1 } (3.289)

where Esi is the tangent material slope:

dfsi
Esi = (3.290)
dεsi

The differential dεsi can be expressed as:

dεsi = {e1 }T · d{εs }αi (3.291)

Substituting Equation 3.291 into 3.289:

d{σsi }αi = ρi Esi ({e1 }T · d{εs }αi ){e1 } (3.292)

= ρi Esi ({e1 } · {e1 }T ) · d{εs }αi (3.293)

= (ρi Esi {e1 } · {e1 }T ) · d{εs }αi (3.294)

The matrix [Dsi ]αi is found by comparing Equation 3.294 with Equation 3.273.

[Dsi ]αi = ρi Esi {e1 } · {e1 }T (3.295)


 
ρE 0 0
 i si 
 
=  0 0 0 (3.296)
 
0 0 0

As before, the expression used for the [Dsi ]αi matrix is similar to that of the secant fixed
105

steel material stiffness shown in Equation 3.268, but the tangent fixed stiffness matrix uses
a tangent material slope Esi instead of the secant material slope E si .

3.4.2 Concrete

Three different expressions for the concrete stiffness matrix [Dc ] are presented: secant fixed,
tangent fixed, and tangent rotating. Where fixed refers to the idea that the rotation of the
principal concrete directions is not considered in the stiffness matrix. The secant fixed
matrix was originally used in the DSFM model, but since it is secant it is not suitable for
use in traditional tangent based solution algorithms. The tangent fixed stiffness is similar
to the secant fixed stiffness matrix but uses the tangent material slopes instead of secant
material slopes, and it is also the initially intuitive first option, when wanting to switch
from secant to tangent stiffness matrix. The tangent rotating stiffness matrix incorporate
the rotating nature of the concrete model. The tangent rotating matrix is a consistent
tangent, and is e one selected to be used in DSFM-TG . An extensive comparison of the
performance of this three type of concrete material stiffness matrices is presented in Chapter
5.

Secant Fixed

The secant fixed stiffness matrix in the XY coordinate system is is defined in the DSFM as:

[Dc ] = [Tstif f (−θ12c )]T · [Dco ]12c · [Tstif f (−θ12c )] (3.297)


106

where
 
E c1 0 0
 
 
[Dco ]12c =  0 E c2 0  (3.298)
 
0 0 G
fc1
E c1 = (3.299)
εc1
fc2
E c2 = (3.300)
εc2
E c1 E c2
G = (3.301)
E c1 + E c2

Tangent Fixed

The tangent fixed stiffness matrix is identical to the secant fixed stiffness matrix with the
exception the tangent fixed stiffness uses tangent material slopes instead of secant material
slopes.
[Dc ] = [Tstif f (−θ12c )]T · [Dco ]12c · [Tstif f (−θ12c )] (3.302)

where
 
Ec1 0 0
 
 
[Dco ]12c =  0 Ec2 0  (3.303)
 
0 0 G
∂fc1
Ec1 = (3.304)
∂εc1
∂fc2
Ec2 = (3.305)
∂εc2
Ec1 Ec2
G = (3.306)
Ec1 + Ec2

Tangent Rotating

During the development of this research it was notices that neither the secant fixed or tan-
gent fixed material stiffness matrix resulted in stable analysis that converged quadratically,
as it would be expected when using a constant tangent. This motivated a derivation of [Dc ]
independent of the work developed in the DSFM. The derivation starts with the definition
107

of [Dc ]:

d{σc } = [Dc ] · d{ε} (3.307)

To determine [Dc ], a differential operatio is applied to Equation 3.253, which defines d{σc }
as a function of ConcreteAverageStressV ector12c :

d{σc } = d([Tstress (−θ12c )] · {σc }12c ) (3.308)

= d[Tstress (−θ12c )] · {σc }12c + [Tstress (−θ12c )] · d{σc }12c (3.309)


0
= −[Tstress (−θ12c )] · {σc }12c dθ12c + [Tstress (−θ12c )] · d{σc }12c (3.310)

The differential dθ12c is not zero as the orientation of the concrete principal stresses is not
constant. This is the fundamental difference between a rotational and a fixed material
model such reinforcing steel. The differential d{σc }12c can be expressed as:

{σc }12c = [R2D ] · {σc3P D } (3.311)

Applying a differential to this expression:

d{σc }12c = [R2D ] · d{σc3P D } (3.312)

(3.313)

and expanding d{σc3P D }:

³
ε
d{σc }12c = [R2D ] · [Dc3P D,Load ] · d{ε3Pc´
D,Load
} + [Dc p ] · d{εp }
f
+ [Dc p ] · d{fp }{bθ12c } · dθ12c (3.314)
108

The following variables are defined to simplify the derivation of [Dc ]:

d{ε3P
c
D,Load
} = [N1 ] · d{ε} (3.315)

d{εp } = [N2 ] · d{ε} (3.316)

d{fp } = [N3 ] · d{ε} (3.317)

Substituting the last three expressions into Equation 3.314:

d{σc }12c = [R2D ] · d{σc3P D } (3.318)


³
ε
= [R2D ] · [Dc3P D,Load ] · [N1 ] +¶ [Dc p ] · [N2 ] (3.319)
f dθ12c
+ [Dc p ] · [N3 ] + {bθ12c } · · d{ε} (3.320)
d{εc }

Finding matrices [N1 ], [N2 ], and [N3 ] is a very complicated task, since multiaxial effects
such as Poisson, compression softening, and confinement need to be incorporated. Initially,
to simplify the derivation of [Dc ], those effects are ignored. This results in:

{ε3P
c
D,Load
} = {ε3P
c
D
} (3.321)

d{ε3P
c
D,Load
} = d{ε3P
c
D
} (3.322)

{εp } = {1, 1, 1}T εco (3.323)

d{εp } = {0, 0, 0}T (3.324)


0
{fp } = {1, 1, 1}T fc (3.325)

d{fp } = {0, 0, 0}T (3.326)

Introducing the above simplifications, Equation 3.314 reduces to:

d{σc }12c = [R2D ] · d{σc3P D } (3.327)


¡ ¢
= [R2D ] · [Dc ]3P D · d{ε3P
c
D
} + {bθ12c } · dθ12c (3.328)

= ([R2D ] · [Dc ]3P D ) · d{ε3P


c
D
} + [R2D ] · {bθ12c } · dθ12c (3.329)

(3.330)
109

where DcoT hreeP D is


 
Ec1 0 0
∂{σc3P D } 


[Dc ]3P D = 3P D
=  0 Ec2 0  (3.331)
∂{εc }  
0 0 Ecz

then the term 3.330 can be written as


 
Ec1 0 0
 
 
[R2D ] · [Dc ]3P D =  0 Ec2 0 = [Dco ]12c (3.332)
 
0 0 0

Substituting 3.332 into 3.330

d{σc }12c = [Dco ]12c · d{ε3P


c
D
} + [R2D ] · {bθ12c } · dθ12c (3.333)

(3.334)

Observing matrix [Dco ]12c it is possible to infer the following Equation:

[Dco ]12c · d{εc }12c = [Dco ]12c · d{ε3P


c
D
} (3.335)

Introducing the last Equation into ??:

d{σc }12c = [Dco ]12c · d{εc }12c + [R2D ] · {bθ12c } · dθ12c (3.336)

The following variables are defined to simplify the derivation

d{εc }12c = [N ] · d{ε} (3.337)

d{εc } = [M ] · d{ε} (3.338)


110

dθ12c
and also writing as:
d{εc }

dθ12c
dθ12c = · d{εc } (3.339)
d{εc }

Introducing the last three Equations into n Equation 3.336:

dθ12c
d{σc }12c = [Dco ]12c · [N ] · d{ε} + [R2D ] · {bθ12c } · · [M ] · d{ε} (3.340)
d{εc }
µ ¶
12c dθ12c
= [Dco ] · [N ] + [R2D ] · {bθ12c } · · [M ] · d{ε} (3.341)
d{εc }

This completes the derivation of an expression for d{σc }12c . Expressions for [M ] and [N ]
need to be found now in order to complete the derivation of [Dc ]. Starting again from the
first term of Equation 3.309 d[Tstress (−θ12c )] · {σc }12c differential:
µ ¶
12c d[Tstress (−θ12c )] d(−θ12c ) dθ12c
d[Tstress (−θ12c )] · {σc } = · d{εc } · {σc }12c
d(−θ12c ) dθ12c d{εc }
(3.342)
µ ¶
0 dθ12c
= −[Tstress (−θ12c )] · d{εc } · {σc }12c (3.343)
d{εc }
µ µ ¶¶
0 dθ12c
= −[Tstress (−θ12c )] · d{εc } · {σc }12c (3.344)
d{εc }
µµ ¶ ¶
0 dθ12c 12c
= −[Tstress (−θ12c )] · · d{εc } {σc } (3.345)
d{εc }
µ ¶
0 dθ12c
= −[Tstress (−θ12c )] · {σc }12c · · d{εc } (3.346)
d{εc }

Substituting Equation 3.338 into Equation 3.346


µ ¶
12c 0 12c dθ12c
d[Tstress (−θ12c )] · {σc } = −[Tstress (−θ12c )] · {σc } · · [M ] · d{ε}
d{εc }
(3.347)
111

Finally substituting 3.347 and 3.341 into 3.310:


µ µ ¶
0 12c dθ12c
d{σc } = −[Tstress (−θ12c )] · {σc } · · [M ]
d{εc }
+ [Tstress (−θ12c )] · [Dc0 ]12c · [N ]

dθ12c
+ [R2D ] · {bθ12c } · · [M ] · d{ε} (3.348)
d{εc }
= ([A1 ] · [M ] + [A2 ] · [N ] + [A3 ] · [M ]) · d{ε} (3.349)

where
µ ¶
0 12c dθ12c
[A1 ] = −[Tstress (−θ12c )] · {σc } · (3.350)
d{εc }
12c
[A2 ] = [Tstress (−θ12c )] · [Dc0 ] (3.351)
dθ12c
[A3 ] = [R2D ] · {bθ12c } · · [M ] (3.352)
d{εc }

The expression found for [Dc ] at this level of the derivation is is equal to

[Dc ] = ([A1 ] · [M ] + [A2 ] · [N ] + [A3 ] · [M ]) (3.353)

dθ12c
as shown in Equation 3.349. However, matrices [M ], [N ] and vector have not been
d{εc }
derived yet. Matrix [N ] is determined using Equation 2.7 from Section 2.3:

d{εc }12c = d ([Tstrain (θ12c )] · {εc }) (3.354)

= d[Tstrain (θ12c )] · {εc } + [Tstrain (θ12c )] · d{εc } (3.355)


µ µ ¶¶
d[Tstrain (θ12c )] dθ12c
= · d{εc } · {εc } + [Tstrain (θ12c )] · d{εc }(3.356)
dθ12c d{εc }
µµ ¶ ¶
0 dθ12c
= [Tstrain (θ12c )] · · d{εc } {εc }
d{εc }
+ [Tstrain (θ12c )] · d{εc } (3.357)
µ ¶
0 dθ12c
= [Tstrain (θ12c )] · {εc } · · d{εc } + [Tstrain (θ12c )] · d{εc } (3.358)
d{εc }
µ µ ¶ ¶
0 dθ12c
= [Tstrain (θ12c )] · {εc } · + [Tstrain (θ12c )] · d{εc } (3.359)
d{εc }
µ µ ¶ ¶
0 dθ12c
= [Tstrain (θ12c )] · {εc } · + [Tstrain (θ12c )] · [M ] · d{ε} (3.360)
d{εc }
112

Comparing 3.360 with 3.337 leads to:


µ µ ¶ ¶
0 ∂θ12c
[N ] = [Tstrain (θ12c )] · {εc } · + [Tstrain (θ12c )] · [M ] (3.361)
∂{εc }

where the matrix [N ] can be expressed as:

[N ] = [A4 ] · [M ] (3.362)
µ ¶
0 ∂θ12c
[A4 ] = [Tstrain (θ12c )] · {εc } · + [Tstrain (θ12c )] (3.363)
∂{εc }

The matrix [M ] is determined using Equation 2.3 from Section 2.4:

d{εc } = d{ε} − d{εslip } (3.364)


∂{εslip }
= d{ε} − · d{ε} (3.365)
∂{ε}
µ ¶
∂{εslip }
= [I] − · d{ε} (3.366)
∂{ε}

∂{εslip }
[M ] = [I] − (3.367)
∂{ε}
dθ12c
where [I] is the 3x3 identity matrix. The derivative can be calculated using Equation
d{εc }
2.8 from Section 2.3:
µ µ ¶¶
1 γcxy
dθ12c = d arctan (3.368)
2 εcx − εcy
n o
− sin(2θ12c ), sin(2θ12c ), − cos(2θ12c )
= · d{εc } (3.369)
2(εc1 − εc2 )
n o
dθ12c − sin(2θ12c ), sin(2θ12c ), − cos(2θ12c )
= (3.370)
d{εc } 2(εc1 − εc2 )

Substituting Equation 3.362 into Equation 3.349 leads to:

d{σc } = ([A1 ] · [M ] + [A2 ] · [A4 ] · [M ] + [A3 ] · [M ]) · d{ε} (3.371)

= ([A1 ] + [A2 ] · [A4 ] + [A3 ]) · [M ] · d{ε} (3.372)


113

The [Dc ] matrix can now be expressed as:

[Dc ] = ([A1 ] + [A2 ] · [A4 ] + [A3 ]) · [M ] (3.373)

Substituting [A1 ], [A2 ],[A3 ] and [A4 ] into the last equation:

[Dc ] = ([Dco ] + [DcstressRot ] + [DcstrainRot ] + [Dcθ12c ]) · [M ] (3.374)

where

(3.375)
T 12c
[Dco ] = [Tstif f (−θ12c )] · [Dc0 ] · [Tstif f (−θ12c )] (3.376)
0T dθ12c
[DcstressRot ] = −[Tstif f (−θ12c )] · {σc } · (3.377)
d{εc }
T 12c 0 dθ12c
[DcstrainRot ] = −[Tstif f (−θ12c )] · [Dc0 ] · [Tstif f (−θ12c )] · {ε} · (3.378)
d{εc }
dθ12c
[Dcθ12c ] = [R2D ] · {bθ12c } · (3.379)
d{εc }

Until this point in this derivation of the simplified [Dc ] matrix (no multiaxial effects), it
becomes clear that there is a series of new terms not included in the original DSFM fixed
version of [Dc ]. The [Dc ] matrix shown in Equation 3.375 is made of five matrices ([Dco ],
[DcstressRot ], [DcstrainRot ], [Dcθ12c ], and [M ]).
Matrix [Dco ] resembles in its construction the secant fixed and tangent fixed matrices,
but without the G term in the [Dco ]12c matrix. The fact that matrix [Dco ] does not have
a G term is in accordance with the definition of {σc }12c and {εc }12c . Both variables are
written in principal concrete directions; therefore, no shear stress can be generated from a
variation in the principal concrete strains.
Matrices [DcstressRot ] and [DcstrainRot ] account for the stiffness due to the rotation of
the concrete stress and strains. Matrix [Dcθ12c ] represents the effect in the tension stiff-
ening models due to the change of angle of the in-plane steel layers with respect to the
crack orientation defined by θ12c . This group of matrices ([DcstressRot ], [DcstrainRot ], and
114

dθ12c
[Dcθ12c ]) account for the rotation of the principal concrete directions captured by the
d{εc }
derivative. This is the fundamental difference with respect to the secant and tangent fixed
approach, where the rotation of the µprincipal concrete
¶ directions is not accounted for in the
dθ12c
construction of the stiffness matrix =0 .
d{εc }
Finally matrix [M ] accounts for the effect of the variation of slip δslip at the crack.
∂{εslip }
Matrix [M ] is a 3 × 3 identity matrix if the derivative is a zero 3 × 3 matrix,
∂{ε}
∂{εslip }
which is the case when the concrete is uncracked. When the concrete is cracked,
∂{ε}
is generally not a zero 3 × 3 matrix and it has effects on the [Dc ] matrix even if there is no
slip, δslip .
The only missing term to complete an expression for the simplified version of the [Dc ]
∂{εslip }
matrix is the term . This term can be found by applying a differential operator to
∂{ε}
{εslip }:

d{εslip } = d([Tstrain (−θ12c )] · {εslip }12c ) (3.380)

= d[Tstrain (−θ12c )] · {εslip }12c + [Tstrain (−θ12c )] · d{εslip }12c (3.381)


µ µ ¶¶
0 ∂θ12c
= −[Tstrain (−θ12c )] · d{εc } · {εslip }12c
∂{εc }

+ [Tstrain (−θ12c )] · d{εslip }12c (3.382)


µµ ¶ ¶
0 ∂θ12c 12c
= −[Tstrain (−θ12c )] · · d{εc } {εslip }
∂{εc }

+ [Tstrain (−θ12c )] · d{εslip }12c (3.383)


µ ¶
0 12c ∂θ12c
= −[Tstrain (−θ12c )] · {εslip } · · d{εc }
∂{εc }

+ [Tstrain (−θ12c )] · d{εslip }12c (3.384)

= [G] · d{εc } + [Tstrain (−θ12c )] · d{εslip }12c (3.385)

= [G] · [M ] · d{ε} + [Tstrain (−θ12c )] · d{εslip }12c (3.386)

(3.387)
115

where µ ¶
0 12c ∂θ12c
[G] = −[Tstrain (−θ12c )] · {εslip } · (3.388)
∂{εc }

Expanding the d{εslip }12c differential:


 

 0  

 

 
d{εslip }12c = d 0 (3.389)

 

 δ  
 slip 
S
µ cr ¶
δslip
= d {e3 } (3.390)
Scr
µ ¶
δslip
= d {e3 } (3.391)
Scr
µ ¶
dδslip δslip dScr
= d − 2
{e3 } (3.392)
Scr Scr

Expanding the dScr differential:

∂Scr ∂θ12c
dScr = · d{εc } (3.393)
∂θ12c ∂{εc }
∂Scr ∂θ12c
= · [M ] · d{ε} (3.394)
∂θ12c ∂{εc }

Expanding the dδslip differential:

∂δslip ∂δslip
dδslip = dvc + dwcr (3.395)
∂vc ∂wcr
116

Expanding the dwcr differential:

dwcr = d(εc1 Scr ) (3.396)

= d({e1 }T · {εc }12c Scr ) (3.397)

= {e1 }T · d{εc }12c Scr + {e1 }T · {εc }12c dScr (3.398)


µ ¶
¡ ¢ ∂Scr ∂θ12c
= {e1 }T · [N ] · d{ε}Scr + {e1 }T · {εc }12c · [M ] · d{ε} (3.399)
∂θ12c ∂{εc }
µ µ ¶¶
∂Scr ¡ ¢ ∂θ12c
= Scr {e1 }T · [N ] + {e1 }T · {εc }12c · [M ] · d{ε} (3.400)
∂θ12c ∂{εc }
= ({p1 } · [N ] + {p2 } · [M ]) · d{ε} (3.401)

where the 1 × 3 vectors {p1 } and {p2 } are:

{p1 } = Scr {e1 }T (3.402)


µ ¶
∂Scr ¡ ¢ ∂θ12c
{p2 } = {e1 }T · {εc }12c (3.403)
∂θ12c ∂{εc }

∂{εslip }
During the intent of deriving an expression for matrix [M ] = [I] − ∂{ε} it becomes clear
that some of the same derivatives obtained in the internal RC state problem described in
Section 3.2 are starting arise again. In order to avoid repeating the long process completed
for the Jacobian matrix, [J] a new approach is taken to derive the the consistant tangent
[Dc ] This approach is described below:

Alternative Derivation of [Dc ] The Jacobian [J] obtained in Section 3.2 contains all
the derivatives needed to find not only the simplified version of [Dc ] (no multiaxial effects),
but also the full version of [Dc ] (with multiaxial effects). The following alternative or indirect
derivation of the full [Dc ] takes advantage of the already available Jacobian derived from
the internal RC state problem.
The idea is to relate the differential d{σc } to the to the d{x} differential. Where {x}
was defined before as the internal RC variable vector. Then d{x} can be related to d{ε}
through the Jacobian matrix [J]. Once those relationships are found, it is possible to find
link d{σc } to d{ε} by a certain expression, which is in fact [Dc ]. Starting the derivation
117

from Equation 3.310:

d{σc } = d[Tstress (−θ12c )] · {σc }12c + [Tstress (−θ12c )] · d{σc }12c (3.404)
0
= −[Tstress (−θ12c )] · {σc }12c dθ12c + [Tstress (−θ12c )] · d{σc }12c (3.405)

The differential dθ12c can be written as:

dθ12c = {e1 }T · d{p} (3.406)

where vector {p} is defined in Equation 3.17 in Section 3.2. The differential d{σc }12c can
be written as:

³
ε
d{σc }12c = [R2D ] · [Dc3P D,Load ] · d{ε3P
c
D,Load
} +´ [Dc p ] · d{εp }
fp
+ [Dc ] · d{fp } + {bθ12c } · {e1 }T · d{p} (3.407)

where vectors {ε3P


c
D,Load
}, {εp }, and {fp } are as defined in Equations 3.15, 3.15, and 3.16
in Section 3.2. The differentials d{ε3P
c
D,Load
}, d{εp }, d{fp } and d{p} can be expressed as a
function of d{x}:

d{ε3P
c
D,Load
} = [E 3P D,Load ] · d{x} (3.408)

d{εp } = [E εp ] · d{x} (3.409)

d{fp } = [E fp ] · d{x} (3.410)

d{p} = [E p ] · d{x} (3.411)

where [E 3P D,Load ], [E εp ], [E fp ], and [E p ] are known 3x15 matrices:

[E 3P D,Load ] = [ [0], [I], [0], [0], [0] ] (3.412)

[E εp ] = [ [0], [0], [I], [0], [0] ] (3.413)

[E fp ] = [ [0], [0], [0], [I], [0] ] (3.414)

[E p ] = [ [0], [0], [0], [0], [I] ] (3.415)


118

and matrix [I] is a 3 × 3 identity matrix, and [0] is 3 × 3 zero matrix. Using the last four
equations in 3.407:

d{σc }12c = [A] · d{x} (3.416)

where [A] is defined as

³ ´
ε f
[A] = [R2D ]· [Dc3P D,Load ] · [E 3P D,Load ] + [Dc p ] · [E εp ] + [Dc p ] · [E fp ] + {bθ12c } · {e1 }T · [E p ]
(3.417)
Replacing equations 3.416 and 3.406 in Equation 3.405:

0
d{σc } = −[Tstress (−θ12c )] · {σc }12c · {e1 }T · d{p} + [Tstress (−θ12c )] · [A] · d{x} (3.418)

And replacing d{p} by [E p ] · d{x} :

d{σc } = [B] · d{x} (3.419)

where the 3 × 15 matrix [B] is

0
[B] = −[Tstress (−θ12c )] · {σc }12c · {e1 }T · [E p ] + [Tstress (−θ12c )] · [A] (3.420)

Equation 3.419 provides a link between d{σc } and d{x}, but d{σc } still needs to be linked
to d{ε} to find [Dc ]. This link is accomplished indirectly by finding a relationship between
d{x} and d{ε}, through the use of the Jacobian [J] matrix from the converged internal RC
state.
The internal RC state problem residual vector {R} for a converged internal RC state
defined by the solution vector {xsol } is

{R({xsol }, {ε})} = {0} (3.421)


119

Where {0} is a 15 × 1 zero vector. Applying a differential operator to the last equation:

∂{R({xsol }, {ε})} ∂{R({xsol }, {ε})}


d{R} = · d{x} + · d{ε} = {0} (3.422)
∂{x} ∂{ε}

The first derivative is the definition of the Jacobian [J] and the second is a known matrix:

∂{R({xsol }, {ε})}
= [J({xsol }, {ε})] (3.423)
∂{x}
∂{R({xsol }, {ε})}
= [E ε ] (3.424)
∂{ε}

Where [E ε ] is a 15 × 3 matrix:

[E ε ] = [ [I], [0], [0], [0], [0] ]T (3.425)

Introducing Equations 3.423 and 3.424 into Equation 3.422:

[J] · d{x} + [E ε ] · d{ε} = {0} (3.426)

where the arguments of [J] have been dropped for clarity. The differential d{x} can be
obtained from the last Equation as:

d{x} = −[J]−1 · [E ε ] · d{ε} (3.427)

Combining equations 3.427 and 3.419

d{σc } = −[B] · [J]−1 · [E ε ] · d{ε} (3.428)

Which defines the full concrete stiffness matrix in the XY coordinate system:

[Dc ] = −[B] · [J]−1 · [E ε ] (3.429)

This an exact expression for [Dc ] including all the multiaxial effects. The interesting part
about this expression is that it reuses a by-product of the internal RC problem, avoiding
120

the need to develop a direct analytical expression for [Dc ]. This task can be very time-
consuming, and it is difficult to modify, especially if more effects are added to the model.

3.5 Summary and Conclusions

The DSFM-TG compatibility, constitutive, and equilibrium governing equations to develop


a set of mathematical expressions that define the internal RC material state. The internal
RC material is uniquely defined by 15 internal variables that include parameters related to
the multiaxial loading effects such as Poisson, confinement, compression softening and out-
of-plane reinforcement. The solution for the internal RC material state requires the solution
of a 15x15 nonlinear system of equations. This Chapter also 3 presents an expression for
the consistent tangent material stiffness matrix which replaces the secant stiffness matrix
used in the original DSFM.
The main conclusions obtained from this Chapter are

• The internal RC material state is defined by a set of equations and parameters con-
densed in a 15 × 15 nonlinear system of equations.

• The solution of the in-plane stresses requires the consideration of the out-of-plane
stresses.

• The original DSFM concrete material stiffness matrix (secant fixed), [Dc ], written in
the tangent form (tangent fixed), does not include the effect of the principal concrete
stress rotation

• The DSFM-TG concrete material stiffness matrix, [Dc ], includes the effect of the
principal concrete stress rotation and is calculated using the Jacobian matrix, [J],
from the internal RC material problem.

The DSFM governing equations that define the internal RC problem are condensed
into condensed a system of 15 nonlinear equations. These equations address the material
compatibility and internal equilibrium, accounting for all the multiaxial effect include in the
model. The number of internal basic variables is 15 which includes the parameters related to
the multiaxial loading effects such as Poisson, confinement, compression softening and out-
of-plane reinforcement. The general or cracked 15x15 Nonlinear system of equations could
121

be reduced to a 5x5 system if none of the multiaxial effects were included. Even though it
may seem a like a good idea to reduce the internal problem size when the multiaxial effect
are not included in the model, no significant process time improvement was observed, and
the process implementation becomes rather cumbersome by adding all the different analysis
combinations cases.
Even though the DSFM-TG models the behavior od 2D plane stresses RC material, the
out-of-plane concrete and steel stresses need to be included into the internal RC problem
since they affect the in-plane response through Poisson effect and confinement.
In order to find a correct expression for [Dc ] it was necessary to incorporate the rotating
nature of RC during the derivation of the tangent material stiffness matrix [Dc ]. This effect
of the principal concrete direction rotation was not incorporated into the original secant
fixed(DSFM) or tangent fixed [Dc ] expressions. The secant fixed and tangent fixed expres-
sions use a G term in the diagonal of the matrix written in the principal concrete directions.
This G term comes from the traditional expression for fixed orthotropic materials, but is
not applicable to a rotating orthotropic material such as concrete. The correct expression
[Dc ] found was called tangent rotating because it uses tangent material stiffnesses, and also
because it recognizes the rotating nature of concrete in this model.
The expression developed for [Dc ] uses an indirect approach, using the [J] found during
the resolution of the internal RC problem rather than extracting an expression directly from
the 15 governing equations. The Jacobian matrix [J] contains all the derivatives needed to
find [Dc ], and reusing greatly simplifies the [Dc ] expression.
122

Chapter 4

DSFM-TG RC MATERIAL MODEL IMPLEMENTATION FOR USE


IN FEM

4.1 Introduction

This Chapter presents the implementation of the DSFM-TG formulation presented in Chap-
ter 2and 3. The objective of the implementation is to design a set of algorithms that will
generate the RC stress vector {σ} and material stiffness [D], given RC strain {ε}, and the
information from a previously converged material state. This implementation considers only
the case of monotonic loading; however, given the approach taken during the implementa-
tion, it could be expanded to cyclic loading. The implementation is done in Matlab.
The main function is called RCmaterial. This function is composed subfunctions or
processes as described in Section 4.2. Process 1 finds the internal RC material state. Process
2 generates the steel and concrete stiffness matrices and stress vectors. Process 3 generates
the RC material stiffness matrix and stress vector. Processes 1, 2, and 3 are presented
in Sections 4.3, 4.4, and 4.5, respectively. The actual MATLAB and additional MATLAB
scripts created for the implementation are provided in Appendix A.

4.2 RC material

The RCmaterial function is the main function and is responsible for generating the variables
{σ}n+1 , [D]n+1 , and RCstaten+1 from the variables {ε}n+1 , RCstaten , M atP rop, and
AnlP ar. The subscript n and n + 1 denote the FEM analysis stage number. The variables
{ε}n+1 , {σ}n+1 , and [D]n+1 are the RC strain, stress vector, and the material stiffness
matrix associated with stage n+1. RCstate is a MATLAB structure variable that stores the
internal RC vector {x} and additional internal RC variables relevant for analysis evaluation
purposes such as the internal error history and material slopes, among others. RCstate
does not store history variables, since they are not used in the 1D models. M atP rop is a
123

structure variable that contains the material properties used in the 1D models. AnlP ar is a
structure variable that contains analysis parameters to be used in the internal RC problem
such as the maximum number of iterations, error tolerances, and type of 1D models to
be used. However, future expansions of the 1D models from monotonic to cyclical loading
could use the variable RCstate to store history variables. The variable RCstate is used
inside different processes and subprocess to store some of their variables not individually
included in their output. The RCmaterial function is called by the finite element software
in the following way:

[{σ}n+1 , [D]n+1 , RCstaten+1 ] = RCmaterial({ε}n+1 , RCstaten , M atP rop, AnlP ar) (4.1)

where the variables in square brackets are the output variables and the variables in paren-
thesis are the input variables.
The function RCmaterial uses three processes to generate the output:

• Process 1: Internal RC state

• Process 2: Concrete and Steel material stiffness matrices and stress vectors

• Process 3: RC material stiffness matrix and stress vector

Process 1 solves the set of nonlinear equations that govern the internal RC material state
using an iterative Newton algorithm. Process 1 is the most complex process out of the three
used, and it is subdivided into several subprocesses. Process 1 is the first process executed
in the RCmaterial function, and it generates input for Process 2, along with the RCstate
variable for the current stage.
Process 2 computes the concrete and steel material stiffness matrices, and is executed
after the internal RC problem has been solved by Process 1. Process 2 is considerably
simpler than Process 1 because it uses explicit expressions to generate its output from the
input provided by Process 1.
Process 3 assembles the concrete ({σc }, [Dc ]), and steel ({σs }, [Ds ]) contributions to
the RC stress vector and stiffness matrix({σ}, [D]). Process 3 is executed last, and it is the
simplest out of the three processes. Process 3 completes the RCmaterial function output.
124

The RCmaterial function and its processes are shown in Figure 4.1.

Figure 4.1: RCmaterial function and processes

Figure 4.1 shows that Process 1 passes to Process 2 more concrete variables({σc3P D },
ε f
[Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, and θ12c ) than that of the steel({fs } and {Es }). This is
a reflection of the large number of variables on which the concrete stresses depend.
Notice that the subscripts n and n + 1 are not used inside the RCmaterial function.
This is because the material model is only concerned with the current (no subscript) and the
previous (subscript 0) stage, not with the specific stage number within the FEM analysis
as defined by n and n + 1.
Presented below are the headers of the functions associated with each process:
125

Process 1:

h i
ε f
[J], {fs }, {Es }, {σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, θ12c , RCstate
³ ´
= findRCstate {ε}, RCstate0 , M atP rop, AnlP ar (4.2)

Process 2:

h i ³
{σs }, [Ds ], {σc }, [Dc ] = getSteelConcStressStiff [J], {fs }, {Es },
´
ε f
{σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, θ12c , M atP rop, AnlP ar (4.3)

where the variables {fs } and {Es } are vectors that store the in-plane steel layers stresses
and material slopes:

{fs } = {fs1 , fs2 , ..., fsnp }T (4.4)

{Es } = {Es1 , Es2 , ..., Esnp }T (4.5)

Process 3:

h i ³ ´
{σ}, [D] = getRCStressStiff {σs }, [Ds ], {σc }, [Dc ] (4.6)

4.3 Process 1: Internal RC State

This process finds the current RC internal state associated with the current RC strain
vector, {ε}, by solving the system of 15 nonlinear equations that govern the behavior of the
RC material model. The process uses an iterative Newton solution algorithm to solve the
system of equations. The header of the function associated with this process is:

h i
ε f
[J], {fs }, {Es }, {σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, θ12c , RCstate
³ ´
= findRCstate {ε}, RCstate0 , M atP rop, AnlP ar (4.7)

(4.8)
126

where the input is the same as that of the RCmaterial function because this is the first
process or function executed inside RCmaterial. The output variables are associated with
the current stage defined by {ε}. RCstate0 contains the information from the previously
converged stage as indicated by the subscript 0. Process 1 is divided into five subprocesses:

• Subprocess 1N LS: Nonlinear System of Equations

• Subprocess 1RJ: Residual and Jacobian

• Subprocess 1CR: Crack Region Models

• Subprocess 1AR: Average Region Models

• Subprocess 1M E: Multiaxial Loading Effect Models

Process 1 and its subprocesses are shown in Figure 4.2. Subprocess 1N SL is responsible
for solving the system of 15 nonlinear equations, and controls the execution of all the other
subprocesses within Process 1. Subprocess 1N SL updates the value of the internal RC
variable vector {x} (or unknown vector) using the residual {R} and the Jacobian matrix [J]
obtained from Subprocess 1RJ. Thus, subprocess 1RJ is called multiple times by subprocess
1N LS. The variable RCstate0 contains vector {x}0 from the previously converged stage,
which is used as the initial value for vector {x}.
Subprocesses 1CR, 1AR, and 1M E return the internal RC variables associated with
the 1D models of the cracking and average regions, and the multiaxial loading effects,
respectively. None of this subproceeses currently uses variables from the previous stage, but
it has been made available for future 1D model expansion to cyclic loading.
Notice that subprocesses 1CR, 1AR, and 1M E return two type of variables: non-
derivative and derivative variables. The non-derivative variables are used to construct the
residual {R}, while a combination of derivative and non-derivative variables are used to
construct the Jacobian [J]. In general, the derivative variables represent a linearization in
an infinitesimal differential of the output/input relationship, which is used as a predictor of
the output/input rate of change. Those derivatives are the base of the Jacobian [J], which
is the predictor used to update {x}.
127

Figure 4.2: Process 1: Internal RC Problem

4.3.1 Subprocess 1N LS: Solve Non-linear System of Equations

This subprocess solves the system of 15 nonlinear equations that defines the internal RC
material state problem, using an iterative Newton solution algorithm. The internal RC
128

vector or unknown vector {x} is defined as:

{x} = {{ε3P
c
D T
} , {ε3P
c
D,Load T
} , {εp }T , {fp }T , {p}T }T (4.9)

{ε3P
c
D
} = {εc1 , εc2 , εcz }T (4.10)

{ε3P
c
D,Load
} = {εload load load T
c1 , εc2 , εcz } (4.11)

{εp } = {εp1 , εp2 , εpz }T (4.12)

{fp } = {fp1 , fp2 , fpz }T (4.13)

{p} = {θ12c , δslip , ∆ε1cr }T (4.14)

Subprocess 1N LS is divided into the following subprocesses:

• Subprocess 1N LS − Xintial: Initial {x}

• Subprocess 1N LS − Convergence: Convergence Criteria

• Subprocess 1N LS − Xupdate: {x} Update

Subprocess 1N LS and its subprocesses are shown in Figure 4.3.


Subprocess 1N LS executes first Subprocess 1N LS − Xinitial to find the initial value of
vector {x}. This initial {x} is the first trial {x}. This trial {x} is sent to Subprocess 1RJ
which returns the residual {R}, the Jacobian [J], and additional variables associated with
the trial {x}. The internal state validity is checked by Subprocess 1N LS − Convergence
by evaluating the error represented by the residual {R}. If the error is under a specified
tolerance, the internal state is called converged, and Subprocess 1N LS is exited. If the
error exceeds the specified tolerance, then {x}, {R} and [J] are sent to subprocess 1N LS −
Xupdate where the value of {x} is updated, and then again sent to Subprocess 1RJ. The
iterations continue until either convergence is achieved, the specified maximum number of
iterations is exceeded, or matrix [J] becomes singular.

Subprocess 1NLS − Xinitial. This subprocess finds the initial value of vector {x} to be
used in the system of 15 nonlinear equations. In general the initial {x} used for the current
129

stage is the {x} from the previously converged stage:

{x}initial = {x}0 (4.15)

where {x}0 is stored in RCstate0 . The exception to the occurs during the execution of
the first FEM analysis stage (n = 1). Since this is the first stage, there is no previously
converged stage, and the {x}initial is set to be

{x}0 = {{ε3P
c
D T
}0 , {ε3P
c
D,Load T
}0 , {εp }T0 , {fp }T0 , {p}T0 }T (4.16)

{ε3P
c
D
}0 = {0, 0, 0}T (4.17)

{ε3P
c
D,Load
}0 = {0, 0, 0}T (4.18)

{εp }0 = {εco , εco , εco }T (4.19)


0 0 0
{fp }0 = {fc , fc , fc }T (4.20)

{p}0 = {0, 0, 0}T (4.21)

where the values of {εp }0 and {fp }0 are not affected by compression softening or confinement
effects.

Subprocess 1NLS − Convergence. This subprocess evaluates the residual {R} by defin-
ing two error measurements:

errorε = |{Rε }| (4.22)

errorσ = |{Rσ }| (4.23)

{Rε } = {{R{ε} }T , {R{ε3P D,Load } }T , {R{εp } }T }T (4.24)


c

{Rσ } = { {R{fp } }T , {Req }T }T (4.25)

{R} = {{R{ε} }T , {R{ε3P D,Load } }T , {R{εp } }T , {R{fp } }T , {Req }T } (4.26)


c

The variable errorε measures the strain components of the residual {R}. The variable
errorσ measures the stress components of the residual {R}. The convergence is defined by
130

the convergeStatus flag, which is defined as:




1, converged state errorε < tolε and errorσ < tolσ
convergeStatus =

0, non-converged state otherwise
(4.27)

where tolε and tolσ are the strain and stress tolerances (both stored in AnlP ar). Vector {x}
is sent to subprocess 1N LS−Xupdate if flag updateStatus is equal to 1, where updateStatus
is:


1, update {x} convergeStatus = 0 and niter < nmax
iter
updateStatus =

0, do not update {x} otherwise
(4.28)

where nmax
iter is the maximum number of iterations (stored in AnlP ar). The variable niter

is a counter of the number of iterations performed and it is incremented by 1 each time


Subprocess 1N LS−Convergence is executed. If updateStatus is 0 then no further iterations
are performed and {x} and the other variables generated are sent to the output of Process
1N LS.

Subprocess 1NLS − Xupdate. Vector {x} is updated using the following equation:

{x}niter +1 = {x}niter − [J]−1


niter · {R}niter (4.29)

If the concrete is uncracked, the residuals for equilibrium in the crack normal and tangential
directions, Rnormal and Rtangent , are not considered because there is no crack. Also, when
the concrete is uncracked the variables δslip and ∆ε1cr equal zero. This reduces the problem
from a 15 system of equations to a 13 system of equations. The unknown vector, residual,
131

and Jacobian of the reduced problem are

{xred } = {εc1 , εc2 , εcz , εload load load T


c1 , εc2 , εcz , εp1 , εp2 , εpz , fp1 , fp2 , fpz , θ12c } (4.30)

{Rred } = { {R{ε} }T , {R{ε3P D,Load } }T , {R{εp } }T , {R{fp } }T , Rout−of −plane }T (4.31)


c

∂{Rred }
[Jred ] = (4.32)
∂{xred }

The reduced Jacobian matrix [Jred ] is the same as the full Jacobian matrix [J], but where
the Rnormal (13) and Rtangent (14) rows, and the δslip (14) and ∆ε1cr (15) columns have been
extracted from the full Jacobian [J]. The {x} is updated as

{x}niter +1 = {{xred }Tniter +1 , 0, 0}T (4.33)

where {xred }niter +1 is found as:

{xred }niter +1 = {xred }niter − [Jred ]niter · {Rred }niter (4.34)

There are situations when the Jacobian ([J] or [Jred ]) becomes singular. This could happen
due to the lost of stiffness or strength in of some resistant component of the RC material.
Examples of cases when the Jacobian [J] becomes singular are fracture of the bars crossing
the crack or yielding of the bars when using a steel model with a flat yielding plateau. If one
of these situations occurs, the Jacobian is not inverted and Subprocess 1N LS − Xupdate
returns the original {x} (not updated) and a flag indicating the solution has failed:


1, [J] or [Jred ] are not singular
solutionStatus =

0, [J] or [J ] are singular
red

(4.35)

The variable solutionStatus is stored in RCstate. The FEM program using the RCmaterial
function could opt to disregard the information from the particular integration point where
the internal RC problem has become unstable. The FEM analysis would not necessarily
132

fail because one integration point has failed to converged internally. The FEM could even
assigned to the integration point a stiffness and a stress zero (consistent with the internal
material failure associated with a singular Jacobian), and that would not necessarily lead
the FEM analysis to fail. The FEM analysis will fail when the structure stiffness matrix
becomes singular, rather than a particular integration point within an element.

4.3.2 Subprocess 1RJ: Residual and Jacobian

This subprocess is the responsible for creating the residual {R} and the Jacobian [J] for the
system of 15 nonlinear equations that define the internal RC problem. The header of the
function associated with this process is

h i
ε f
{R}, [J], {fs }, {Es }, {σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, θ12c , RCstate
³ ´
= getRJ {ε}, {x}, RCstate0 , M atP rop, AnlP ar

(4.36)

Subprocess 1RJ is divided into the following subprocesses:

• Subprocess 1RJ-Construction: Residual and Jacobian Construction

• Subprocess 1RJ-CR: Crack Region

• Subprocess 1RJ-AR: Average Region

• Subprocess 1RJ-M E: Multiaxial Loading Effects

Subprocess 1RJ and its subprocesses are shown in Figure 4.4. The subprocesses 1RJ-CR,
1RJ-AR, 1RJ-M E prepare the input to call subprocesses 1CR, 1AR, and 1M E. Subpro-
cesses 1RJ-CR, 1RJ-AR, and 1RJ-M E are executed in that order because some of the
input needed in the later is generated by the previous subprocess. The variable RCstate is
passed between subprocesses 1CR, 1AR, and 1M E to store variables that are not directly
included in the output. The output from subprocesses 1CR, 1AR, and 1M E is used by
subprocess 1RJ-Construction to construct the residual {R} and the Jacobian [J].
133

Subprocess 1RJ-Construction . This subprocess constructs the residual, {R}, and the
Jacobian, [J], using the equations derived in Section 3.2. The derived equations needed for
the implementation are presented again (excluding the derivation of them). The residual,
{R}, equations are re-written below:

{R} = {{R{ε} }T , {R{ε3P D,Load } }T , {R{εp } }T , {R{fp } }T , {Req }T }T (4.37)


c

{R{ε} } = {ε} − [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ) (4.38)

{R{ε3P D,Load } } = {ε3P


c
D,Load
} − [P ] · {ε3P
c
D
} (4.39)
c

{R{εp } } = {εp } − {β ε }εco (4.40)


0
{R{fp } } = {fp } − {β σ }fc (4.41)

{Req } = { Rnormal , Rtangent , Rout−of −plane }T (4.42)


np
X np
X
Rnormal = fc1 − ρi (fscri − fsi ) cos2 (φi ) − ρi fdcri cos(φi ) sin(φi ) (4.43)
i=1 i=1
np
X np
X
Rtangent = vc − ρi (fscri − fsi ) sin(φi ) cos(φi ) + ρi fdcri cos2 (φi ) (4.44)
i=1 i=1
Rout−of −plane = ρz fsz + fcz (4.45)

where
 
cos2 (ω) sin2 (ω) cos(ω) sin(ω)
 
 
[Tstrain (ω = −θ12c )] =  sin2 (ω) cos2 (ω) − cos(ω) sin(ω) (4.46)
 
−2 cos(ω) sin(ω) 2 cos(ω) sin(ω) cos2 (ω) − sin2 (ω)
δslip T
{εslip }12c = {0, 0, } (4.47)
Scr
{εc }12c = {εc1 , εc2 , 0}T = [R2D ] · {ε3P
c
D
} (4.48)
 
1 0 0
 
 
[R2D ] = 0 1 0 (4.49)
 
0 0 0

The origin of the variables needed to construct the residual {R} is shown in Table 4.1.
134

Table 4.1: Residual Variables


Variable Origin
{ε} Input
θ12c {x}
δslip {x}
Scr 1M E
{ε3P
c
D
} {x}
3P D,Load
{εc } {x}
{εp } {x}
{fp } {x}
{β ε } 1M E
{β σ } 1M E
fc1 Extracted from {σc3P D } from 1AR
fscri Extracted from {fscr } from 1AR
fsi Extracted from {fs } from 1AR
fdcri Extracted from {fdcr } from 1CR
vc 1CR
fcz Extracted from {σc3P D } from 1AR
fsz 1AR

The Jacobian [J] equations are re-written below:


2 3
∂{R{ε} } ∂{R{ε} } ∂{R{ε} } ∂{R{ε} } ∂{R{ε} }
6 , , , , 7
6 ∂{ε3Pc
D}
∂{ε3P
c
D,Load
} ∂{εp } ∂{fp } ∂{p} 7
6 ∂{R ∂{R{ε3P D,Load } } 7
6 3P D,Load } ∂{R{ε3P D,Load } } ∂{R{ε3P D,Load } } ∂{R{ε3P D,Load } } 7
6 {εc } c c c c 7
6 , , , , 7
6 ∂{ε3Pc
D}
∂{ε3P
c
D,Load
} ∂{εp } ∂{fp } ∂{p} 7
6 ∂{R{εp } } ∂{R{εp } } ∂{R{εp } } ∂{R{εp } } ∂{R{εp } } 7
6 7
[J] = 6 , , , , 7
6 ∂{ε3P D}
∂{ε3P D,Load
} ∂{εp } ∂{fp } ∂{p} 7
6 c c 7
6 ∂{R{fp } } ∂{R{fp } } ∂{R{fp } } ∂{R{fp } } ∂{R{fp } } 7
6 , , , , 7
6 7
6
6
∂{ε3Pc
D}
∂{ε3P
c
D,Load
} ∂{εp } ∂{fp } ∂{p} 7
7
4 ∂{Req } ∂{Req } ∂{Req } ∂{Req } ∂{Req } 5
, , , ,
∂{ε3Pc
D}
∂{ε3P
c
D,Load
} ∂{εp } ∂{fp } ∂{p}
(4.50)
135

The Jacobian submatrices associated with {R{ε} } are

∂{R{ε} }
= −[Tstrain (−θ12c )] · [R2D ] (4.51)
∂{ε3P
c
D}

∂{R{ε} }
= [0] (4.52)
∂{ε3P
c
D,Load
}
∂{R{ε} }
= [0] (4.53)
∂{εp }
∂{R{ε} }
= [0] (4.54)
∂{fp }
∂{R{ε} } 0
= [Tstrain (−θ12c )] · ({εslip }12c + {εc }12c ) · {e1 }T − [Tstrain (−θ12c )] · [H1 ]
∂{p}
(4.55)

The variables {e1 } and [H1 ] are

{e1 } = {1, 0, 0}T (4.56)

{e2 } = {0, 1, 0}T (4.57)

{e3 } = {0, 0, 1}T (4.58)


1 δslip dScr
[H1 ] = {e3 } · {e2 }T − 2 {e3 } · {e1 }T (4.59)
Scr Scr dθ12c

The Jacobian submatrices associated with {R{ε3P D,Load } } are


c

∂{R{ε3P D,Load } }
c
= −[P ] (4.60)
∂{ε3P
c
D}

∂{R{ε3P D,Load } }
c
= [I] − [Q1 ] (4.61)
∂{ε3P
c
D,Load
}
∂{R{ε3P D,Load } }
c
= −[Q2 ] (4.62)
∂{εp }
∂{R{ε3P D,Load } }
c
= [0] (4.63)
∂{fp }
∂{R{ε3P D,Load } }
c
= [0] (4.64)
∂{p}
136

The variable [I] is 3x3 identity matrix, and [Q1 ] and [Q2 ] are

X ∂[P ] ∂{ν}
[Q1 ] = · {ε3P
c
D
} · {ej }T · (4.65)
j=1,2,z
∂νj ∂{εc D,Load }
3P

X ∂[P ] ∂{ν}
[Q2 ] = · {ε3P
c
D
} · {ej }T · (4.66)
∂νj ∂{εp }
j=1,2,z

The Jacobian submatrices associated with {R{εp } } are:

∂{R{εp } }
= −[S1 ] · [S4 ]εco (4.67)
∂{ε3P
c
D}

∂{R{εp } }
= −([S2 ] · [S5 ] · [Dc3P D,Load ] + [S3 ])εco (4.68)
∂{ε3P
c
D,Load
}
∂{R{εp } } ε
= [I] − [S2 ] · [S5 ] · [Dc p ]εco (4.69)
∂{εp }
∂{R{εp } } f
= −[S2 ] · [S5 ] · [Dc p ]εco (4.70)
∂{fp }
∂{R{εp } }
= −[S2 ] · [S5 ] · {bθ12c } · {e1 }T εco (4.71)
∂{p}
(4.72)

The Jacobian submatrices associated with {R{fp } } are:

∂{R{fp } } 0
= −[S6 ] · [S8 ]fc (4.73)
∂{ε3P
c
D}

∂{R{fp } } 0
= −[S7 ] · [S9 ] · [Dc3P D,Load ]fc (4.74)
∂{ε3P
c
D,Load
}
∂{R{fp } } ε 0
= −[S7 ] · [S9 ] · [Dc p ]fc (4.75)
∂{εp }
∂{R{fp } } f 0
= [I] − [S7 ] · [S9 ] · [Dc p ]fc (4.76)
∂{fp }
∂{R{fp } } 0
= −[S7 ] · [S9 ] · {bθ12c } · {e1 }T fc (4.77)
∂{p}
(4.78)
137

The Jacobian submatrices associated with {Req } are:

∂{Req }
= ρz Esz [E3 ] + {e2 }{a1 } (4.79)
∂{ε3P
c
D}

∂{Req }
= ([E1 ] + [E3 ]) · [Dc3P D,Load ] (4.80)
∂{ε3P
c
D,Load
}
∂{Req } ε
= ([E1 ] + [E3 ]) · [Dc p ] (4.81)
∂{εp }
∂{Req } f
= ([E1 ] + [E3 ]) · [Dc p ] (4.82)
∂{fp }
∂{Req }
= ([E1 ] + [E3 ]){bθ12c } · {e1 }T + {e1 } · {gnormal } + {e2 }({a2 } + {gtangent })
∂{p}
(4.83)

The variables [E1 ], [E3 ],{a1 }, {a2 }, {gnormal }, and {gtangent } are

[E1 ] = {e1 } · {e1 }T (4.84)

[E3 ] = {e3 } · {e3 }T (4.85)


∂vc ∂wcr
{a1 } = {e1 }T (4.86)
∂wcr ∂εc1
∂vc ∂vc ∂wcr dScr
{a2 } = {e2 }T + {e1 }T (4.87)
∂δslip ∂wcr ∂Scr dθ12c
Xnp
¡
{gnormal } = − ρi Escri cos2 (φi ){c1i } + (fscri − fsi ){c3i }
i=1
+Edcri cos(φi ) sin(φi ){c2i } + fdcri {c4i }) (4.88)
np
X
{gtangent } = − ρi (Escri sin(φi ) cos(φi ){c1i }
i=1
¢
+(fscri − fsi ){c4i } − Edcri cos2 (φi ){c2i } − fdcri {c3i } (4.89)

{c1i } = cos2 (φi ){e3 }T − 2∆ε1cr cos(φi ) sin(φi ){e1 }T (4.90)

{c2i } = cos(φi ){e2 }T − δslip sin(φi ){e1 }T (4.91)

{c3i } = −2 cos(φi ) sin(φi ){e1 }T (4.92)

{c4i } = (cos2 (φi ) − sin2 (φi )){e1 }T (4.93)


138

Table 4.2: Jacobian Variables

Variable Origin
dScr ∂wcr ∂vc
, , dwcrdScr, dvcdwcr, 1CR
dθ12c ∂εc1 ∂δslip
[P ] 1M E
∂{ν}
1M E
∂{εp }
∂{ν}
1M E
∂{ε3P
c
D,Load
}
[S1 ]to[S9 ] 1M E
[Dc3P D,Load ] 1AR
ε
[Dc p ] 1AR
f
[Dc p ] 1AR
{bθ12c } 1AR

Table 4.2 presents the origin of the variables needed to construct the Jacobian that were
not already shown in Table 4.1. As shown in Tables 4.1 and 4.2, all the variables needed
to construct the residual and the Jacobian are provided by subprocesses 1CR, 1AR, and
1M E.

Subprocess 1RJ-CR. This subprocess determines the crack status of the concrete. If
the concrete principal load strain, εload
c1 , exceeds the crack strain, εccr ,then the concrete is

considered cracked:


1 concrete is cracked εload
c1 > εccr
crackStatus =

0 concrete is uncracked εload
c1 ≤ εccr

(4.94)

If the concrete is cracked (crackStatus = 1), Subprocess 1RJ-CR calls Subprocess 1CR to
compute the variables associated to the cracking region. This subprocess is called as follows:

h dScr ∂wcr ∂wcr ∂vc ∂vc i


Scr , , wcr , , , vc , , , {fdcr }, {Edcr }, RCstate =
dθ12c ∂εc1 ∂Scr ∂δslip ∂wcr
³ ´
crackingRegionM odels θ12c , εc1 , δslip , {∆dcr }, RCstate0 , M atP rop, AnlP ar (4.95)
139

The Subprocess 1CR input variable {∆dcr } is prepared by Subprocess 1RJ-CR using the
following equations:

{∆dcr } = {∆dcr1 , ∆dcr2 , ..., ∆dcrnp }T (4.96)

∆dcri = δslip cos(φi ) (4.97)

{φ} = {φ1 , φ2 , ..., φnp }T (4.98)

φi = θ12c − αi (4.99)

The Subprocess 1CR output variables {fdcr } and {Edcr } are

{fdcr } = {fdcr1 , fcr2 , ..., fcrnp }T (4.100)

{Edcr } = {Edcr1 , Edcr2 , ..., Edcrnp }T (4.101)

If the concrete is uncracked (crackStatus = 0), Subprocess 1RJ-CR uses the following
default values instead of calling Subprocess 1CR:

Scr = 1 (4.102)
dScr ∂wcr ∂wcr ∂vc ∂vc
= wcr = = = vc = = =0 (4.103)
dθ12c ∂εc1 ∂Scr ∂δslip ∂wcr
{fdcr } = {Edcr } = {0, 0, ..., 0}T (4.104)

δslip
The variable Scr value is set to 1 to avoid the indefiniteness of the term which is
Scr
computed as part of the Jacobian, [J]. This term is part of the Jacobian [J]. The actual
δslip
value of is irrelevant when the concrete is uncracked because the terms of the Jacobian
Scr
δslip
[J] affected by are eliminated during the construction of [Jred ].
Scr
140

Subprocess 1RJ-AR. This subprocess calls Subprocess 1AR as follows:

h
{fs }, {Es }, {fscr }, {Escr }, fsz , Esz ,
i
ε f
{σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, RCstate =
³ dScr
averageRegionM odels {εs }, {εscr }, εsz , {ε3P c
D,Load
}, {εp }, {fp }, {φ}, Scr , ,
dθ12c
´
RCstate, RCstate0 , M atP rop, AnlP ar (4.105)

The Subprocess 1AR input variables {εs }, {εscr }, and εsz are prepared by Subprocess
1RJ-CR using the following equations:

{εs } = {εs1 , εs2 , ..., εsnp }T (4.106)

{εscr } = {εscr1 , εscr2 , ..., εscrnp }T (4.107)

εsi = {e1 }T · ([Tstrain (αi )] · {ε}) (4.108)

εscri = εsi + ∆ε1cr cos2 (φi ) (4.109)

(4.110)

Subprocess 1CR output variables {fscr } and {Escr } are analogous to the previously defined
vectors {fs }, and {Es }:

{fscr } = {fscr1 , fscr2 , ..., fscrnp }T (4.111)

{Escr } = {Escr1 , Escr2 , ..., Escrnp }T (4.112)

Subprocess 1RJ-ME. This subprocess calls Subprocess 1M E as follows:

h ∂{ν} ∂{ν} ∂[P ] ∂[P ] ∂[P ] i


, , , , , RCstate =
∂{ε3P
c
D,Load
} ∂{εp } ∂ν1 ∂ν2 ∂νZ
multiaxialLoadingEffectModels({ε3P
c
D
}, {σc3P D }, {ε3P
c
D,Load
}, {εp },
´
RCstate, RCstate0 , M atP rop, AnlP ar (4.113)
141

Notice that Subprocess 1RJ-M E uses vector {σc3P D } found by Subprocess 1AR to complete
the input for Subprocess 1M E.

4.3.3 Subprocess 1CR: Crack Region Models

This subprocess calls the 1D cracking region models. The header of the function associated
with subprocess 1CR is

h dScr ∂wcr ∂wcr ∂vc ∂vc i


Scr , , wcr , , , vc , , , {fdcr }, {Edcr }, RCstate =
dθ12c ∂εc1 ∂Scr ∂δslip ∂wcr
³ ´
crackingRegionModels θ12c , εc1 , δslip , {∆dcr }, RCstate0 , M atP rop, AnlP ar

(4.114)

The Subprocess 1CR is shown in Figure 4.5.


The 1D cracking region models are called as

h dScr i
Scr , = crackSpacingModel(θ12c , M atP rop) (4.115)
dθ12c
h ∂wcr ∂wcr i
wcr , , = crackWidthModel(Scr , εc1 ) (4.116)
∂εc1 ∂Scr
h ∂vc ∂vc i
vc , , = aggregateInterlockModel(wcr , δslip , M atP rop) (4.117)
∂δslip ∂wcr
h i
{fdcr }, {Edcr } = dowelActionModel(∆dcr , δslip , M atP rop) (4.118)

(4.119)

The dowel action model is called for each value ∆dcri associated to the in-plane layer i.
The variable ∆dcri is extracted from the vector {∆dcr }. The scalar results fdcri and Edcri
are stored in vectors {fdcr } and {Edcr }. The 1D cracking models equations necessary to
generated their output were presented in Section 2.6.
142

4.3.4 Subprocess 1AR: Average Region Models

This subprocess calls the 1D average region models. The header of the function associated
with this subprocess is

h i
ε f
{fs }, {Es }, {fscr }, {Escr }, fsz , Esz , {σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, RCstate =
³ dScr
averageRegionModels {εs }, {εscr }, εsz , {ε3P c
D,Load
}, {εp }, {fp }, {φ}, Scr , ,
dθ12c
´
RCstate, RCstate0 , M atP rop, AnlP ar (4.120)

Subprocess 1AR is divided into two subprocesses:

• Subprocess 1AR-M S: Multiple Steel Longitudinal Response

• Subprocess 1AR-M C : Multiple Concrete Loading Direction

Subprocess 1AR and its subprocesses are shown in Figure 4.6. Subprocess 1AR-M S calls
multiple times the longitudinal steel 1D model to obtain the longitudinal stresses for each in-
plane layer (average and at crack), and the single out-of-plane layer. Subprocess 1AR-M C
calls the concrete 1D models multiple times to calculate stresses in the principal directions
fcj (j = 1, 2, z).

Subprocess 1AR-MS. This subprocess calls multiple times the longitudinal steel 1D
model to obtain the longitudinal stresses for each in-plane layer (average and at crack), and
the single out-of-plane layer. This subprocess is called by Subprocess 1AR as:

h i
{fs }, {Es }, {fscr }, {Escr }, fsz , Esz , RCstate =
³ ´
multipleSteel {εs }, {εscr }, εsz , RCstate, RCstate0 , M atP rop, AnlP ar

(4.121)

Subprocess 1ARJ-M S calls the 1D steel longitudinal model as

h i
fs , Es = steelLongitudinalModel(εs , M atP rop, AnlP ar) (4.122)
143

The steel longitudinal model is called for each value εsi and εscri (extracted from the vec-
tors {εs } and {εscr })associated to the in-plane layer i and the out-of-plane strain εsz . The
variables fsi , fscri , Esi , and Escri are stored in the vectors {fs }, {Es }, {fscr }, and {Escr }, re-
spectively. The variables fsz and Esz are returned directly as scalars. The steel longitudinal
1D models equations necessary to find fs and Es are presented in Section 2.5.

Subprocess 1AR-MC. This subprocess calls the concrete 1D models multiple times to
calculate stresses in the principal directions fcj (j = 1, 2, z). This subprocess is called by
Subprocess 1AR as:

h i
ε f
{σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, RCstate =
³ dScr ´
multipleConcrete {ε3P c
D,Load
}, {ε p }, {fp }, {φ}, Scr , , RCstate, RCstate0 , M atP rop
dθ12c
(4.123)

Subprocess 1RJ-M C calls the compressive or tensile 1D concrete models depending in the
value of εload
c associated with the concrete principal stress direction (j =1,2 z). If εload
c <0
concrete is in compression, then output is found using the following equations:

h ∂fc ∂fc i
fc , EcLoad , , = concreteCompM odel(εload
c , εp , fp , M atP rop, AnlP ar)
∂εp ∂fp
(4.124)
∂fc
= 0 (4.125)
∂θ12c

If εload
c ≥ 0 concrete is in tension, then the output is found using the following equations:

h ∂fc ∂fc i dScr


fc , EcLoad , , = concreteT ensM odel(εload
c , {φ}, Scr , , M atP rop, AnlP ar)
∂{φ} ∂Scr dθ12c
(4.126)
∂fc ∂fc
= =0 (4.127)
∂εp ∂fp
np
X
∂fc ∂fc dScr ∂fc
= + (4.128)
∂θ12c ∂Scr dθ12c ∂φi
i=1
144

The concrete 1D models are called for each fcj (j = 1, 2, z). The values of εload
c , εp , fp are
extracted from the vectors {ε3P
c
D,Load
}, {εp } and {fp }, respectively. These scalar output
ε f
variables are stored in the variables {σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], and {bθ12c }. Where
these variables were defined previously as:
 
∂fc1
 ∂εload 0 0 
 c1 
 ∂fc2 
3P D,Load
[Dc ] = 
 0 0 
 (4.129)
 ∂εload
c2 
 ∂fcz 
0 0
∂εload
cz
 ∂f 
c1
0 0
 ∂εp1 
 ∂fc2 
ε  0 
[Dc p ] =  0  (4.130)
 ∂εp2 
 ∂fcz 
0 0
∂εpz
 ∂f 
c1
0 0
 ∂fp1 
 ∂fc2 
f  0 
[Dc p ] =  0  (4.131)
 ∂fp2 
 ∂fcz 
0 0
∂fpz
 

 ∂fc1 


 

 ∂θ12c 
 
{bθ12c } = ∂fc2 (4.132)
 ∂θ12c 
 

 

 0 
 

(4.133)

4.3.5 Subprocess 1M E: Multiaxial Loading Effect Models

This subprocess calls the multiaxial loading effect 1D models. The header of the function
associated with the subprocess is
" #
∂{ν} ∂{ν} ∂[P ] ∂[P ] ∂[P ]
{β σ }, {β ε }, [S1 ]...[S9 ], {ν} , , [P ], , , , RCstate
∂{ε3P
c
D,Load
} ∂{εp } ∂ν1 ∂ν2 ∂νZ

= multiaxialLoadingEffectModels({ε3Pc
D
}, {σc3P D }, {ε3P
c
D,Load
}, {εp },
´
RCstate, RCstate0 , M atP rop, AnlP ar (4.134)
145

Subprocess 1M E is divided into four subprocesses:

• Subprocess 1M E-M CS: Multiple Compression Softening Directions

• Subprocess 1M E-M CF : Multiple Confinement Directions

• Subprocess 1M E-COM B: Compression Softening and Confinement Combination

• Subprocess 1M E-M P : Multiple Poisson Effect Directions

Subprocess 1M E and its subprocesses are shown in Figure 4.7. Subprocesses 1M E-M CS
and 1M E-M CF finds the compression softening and confinement effect in the three princi-
pal concrete stress directions. After this is completed subprocesses 1M E-COM B combines
the compression softening and confinement effects. Finally, Subprocess 1M E-M P finds the
Poisson Modulus in the three principal concrete stress directions. Subprocess 1M E calls
Subprocess 1M E-M CS as:

h i
{βdσ }, {βdε }, [S4 ], [S8 ], RCstate =
³ ´
multipleCompressionSoftening {ε3P c
D
}, RCstate, RCstate0 , M atP rop, AnlP ar

If the concrete is uncracked (crackStatus = 1) or the compression softening effects are


not included (specified in AnlP ar) then Subprocess 1M E uses the following default values
instead of calling Subprocess 1M E-M CS:

{βdσ } = {1, 1, 1}T (4.135)

{βdε } = {1, 1, 1}T (4.136)

[S4 ] = [0] (4.137)

[S8 ] = [0] (4.138)

where [0] is a 3x3 zero matrix. Subprocess 1M E calls Subprocess 1M E-M CF as:

h i
{βlσ }, {βlε }, [S5 ], [S9 ], RCstate =
³ ´
multipleConfinement {σc3P D }, RCstate, RCstate0 , M atP rop, AnlP ar
146

If the confinement effect is not included (specified in AnlP ar) then Subprocess 1M E uses
the following default values instead of calling Subprocess 1M E-M CF :

{βlσ } = {1, 1, 1}T (4.139)

{βlε } = {1, 1, 1}T (4.140)

[S5 ] = [0] (4.141)

[S9 ] = [0] (4.142)

Subprocess 1M E calls Subprocess 1M E-COM B as:

h i
{β σ }, {β ε }, [S1 ], [S2 ], [S3 ], [S6 ], [S7 ], RCstate =
³
multipleCSandCFcombination {ε3P c
D,Load
}, {βdε }, {βlε }, {βdσ }, {βlσ }, RCstate,
´
RCstate0 , M atP rop, AnlP ar (4.143)

where Subprocess 1M E-COM B combines the compression softening(CS) and confinement


effect(CF). Subprocess 1M E calls Subprocess 1M E-M P as:
" #
∂{ν} ∂{ν} ∂[P ] ∂[P ] ∂[P ]
{ν} , , [P ], , , , RCstate =
∂{ε3P
c
D,Load
} ∂{εp } ∂ν1 ∂ν2 ∂νZ
³ ´
multipleP oisson {ε3P c
D,Load
}, {εp }, RCstate, RCstate0 , M atP rop, AnlP ar

If the Poisson effect is not included (specified in AnlP ar) then Subprocess 1M E uses the
following default values instead of calling Subprocess 1M E-M P :

{ν} = {1, 1, 1}T (4.144)

[P ] = [I] (4.145)
∂{ν} ∂{ν} ∂[P ] ∂[P ] ∂[P ]
3P D,Load
= = = = = [0] (4.146)
∂{εc } ∂{εp } ∂ν1 ∂ν2 ∂νZ

where [I] is a 3 × 3 identity matrix.


147

Table 4.3: Compression Softening Strains

Direction εc εperp
c βdε βdσ
ε σ
1sc εc1 εc2 βd1 βd1
ε σ
2sc εc2 εc1 βd2 βd2
ε σ
z εcz εc1 βdz βdz

Subprocess 1ME-MCS. This subprocess finds the compression softening effects in the
three principal concrete stress directions. The header of the function associated with this
subprocess is

h i
{β σ }, {β ε }, [S4 ], [S8 ], RCstate =
³ ´
multipleCompressionSof tening {ε3P c
D
}, RCstate, RCstate0 , M atP rop, AnlP ar

(4.147)

Subprocess 1M E-M CS calls the compression softening 1D model as:

h ∂β ε ∂β σ ∂βdε ∂βdσ i ³ ´
βdε , βdσ , d , d , perp , perp = compressionSof teningM odel εc , εperp
c , M atP rop
∂εc ∂εc ∂εc ∂εc
(4.148)

The 1D model is called for each concrete principal direction. Table 4.3 shows what com-
bination of strains used to obtain the output in the three principal concrete directions.
148

The scalar output variables are stored in variables {βdσ }, {βdε }, [S4 ], [S8 ]:

{βdε } = {βd1
ε ε
, βd2 ε T
, βdz } (4.149)

{βdσ } = {βd1
σ
, β σ , β σ }T (4.150)
 ε d2 dzε 
∂βd1 ∂βd1
 ∂εc1 ∂εc2 0 
 ∂β ε ∂βd1ε 
 
[S4 ] =  d2 0  (4.151)
 ∂εc1 ∂εc2 
 ∂β ε ε 
∂βdz
dz
0
∂εc1 ∂εcz
 σ σ 
∂βd1 ∂βd1
 ∂εc1 ∂εc2 0 
 ∂β σ ∂β σ 
 
[S8 ] =  d2 d1
0  (4.152)
 ∂εc1 ∂εc2 
 ∂β σ ∂β σ 
dz dz
0
∂εc1 ∂εcz

Subprocess 1ME-MCF. This subprocess finds the compression softening and confine-
ment effect in the three principal concrete stress directions. The header of the function
associated with this subprocess is

h i
{βlσ }, {βlε }, [S5 ], [S9 ], RCstate =
³ ´
multipleConf inement {σc3P D }, RCstate, RCstate0 , M atP rop, AnlP ar

Subprocess 1M E-M CF calls the confinement 1D model as:

h ∂βlε ∂βlσ ∂βlε ∂βlσ i ³ ´


βlε , βlσ , min , min , max , max = conf inementM odel fcmin , fcmax , M atP rop
∂fc ∂fc ∂fc ∂fc
(4.153)

The 1D model is call for each concrete principal direction. Table 4.4 shows the combination
of confining stresses used to obtain the output in the three principal concrete directions.
The scalar output variables are stored in variables {βlσ }, {βlε }, [S5 ], [S9 ]:
149

Table 4.4: Confinement Stresses


Direction fcmin fcmax βlε βlσ
ε σ
1sc min(fc2 , fcz ) max(fc2 , fcz ) βl1 βl1
ε σ
2sc min(fc1 , fcz ) max(fc1 , fcz ) βl2 βl2
ε σ
z min(fc1 , fc2 ) max(fc1 , fc2 ) βlz βlz

{βlε } = {βl1
ε ε
, βl2 ε T
, βlz } (4.154)

{βlε } = {βl1
σ σ
, βl2 σ T
, βlz } (4.155)
 ε ε 
∂βl1 ∂βl1
0
 ∂f c2 ∂fcz 
 ∂β ε ε 
∂βl2
 l2 
[S5 ] =  0  (4.156)
 ∂fc1 ∂fcz 
 ∂β ε ∂β ε 
lz lz
0
∂fc1 ∂fc2
 σ σ 
∂βl1 ∂βl1
0
 ∂fc2 ∂fcz 
 ∂β σ σ 
∂βl2
 
[S9 ] =  l2 0  (4.157)
 ∂fc1 ∂fcz 
 ∂β σ σ
∂βlz 
lz
0
∂fc1 ∂fc2

Subprocess 1ME-COMB. This subprocess combines the compression softening and


confinement effects. The header of the function associated with this subprocess is

h i
{β σ }, {β ε }, [S1 ], [S2 ], [S3 ], [S6 ], [S7 ], RCstate =
³
multipleCSandCF combination {ε3P c
D,Load
},
´
{βdε }, {βlε }, {βdσ }, {βlσ }, RCstate, RCstate0 , M atP rop, AnlP ar

(4.158)
150

The output variables {β σ }, [S6 ], and [S7 ], which are associated with the modification of the
0
peak compression strength, fc , are defined as

{β σ } = {βd1
σ σ σ σ
βl1 , βl2 σ σ T
βd2 , βlz βdz } (4.159)
 
βσ 0 0
 l1 
 σ 
[S6 ] =  0 βl2 0 (4.160)
 
0 0 βlz σ
 
σ
βd1 0 0
 
 σ 
[S7 ] =  0 βd2 0  (4.161)
 
0 0 βdz σ

(4.162)

The output variables {β ε }, [S1 ], and [S2 ], and [S3 ], which are associated with the modifica-
tion of the strain at peak, εco , are defined as

{β ε } = {β1ε , β2ε , β3ε , }T (4.163)


 
∂β1ε
 ∂β ε 0 0 
 d1 ε 
 ∂β 
[S1 ] = 
 0 ε
2
0   (4.164)
 ∂βd2 
 ε
∂βz 
0 0 ε
∂βdz
 ε 
∂β1
 ∂β ε 0 0 
 l1 

 ∂β2ε 
[S2 ] =  0 ε 0   (4.165)
 ∂βl2 
 ∂βzε 
0 0 ε
∂βlz
 
∂β1ε
 ∂εload 0 0 
 c1 
 ∂β2ε 

[S3 ] =  0 0  (4.166)
∂εload 
 c2 ε

 ∂βz 
0 0
∂εload
cz
(4.167)
151

The scalars β1ε , β2ε , and β3ε , which defined the strain at peak, εco , modification factors, are
calculated using the following equation:



βdε βlε
 βdε βlε ≥ 1 or εLoad
c ≥ βdε βlε εco

 Load
ε εc
β = εc < βdε βlε εco (4.168)

 εco



1 εLoad
c < εco

∂β1ε ∂β2ε ∂βzε


The scalars ε , ε , and ε , are calculated using the following equation:
∂βd1 ∂βd2 ∂βdz




βlε βdε βlε ≥ 1 or εLoad
c ≥ βdε βlε εco

∂β ε 
= 0 εLoad < βdε βlε εco (4.169)
∂βdε 

c



0 εLoad
c < εco

∂β1ε ∂β2ε ∂βzε


The scalars ε , ε , and ε are calculated using the following equation:
∂βl1 ∂βl2 ∂βlz


 ε

βd βdε βlε ≥ 1 or εLoad
c ≥ βdε βlε εco


∂β ε
= 0 εLoad < βdε βlε εco (4.170)
∂βlε 
c



0 εLoad
c < εco

∂β1ε ∂β2ε ∂βzε


The scalars , , and are calculated using the following equation:
∂εload
c1 ∂εload
c2 ∂εload
cz




βdε
 βdε βlε ≥ 1 or εLoad
c ≥ βdε βlε εco


∂β ε 1
= εLoad
c < βdε βlε εco (4.171)
∂εLoad
c 
 ε co



0 εLoad
c < εco

The above equations are used for each concrete principal direction. Table 4.5 shows the
combination of variables used
152

Table 4.5: β ε variables

Direction βlε βdε βε


ε ε
1sc βl1 βd1 β1ε
ε ε
2sc βl2 βd2 β2ε
ε ε
z βlz βdz βzε

Subprocess 1ME-MP. This subprocess finds the Poisson Modulus in the three principal
concrete stress directions. The header of the function associated with this subprocess is
" #
∂{ν} ∂{ν} ∂[P ] ∂[P ] ∂[P ]
{ν} , , [P ], , , , RCstate =
∂{ε3P
c
D,Load
} ∂{εp } ∂ν1 ∂ν2 ∂νZ
³ ´
multipleP oisson {ε3P c
D,Load
}, {εp }, RCstate, RCstate 0 , M atP rop, AnlP ar

Subprocess 1M E-M P calls the compressive or tensile 1D Poisson models depending on the
value of εload
c . If εload
c < 0 concrete is in compression and the compressive Poisson model is
used:

h ∂ν ∂ν i
ν, , = poissonCompM odel(εload
c , εp , M atP rop) (4.172)
∂εLoad
c ∂ε p

If εload
c ≥ 0 concrete is in tension and the tensile Poisson model is used:

h ∂ν i
ν, = poissonT ensM odel(εload
c , M atP rop) (4.173)
∂εLoad
c
∂ν
= 0 (4.174)
∂εp

The Poisson 1D models are called for each principal concrete direction. The εload
c and εp are
extracted from the vectors {ε3P
c
D,Load
} and {εp }, respectively. Table 4.6 shows the variables
used to calculate each Poisson modulus.
∂{ν} ∂{ν}
The scalar output variables are stored in the variables {ν}, , and .
∂{ε3P
c
D,Load
} ∂{εp }
153

Table 4.6: ν variables


Direction εload
c εp ν
1sc εload
c1 εp1 ν1
2sc εload
c2 εp2 ν2
z εload
cz εpz νZ

These variables were defined previously as:

{ν} = {ν1 , ν2 , νZ }T (4.175)


 
∂ν1
 ∂εLoad 0 0 
 c1 
∂{ν}  ∂ν2 
=  0 0  (4.176)
3P D,Load  ∂ε Load 
∂{εc }  c2 
 ∂νz 
0 0
∂εLoad
cz
 ∂ν 
1
0 0
 ∂εp1 
∂{ν}  ∂ν2 
 0 
=  0  (4.177)
∂{εp }  ∂εp2 
 ∂νz 
0 0
∂εpz
(4.178)

∂[P ] ∂[P ] ∂[P ]


The matrices [P ], , , and are constructed using the following equations:
∂ν1 ∂ν2 ∂νZ
 
1 − νZ ν2ν2 (1 + νZ ) νZ (ν2 + 1)
1



[P ] = ν1 (1 + νZ ) 1 − ν1 νZ νZ (1 + ν1 ) (4.179)
a 
ν1 (ν2 + 1) ν2 (1 + ν1 ) 1 − ν1 ν2
a = 1 − νZ ν2 − ν1 ν2 − 2ν1 νZ − ν1 νZ (4.180)

where the Poisson matrix [P ] is used to find the load concrete strains from the total concrete
strains.
154

Figure 4.3: Subprocess 1NLS


155

Figure 4.4: Subprocess 1RJ


156

Figure 4.5: Subprocess 1CR


157

Figure 4.6: Subprocess 1AR


158

Figure 4.7: Subprocess 1ME


159

4.4 Process 2: Concrete and Steel Material Stiffness Matrices and Stress Vec-
tors

Process 2 is responsible for generating the concrete and steel material stiffness and stress
vectors. The header of the function associated with this process is:

h i ³
{σs }, [Ds ], {σc }, [Dc ] = getSteelConcStressStiff {fs }, {Es },
´
ε f
{σc3P D }, [Dc3P D,Load ], [Dc p ], [Dc p ], {bθ12c }, θ12c , RCstatus (4.181)

Process 2 is divided into two subprocesses:

• Subprocess 2C: Concrete material stiffness and stress vector

• Subprocess 2S: Steel material stiffness and stress vector

Figure 4.8 shows Process 2 and its subprocesses. If the solution from Process 2 is not stable
(solutionStatus = 0), Process 2 uses the following default values

{σs } = {σc } = {0, 0, 0}T (4.182)

[Ds ] = [Dc ] = [0] (4.183)

where [0] is a 3x3 matrix.

4.4.1 Subprocess 2C: Concrete material stiffness and stress vector

Subprocess 2C finds the concrete stress vector {σc } using the following equations obtained
in Section 3.3:

{σc } = [Tstress (−θ12c )] · {σc }12c (4.184)

{σc }12c = [R2D ] · {σc3P D } (4.185)


 
cos2 (ω) sin2 (ω) 2 cos(ω) sin(ω)
 
 2 2 
[Tstress (ω = −θ12c )] =  sin (ω) cos (ω) −2 cos(ω) sin(ω) (4.186)
 
2 2
− cos(ω) sin(ω) cos(ω) sin(ω) cos (ω) − sin (ω)
160

Figure 4.8: Process 2: Concrete and Steel Material Stiffness Matrices and Stress Vectors
161

Subprocess 2C finds the concrete material stiffness matrix [Dc ] using the following equations
obtained in Section 3.4:

[Dc ] = −[B] · [J]−1 · [E ε ] (4.187)


0
[B] = −[Tstress (−θ12c )] · {σc }12c · {e1 }T · [E p ] + [Tstress (−θ12c )] · [A] (4.188)
³
ε
[A] = [R2D ] · [Dc3P D,Load ] · [E 3P D,Load ] + [Dc p ] · [E εp ]
´
f
+ [Dc p ] · [E fp ] + {bθ12c } · {e1 }T · [E p ] (4.189)

where matrices [E ε ], [E 3P D,Load ], [E fp ], and [E p ] are

[E ε ] = [ [I], [0], [0], [0], [0] ]T (4.190)

[E 3P D,Load ] = [ [0], [I], [0], [0], [0] ] (4.191)

[E εp ] = [ [0], [0], [I], [0], [0] ] (4.192)

[E fp ] = [ [0], [0], [0], [I], [0] ] (4.193)

[E p ] = [ [0], [0], [0], [0], [I] ] (4.194)

and matrices [I] and [0] are the 3 × 3 identity and zero matrix. Matrices [Tstress (−θ12c )]
0
and [Tstress (−θ12c )] are associated with the transformation of the stresses from coordinate
system to another, and are provided by the auxiliary function transf ormation. If the status
of the concrete is uncracked (crackStatus = 0), then the internal RC material problem is
reduced from 15 to 13 equations. This also requires the modification of the Jacobian matrix,
[J], from a 15 × 15 size to a reduced 13 × 13 Jacobian matrix, [J]red . When the concrete is
uncracked then, the matrix [Dc ] calculated as:

[Dc ] = −[B]red · [J]−1 ε


red · [E ]red (4.195)
0
[B]red = −[Tstress (−θ12c )] · {σc }12c · {e1 }T · [E p ]red + [Tstress (−θ12c )] · [A]

(4.196)
³
ε
[A]red = [R2D ] · [Dc3P D,Load ] · [E 3P D,Load ]red + [Dc p ] · [E εp ]red
´
f
+ [Dc p ] · [E fp ]red + {bθ12c } · {e1 }T · [E p ]red (4.197)
162

where [E 3P D,Load ]red , [E εp ]red , [E fp ]red , [E p ]red are the same as [E 3P D,Load ], [E εp ], [E fp ],
[E p ] but with columns 14 and 15 extracted. Matrix [E ε ]red is the same as matrix [E ε ] but
with rows 13 and 14 extracted.

4.4.2 Subprocess 2S: Steel material stiffness and stress vector

Subprocess 2S finds the concrete stress vector {σs } using the following equations obtained
in Section 3.3:

np
X
{σs } = {σsi } (4.198)
i=1
{σsi } = [Tstress (−αi )] · {σsi }αi (4.199)

{σsi }αi = {ρi fsi , 0, 0}T (4.200)

where fsi is stored as the ith component of vector {fs }. Subprocess 2S finds the steel
material stiffness matrix [Ds ] using the following equations obtained in Section 3.4:

[Dsi ] = [Tstif f (−αi )]T · [Dsi ]αi · [Tstif f (−αi )] (4.201)


 
ρE 0 0
 i si 
αi  
[Dsi ] =  0 0 0 (4.202)
 
0 0 0
 
cos2 (−αi ) sin2 (−αi ) − cos(−αi ) sin(−αi )
 
 2 2 
[Tstif f (−αi )] =  sin (−αi ) cos (−αi ) cos(−αi ) sin(−αi ) 
 
2 2
2 cos(−αi ) sin(−αi ) −2 cos(−αi ) sin(−αi ) cos (−αi ) − sin (−αi )
(4.203)

where Esi is stored as the ith component of vector {Es }. The angles αi and the steel ratios
ρi are provided by the input variable M atP rop.
163

Figure 4.9: Process 3

4.5 Process 3: RC material stiffness matrix and stress vector

This process is responsible for assembling the concrete and steel contributions to the RC
stress vector and material stiffness matrix. The header of the function associated with this
process is:

h i ³ ´
{σ}, [D] = getRCStressStiff {σs }, [Ds ], {σc }, [Dc ] (4.204)

The process finds the RC stress vector and material stiffness by adding that of the concrete
and the steel:

{σ} = {σc } + {σs } (4.205)

[D] = [Dc ] + [Ds ] (4.206)

Figure 4.9 shows Process 3.


164

4.6 Summary

Chapter 4 presents the implementation of the theory developed in Chapter 3. The imple-
mentation of the DSFM-TG formulation consists of Matlab code that accomplishes two main
tasks. The first task is to solve the internal RC material state problem for a given strain
state (Process 1). The second task is to assemble the the RC stress vector and material
stiffness matrix (Processes 2 and 3).
Process 1 calls the 1D models, several times and in different directions for the case of
concrete and for different layers in the case of the steel. The 1D models return non-derivative
(stress for instance) and derivative variables (stiffness for instance). This scalar output is
assemble into vectors and matrices as a way to condensed the large number of variables.
Ultimately, all these vectors and matrices are assembled into the residual, {R}, and the
Jacobian, J. Even though this condensation process makes the solution algorithm easier to
implement and understand, it is still quite complicated. A good portion of the complexity
of the implementation comes from the use of 1D models for a 3D material.
During the execution of Process 1, the Jacobian matrix [J] could become singular. A
singular Jacobian is a reflect of a loss of stiffness at the crack which makes it impossible to
balance the stresses in the average region. The loss of stiffness can occur due to flat yielding
or fracture of the bars crossing the crack. When this occurs the internal problem cannot be
solved and a non-solved problem flag is returned.
The coding of the implementation presented in this Chapter was carried out in Matlab,
and it is presented in Appendix .
The main aspects of the implementation are condensed as follows

• The implementation is divided into three processes: Process 1 finds the internal RC
material process; Process 2 finds the concrete and steel stress vectors and material
stiffness matrices; and Process 3 assembles and obtains the RC stress vector and
stiffness matrices by assembling the concrete and steel contributions.

• Process 1 is the most time consuming since it is the only one that requires an iterative
algorithm (Newton) to generate its output.
165

• Processes 2 and 3 use explicit expressions to generate their output, which makes them
faster than Process 1 to execute.

• The internal RC problem cannot be solved when the Jacobian matrix, [J], used in the
system of equations becomes singular. This singularity of the Jacobian is associated
with the loss of stiffness due to steel failure at the crack.
166

Chapter 5

MODEL EVALUATION AND VALIDATION

5.1 Introduction

The purpose of the research presented in this Chapter is to evaluate and validate the DSFM-
TG formulation and implementation by performing a series of tests on it. Three type of
tests are performed as part of the evaluation and validation process:

• 1D model tests

• RC material model tests

• RC material model in FEM analysis tests

The tests on the 1D models are intended to detect and correct mistakes in the equations used
in the implementation of DSFM-TG . The tests on the RC material evaluate the numerical
(i.e convergence and stability of the solution) and physical aspects of the DSFM-TG . The
last group of tests applies the RC material model within an FEM analysis and compares
results simulated using the model with experimental results and Vector2 results for a series
of RC panels tested by Vecchio and Collins (1986).

5.2 1D model tests

Tests are performed on each of the 1D models in the DSF M −T G as well as the models that
define parameters that modify the 1D models. Special attention is given to the derivatives
returned by the 1D models as these can have a significant effect on the solution of the
internal RC state problem as well as the external equilibrium problem . The derivatives are
checked by comparting the curves obtained directly from the 1D model with that obtained
using only the derivatives and the Euler’s method. The 1D models tested are

• Crack spacing model: Basic crack spacing model (Vecchio (2000))


167

• Crack width model: Basic crack width model (Vecchio (2000))

• Aggregate interlock model: Walraven model (Walraven (1981))

• Dowel action model: He-Kwan model (He-Kwan (?))

• Longitudinal Steel 1D models: Bilinear, Trilinear, and Raynor (Raynor (2000))model

• Compressive Concrete 1D models: Hognestad , and PopovicHS (Popovic (1973))model

• Tensile Concrete 1D models: Vecchio1982 (Vecchio and Collins (1986)), and LNR-
Bentz model (Vecchio (2000))

• Compression softening 1D models: Vecchio 1986 (Vecchio and Collins (1986)), Vecc-
chio 1992A (Vecchio (2000))

• Confinement 1D model: Kupfer-Richart model (Kupfer (1969), Richart (1928))

• Poisson compressive 1D model: Variable Poisson’s ratio-Kupfer (Kupfer (1973))

• Poisson tensile 1D model: Linear Poisson Model (Vector2 User’s Manual (2002))

5.2.1 Crack spacing model

The basic crack spacing model determines the crack spacing, Scr , based on the angle of
orientation of the crack with respect to the x-axis, θ12c , and spacing in the x and y directions,
Smx and Smy . The basic crack spacing model was tested for the cases shown in Table 5.1.

Table 5.1: Basic crack spacing model test cases

Case Model θ12c Smx Smy


[deg] [mm] [mm]
1 Basic -360 : 360 50 50
2 Basic -360 : 360 40 60

The idea Figure 5.1 shows that Scr matches Smx and Smy at horizontal and vertical
dScr dScr
orientations of the angle θ12c . Figure 5.2(a) shows the derivative . When is
dθ12c dθ12c
integrated over the θ12c domain an approximated Screuler is obtained. The difference between

euler is negligible as shown in Figure 5.2(b), which indicates that the expression
Scr and Scr
168

dScr dScr
used for is correct. An error in the expression used for the derivative for would
dθ12c dθ12c
euler considerably different from the S curve.
have results in a Scr cr

Case 1: S = 50mm, S == 50 mm
mx my
Case 2: S = 40mm, S == 60 mm
mx my

−315 −270 −225 −180 −135 −90 −45 0 45 90 135 180 225 270 315 360
θ [Deg]
12c

Figure 5.1: Basic Crack Spacing Model, Scr


169

−300 −200 −100 0 100 200 300


θ [Deg]
12c

(a)

Case 1: S = 50mm, S == 50 mm
mx my
Case 1 − Euler
Case 2: S = 40mm, S == 60 mm
mx my
Case 2 − Euler
−300 −200 −100 0 100 200 300
θ [Deg]
12c

(b)

euler and dScr


Figure 5.2: Basic Crack Spacing Model, Scr
dθ12c
170

5.2.2 Crack width model

The crack width model determines the crack width, wcr , as the product of the crack spacing,
Scr , and the maximum principal concrete strain, εc1 . The basic crack width model was tested
for the cases shown in Table 5.2.

Table 5.2: Basic crack width model test cases

Case Model εc1 Scr


[mm/mm] [mm]
1 Basic 0-2(10)−3 40
−3
2 Basic 0-2(10) 50

Figure 5.3 shows the expected linear relationship between wcr and εc1 defined by the
basic cracking model (wcr = Scr εc1 ). Figures 5.3(a) and (b) show that the derivatives
∂wcr ∂wcr
returned by the model, ( = Scr and = εc1 , are correct.
∂εc1 ∂Scr
171

60

55

d w / d ε [mm]
50

c1
45
cr
40

35

30
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−3
c1 c1

(a) (b)

Case 1: S = 40mm
cr
Case 2: S = 50mm
cr

0.5 1 1.5 2
ε [mm/mm] −3
x 10
c1

(c)

∂wcr ∂wcr
Figure 5.3: Basic Crack width model, wcr , , and
∂εc1 ∂Scr
172

5.2.3 Aggregate interlock model

The Walraven model determines the aggregate interlock stress at the crack vc , from the
crack wdith, wcr , the slip, δslip , and the cubic concrete strength, fcc . The Walraven model
was tested for the cases shown in Table 5.3.

Table 5.3: Walraven model test cases

Case Model wcr δslip fcc


[mm] [mm] [MPa]
1 Walraven 0-2 1 -20
2 Walraven 0-2 2 -20
3 Walraven 0-2 1 -40
4 Walraven 0-2 2 -40
5 Walraven 1 0-2 -20
6 Walraven 2 0-2 -20
7 Walraven 1 0-2 -40
8 Walraven 2 0-2 -40

∂vc ∂vc
Figure 5.4 shows cases 1 to 4. Figure 5.5 shows , , and vceuler under a variable
∂wcr ∂δslip
∂vc
wcr . The variable vceuler is obtained by integrating over the wcr domain with δslip
∂wcr
constant. The difference between vc and vceuler shown in Figure 5.5(c) is minimal, which
∂vc
indicates that the expression used for is correct.
∂wcr
∂vc ∂vc
Figure 5.6 shows cases 4 to 8. Figure 5.7 shows , , and vceuler under a variable
∂δslip ∂wcr
∂vc
δslip . The variable vceuler is obtained by integrating over the δslip domain with wcr
∂δslip
constant. The difference between vc and vceuler shown in Figure 5.7(c) is minimal, which
∂vc
indicates that the expression used for is correct.
∂δslip
173

Case 1:δ =1 mm,f =20 MPa


slip cc

Case 2:δslip=2 mm,fcc=20 MPa

Case 3:δslip=1 mm,fcc=40 MPa

Case 4:δslip=2 mm,fcc=40 MPa

0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2


wcr [mm]

Figure 5.4: Walraven model, vc under variable wcr

60
∂ vc/ ∂ δslip [MPa/mm]

50
40
30
20
10
0
0.2 0.4 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
w [mm] w [mm]
cr
cr
(a) (b)

Case 1:δ =1 mm,f =20 MPa


slip cc

Case 1−Euler
Case 2:δ =2 mm,f =20 MPa
slip cc

Case 2−Euler
Case 3:δ =1 mm,f =40 MPa
slip cc

Case 3−Euler
Case 4:δslip=2 mm,fcc=40 MPa

Case 4−Euler

0.2 0.25 0.3 0.35 0.4 0.45


wcr [mm]
(c)

Figure 5.5: Walraven model, vceuler and derivatives under variable wcr
174

Case 5:wcr=1 mm,fcc=20MPa

Case 6:wcr=2 mm,fcc=20MPa

Case 7:wcr=1 mm,fcc=40MPa

Case 8:wcr=2 mm,fcc=40MPa

0.5 1 1.5 2
δslip [mm]

Figure 5.6: Walraven model, vc under variable δslip

0
−2
∂ v /∂ w [MPa/mm]

−4
−6
−8
cr

−10
−12
c

−14
−16
−18
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
δ [mm] δ lip [mm]
slip s

(a) (b)

Case 5:w =1 mm,f =20MPa


cr cc
Case 5−Euler
Case 6:w =2 mm,f =20MPa
cr cc
Case 6−Euler
Case 7:w =1 mm,f =40MPa
cr cc
Case 7−Euler
Case 8:w =2 mm,f =40MPa
cr cc
Case 8−Euler

0.5 1 1.5 2
δ [mm]
slip

(c)

Figure 5.7: Walraven model,vceuler and derivatives under variable δslip


175

5.2.4 Dowel action model

The He-Kwan model dowel action model determines the dowel stress, fdcr , from the dowel
deformation at the crack ∆dcr . The He-Kwan model was tested for the cases shown in Table
5.4.

Table 5.4: He-Kwan dowel model test cases


0
Case Model ∆d c1 db fc εsy fsy
[mm] [mm] [MPa] [mm/mm] [MPa]
1 He-Kwan 0-2.5 0.6 5 -20 2.00E-03 400
2 He-Kwan 0-2.5 0.8 5 -20 2.00E-03 400
3 He-Kwan 0-2.5 0.6 8 -20 2.00E-03 400
4 He-Kwan 0-2.5 0.6 5 -30 2.00E-03 400
5 He-Kwan 0-2.5 0.6 5 -20 2.50E-03 500

dfdcr
Figure 5.8 shows the results of the cases in Table 5.4. Figure 5.9 shows Edcr =
d∆dcr
euler . The variable f euler is obtained by integrating E
and fdcr dcr dcr over the ∆dcr domain. The
euler is minimal, which indicates that the expression used for
difference between fdcr and fdcr
Edcr is correct.
176

Case 1: c1=0.6,db=5mm,f’ c = −20MPa,εsy=2e−3,fy = 400MPa


Case 2: c1=0.8,db=5mm,f’ c = −20MPa,εsy=2e−3,fy = 400MPa
Case 3: c1=0.6,db=8mm,f’ c = −20MPa,εsy=2e−3,fy = 400MPa
Case 4: c1=0.6,db=5mm,f’ c = −30MPa,εsy=2e−3,fy = 400MPa
Case 5: c1=0.6,db=5mm,f’ c = −20MPa,εsy=2.5e−3,fy = 500MPa

0.5 1 1.5 2 2.5


∆dcr[mm/mm]

Figure 5.8: He-Kwan model

Case 1
Case 1−Euler
Case 2
0.5 1 1.5 2 2.5 Case 2−Euler
∆dcr [mm]
Case 3
(a) Case 3−Euler
Case 4
Case 4−Euler
Case 5
Case 5−Euler

0.5 1
∆dcr [mm]
(b)

Figure 5.9: He-Kwan model, Ed and fdeuler


177

5.2.5 Longitudinal Steel 1D models

The longitudinal steel 1D models calculate the longitudinal steel stress, fs , from the longi-
tudinal steel strain, εs . The longitudinal steel 1D models were tested for the cases shown
in Table 5.5.

Table 5.5: Longitudinal steel model test cases

Case Model εy fy εsh fsh εsu fu nsh


[mm/mm] [MPa] [mm/mm] [MPa] [mm/mm] [MPa]
1 Bilinear 2.00E-03 200 - - - - -
2 Trilinear 2.00E-03 200 1.00E-02 - 1.50E-02 220 -
3 Raynor 2.00E-03 200 1.00E-02 205 1.50E-02 220 1
4 Raynor 2.00E-03 200 1.00E-02 205 1.50E-02 220 3
5 Raynor 2.00E-03 200 1.00E-02 205 1.50E-02 220 6

Figure 5.10 shows the fs vs. εs plots for the cases presented in Table 5.5. The cases
presented were already discussed in Section 2.5 to impact of using different models and
fs
material parameters. Figure 5.11 shows Es = and fseuler . The variable fseuler is obtained
εs
by integrating Es over the εs domain. The difference between fs and fseuler is minimal,
which indicates that the expressions used for Es is correct.
178

250

200

150
Case 1: Bilinear
Case 2: Trilinear
Case 3: Raynor, nsh = 1
Case 4: Raynor, nsh = 3
100 Case 5: Raynor, n = 6
sh

50

0
0 0.005 0.01 0.015
ε [mm/mm]
s

Figure 5.10: Longitudinal steel models, fs


179

15000

10000

5000

Case 1: Bilinear
0
Case1−Euler
Case 2: Trilinear
0.01 0.012 0.014
εs [mm/mm] Case1−Euler
Case 3: Raynor, nsh = 1
(ab) Case 3−Euler
Case 4: Raynor, nsh = 3

220 Case 4−Euler


Case 5: Raynor, nsh = 6
215 Case 5−Euler

210

205

200

0.01 0.012 0.014


εs [mm/mm]

(b)

Figure 5.11: Longitudinal steel models, Es and fseuler


180

5.2.6 Compressive Concrete 1D model

The compressive concrete 1D models determine the compressive stress, fc < 0, from the
compressive strain, εc < 0, and the modified peak stress parameters, fp and εp . The
compressive concrete 1D model cases are shown in Table 5.6.

Table 5.6: Compression 1D models test cases

Case Model εc εp fp
[mm/mm] [mm/mm] [MPa]
1 PopovicHS [0,−7(10)−3 ] -0.003 -10
−3
2 PopovicHS [0,−7(10) ] -0.003 -30
3 PopovicHS [0,−7(10)−3 ] -0.003 -50
−3 −3
4 PopovicHS -0.003 [−2(10) ,−7(10) ] -30
5 PopovicHS -0.003 -0.003 -10 to -50
6 Hognestad [0,−7(10)−3 ] -0.003 -10
−3
7 Hognestad [0,−7(10) ] -0.003 -30
8 Hognestad [0,−7(10)−3 ] -0.003 -50
−3 −3
9 Hognestad -0.003 [−2(10) ,−7(10) ] -30
10 Hognestad -0.003 -0.003 -10 to -50

Figure 5.12 shows cases 1, 2, 3, 6, 7, and 8. These cases show the variation of fc for a
variable εc and a constant εp and fp . These cases were already discussed in Section 2.5 to
illustrate the impact of different different models and parameters. Figure 5.13 shows Ec ,
∂fc ∂fc
fceuler , , and for a variable εc . The variable fceuler is obtained by integrating Ec over
∂εp ∂fp
the εc domain with εp and fp constant. The difference between fc and fceuler is minimal,
which indicates that the expressions used for Ec are correct.
Figure 5.14 shows cases 4 and 9. These cases show the variation of fc for a variable
∂fc euler ∂fc
εp and a constant εc and fp . Figure 5.15 shows ,fc , Ec , and , for a variable
∂εp ∂εp
∂fc
εp . The variable fceuler is obtained by integrating over the εp domain (other variable
∂εp
are constant). The difference between fc and fceuler is minimal, which indicates that the
181

∂fc
expressions used for are correct.
∂εp
0
Figure 5.16 shows cases 5 and 10. These cases show the variation of fc for a variable fp
∂fc euler ∂fc
and constant εc and εp . Figure 5.17 shows ,fc , Ec , and , for a variable fp . The
∂fp ∂εp
∂fc
variable fceuler is obtained by integrating over the fp domain with εc and εp constant.
∂fp
The difference between fc and fceuler is minimal, which indicates that the expressions used
∂fc
for are correct.
∂fp
182

−5

−10

−15

−20

−25

−30

−35 Case 1:PopovicHS


Case 2:PopovicHS
−40 Case 3:PopovicHS
Case 6:Hognestad
Case 7:Hognestad
−45
Case 8:Hognestad

−50
−7 −6 −5 −4 −3 −2 −1 0
ε [mm/mm] −3
x 10
c

Figure 5.12: Compressive concrete model, fc under variable εc


183

−10

−20
fc [MPa]

−30

−40

−50

−60
−8 −6 −4 −2 0
−5 −4 −3 −2 −1 0 ε [mm/mm] −3
ε [mm/mm] −3 c x 10
c x 10 (b)
(a)

1.2 Case 1

1 Case 1−Euler
Case 2
0.8
Case 2−Euler
0.6
Case 3
0.4
p

Case 3−Euler
∂f /∂f

0.2
Case 6
c

0
Case 6−Euler
−0.2 Case 7
−0.4 Case 7−Euler
−0.6 Case 8
−0.8 Case 8−Euler
−5 −4 −3 −2 −1 0 −7 −6 −5 −4 −3 −2 −1 0
εc [mm/mm] −3
x 10 εc [mm/mm] x 10
−3

(c) (d)

Figure 5.13: Compressive concrete model, fceuler and derivatives under variable εc
184

−22

Case 4:PopovicHS,fp= −30 MPa,εc=−3e−3


−23
Case 9:Hognestad,fp= −30 MPa,εc=−3e−3

−24

−25

−26

−27

−28

−29

−30
−4 −3.8 −3.6 −3.4 −3.2 −3 −2.8 −2.6 −2.4 −2.2 −2
εp [mm/mm] −3
x 10

Figure 5.14: Compressive concrete model, fc under variable εp

4
x 10
2.5 −22

2 −24
1.5
−26
fc [MPa]

1
−28
0.5
−30
0

−0.5 −32
−4 −3.5 −3 −2.5 −2 −4 −3 −2
εp [mm/mm] −3 εp [mm/mm] −3
x 10
x 10
(a) (b)

4000 1
2000
0.9
0
−2000 0.8 Case 4
−4000
p

0.7 Case 4−Euler


∂f /∂f

−6000 Case 9
c

−8000 0.6
Case 9−Euler
−10000 0.5
−12000
0.4
−14000
−16000
−4 −3.5 −3 −2.5 −2 −4 −3.5 −3 −2.5 −2
εp [mm/mm] x 10
−3 εp [mm/mm] x 10
−3

(c) (d)

Figure 5.15: Compressive concrete model, fceuler and derivatives under variable εp
185

Case 5:PopovicHS ,εp=−3e−3,εc=−3e−3

Case 10:Hognestad,ε =−3e−3,ε =−3e−3


p c

−45 −40 −35 −30 −25 −20 −15 −10


fp [mm/mm]

Figure 5.16: Compressive concrete model, fc under variable fp

−10

−15

−20

−25
fc [MPa]

−30

−35

−40

−45

−50
−50 −40 −30 −20 −10
−40 −35 −30 −25 −20 −15 −10 f [mm/mm]
fp [mm/mm] p
(b)
(a)

1 Case 5:PopovicHS
Case 5−Euler
Case 10:Hognestad
0.5 Case 10−Euler
p
∂ f / ∂ε

0
c

−0.5

−1
−40 −35 −30 −25 −20 −15 −10 −50 −45 −40 −35 −30 −25 −20 −15 −10
fp [mm/mm] fp [mm/mm]
(c) (d)

Figure 5.17: Compressive concrete model, fceuler and derivatives under variable fp
186

5.2.7 Tensile Concrete 1D model

The tensile concrete 1D models determine the tensile concrete stress, fc > 0) for a tensile
strain, ε > 0 and the additional variable parameters such the orientation of the steel with
respect to crack defined by the angle, φ, and the crack spacing, Scr . The tensile concrete
1D model cases are shown in Table 5.7.

Table 5.7: Tension concrete 1D models


0
Case Model εc Scr φ ρ db fc
[mm/mm] [mm] [MPa] % [mm] [MPa]
1 LNRBentz varies 40 0 1 5 -20
2 LNRBentz varies 60 0 1 5 -20
3 LNRBentz varies 40 10 1 5 -20
4 LNRBentz varies 40 0 5 5 -20
5 LNRBentz varies 40 0 1 3 -20
6 LNRBentz 1.00E-03 varies 0 1 5 -20
7 LNRBentz 3.00E-03 40 varies 1 5 -20
8 Vecchio1982 varies 40 0 1 5 -20

Figure 5.18 shows the results for cases 1 to 5, and 8. These cases show the variation
of fc for a variable εc (all other variables in Table 5.7 are held). These cases were already
discussed in Section 2.5 to illustrate the impact of using different models and parameters.
∂fc euler ∂fc ∂fc
Figure 5.19 shows Ec = , fc , , and for a variable εc . The variable fceuler is
∂εc ∂φ ∂Scr
obtained by integrating Ec over the εc domain, with all other variables in Table 5.7 held.
The difference between fc and fceuler is minimal, which indicates that the expressions used
for Ec are correct.
Figure 5.20 shows results for case 6, which shows the variation of fc with Scr . Figure
∂fc ∂fc
5.21 shows Ec , fceuler , , and for a variable Scr . The variable fceuler is obtained by
∂φ ∂Scr
∂fc
integrating over the Scr domain (other variables are constant). The difference between
∂Scr
∂fc
fc and fceuler is minimal, which indicates that the expressions used for are correct.
∂Scr
187

Figure 5.22 shows case 7, which shows the variation of f c with φ. Figure 5.21 shows Ec ,
∂fc ∂fc ∂fc
fceuler , , and for a variable φ. The variable fceuler is obtained by integrating
∂φ ∂Scr ∂φ
over the φ , with all other variables in Table 5.7 held. The difference between fc and fceuler
∂fc
is minimal, which indicates that the expressions used for are correct.
∂φ

1.5

0.5
Case 1
Case 2
Case 3
Case 4
Case 5
Case 8

0
0 0.5 1 1.5 2 2.5 3 3.5
εc [mm/mm] −3
x 10

Figure 5.18: Tensile concrete models, fc under variable εc


188

200 1.5

fc [MPa]
0
1

−200

0.5
0 0.5 1 1.5 2 2.5 3 3.5 0 1
εc [mm/mm] −3 εc [mm/mm] x 10
−3
x 10
(a) (b)

0 1
−0.05 0.8 Case 1:LNRBentz
−0.1 0.6 Case 1−Euler
Case 2:LNRBentz
−0.15 0.4
∂ fc/ ∂ Scr Case 2−Euler
−0.2 0.2
Case 3:LNRBentz
−0.25 0
Case 3−Euler
−0.3 −0.2 Case 4:LNRBentz
−0.35 −0.4 Case 4−Euler
−0.4 −0.6 Case 5:LNRBentz
−0.45 −0.8 Case 5−Euler
−0.5 −1 Case 8:Vecchio1982
0 0.5 1 1.5 2 2.5 3 3.5 0 0.5 1 1.5 2 2.5 3 3.5
εc [mm/mm] −3 εc [mm/mm] x 10−3 Case 8−Euler
x 10
(c) (d)

Figure 5.19: Tensile concrete models, fceuler and derivatives under variable εc

Case 6:LNRBentz, , φ = 0 Deg,ρ =1%, db=5mm

42 44 46 48 50 52 54 56 58 60
Scr [mm]

Figure 5.20: Tensile concrete models, fc under variable Scr


189

1
1.1
0.5
1.05

fc [MPa]
0 1

0.95
−0.5
0.9

−1 0.85
40 45 50 55 60 40 42 44
Scr [mm] S [mm] (b)
cr Case 6:LNRBentz
(a)
Case 6−Euler
−165.5 1

−166
0.5

−166.5
∂ fc/ ∂φ

0
−167

−0.5
−167.5

−168 −1
40 45 50 55 60 40 45 50 55 60
Scr [mm] Scr [mm]
(c) (d)

Figure 5.21: Tensile concrete models, fceuler and derivatives under variable Scr

Case 7

5 10 15 20 25 30 35 40 45
φ [Deg]

Figure 5.22: Tensile concrete models, fc under variable φ


190

1 0.78
0.8
0.6
0.4
0.2
fc [MPa]

0 0.77
−0.2
−0.4
−0.6
−0.8
−1 0.76
0 5 10 15 20 25 30 35 40 45 0 5 10 15
φ [Deg] φ [Deg]
(a) (b)

−61.34 0
−61.36 −0.02
−0.04 Case 7:LNRBentz
−61.38
−0.06 Case 7−Euler
−61.4 −0.08
∂ fc/ ∂φ

−61.42 −0.1
−61.44 −0.12
−0.14
−61.46
−0.16
−61.48 −0.18
−61.5 −0.2
0 5 10 15 20 25 30 35 40 45 0 10 20 30 40 50
φ[Deg] φ[Deg]
(c) (d)

Figure 5.23: Tensile concrete models, fceuler and derivatives under variable φ
191

5.2.8 Compression softening 1D model

The compression softening 1D models test determine the compression softening modification
factor, βd , for a given concrete compressive strain, εc < 0, in the direction under considera-
tion, and a strain in the perpendicular direction, εperp
c > 0. The compression softening 1D
models cases are shown in Table 5.8.

Table 5.8: Compression Softening 1D models

Case Model εc εperp


c εperp
c /εc
[mm/mm] [mm/mm]
1 Vecchio 1992A - - varies
2 Vecchio 1992A varies 1.00E-03 -
3 Vecchio 1992A varies 2.00E-03 -
4 Vecchio 1992A -2.00E-03 varies -
5 Vecchio 1986 - - varies
6 Vecchio 1986 varies 1.00E-03 -
7 Vecchio 1986 varies 2.00E-03 -
8 Vecchio 1986 -2.00E-03 varies -

Figure 5.24 shows cases 1 and 5. These cases were already discussed in Section 2.8 to
illustrate the difference between the Vecchio 1992A and Vecchio 1986 models.
Figure 5.25 shows the results for cases 2,3,6 and 7. These cases show the variation of βd
∂βd ∂βd
with respect to εc . Figure 5.26 show , , and βdeuler for a variable εc . The variable
∂εc ∂εperp
c
euler ∂βd
βd is obtained by integrating over the εc domain. The difference between βd and
∂εc
∂βd
βdeuler is minimal, which indicates that the expressions used for are correct.
∂εc
Figure 5.27 shows the results for cases 4, and 8. These cases show the variation of βd with
∂βd ∂βd
respect to εperp
c . Figure 5.28 show , , and βdeuler for a variable εperp
c . The variable
∂εc ∂εperp
c
∂βd
βdeuler is obtained by integrating over the εperp
c domain. The difference between βd
∂εperp
c
∂βd perp
and βdeuler is minimal, which indicates that the expressions used for are correct.
∂εc
192

Case 1: Vecchio1992A
Case 5: Vecchio1986

5 10 15 20 25 30 35 40 45 50
perp
−εc /εc

εperp
c
Figure 5.24: Compression softening models, βd under variable −
εc
193

Case 2: Vecchio1992A, εperp = 1e−3, εc0=−3e−3

Case 3: Vecchio1992A, εperp = 2e−3, ε =−3e−3


c0

Case 6: Vecchio1986, εperp = 1e−3, εc0=−3e−3

Case 7: Vecchio1986, εperp = 2e−3, εc0=−3e−3

−7 −6 −5 −4 −3 −2 −1 0
εc [mm/mm] −3
x 10

Figure 5.25: Compression softening models, βd under variable εc

−50
dβd/εperp
c

−100

−150

−200

−250
−2 −1.5 −1 −0.5 0 −8 −7 −6 −5 −4 −3 −2 −1 0
ε [mm/mm] −3 ε [mm/mm] −3
c x 10 c x 10

(a) (b)

Case 2: Vecchio1992A, εperp = 1e−3, εc0=−3e−3

Case 2−Euler
Case 3: Vecchio1992A, εperp = 2e−3, εc0=−3e−3

Case 3−Euler
Case 6: Vecchio1986, εperp = 1e−3, εc0=−3e−3

Case 6−Euler
perp
Case 7: Vecchio1986, ε = 2e−3, εc0=−3e−3

Case 7−Euler

−4 −2 0
εc [mm/mm] −3
x 10

(c)

Figure 5.26: Compression softening models, βdeuler and derivatives under variable εc
194

Case 4: Vecchio1992A, ε = −2e−3, ε =−3e−3


c c0
Case 8: Vecchio1986, εc = −2e−3, εc0=−3e−3

1 2 3 4 5 6
εperp [mm/mm] −3
x 10
c

Figure 5.27: Compression softening models, βd under variable εperp


c

0
−10
−20
∂βd/∂εperp

−30
c

−40
−50
−60
−70
−80
2 3 4 5 6 −90
perp
εc [mm/mm] x 10
−3
−100
0 1 2 3 4 5 6
(a)
εc [mm/mm] −3
x 10
(b)

Case 4: Vecchio1992A, εc = −2e−3, εc0=−3e−3

Case 4−Euler
Case 8: Vecchio1986, ε = −2e−3, ε =−3e−3
c c0

Case 8−Euler

1 2 3
εperp [mm/mm] −3
x 10
c

(c)

Figure 5.28: Compression softening models, βdeuler and derivatives under variable εperp
c
195

5.2.9 Confinement 1D model

The confinement 1D models determine the confinement modification factor,βl , from the
maximum and minimum confinement stresses, fcmax and fcmin . Notice the the fcmin is the
most compressive confining stress, since it the most negative one. The cases performed on
the 1D confinement tests are

Table 5.9: Confinement models

Case Model fcmin fcmax fcmin /fcmax f’c


[MPa] [MPa] Mpa
1 Kupfer-Richart varies - 1.5 -20
2 Kupfer-Richart varies - 1 -20
3 Kupfer-Richart varies - -0.5 -20
4 Kupfer-Richart varies - 1.5 -30
5 Kupfer-Richart varies - 1 -30
6 Kupfer-Richart varies - -0.5 -30
7 Kupfer-Richart varies -2 - -20
8 Kupfer-Richart -2 varies - -20

Figure 5.29 shows the results of cases cases 1 to 6. These cases show the variation of
f min 0
βl for a constant cmax and fc . Under similar confining stress the value of βl is higher for
fc
0
lower strength concretes fc . This can be seen in Figure 5.29 by comparing Cases 1 and 4, 2
and 5, and 3 and 6. Figure 5.29 also shows that the factor if the maximum confining stress
f min
(least negative) fcmax ( c < 0) is positive the increase in the confining factor βl due to
fc max
more negative fcmin is lower. This can be seen in 5.29 by comparing the slope of cases 1, 2,
4, and 5 to that of cases 3 and 6.
Figure 5.30 shows the results of Case 7. This case shows the variation of βl with respect
0 ∂βl ∂βl
to fcmin for a constant fc and fcmax . Figure 5.31 shows , , and βleuler for a
∂fcmin ∂fcmax
∂βl
variable fcmin . The variable βleuler is obtained by integrating over the fcmin domain.
∂fcmin
The difference between βl and βleuler is minimal, which indicates that the expressions used
196

∂βl
for are correct.
∂fcmin
Figure 5.32 shows the results of Case 8. This case shows the variation of βl with respect
0 ∂βl ∂βl
to fcmax for a constant fc and fcmin . Figure 5.32 shows max
, , and βleuler for a
∂fc ∂fcmin
∂βl
variable fcmax . The variable βleuler is obtained by integrating over the fcmax domain.
∂fcmax
The difference between βl and βleuler is minimal, which indicates that the expressions used
∂βl
for are correct.
∂fcmax

Case 1:f min / f max = 1.5, fc = −20 MPa



c c
min max ’
Case 2:fc / fc = 1.0, fc = −20 MPa

Case 3:fcmin / fcmax = −0.5, f’c = −20 MPa


min max ’
Case 4:fc / fc = 1.5, fc = −30 MPa
min max
Case 5:fc / fc = 1.0, f’c = −30 MPa

Case 6:f min / f max = −0.5, f’ = −30 MPa


c c c

−3.8 −3.6 −3.4 −3.2 −3 −2.8 −2.6 −2.4 −2.2 −2


min
fc [MPa]

fcmin
Figure 5.29: Confinement model, βl under variable fcmin and constant
fcmax
197

Case 7: f max = −2 MPa, f’ = −20 MPa


c c

−3.8 −3.6 −3.4 −3.2 −3 −2.8 −2.6 −2.4 −2.2 −2


min
fc [MPa]

Figure 5.30: Confinement model, βl under variable fcmin

−0.156

−0.158

−0.16
dβl/fcmax

−0.162

−0.164

−0.166

−0.168
−3.6 −3.4 −3.2 −3 −2.8 −2.6 −2.4 −2.2 −2 −4 −3.8 −3.6 −3.4 −3.2 −3 −2.8 −2.6 −2.4 −2.2 −2
f min [MPa] f min [MPa]
c c

(a) (b)

Case 7: fcmax = −2 MPa, f’ = −20 MPa


c
Case 7 − Euler

−3.5 −3 −2.5 −2
f min [MPa]
c

(c)

Figure 5.31: Confinement model, βleuler and derivatives under variablefcmin


198

Case 7: f max = −2 MPa, f’ = −20 MPa


c c

−1.5 −1 −0.5 0 0.5 1 1.5 2


max
fc [MPa]

Figure 5.32: Confinement model, βl under variable fcmax

0
−0.02
−0.04
−0.06
dβ /f max

−0.08
l c

−0.1
−0.12
−0.14
−0.16
−0.18
−1 −0.5 0 0.5 1 1.5 2 −2 −1.5 −1 −0.5 0 0.5 1 1.5 2
f max [MPa] f max [MPa]
c c

(a) (b)

Case 8: fcmin = −2 MPa, f’c = −20 MPa


Case 8 − Euler

−1 0 1 2
f max [MPa]
c

(c)

Figure 5.33: Confinement model, βleuler and derivatives under variable fcmax
199

5.2.10 Poisson compressive 1D model

The Poisson compressive 1D models determine the Poisson’s ratio,ν, for a compressive load
strain, εload
c < 0, and the modified peak strength, εp . The tests performed on the 1D Poisson
compressive models are

Table 5.10: Poisson Compressive models

Case Model εload


c εp εload
c /εp
1 Kupfer - - varies
2 Kupfer varies -3.00E-03 -
3 Kupfer -3.00E-03 varies -

Figure 5.34 shows the results of Case 1. This case shows the variation of the Poisson
εload
modulus ν for a variable c ratio, where εload
c and εp are negative. The variable ν varies
εp
load
from 0.2 when εload = 0 to 0.5 when ε load reaches ε ( εc = 1).
c c p
εp
Figure 5.35 shows the result of Case 2. This case shows the variation of ν with respect
∂ν ∂ν
to εload
c for a constant εp . Figure 5.36 shows load
, , and ν euler for a variable εload
c .
∂εc ∂εp
∂ν
The variable ν euler is obtained by integrating over the εload
c domain. The difference
∂εload
c
∂ν
between ν and ν euler is minimal, which indicates that the expressions used for are
∂εload
c
correct.
Figure 5.37 shows the results of Case 3. This case shows the variation of ν with respect
∂ν ∂ν
to εp for a constant εload
c . Figure 5.38 shows , , and ν euler for a variable εp . The
∂εp ∂εload
c
∂ν
variable ν euler is obtained by integrating over the εp domain. The difference between
∂εp
∂ν
ν and ν euler is minimal, which indicates that the expressions used for are correct.
∂εp
200

Case 1: Kupfer

0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2


εload/ε
c p

εload
c
Figure 5.34: Poisson Compressive Model, ν under variable
εp

Case 2: Kupfer, ε = −3e−3


p

−4.5 −4 −3.5 −3 −2.5 −2 −1.5 −1 −0.5 0


load
ε [mm/mm] x 10
−3
c

Figure 5.35: Poisson Compressive Model, ν under variable εload


c
201

100 500

0 400

−100 300

p
∂ν/∂ε
−200 200

−300 100

−400 0

−500 −100
−5 −4.5 −4 −3.5 −3 −2.5 −2 −1.5 −1 −0.5 0 −5 −4.5 −4 −3.5 −3 −2.5 −2 −1.5 −1 −0.5 0
εc load[mm/mm] x 10
−3
εload [mm/mm] x 10
−3
c
(a) (b)

Case 2: Kupfer, ε = −3e−3


0.5 p

0.4

0.3

0.2

0.1

0
−5 −4 −3 −2 −1 0
load
εc [mm/mm] x 10
−3

(c)

Figure 5.36: Poisson Compressive Model, ν euler and derivatives under variable εload
c

Case 2: Kupfer, εload


c
= −3e−3

−6.5 −6 −5.5 −5 −4.5 −4 −3.5 −3 −2.5 −2


εp [mm/mm] x 10
−3

Figure 5.37: Poisson Compressive Model, ν under variable εp


202

500

400

300
∂ν/∂εp

200

100

−100
−5.5 −5 −4.5 −4 −3.5 −3 −2.5 −2 −7 −6.5 −6 −5.5 −5 −4.5 −4 −3.5 −3 −2.5 −2
ε [mm/mm] −3 ε [mm/mm] −3
p x 10 p x 10
(a) (b)

Case 2: Kupfer, εload


c
= −3e−3

−5 −4 −3 −2
εp [mm/mm] −3
x 10
(c)

Figure 5.38: Poisson Compressive Model, ν euler and derivatives under variable εp
203

5.2.11 Poisson tensile 1D model

The Poisson tensile 1D model determines the Poisson’s ratio, ν, for a tensile load strain,
εload
c > 0, and the concrete cracking strain, εccr . The tests performed on the 1D Poisson
tensile models are

Table 5.11: Poisson Tensile models

Case Model εload


c εccr εload
c /εccr
1 V2 - - varies
2 V2 varies 6.60E-05 -

Figure 5.39 shows Case 1. This case shows the variation of the Poisson modulus ν for a
εload
variable c ratio, where εload
c is positive. The variable ν varies from 0.2 when εload c =0
εccr
εload
to 0 when εload
c reaches 2εccr ( c = 2).
εccr
Figure 5.40 shows Case 2. This case shows the variation of ν with respect to εload c for a
∂ν
constant εccr . Figure 5.41 shows , and ν euler for a variable εload
c . The variable ν euler
∂εload
c
∂ν
is obtained by integrating load over the εload c domain. The difference between ν and ν euler
∂εc
∂ν
is minimal, which indicates that the expressions used for load are correct.
∂εc
204

0.25
Case 1: V2

0.2

0.15

0.1

0.05

0
0 0.5 1 1.5 2 2.5 3
load
εc /εccr

εload
c
Figure 5.39: Poisson Tension Model, nu under variable ratio
εccr

0.25
Case 2: V2, εccr = 6.6e−5

0.2

0.15

0.1

0.05

0
0 1 2
εload
c
[mm/mm] x 10
−4

Figure 5.40: Poisson Tension Model, nu under variable εload


c
205

0
−500
−1000
−1500
−2000
−2500
−3000
−3500
−4000
0 1 2
load
ε [mm/mm] −4
x 10
c
(a)

0.25

0.2 Case 2: V2, ε = 6.6e−5


ccr
0.15 Case2 − Euler

0.1

0.05

0
0 1 2
εload [mm/mm] −4
x 10
c
(b)

Figure 5.41: Poisson Tension Model, nueuler and derivatives under variable εload
c
206

5.3 RC material model tests

The RC material tests evaluate the response of the DSFM-TG using four groups of test:

• Compression tests

• Tension tests

• Shear tests

• Experimental tests

The first three groups of tests evaluate physical and numerical aspects of the DSFM-
TG using different types of 1D constitutive models and concrete stiffness matrices. The last
group compares the DSFM-TG with experimental results obtained from the testing of a set
of RC panels by Vecchio and Collins (1982).

5.3.1 Compression tests

The compression tests performed are listed in Table 5.12.

Table 5.12: RC material compression tests

Cases Control Type Load Pattern Effects Material [Dc ] Type


Stiffness
CC1-TR-Strain Strain Driven εx : εy : γxy 0:-1:0 - Tangent Rotating
CC2-TR-Strain Strain Driven εx : εy : γxy 0:-1:0 P Tangent Rotating
CC3-TR-Strain Strain Driven εx : εy : γxy 0:-1:0 P, OP Tangent Rotating
CC4-TR-Strain Strain Driven εx : εy : γxy 0:-1:0 P,OP,CF Tangent Rotating
CC1-SF-Stress Stress Driven σx : σy : τxy 0:-1:0 - Secant Fixed
CC1-TF-Stress Stress Driven σx : σy : τxy 0:-1:0 - Tangent Fixed
CC1-TR-Stress Stress Driven σx : σy : τxy 0:-1:0 - Tangent Rotating
CC2-TR-Stress Stress Driven σx : σy : τxy 0:-1:0 P Tangent Rotating
CC3-TR-Stress Stress Driven σx : σy : τxy 0:-1:0 P, CF Tangent Rotating
CC4-TR-Stress Stress Driven σx : σy : τxy 0:-1:0 P,CF,OP Tangent Rotating

The type of material stiffness used in the tests shown in Table 5.12 could be tangent or
secant. The tangent stiffness is used in the DSFM-TG , while the secant stiffness is used in
207

the original version of the DSFM. The concrete material stiffness matrix could be Rotating
or Fixed. The rotating type includes the effect of the principal concrete direction rotation
and it is the type used in the DSFM-TG . The fixed type does not include the effect of the
rotation of the principal concrete directions, and it is the type used in the original version
of the DSFM.
Test cases CC1-TR-Strain, CC2-TR-Strain, CC3-TR-Strain, and CC4-TR-Strain are
strain driven and are intended to evaluate the predicted stress response and convergence of
the internal RC state problem. The effect of adding Poisson (P), out-of-plane steel (OP),
and confinement (CF) is also included. The evaluation is performed on the physical aspects
of the predicted for the different effects and models used response as well as the convergence
rate and time.
Test cases CC1-SF-Stress, CC1-TF-Stress, and CC1-TR-Stress are stress driven and
they are intended to asses the impact of using different types of material stiffness (tan-
gent or secant) and concrete material matrix [Dc ] types (rotating or fixed) on the problem
convergence.
Tests CC2-TR-Stress, CC3-TR-Stress, and CC4-TR-Stress extend case CC1-TR-Stress
by adding the Poisson, confinement and out-of-plane steel effects. Physical and numerical
comparisons are made between cases CC1-TR-Stress, CC2-TR-Stress, CC3-TR-Stress, and
CC4-TR-Stress.
Material and analysis properties other than those listed in Table 5.12, are identical for
all of the test cases and are shown listed Table 5.13.
208

Table 5.13: RC material compression test properties

Test Analysis Parameters


Number of steps 25
Error tolerance 1.00E-06
RC material Analysis Parameters
Concrete 1D model PopovicHS
Steel Longitudinal 1D model Raynor
Confinement model Kupfer-Richart
Poisson Compression 1D model Kupfer-Variable
Internal Strain tolerance, tolε 1.00E-16
ε
Internal Stress tolerance, tol 1.00E-12
Maximum internal iterations, nmax
iter 100
Material Properties
0
fc -20 MPa
εco -0.003
ν0 0.2
fy 400 MPa
εy 0.002
fsh 440 MPa
εsh 0.012
fu 500 MPa
εu 0.15
nsh 6
ρ1 =ρx , ρ2 = ρy 1%
ρz 0.2%
α1 0 Deg
α2 90 Deg
Smx , Smy 75 mm
209

Figure 5.42 shows the global RC response for the test cases CC1-TR-Strain , CC2-
TR-Strain (P), CC3-TR-Strain (P+OP), and CC4-TR-Strain (P+OP+CF). Figure 5.42a)
shows the RC compressive stress σy vs. εy . The maximum strength is achieved for the
CC4-TR-Strain case, which is the only case that includes confinement. All the other cases
reach a lower equal strength because the maximum concrete compressive stress, fp , is not
0
equal to fc , as an not greater as in the CC4-TR-Strain case. Figure 5.42(b) shows that
stresses develop in the X direction (εx = 0) for all but the test case CC1-TR-Strain, which
does not include Poisson effect. Figures 5.42(c) and 5.42(d) show that the shear stress τxy
and the angle θ12c are nearly zero for all the test cases, as expected.
The inclusion of the Poisson effect generates compressive stresses in the x-direction since
the strain in the x-direction is restrain(εx = 0). The Poisson effect increases the pre-peak
stresses for a given εy , but increases the post-peak decay. This can be explained by the
increase of εload
c2 for a given εy when the Poisson effect is included (Figure 5.42 (c)). The
out-of-plane steel included in case CC3-TR-Strain does not affect the response significantly
with respect to what it is observed in cases CC2-TR-Strain. The shear stress τxy and angle
θ12c are nearly zero for all the cases, as expected.
210

−5

−10
σx [MPa]

−15

−20

−25

−30
−6 −5 −4 −3 −2 −1 0 −8 −7 −6 −5 −4 −3 −2 −1 0
εy [mm/mm] −3 εy [mm/mm] −3
x 10 x 10
(a) (b)

1
CC1−TR−Strain
CC2−TR−Strain
CC3−TR−Strain
0.5 CC4−TR−Strain
θ12c [rad]

−0.5

−1
−6 −5 −4 −3 −2 −1 0 −8 −6 −4 −2 0
εy [mm/mm] −3
x 10 εy [mm/mm] x 10
−3

(c) (d)

Figure 5.42: RC response for test cases CC(1-4)-TR-Strain


211

The global response is dominated by the concrete behavior which can be observed by
comparing Figures 5.42(a) and 5.43(b)
Figure 5.43 shows the principal concrete compressive response, fc2 , for the test cases
CC1-TR-Strain , CC2-TR-Strain (P), CC3-TR-Strain (P+OP), and CC4-TR-Strain (P+OP+CF).
Figure 5.43(a) shows that εc2 is equal to εy for all the cases. This is expected because the
test cases are strain driven. Figure 5.43(b) shows the relationship between εc2 and fc2 .
Figure 5.43(d) shows that the response εload
c2 -fc2 is identical for all the cases, except CC4-

TR-Strain, which included confinement (more compressive fp ). However, the relationship


εc2 -fc2 is different for cases CC1-TR-Strain, CC2-TR-Strain, and CC3-TR-Strain, which
0
include the Poisson effect. This is because they all used the same fp = fc .
The combination of out-of-plane stress, Poisson and confinement increases the fc2 peak
response significantly from a value of about -20MPa to a value of nearly -28MPa. However,
the post-peak decay is very sharp. This is an artefact of the PopovicHS model, which
predicts that concrete with higher strength (higher fp ), exhibits more rapid strength loss
than does concrete with lower strength (lower fp ).
212

−5

−10
[MPa]
−15
c2
f

−20

−25

−30
−7 −6 −5 −4 −3 −2 −1 0 −8 −7 −6 −5 −4 −3 −2 −1 0
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−3
y c2
(a) (b)

0
CC1−TR−Strain
−5 CC2−TR−Strain
CC3−TR−Strain
CC4−TR−Strain
−10
[MPa]

−15
c2
f

−20

−25

−30
−7 −6 −5 −4 −3 −2 −1 0 −0.012 −0.01 −0.008 −0.006 −0.004 −0.002 0
ε [mm/mm] x 10
−3
εLoad [mm/mm]
y c2
(c) (d)

Figure 5.43: Concrete principal compressive, fc2 , response for test cases CC(1-4)-TR-Strain
213

Figure 5.44 shows the principal concrete tensile fc1 response for the test cases CC1-TR-
Strain , CC2-TR-Strain (P), CC3-TR-Strain (P+OP), and CC4-TR-Strain (P+OP+CF).
Figure 5.44(a) shows that εc1 is nominally equal to zero for all cases, which is expected
given that εc1 = εx = 0. However, εc1 = 0 does not imply fc1 = 0, since the εload
c1 is not
zero as shown in Figure 5.44(c). Figure 5.44(d) shows the εload
c1 -fc1 relationship in which

the only different curve (compression) is CC4-TR-Strain, which includes confinement.

−23
0

−5

−10

[MPa]
−15

c1
f −20

−25

−30
−7 −6 −5 −4 −3 −2 −1 0 −2 0 2 4 6 8 10 12 14
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−23
y c1
(a) (b)

0
CC1−TR−Strain
−5 CC2−TR−Strain
CC3−TR−Strain
CC4−TR−Strain
−10
[MPa]

−15
c1
f

−20

−25

−30
−7 −6 −5 −4 −3 −2 −1 0 −6 −5 −4 −3 −2 −1 0
ε [mm/mm] x 10
−3
εLoad [mm/mm] x 10
−3
y c1
(c) (d)

Figure 5.44: Concrete principal tensile, fc1 ,response for test cases CC(1-4)-TR-Strain
214

Figure 5.45 shows the concrete out-of-plane, fcz , response. Figure 5.45(a) shows that
only the cases that include Poisson effect (CC(2-4)-TR-Strain) results in out-of-plane ex-
pansion εcz > 0. This out-of-plane expansion becomes out-of-plane compression only when
the out-of-plane steel is included, such as in the cases CC(3-4)-TR-Strain. When this hap-
pens the out-of-plane load strain, εload
cz , becomes negative (Figure 5.45(c)) and out-of-plane

compression stresses are generated (Figure 5.45(b) and 5.45(d)). Note that the fcz − εcz
response in Figure 5.45(b) shows a εcz -fcz curve that resembles the steel stress-strain curve
(Figure 5.46(f). This is a product of the out-of-plane equilibrium requirement that implies
fsz
fcz = − .
ρz

CC1−TR−Strain
−0.2 CC2−TR−Strain
CC3−TR−Strain
−0.4
fcz [MPa]

CC4−TR−Strain

−0.6

−0.8

−1
−6 −5 −4 −3 −2 −1 0 0 2 4 6 8
εy [mm/mm] −3 εcz [mm/mm] −3
x 10 x 10
(a) (b)

0
CC1−TR−Strain
CC2−TR−Strain
−0.2
CC3−TR−Strain
CC4−TR−Strain
−0.4
fcz [MPa]

−0.6

−0.8

−1
−6 −5 −4 −3 −2 −1 0 −1.2 −1 −0.8 −0.6 −0.4 −0.2 0
εy [mm/mm] −3
x 10 εLoad [mm/mm] −4
x 10
cz
(c) (d)

Figure 5.45: Concrete principal out-of-plane, fcz , response for test cases CC(1-4)-TR-Strain
215

Figure 5.46 shows the steel stress response for cases CC(1-4)-TR-Stain. As defined
in Table 5.13 in-plane steel layers 1 and 2 are with the x- and y-axes, respectively. The
response of these in-plane steel layers response is not affected by the Poisson effect, out-of-
plane steel, and the presence of confinement, because the strain is calculated directly from
the RC average strains, ε, and the steel constitutive model is not subjected to modifications
as the concrete compressive model is. This explains why Figures 5.46(a) to (d) show equal
response for all the cases. The out-of-plane steel layer strain, εsz , however equals the out-
of-plane concrete strain, εcz , (εsz = εcz ) and is affected by the concrete response shown in
Figures 5.46(d) and (f).

0
−100
[MPa]

−200
−300
s2
f

−400
−500
−7 −6 −5 −4 −3 −2 −1 0 −8 −7 −6 −5 −4 −3 −2 −1 0
ε [mm/mm] x 10
−3 ε [mm/mm] −3
x 10
y s2
(a) (b)
1

0.5
[MPa]

0
s1
f

−0.5

−1
−7 −6 −5 −4 −3 −2 −1 0 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1
ε [mm/mm] x 10
−3 ε [mm/mm]
y s1
(c) (d)
600

400
[MPa]

CC1−TR−Strain
CC2−TR−Strain
sz

200 CC3−TR−Strain
f

CC4−TR−Strain
0
−7 −6 −5 −4 −3 −2 −1 0 0 2 4 6 8
ε [mm/mm] x 10
−3 ε [mm/mm] −3
x 10
y sz
(e) (f)

Figure 5.46: Steel stress response, CC(1-4)-TR-Strain


216

Figure 5.47 shows the number of internal required to solve the internal RC state problem
for the test cases CC(1-4)-TR-Strain. The data shown in Figure 5.47 indicate that more
iterations are required as more ”effects” are added to this model. Stage in Figure 5.47 refers
to the the analysis stage number within the monotonically increasing strain-driven analysis.

CC1−TR−Strain
CC2−TR−Strain
CC3−TR−Strain
CC4−TR−Strain

5 10 15 20 25 30
Stage

Figure 5.47: Number of Internal Iterations, CC(1-4)-TR-Strain

The analysis times for the test cases shown in Figure 5.47 are listed in Table 5.14.

Table 5.14: Analysis Times

Analysis Case Time [s]


CC1-TR-Strain 0.66
CC2-TR-Strain 0.89
CC3-TR-Strain 1.66
CC4-TR-Strain 1.45
217

ε , for the different analysis cases are shown in Figure


The internal strain error, errorint
5.48. The data in these plots indicate that the rate of convergence is quadratic.

0
10
CC1−TR−Strain CC2−TR−Strain

−5
10

int
errorε
−10
10

−15
10

−20
10
10 20 30 40 50 60 0 20 40 60 80 100
Internal Iterations Internal Iterations
(a) (b)

0
10
CC3−TR−Strain CC4−TR−Strain

−5
10
int
errorε

−10
10

−15
10

−20
10
20 40 60 80 100 0 50 100 150
Internal Iterations Internal Iterations
(c) (d)

Figure 5.48: Internal Strain Error, CC(1-4)-TR-Strain


218

Figure 5.49 shows the plot of stress driven test cases CC1-SF-Stress, CC1-TF-Stress,
and CC1-TR-Stress. The response are all similar other than the small errors of case CC1-
TR-Stress in Figure 5.49(b) and 5.49(c).

−32
x 10
1.4

1.2

ε [mm/mm]
0.8

0.6

x
0.4

0.2

0
−20 −15 −10 −5 0 −25 −20 −15 −10 −5 0
σ [MPa] σ [MPa]
y y
(a) (b)
−17 −15
x 10
2

0
[rad]

CC1−SF−Stress
−2 CC1−TF−Stress
12c

CC1−TR−Stress
θ

−4

−6
−20 −15 −10 −5 0 −25 −20 −15 −10 −5 0
σ [MPa] σ [MPa]
y y
(c) (d)

Figure 5.49: RC response, CC1-SF-Stress, CC2-TF-Stress, and CC3-TR-Stress


219

Figure 5.51 shows the number of external iterations for the test cases CC1-SF-Stress,CC1-
TF-Stress, and CC1-TR-Stress. As the analysis approaches the peak stress and the slope
becomes flatter, the secant stiffness (CC1-SF-Stress) becomes an increasingly worse predic-
tor. There is no significant difference between CC1-TF-Stress and CC1-TR-Strain, both
require the same number of iterations.

CC1−SF−Stress
CC1−TF−Stress
CC1−TR−Stress

5 10 15 20 25 30
Stage

Figure 5.50: Number of External Iterations, CC1-SF-Stress, CC1-TF-Stress, and CC1-TR-


Stress

The analysis times for the test cases shown in Figure 5.50 are listed in Table 5.15.

Table 5.15: Analysis Times

Analysis Case Time [s]


CC1-SF-Stress 12.97
CC1-TF-Stress 2.14
CC1-TR-Stress 2.13
220

The error is plotted in Figure 5.51, which clearly shows the increasingly worse perfor-
mance of CC1-SF-Stress case. The convergence of CC1-TF-Stress and CC1-TR-Stress cases
is quadratic as the error diminishes quadratically between iterations. The CC1-SF is linear
because the error diminishes linearly between iterations.

0
10

−5
10

ext
errorσ
−10
10

CC1−SF−Stress CC1−TS−Stress
−15
10
100 200 300 400 0 20 40 60 80
External Iterations External Iterations
(a) (b)

CC1−TR−Stress

20 40 60 80
External Iterations
(c)

Figure 5.51: External error, CC1-SF-Stress, CC1-TF-Stress, andCC1-TR-Stress


221

Figure 5.52 shows the RC response for test cases CC1-TR-Stress, CC2-TR-Stress(P),
CC3-TR-Stress(P+OP), and CC4-TR-Stress(P+OP+CF). The confinement makes the con-
crete stiffer which means that less εy is needed to generate a given stress as shown in Figure
5.52(a), where the test case CC4-TR-Stress (only one that includes confinement) reaches a
lower strain εy than that of the other cases, for the same stress σy . The expansion in the
x-direction shown in Figure 5.52(b) is caused by the Poisson effect. . As expected the shear
strain γxy and the angle θ12c , shown in Figures 5.52(c) and (d), are nominally zero.

−3
x 10
1.2

0.8
ε [MPa]
0.6
x

0.4

0.2

0
−20 −15 −10 −5 0 −25 −20 −15 −10 −5 0
σ [MPa] σ [MPa]
y y
(a) (b)
−19 −17
x 10
1.4

1.2

1
[MPa]

0.8
12c

0.6 CC1−TR−Stress
θ

CC2−TR−Stress
0.4 CC3−TR−Stress
CC4−TR−Stress
0.2

0
−20 −15 −10 −5 0 −25 −20 −15 −10 −5 0
σ [MPa] σ [MPa]
y y
(c) (d)

Figure 5.52: RC response, CC(1-4)-TR-Stress


222

Figure 5.53 shows the concrete fc2 response for test cases CC(1-4)-TR-Stress. The
impact of the confinement and the out-out-of-plane reinforcement can be seen in the stiffer
response of CC4-TR-Stress.

0
−2
−4
−6

fc2 [MPa]
−8
−10
−12
−14
−16
−18
−20
−20 −15 −10 −5 0 −3 −2.5 −2 −1.5 −1 −0.5 0
σy [MPa] εc2 [mm/mm] −3
x 10

(a) (b)

0 CC1−TR−Stress
CC2−TR−Stress
CC3−TR−Stress
−5
CC4−TR−Stress
fc2 [MPa]

−10

−15

−20
−20 −15 −10 −5 0 −3 −2.5 −2 −1.5 −1 −0.5 0
σy [MPa] εLoad [mm/mm]
−3
x 10
c2
(c)
(d)

Figure 5.53: Concrete fc2 stress response, CC(1-4)-TR-Stress


223

Figure 5.54 shows the concrete fc1 stress response for test cases CC(1-4)-TR-Stress. The
concrete is in compression in the x direction because the horizontal reinforcement restrains
its expansion in that direction as can bee seen in Figure 5.54(c). This is an example of
the need of using εload
c1 (< 0) instead of εc1 (> 0), to calculate fc1 (< 0). CC1-TR-Stress test

case does not results in stresses fc1 because it does not include Poisson effect, which is the
responsible for causing the expansion in the direction normal to the externally applied stress
σy .

−3
x 10
1.2 0.5

1 0

0.8 [MPa] −0.5

0.6 −1
c1
f

0.4 −1.5

0.2 −2

0 −2.5
−25 −20 −15 −10 −5 0 0 0.2 0.4 0.6 0.8 1 1.2
σy [MPa] εc1 [mm/mm] −3
x 10
(a) (b)
−5
x 10
2 0.5
0
−2 0
−4 −0.5
[MPa]

−6
−8 −1
c1

−10 CC1−TR−Stress
f

−12 −1.5 CC1−TR−Stress


−14 CC1−TR−Stress
−2
−16 CC1−TR−Stress
−18 −2.5
−25 −20 −15 −10 −5 0 −18 −16 −14 −12 −10 −8 −6 −4 −2 0 2
σ [MPa] Load
εc1 [mm/mm] x 10
−5
y
(c) (b)

Figure 5.54: Concrete fc1 stress response, CC(1-4)-TR-Stress


224

Figure 5.55 shows the concrete fcz stress response for test cases CC(1-4)-TR-Stress. The
of plane strain, εcz = εz , arises when the Poisson effect is included, which is the case for
all the cases but test case CC1-TR-Stress (Figure ??. The out-of-plane concrete stress, fcz ,
arises when the out-of-plane steel is added as occurs for cases CC(3-4)-TR-Stress (Figure
5.55(b) to (d)).

−3
x 10
1.4 0
−0.05
1.2
−0.1
1 −0.15

[MPa]
0.8 −0.2
−0.25
0.6

cz
−0.3

f
0.4 −0.35
−0.4
0.2
−0.45
0 −0.5
−25 −20 −15 −10 −5 0 0 0.2 0.4 0.6 0.8 1 1.2 1.4
σy [MPa] εcz [mm/mm] −3
x 10
(a) (b)
−5
x 10 CC1−TR−Stress
0 0
CC2−TR−Stress
−0.5 CC3−TR−Stress
−0.1
−1 CC4−TR−Stress
−1.5
[MPa]

−0.2
−2
cz

−2.5 −0.3
f

−3
−0.4
−3.5
−4 −0.5
−25 −20 −15 −10 −5 0 −4 −3 −2 −1 0
σy [MPa] Load
εcz [mm/mm] x 10
−5

(c) (d)

Figure 5.55: Concrete fcz stress response, CC(1-4)-TR-Stress


225

Figure 5.56 shows the steel stress response for test cases CC(1-4)-TR-Stress. The in-
plane steel layer 2 (y- direction) is in compression due to the external stress σy applied in
the same direction as this in-plane steel layer (Figures 5.56(a) nd (b)). Notice that the
in-plane steel layer 1 (x- direction) is in tension due to the concrete Poisson expansion
(CC1(2-4)-TR-Stress) in the x-direction as shown in Figure 5.56(c) and (d). The out-of-
plane steel stress, fsz , gets activated when the out-of-plane effect is included as in cases
CC1(3-4)-TR-Stress, which is shown in Figures 5.56(e) and (f).

0
−100

fs2 [MPa]
−200
−300
−400
−500
−2 −1.5 −1 −0.5 0 −3 −2.5 −2 −1.5 −1 −0.5 0
εy [mm/mm] −3 εs2 [mm/mm] −3
x 10 x 10
(a) (b)

250
200
fs1 [MPa]

150
100
50
0
−2 −1.5 −1 −0.5 0 0 0.2 0.4 0.6 0.8 1 1.2
εy [mm/mm] −3
x 10 εs1 [mm/mm] −3
x 10
(c) (d)

250

200
fsz [MPa]

150 CC1−TR−Stress
CC2−TR−Stress
100
CC3−TR−Stress
50 CC4−TR−Stress
0
−2 −1.5 −1 −0.5 0 0 0.2 0.4 0.6 0.8 1 1.2 1.4
εy [mm/mm] −3
x 10 εsz [mm/mm] −3
x 10
(e) (f)

Figure 5.56: Steel stress response, CC(1-4)-TR-Stress


226

The number of external iterations or the number of iterations needed to balance the
internal and the external stresses are shown in Figure 5.57. The number of iterations
increases close to the peak compressive stress.

CC1−TR−Stress
5.5
CC2−TR−Stress
CC3−TR−Stress
5 CC4−TR−Stress

4.5

3.5

2.5

1.5

1
0 5 10 15 20 25 30
Stage

Figure 5.57: Number of External Iterations, CC(1-4)-TR-Stress

The analysis times for the test cases shown in Figure 5.57 are listed in Table 5.16.

Table 5.16: Analysis Times

Analysis Case Time [s]


CC1-TR-Stress 2.13
CC2-TR-Stress 2.63
CC3-TR-Stress 3.50
CC4-TR-Stress 3.73
227

Figure 5.58 shows the stress error for test cases CC(1-4)-TR-Stress. It is interesting to
note that the error behavior does not change considerably as more effects are added. This
indicates that the [Dc ] matrix returned by the material model is working properly.

5 5
10 10
CC1−TR−Stress CC2−TR−Stress

0 0
10 10

error[MPa]
−5
10

−10
−10 10
10

−15
10
0 20 40 60 80 0 20 40 60 80
External Iterations External Iterations
(a) (b)
5
10
CC3−TR−Stress CC4−TR−Stress
0 0
10 10
error[MPa]

−5
10 −5
10

−10
10 −10
10

−15
10
0 20 40 60 80 0 20 40 60 80
External Iterations External Iterations
(c) (d)

Figure 5.58: External error, CC(1-4)-TR-Stress


228

5.3.2 Tension tests

The tension tests performed are listed in Table 5.17.

Table 5.17: RC material tension tests

Cases Control Type Load Pattern Effects Material [Dc ] Type


Stiffness
CT1-TR-Strain Strain Driven εx : εy : γxy 0:1:0 - Tangent Rotating
CT2-TR-Strain Strain Driven εx : εy : γxy 0:1:0 P Tangent Rotating
CT1-SF-Stress Stress Driven σx : σy : τxy 0:1:0 - Secant Fixed
CT1-TF-Stress Stress Driven σx : σy : τxy 0:1:0 - Tangent Fixed
CT1-TR-Stress Stress Driven σx : σy : τxy 0:1:0 - Tangent Rotating
CT2-TR-Stress Stress Driven σx : σy : τxy 0:1:0 P Tangent Rotating

Cases CT1-TR-Strain and CT2-TR-Strain are strain driven and are intended to asses
the stress response and the rate of convergence internal RC state problem. Test cases
CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-Stress are stress driven and are intended to
asses the impact on the rate of convergence of the different material stiffnesses (tangent
or secant) and [Dc ] types (fixed or rotating). Test CT2-TR-Stress is an extension of the
CT1-TR-Stress case in which the Poisson effect is added to the model
The material and the analysis properties other than the ones shown on Table 5.17 are
identical for all the tests cases and are shown on Table 5.18
229

Table 5.18: RC material tension test properties

Test Analysis Parameters


Number of steps 50
Error tolerance 1.00E-06
RC material Analysis Parameters
Concrete 1D model PopovicHS
Steel Longitudinal 1D model Raynor
Confinement model Kupfer-Richart
Poisson Compression 1D model Kupfer-Variable
Internal Strain tolerance, tolε 1.00E-16
ε
Internal Stress tolerance, tol 1.00E-12
Maximum internal iterations, nmax
iter 100
Material Properties
0
fc -20 MPa
εco -0.003
ν0 0.2
fy 400 MPa
εy 0.002
fsh 440 MPa
εsh 0.012
fu 500 MPa
εu 0.15
nsh 6
ρ1 =ρx 1%
ρ2 = ρy 1.3%
α1 0 Deg
α2 90 Deg
Smx , Smy 75 mm
230

Figure 5.59 shows the RC response for the test cases CT(1-2)-TR-Strain. Figure 5.59(a)
shows that the RC response is dominated by the steel . The presence of the Poisson effect
generates tensile stresses in the x-direction, which disappear as the concrete strain in the
π
vertical direction increases. The shear τxy is nearly zero as expected. The angle θ12c is ,
2
which is correct.

0.35

0.3

0.25

σ [MPa]
0.2

0.15

x
0.1

0.05

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−3
y y
(a) (b)
−16
2
CT1−TR−Strain
CT2−TR−Strain
1.5
[rad]

1
12c
θ

0.5

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.5 1 1.5 2
ε [mm/mm] x 10
−3 ε [mm/mm] −3
x 10
y y
(c) (d)

Figure 5.59: RC response, CT1-TR-Strain and CT2-TR-Strain


231

Figure 5.59 shows the concrete fc1 stress response for the test cases CT1-TR-Strain and
CT2-TR-Strain. The Poisson effect is not significant and it is limited to the first stages.
This is reflected in Figure 5.60, where εload
c1 is equal to the εc1 = εy

1.5

[MPa]
c1
f
0.5

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−3
y c1
(a) (b)

1.5
CT1−TR−Strain
CT2−TR−Strain

1
[MPa]
c1
f

0.5

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.5 1 1.5 2
ε [mm/mm] x 10
−3
εLoad [mm/mm] −3
x 10
y c1
(c) (d)

Figure 5.60: Concrete fc1 response, CT1-TR-Strain and CT2-TR-Strain


232

Figure 5.61 shows the concrete fc2 response for the test cases CT1-TR-Strain and CT2-
TR-Strain. The strain εc2 is nearly zero as shown in Figure 5.61(a). The strain εload
c2

increases until the Poisson effect vanishes due to greater εload


c1 strain values.

−22
0.3

0.25

0.2

[MPa]
0.15

0.1

c2
f
0.05

−0.05
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 −2 0 2 4 6 8 10 12 14
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−22
y c2
(a) (b)

0.3
CT1−TR−Strain
0.25
CT2−TR−Strain
0.2
[MPa]

0.15

0.1
c2
f

0.05

−0.05
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 −5 0 5 10 15
ε [mm/mm] x 10
−3
εLoad [mm/mm] x 10
−6
y c2
(c) (d)

Figure 5.61: Concrete fc2 response, CT1-TR-Strain and CT2-TR-Strain


233

Figure 5.62 shows the in-plane steel layer 2 (y direction) stress response for cases CT1-
TR-Strain and CT2-TR-Strain. The strain εs2 is identical to εy . The cracking strain εscr2
is zero before cracking, and greater than zero after cracking as shown in Figure 5.62(c). The
steel strain increases considerably after yielding as shown in Figures 5.62(c) and 5.62(d).

400
CT1−TR−Strain
CT2−TR−Strain
300

[MPa]
200

s2
f
100

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.5 1 1.5 2
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−3
y s2
(a) (b)

500

400 CT1−TR−Strain
CT2−TR−Strain
[MPa]

300
scr2

200
f

100

0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.005 0.01 0.015 0.02 0.025 0.03
ε [mm/mm] x 10
−3 ε [mm/mm]
y scr2
(c) (c)

Figure 5.62: In-plane steel layer 2 stress response, CT1-TR-Strain and CT2-TR-Strain
234

Figure 5.63 shows the in-plane steel layer 1 (x directionl) stress response for cases CT1-
TR-Strain and CT2-TR-Strain. In the average region, the horizontal steel is not affected
by the RC strains (ε1 = εx = 0. The steel stress at the crack is also zero even though it
gets activated due to the fact that the angle θ12c is not exactly zero. The error though is
negligible (fscr1 ∼ 10−29 ).

1
0.8
0.6
0.4

[MPa]
0.2
0

s1
−0.2

f
−0.4
−0.6
−0.8
−1
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1
ε [mm/mm] x 10
−3 ε [mm/mm]
y s1
(a) (b)
−34 −29
x 10
2
1.8
1.6
1.4
[MPa]

1.2
1
scr1

0.8
f

0.6
0.4
0.2
0
0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−34
y scr1
(c) (d)

Figure 5.63: In-plane steel layer 1 stress response, CT1-TR-Strain and CT2-TR-Strain
235

Figure 5.64 shows that the number of internal iterations is higher initially for the CT2-
TR-Strain case, but this difference disappears as the Poisson effect vanishes.

CT1−TR−Strain
CT2−TR−Strain

10 20 30 40 50 60
Stage

Figure 5.64: Number of Internal Iterations, CT1-TR-Strain and CT2-TR-Strain

The analysis times for the test cases shown in Figure 5.64 are listed in Table 5.19.

Table 5.19: Analysis Times

Analysis Case Time [s]


CT1-TR-Strain 1.77
CT2-TR-Strain 1.89
236

The internal error for the CT1-TR-Strain and CT2-TR-Strain cases are shown in Figure
5.65. Out of the two type of errors the errorσ is the controlling one. This is predicatable
as the cracking introduces the Rnormal residual which plays a dominant role in the internal
RC problem.

0
10
CT1−TR−Strain CT2−TR−Strain
−5
10

−10
10

int
errorε
−15
10

−20
10

−25
10
50 100 150 200 0 50 100 150 200
Internal Iterations Internal Iterations
(a) (b)

5
10
CT1−TR−Stress CT2−TR−Stress
0
10

−5
10
int
errorσ

−10
10

−15
10

−20
10
50 100 150 200 0 50 100 150 200
Internal Iterations Internal Iterations
(c) (d)

Figure 5.65: Internal Strain (errorε ) and stress (errorσ ) internal errors, CT1-TR-Strain
and CT2-TR-Strain
237

Figure 5.66 compares cases CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-Stress (all


stress driven). Cases CT1-SF-Stress and CT1-TR-Stress can complete the analysis, while
case CT1-TF-Stress crashes at an early stage.

−10
x 10

0.5

ε [mm/mm]
0

x
−0.5

1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
σ [MPa] σ [MPa]
y y
(a) (b)
−10
0
CT1−SF−Stress
CT1−TF−Stress
−0.5 CT1−TR−Stress
[rad]

−1
12c
θ

−1.5

−2
1 2 3 4 5 6 7 0 2 4 6 8
σ [MPa] σ [MPa]
y y
(c) (d)

Figure 5.66: RC response, CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-Stress


238

Figure 5.67 shows how the CT1-TR-Stress keeps the number of iterations much lower
than the other two analysis cases. The CT1-SF-Stress case is able to go through, but the
number of iterations required are much higher. The CT1-TF-Stress crashes after stage 21.

CT1−SF−Stress
CT1−TF−Stress
CT1−TR−Stress

10 20 30 40 50 60
Stage

Figure 5.67: Number of external iterations, CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-


Stress

The analysis times for the test cases shown in Figure 5.67 are listed in Table 5.20.

Table 5.20: Analysis Times

Analysis Case Time [s]


CT1-SF-Stress 64.33
CT1-TF-Stress 90.41
CT1-TR-Stress 4.22
239

Figure 5.68 shows the external error. CT1-TR-Stress is the only one that exhibits
quadratic convergence.

5
10
CT1−TS−Stress
0
10

−5
10

ext
errorσ
CT1−SF−Stress −10
10

−15
10

−20
10
100 200 300 400 500 600 0 1000 2000 3000 4000
External Iterations External Iterations
(a) (b)

CT1−TR−Stress

20 40 60 80 100
External Iterations
(c)

Figure 5.68: External error, CT1-SF-Stress, CT1-TF-Stress, and CT1-TR-Stress


240

Figure 5.69 shows the RC response for the test cases CT1-TR-Stress, and CT2-TR-
Stress.

−6
x 10
2

−2

ε [MPa]
−4

−6

x
−8

−10

−12
1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
σ [MPa] σ [MPa]
y y
(a) (b)
−18
2
CT1−TR−Stress
CT2−TR−Stress
[MPa] 1

0
12c
θ

−1

−2
1 2 3 4 5 6 7 0 2 4 6 8
σ [MPa] σ [MPa]
y y
(c) (d)

Figure 5.69: RC response, CT1-TR-Stress, and CT2-TR-Stress


241

Figure 5.70 shows the concrete fc1 stress response for the test cases CT1-TR-Stress, and
CT2-TR-Stress.

1.4

1.2

[MPa]
0.8

0.6

c1
f
0.4

0.2

0
1 2 3 4 5 6 7 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
σ [MPa] ε [mm/mm] x 10
−3
y c1
(a) (b)

0.025
CT1−TR−Stress
0.02 CT2−TR−Stress

0.015
[MPa]

0.01
c1
f

0.005

−0.005
1 2 3 4 5 6 7 −5 0 5 10 15
σy [MPa] εLoad [mm/mm] −7
x 10
(c) c1
(d)

Figure 5.70: Concrete fc1 response, CT1-TR-Stress, and CR2-TR-Stress


242

Figure 5.71 shows the concrete fc2 stress response for the test cases CT1-TR-Stress, and
CT2-TR-Stress. The εc2 strain is negative until the Poisson effect vanishes.

0.025

0.02

0.015

[MPa]
0.01

c2
f
0.005

−0.005
1 2 3 4 5 6 7 −12 −10 −8 −6 −4 −2 0 2
σ [MPa] ε [mm/mm] −6
x 10
y c2
(a) (b)

0.025

0.02

0.015
[MPa]

0.01
c1
f

0.005

−0.005
1 2 3 4 5 6 7 −2 0 2 4 6 8 10 12
σ [MPa] εLoad [mm/mm] −7
x 10
y c1
(c) (b)

Figure 5.71: Concrete fc2 response, CT1-TR-Stress, and CR2-TR-Stress


243

Figure 5.72 shows the in-plane steel layer 2 (y direction) for the test cases CT1-TR-
Stress, and CT2-TR-Stress. Figure 5.72(b) shows that the steel layer yields at the crack,
but not at the average zone.

400
350
300
250

[MPa]
200

s2
150

f
100
50
0
1 2 3 4 5 6 7 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8 2
σ [MPa] ε [mm/mm] x 10
−3
y s2
(a) (b)

500

400
CT1−TR−Stress
CT2−TR−Stress
[MPa]

300
scr2

200
f

100

0
1 2 3 4 5 6 7 0 0.005 0.01 0.015 0.02 0.025
σ [MPa] ε [mm/mm]
y scr2
(c) (b)

Figure 5.72: Steel fs2 and fscr2 , CT1-TR-Stress, and CR2-TR-Stress


244

Figure 5.73 shows the in-plane steel layer 1 (x direction) for the test cases CT1-TR-Stress,
and CT2-TR-Stress. Figure 5.73(a) shows that the in-plane steel layer 1 is in compression
until the Poisson effect vanishes.

0.5

−0.5

[MPa]
−1

s1
f
−1.5

−2

−2.5
1 2 3 4 5 6 7 −12 −10 −8 −6 −4 −2 0 2
σ [MPa] ε [mm/mm] x 10
−6
y s1
(a) (b)
−33 −28
x 10
9
8
7
6
[MPa]

5
4
scr1

3
f

2
1
0
−1
1 2 3 4 5 6 7 −0.5 0 0.5 1 1.5 2 2.5 3 3.5 4 4.5
σ [MPa] ε [mm/mm] x 10
−33
y scr1
(c) (d)

Figure 5.73: Steel fs1 and fscr1 , CT1-TR-Stress, and CT2-TR-Stress


245

Figure 5.67 shows the number of external iterations for the test cases CT1-TR-Stress,
and CT2-TR-Stress. The iterations needed are similar for both cases.

CT1−TR−Stress
CT2−TR−Stress

10 20 30 40 50 60
Stage

Figure 5.74: Number of External Iterations, CT1-TR-Stress, and CT2-TR-Stress

The analysis times for the test cases shown in Figure 5.74 are listed in Table 5.21.

Table 5.21: Analysis Times

Analysis Case Time [s]


CT1-TR-Stress 4.22
CT2-TR-Stress 9.22
246

Figure 5.68 shows the external error for test cases CT1-TR-Stress, and CT2-TR-Stress.
The error history is identical, expect at the very beginning where the Poisson effect makes
the convergence for CT2-TR-Stress a bit slower.

CT1−TR−Stress

10 20 30 40 50 60 70 80 90 100
External Iterations
(a)

CT2−TR−Stress

10 20 30 40 50 60 70 80 90 100
External Iterations

Figure 5.75: External Error, CT1-TR-Stress, and CT2-TR-Stress


247

5.3.3 Shear tests

The test cases shown in Table 5.22 will be performed The material properties and analysis

Table 5.22: RC material shear tests

Cases Control Type Load Pattern Effects Material [Dc ] Type


Stiffness
CS1-TR-Strain Strain Driven εx : εy : γxy 0:0:1 - Tangent Rotating
CS2-TR-Strain Strain Driven εx : εy : γxy 0:0:1 CS Tangent Rotating
CS3-TR-Strain Strain Driven εx : εy : γxy 0:0:1 CS+DA Tangent Rotating
CS4-TR-Strain Strain Driven εx : εy : γxy 0:0:1 CS+DA+P Tangent Rotating
CS1-SF-Stress StressDriven σx : σy : τxy 0:0:1 - Secant Fixed
CS1-TF-Stress StressDriven σx : σy : τxy 0:0:1 - Tangent Fixed
CS1-TR-Stress StressDriven σx : σy : τxy 0:0:1 - Tangent Rotating
CS2-TR-Stress σx : σy : τxy 0:0:1 CS Tangent Rotating
CS3-TR-Stress σx : σy : τxy 0:0:1 CS+DA Tangent Rotating
CS4-TR-Stress σx : σy : τxy 0:0:1 CS+DA+P Tangent Rotating

parameters used for the cases shown on Table 5.22 are shown on Table 5.23. Test cases CS1-
TR-Strain, CS2-TR-Strain, CS3-TR-Strain, and CS4-TR-Strain are strain driven tests, and
have as a purpose to analyze the stress response and the internal RC state problem conver-
gence. Test cases CS1-SF-Stress , CS1-TF-Stress, and CS2-TR-Stress are stress driven and
they have as a purpose to analyze the different material stiffnesses and [Dc ] types in terms
of convergence. Test cases CS2-TR-Stress, CS3-TR-Stress, and CS4-TR-Stress, are an ex-
tension of the CS1-TR-Stress case to include compression softening (CS), dowel action(DA),
and Poisson Effect(P).
248

Table 5.23: RC material shear test properties

Test Analysis Parameters


Number of steps 50
Error tolerance 1.00E-06
RC material Analysis Parameters
Concrete Compressive model PopovicHS
Concrete Tensile Model LNRBentz
Steel Longitudinal model Raynor
Confinement model Kupfer-Richart
Compression softening Vecchio 1992A
Poisson Compressive model Kupfer-Variable
Poisson Tensile model Linear
Internal Strain tolerance, tolε 1.00E-16
ε
Internal Stress tolerance, tol 1.00E-12
Maximum internal iterations, nmax
iter 100
Material Properties
0
fc -20 MPa
c 0.8
fy 400 MPa
εy 0.002
fsh 440 MPa
εsh 0.012
fu 500 MPa
εu 0.15
nsh 6
ρ1 3.5%
ρ2 4%
α1 0 Deg
α2 90 Deg
Smx , Smy 75 mm
249

Figure 5.76 shows the RC response for the test cases CS1-TR-Strain, CS2-TR-Strain,
, CS3-TR-Strain, , and CS4-TR-Strain. The test cases that include compression softening
are weaker than CS1-TR-Strain, and there is not significant difference between them.

0
−1
−2
−3

σ [MPa]
−4
−5

x
−6
−7
−8
−9
−10
0.005 0.01 0.015 0 0.005 0.01 0.015
γ [mm/mm] γ [mm/mm]
xy xy
(a) (b)

0.8
CS1−TR−Strain
CS2−TR−Strain
[rad]

0.6 CS3−TR−Strain
CS4−TR−Strain
12c

0.4
θ

0.2

0
0.005 0.01 0.015 0 0.005 0.01 0.015
γ [mm/mm] γ [mm/mm]
xy xy
(c) (d)

Figure 5.76: RC response, CS1, CS2, CS3, and CS4-TR-Strain


250

Figure 5.77 shows the concrete response fc1 and fc2 stress response for the test cases
CS1-TR-Strain, CS2-TR-Strain, , CS3-TR-Strain, , and CS4-TR-Strain. It is clear how the
the concrete fc2 stress gets weakened due to the incorporation of the compression softening
(5.77(d) for cases CS2, CS3, and CS4-TR-Strain.

1.4

1.2

[MPa]
0.8

0.6

c1
f
0.4

0.2

0
0.005 0.01 0.015 0 1 2 3 4 5 6 7 8
γ [mm/mm] ε [mm/mm] x 10
−3
xy c1
(a) (b)

0
CS1−TR−Strain
CS2−TR−Strain
−5 CS3−TR−Strain
CS4−TR−Strain
[MPa]

−10
c2
f

−15

−20
0.005 0.01 0.015 −8 −6 −4 −2 0
γ [mm/mm] ε [mm/mm] x 10
−3
xy c2
(c) (d)

Figure 5.77: Concrete response, CS1, CS2, CS3, and CS4-TR-Strain


251

Figure 5.78 shows the concrete shear response for the test cases CS1-TR-Strain, CS2-
TR-Strain, , CS3-TR-Strain, and CS4-TR-Strain. The δslip is lower and vc is lower for the
test cases that include dowel action.

0
−0.02
−0.04
−0.06

v [MPa]
−0.08

c
−0.1
−0.12
−0.14
−0.16
0.005 0.01 0.015 −8 −7 −6 −5 −4 −3 −2 −1 0
γ [mm/mm] δ [mm] x 10
−3
xy slip
(a) (b)

0
CS1−TR−Strain
CS2−TR−Strain
−5 CS3−TR−Strain
CS4−TR−Strain
[MPa]

−10
c2
f

−15

−20
0.005 0.01 0.015 −8 −6 −4 −2 0
γ [mm/mm] εLoad [mm/mm] x 10
−3
xy c2
(c) (d)

Figure 5.78: Concrete shear response, CS1, CS2, CS3, and CS4-TR-Strain
252

Figure 5.79 shows the in-plane steel layer 2 (y direction) stress response for the test cases
CS1-TR-Strain, CS2-TR-Strain, CS3-TR-Strain, and CS4-TR-Strain. The steel stress at the
crack curve is similar to that of cracked concrete in tension.

−18 −13
x 10
0
CS1−TR−Strain
CS2−TR−Strain
CS3−TR−Strain
−0.5
CS4−TR−Strain

[MPa]
−1

s2
f
−1.5

−2
0.005 0.01 0.015 −1.2 −1 −0.8 −0.6 −0.4 −0.2 0
γ [mm/mm] ε [mm/mm] x 10
−18
xy s2
(a) (b)

40

30
[MPa]

20 CS1−TR−Strain
scr2

CS2−TR−Strain
f

CS3−TR−Strain
10 CS4−TR−Strain

0
0.005 0.01 0.015 0 1
γ [mm/mm] ε [mm/mm] x 10
−4
xy scr2
(c) (c)

Figure 5.79: Steel response, fs2 and fscr2 , CS1, CS2, CS3, and CS4-TR-Strain
253

Figure 5.80 shows the in-plane steel layer 1 (x direction)stress response for the test cases
CS1-TR-Strain, CS2-TR-Strain, CS3-TR-Strain, and CS4-TR-Strain. Notice that the steel
is in tension only in the cracking region.

1
0.8
0.6
0.4

[MPa]
0.2
0

s1
−0.2

f
−0.4
−0.6
−0.8
−1
0.005 0.01 0.015 −1 −0.8 −0.6 −0.4 −0.2 0 0.2 0.4 0.6 0.8 1
γ [mm/mm] ε [mm/mm]
xy s1
(a) (b)

35

30

25
[MPa]

20
scr1

15
f

10

0
0.005 0.01 0.015 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
γ [mm/mm] ε [mm/mm] x 10
−4
xy scr1
(c) (d)

Figure 5.80: Steel response, fs1 and fscr1 , CS1, CS2, CS3, and CS4-TR-Strain
254

Figure 5.81 shows the number of internal iterations for test cases CS1-TR-Strain, CS2-
TR-Strain,, CS3-TR-Strain, and CS4-TR-Strain. The number of iterations are higher ini-
tially after cracking, but they go down for later stages.

CS1−TR−Strain
CS2−TR−Strain
CS3−TR−Strain
CS4−TR−Strain

10 20 30 40 50 60
Stage

Figure 5.81: Number of Internal Iterations, CS1, CS2, CS3, and CS4-TR-Strain

The analysis times for the test cases shown in Figure 5.81 are listed in Table 5.24.

Table 5.24: Analysis Times

Analysis Case Time [s]


CS1-TR-Strain 2.66
CS2-TR-Strain 2.73
CS3-TR-Strain 2.66
CS4-TR-Strain 2.86
255

Figure 5.82 shows the internal errors for test cases CS1-TR-Strain and CS2-TR-Strain.

0
10
CS1−TR−Strain CS2−TR−Strain
−5
10

−10
10

int
errorε
−15
10

−20
10

−25
10
50 100 150 200 250 0 50 100 150 200 250
Internal Iterations Internal Iterations
(a) (b)

5
10
CS1−TR−Stress CS2−TR−Stress
0
10

−5
10
int
errorσ

−10
10

−15
10

−20
10
50 100 150 200 250 0 50 100 150 200 250
Internal Iterations Internal Iterations
(c) (d)

Figure 5.82: Internal Error, CS1, and CS2-TR-Strain


256

Figure 5.83 shows the internal errors for test cases CS3-TR-Strain and CS4-TR-Strain.
The convergence for all the test cases is quadratic.

0
10
CS3−TR−Strain CS4−TR−Strain

−5
10

int
errorε
−10
10

−15
10

−20
10
50 100 150 200 250 0 50 100 150 200 250
Internal Iterations Internal Iterations
(a) (b)

5
10
CS3−TR−Stress CS4−TR−Stress
0
10

−5
10
int
errorσ

−10
10

−15
10

−20
10
50 100 150 200 250 0 50 100 150 200 250
Internal Iterations Internal Iterations
(c) (d)

Figure 5.83: Internal Error, CS3, and CS4-TR-Strain


257

Figure 5.84 shows the RC response for the test cases CS1-SF-Stress, CS1-TF-Stress,
and CS1-TR-Stress. Only the CS1-SF-Stress and CS1-TR-Stress can complete the analysis
correctly while the CS1-TF-Stress fails to converge.

−4
x 10
12

10

ε [mm/mm]
6

x
2

−2
2 3 4 5 6 7 8 9 0 1 2 3 4 5 6 7 8 9
τ [MPa] τ [MPa]
xy xy
(a) (b)

6
CS1−SF−Stress
[rad]

CS1−TF−Stress
4 CS1−TR−Stress
12c
θ

0
0.0020.0040.0060.008 0.01 0.0120.0140.016 0 2 4 6 8 10
γ [mm/mm] τ [MPa]
xy xy
(d)

Figure 5.84: RC response, CS1-SF-Stress, CS1-TF-Stress, CS1-TR-Stress


258

Figure 5.84 shows the number if external iterations for test cases CS1-SF-Stress, CS1-
TF-Stress, and CS1-TR-Stress. The number of iterations for CS1-TR-Stress are far lower
than for the other two cases.

CS1−SF−Stress
CS1−TF−Stress
CS1−TR−Stress

10 20 30 40 50 60
Stage

Figure 5.85: Number of Internal Iterations, CS1-SF-Stress, CS1-TF-Stress, CS1-TR-Stress

The analysis times for the test cases shown in Figure 5.85 are listed in Table 5.25.

Table 5.25: Analysis Times

Analysis Case Time [s]


CS1-SF-Stress 92.59
CS1-TF-Stress 233.97
CS1-TR-Stress 7.13
259

Figure 5.87 shows the error for test cases CS1-SF-Stress, CS1-TF-Stress, and CS1-TR-
Stress. The only test case that converges quadratically is CS-TR-Stress shown in Figure
5.86(c).

5
10
CS1−SF−Stress CS1−TF−Stress

0
10

ext
errorσ
−5
10

−10
10
200 400 600 800 0 500 1000 1500 2000 2500
External Iterations External Iterations
(a) (b)

CS1−TR−Stress

20 40 60 80 100 120
External Iterations
(c)

Figure 5.86: Internal Error, CS1-SF-Stress, CS1-TF-Stress, CS1-TR-Stress


260

Figure 5.87 shows the RC response for the test cases CS1-TR-Stress, CS2-TR-Stress(CS),
CS3-TR-Stress(CS+DA), and CS4-TR-Stress(CS+DA+P). The softening caused by the
compression softening effect can be observed in Figure 5.87.

−4
x 10
12

10

ε [MPa]
6

x
2

−2
2 3 4 5 6 7 8 9 0 1 2 3 4 5 6 7 8 9
τ [MPa] τ [MPa]
xy xy
(a) (b)

0.8

CS1−TR−Stress
CS2−TR−Stress
0.6
CS3−TR−Stress
CS4−TR−Stress
[rad]

0.4
12c
θ

0.2

0
0.002 0.004 0.006 0.008 0.01 0.012 0.014 0 2 4 6 8 10
γ [mm/mm] τ [MPa]
xy xy
(c) (d)

Figure 5.87: RC response, CS1, CS2, CS3, CS4-TR-Stress


261

Figure 5.88 shows the concrete response for the test cases CS1-TR-Stress, CS2-TR-
Stress, CS3-TR-Stress, and CS4-TR-Stress. The effect of the compression softerning on the
concrete compressive is shown in Figure 5.88(c).

1.5

[MPa]
c1
f
0.5

0
2 3 4 5 6 7 8 9 0 1 2 3 4 5 6 7 8
τ [MPa] ε [mm/mm] x 10
−3
xy c1
(a) (b)

0
CS1−TR−Stress
CS2−TR−Stress
−5 CS3−TR−Stress
CS4−TR−Stress
[MPa]

−10
c2
f

−15

−20
2 3 4 5 6 7 8 9 −6 −5 −4 −3 −2 −1 0
τ [MPa] ε [mm/mm] x 10
−3
xy c2
(c) (d)

Figure 5.88: Concrete response, CS1, CS2, CS3, CS4-TR-Stress


262

Figure 5.89 shows the concrete shear response for the test cases CS1-TR-Stress, CS2-
TR-Stress, CS3-TR-Stress, and CS4-TR-Stress. The incorporation of compression softening
increses the slip δslip and the cracking width wcr .

−0.02

−0.04

vc [MPa]
−0.06

−0.08

−0.1

−0.12
2 3 4 5 6 7 8 9 −8 −7 −6 −5 −4 −3 −2 −1 0
τxy [MPa] δ [mm] −3
x 10
slip
(a) (b)

0
CS1−TR−Stress
CS2−TR−Stress
−5 CS3−TR−Stress
CS4−TR−Stress
[MPa]

−10
c2
f

−15

−20
2 3 4 5 6 7 8 9 −6 −5 −4 −3 −2 −1 0
τ [MPa] εLoad [mm/mm] −3
x 10
xy c2
(c) (d)

Figure 5.89: Concrete shear response, CS1, CS2, CS3, CS4-TR-Stress


263

Figure 5.90 shows the steel response of in-plane steel layer 2 (y direction) for test cases
CS1-TR-Stress, CS2-TR-Stress, CS3-TR-Stress, and CS4-TR-Stress. It is possible to see in
Figure 5.90(c) the increase in the steel strains in the cracking region after cracking occurs.

200

150

[MPa]
100

s2
50

f
0

−50
2 3 4 5 6 7 8 9 −2 0 2 4 6 8 10
τ [MPa] ε [mm/mm] −4
x 10
xy s2
(a) (b)

250

200
[MPa]

150
scr2

100
f

50

0
2 3 4 5 6 7 8 9 0 0.2 0.4 0.6 0.8 1 1.2
τ [MPa] ε [mm/mm] −3
x 10
xy scr2
(c) (b)

Figure 5.90: Steel response,fs2 and fscr2 , CS1, CS2, CS3, CS4-TR-Stress
264

Figure 5.91 shows the steel response of in-plane steel layer 1 (x direction) for test cases
CS1-TR-Stress, CS2-TR-Stress, CS3-TR-Stress, and CS4-TR-Stress. It is possible to see in
Figure 5.91(c) the increase in the steel strains in the cracking region after cracking occurs.

250

200

150

[MPa]
100

s1
f
50

−50
2 3 4 5 6 7 8 9 −2 0 2 4 6 8 10 12
τ [MPa] ε [mm/mm] x 10
−4
xy s1
(a) (b)

250

200
[MPa]

150
scr1

100
f

50

0
2 3 4 5 6 7 8 9 0 0.2 0.4 0.6 0.8 1 1.2 1.4
τ [MPa] ε [mm/mm] x 10
−3
xy scr1
(c) (b)

Figure 5.91: Steel response,fs1 and fscr1 , CS1, CS2, CS3, CS4-TR-Stress
265

Figure 5.85 shows the number of external iterations for test cases CS1-TR-Stress, CS2-
TR-Stress, CS3-TR-Stress, and CS4-TR-Stress. The number of iterations tend to increase
as more effects are added.

CS1−TR−Stress
CS2−TR−Stress
CS3−TR−Stress
CS4−TR−Stress

10 20 30 40 50 60
Stage

Figure 5.92: Number of External Iterations, CS1, CS2, CS3, CS4-TR-Stress

The analysis times for the test cases shown in Figure 5.92 are listed in Table 5.26.

Table 5.26: Analysis Times

Analysis Case Time [s]


CS1-TR-Stress 7.13
CS2-TR-Stress 9.36
CS3-TR-Stress 10.08
CS4-TR-Stress 10.03
266

Figure 5.86 shows the errors for the test cases CS1-TR-Stress, CS2-TR-Stress, CS3-TR-
Stress, and CS4-TR-Stress. The convergence is quadratic for all of them.

0
10
CS1−TR−Stress CS2−TR−Stress

−5
10

error[MPa]
−10
10

−15
10
20 40 60 80 100 120 0 50 100 150
External Iterations External Iterations
(a) (b)

0
10
CS3−TR−Stress CS4−TR−Stress

−5
10
error[MPa]

−10
10

−15
10
50 100 150 0 50 100 150
External Iterations External Iterations
(c) (d)

Figure 5.93: External Error, CS1, CS2, CS3, CS4-TR-Stress


267

5.4 RC material model in FEM analysis tests

To evaluate the accuracy with which the DSFM-TG predicts the response of RC elements,
simulated and measured results were compared for three panels tested by Vecchio and Collins
(1982) at the University of Toronto. The panels dimensions are 890 mm × 890 mm × 70
mm and were reinforced with uniformly distributed wire mesh reinforcement. The panels
were subjected to pure shear and and pure uniaxial compression under load control. The
panels were identified by Vecchio and Collins as

• Panel PV10

• Panel PV16

• Panel PV17

Panel PV10 and PV16 were subjected to pure shear, while panel PV17 was tested under
pure uniaxial compression. To simulate these tests, the DSFM-TG was used with a single
isoparametric four-node finite element with four integration points. The external nodal
forces were applied to match the stress field used in the testing. Figures 5.94 and 5.95 show
the boundary condition for panels PV10, PV16, and PV17.

Figure 5.94: Panels PV10 Sand 16 boundary conditions an external loading

Load was applied using an Arch Length. Using an cylindrical Arch Length method
instead of monotonically load control allows the analysis to overcome zones with negative
268

Figure 5.95: Panel PV17 boundary conditions an external loading

stiffness such as right after cracking or post-peak response. The parameters used in the
Arch Length method analysis are shown in Table 5.27

Table 5.27: Arch Length Method Analysis Parameters

Panel Number Number of ∆l ∆ref


l
ined
Tolerance
of Steps Refined steps
PV10 40 20 1 0.05 1.00E-03
PV16 75 20 0.5 0.05 1.00E-03
PV17 50 10 0.25 0.05 1.00E-03

where ∆ref
l
ined
is the step size in the first number of refined steps, and ∆l is the step
size in the subsequent steps. The different ∆ref
l
ined
steps intent to capture phenomenon
that occurs at load levels of loading such as cracking. The maximum number of iterations
refer to the maximum number of iterations used to solve the external equilibrium problem.
The analysis parameters used for the internal RC problem are the same for all the panel
analysis and they are shown in Table 5.28
The experimental and DSFM-TG results were also compared with Vector2 results. The
Vector2 analysis was performed using the default constitutive models and analysis options.
269

Table 5.28: Internal RC Problem Analysis Parameters

tolε 1.00E-16
tolσ 1.00E-12
nmax
iter 100
270

5.4.1 Panel PV10

Panel PV10 was subjected to pure monotonic shear. The panel material properties presented
by Vecchio and Collins (1982) are shown in Table 5.29

Table 5.29: Panel PV10 Material Properties


0
fc -14.5 Mpa
εco -2.7e-3
fyy 276 MPa
fyx 276 MPa
ρ1 = ρx 1.7%
ρ2 = ρy 1%

where x and y-directions are as shown in Figure 5.94. The experimental results show that
Panel PV10 developed initially cracks oriented 45◦ from x axis (θ12c ≈ 45◦ ), spaced between
75 and 100 mm at a shear stress of ∼ 1.86 MPa. As the test approached the ultimate
shear stress of ∼ 3.87 MPa, the in-plane steel layer 2 (y-direction) yielded, and the crack
orientation shifted to be more acute to the x -direction . The experiment was ended one the
panel failed in sliding shear prior to yielding of the in-plane steel layer 1 (x -direction).
The first group of tests of PV10 was intended to compare the performance of the secant
versus tangent material stiffness and fixed versus rotating concrete material stiffness matrix
[Dc ]. Table 5.30 shows the tests included in this group and the analysis times for each. where

Table 5.30: Panel PV10, First group of tests

Cases [Dc] type Models Analysis time [s]


PV10-SF-Arch Secant Fixed PopovicHS, LNRBentz, Raynor 214.5
PV10-TF-Arch Tangent Fixed PopovicHS, LNRBentz, Raynor 1.8
PV10-TR-Arch Tangent Rotating PopovicHS, LNRBentz, Raynor 30.5

the term ”Arch” is used for the case labeling to indicate that an ”Arch Length” method was
271

used to perform the test cases. The PopovicHS, LNRBentz, and Raynor are the concrete
compressive, concrete tensile, and longitudinal steel constitutive models, respectively. The
only test case able to complete the analysis is PV10-TR-Arch. Test cases PV10-SF-Arch
and PV10-TF-Arch crashed after steel yielding and cracking, respectively. The analysis
time required by test case PV10-TR-Arch is significantly lower than that of test case PV10-
SF-Arch. Test case PV10-TF-Arch failed much earlier than the other two test cases, and
could not continue beyond cracking. The shear stress-strain response simulated using these
DSFM-TG results and as computed from experimental data symbols are presented in Figure
5.96.

4.5

3.5

3
PV10−SF−Arch
2.5
PV10−TF−Arch
2 PV10−TR−Arch
1.5 Experimental
Vector2
1

0.5

0
0 0.005 0.01 0.015 0.02 0.025
γxy [mm/mm]

Figure 5.96: Panel PV10 Shear Response, First group of tests

The only test case that could be continued beyond yield was PV10-TR-Arch. Test case
PV10-SF-Arch failed to continue after yielding of in-plane steel layer 1, and PV10-TF-Arch
failed to continue after cracking. Test case PV10-TR-Arch reaches a higher peak shear stress
than experimental results and Vector2. This first group of tests probes that using tangent
material stiffness in combination with a rotating [Dc ] matrix, as test case PV10TR-Arch
does, is the the most stable alternative to perform the analysis of this panel.
272

The second group of tests are an extension of the PV10-TR-Arch and were intended to
asses the impact of compression softening(CS), dowel action(DA), and Poisson effect(P) on
the simulated response. The objective is to compare the impact of this additional effects on
improving the results from the analysis. The test cases and the analysis times required for
each are listed in Table 5.31.

Table 5.31: Panel PV10, Second group of tests

Cases Additional Effects Model s Analysis time [s]


PV10A1-TR-Arch CS PopovicHS, LNRBentz, 37.0
Vecchio1992A
PV10B1-TR-Arch CS + DA Hognestad, LNRBentz, 33.8
Vecchio1992A, He-Kwan
PV10C1-TR-Arch CS + DA + P PopovicHS, LNRBentz, 31.5
Vecchio1992A, He-Kwan,
Variable Kupfer

The PopovicHS, LNRBentz, and Vecchio1992A, and He-Kwan models are the concrete
compressive, concrete tensile, and compression softening, and dowel action constitutive mod-
els, respectively. The shear stress-strain response simulated using these DSFM-TG results
and as computed from experimental data symbols are presented in Figure 5.97.
The inclusion of compression softening, dowel action, and Poisson effect had a marginal
effect in the peak shear stress with respect to the peak shear stress obtained previously
from test case PV10-TR-Arch. However, test case PV10C1-TR-Arch results in a higher
shear strain capacity. The analysis times stay fairly close even though additional effects
were added.
273

4.5

3.5

3
PV10−TR−Arch
2.5 PV10A1−TR−Arch
2 PV10B1−TR−Arch
PV10C1−TR−Arch
1.5 Experimental
1 Vector2

0.5

0
0 0.005 0.01 0.015 0.02 0.025
γxy [mm/mm]

Figure 5.97: Panel PV10 Shear Response, Second group of tests


274

The third group of tests included the same effects included in the test case PV10B1-TR-
Arch (compressions softening and dowel action), but using different constitutive models.
PV10B1-TR-Arch was selected as the base case for the third group of tests because com-
pression softening and dowel action could be excepted to affect the results. Thus, a different
set of constitutive models could be expected to significantly affect the results. Compression
softening is important because a pure shear stress creates cracking parallel to the compres-
sion concrete principal direction. Dowel action is important because shearing stresses at
the crack arise due to the different ratio of steel between the x and y-direction.
The test cases are and their analysis times are listed in Table 5.32. where the letter ”B”

Table 5.32: Panel PV10, Third group of tests

Cases Additional Effects Model s Analysis time [s]


PV10B1-TR-Arch CS + DA PopovicHS, LNRBentz, 33.8
Vecchio1992A
PV10B2-TR-Arch CS + DA Hognestad, LNRBentz, 38.6
Vecchio1992A
PV10B3-TR-Arch CS + DA PopovicHS, Vecchio1982, 31.5
Vecchio1992A
PV10B4-TR-Arch CS + DA PopovicHS, LNRBentz, 40.2
Vecchio1986
PV10B5-TR-Arch CS + DA Hognestad, LNRBentz, 35.1
Vecchio1992A,
Reduced tension stiffening

has been used for test cases PB10B2-TR-Arch to PB10B5-TR-Arch to denote that they
represent variations (in terms of consecutive models) of the PV10B1-TR-Arch case. The
shear stress-strain response simulated using these DSFM-TG results and as computed from
experimental data symbols are presented in Figure 5.98.
The only two test cases that differ significantly from PV10B1-TR-Arch are PV10B4-
TR-Arch and PV10B5-TR-Arch. Test case PV4-TR-Arch uses the Hognestad instead of
the PopovicHS model for the simulation of the 1D concrete response in compression. The
Hognestad model predicts more rapid strength loss than does the PopovicHS.
The test PV10B5-TR-Arch predicts the experimental response much better than the
other test cases. This test reduces the tensile stiffness portion of the concrete tensile stress
275

4.5

3.5

3
PV10B1−TR−Arch
2.5
PV10B2−TR−Arch
2 PV10B3−TR−Arch
PV10B4−TR−Arch
1.5 PV10B5−TR−Arch
Experimental
1 Vector2

0.5

0
0 0.005 0.01 0.015 0.02 0.025
γxy [mm/mm]

Figure 5.98: Panel PV10 Shear Response, Third group of tests

response. The lower tension stresses trigger earlier yielding of the steel (same total stress
need to maintain equilibrium), which results in earlier softening of the entire RC element.
None of models available in DSFM-TG incorporate the option of directly reducing the
post-cracking tension response. However this could be achieved indirectly in with LNRBentz
model by increasing the bar diameter, which reduces the concrete and steel bonding, and
consequently the computed the tension stiffening response. This is what was done in test
case PV10B5-TR-Arch.
276

Figures 5.99 to 5.104 present additional response data for the third group of tests. Figure
5.99 shows the average RC stress-strain response for the third group of tests.

4.5
4
3.5
3

[MPa]
2.5
2

xy
τ
1.5
1
0.5
0
0 1 2 3 4 5 6 −2 0 2 4 6 8 10 12
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−3
x y
(a) (b)
−3
x 10
5

0
σ [MPa]

−5
PV10B1−TR−Arch
PV10B2−TR−Arch
y

−10 PV10B3−TR−Arch
PV10B4−TR−Arch
−15 PV10B5−TR−Arch
Experimental
Vector2
−20
0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 1 2 3 4 5
ε [mm/mm] ε [mm/mm]
x x
(c) (d)

Figure 5.99: Panel PV10 RC Response, Third group of tests

Figures 5.99(a) and (b) show that the shear stress-strain are accompanied by non-zero
εx and εy . Figures 5.99(c) and (d) show σx and σy versus εx and εy , respectively. Since the
panels are loaded in pure shear, equilibrium requires that these remain zero. The Vector2
results, however, present some important deviation from the theoretically correct results at
certain point of the analysis, while the DSFM-TG results show much smaller errors.
277

Figure 5.100 shows the concrete stress angle, θ12c , and RC average strain angle, θε ,
versus τxy for the third group of tests. The angle θ12c and θε are equal before cracking
(τxy < 1.25MPa) because there is no slip, δslip . After cracking the value of θε differs from
θε as the crack experiments slip, δslip . Also, these angles vary during the testing because
the slip, δslip , does not remain constant. The DSFM-TG assumes that θsc and θε are zero
when the material is unloaded (τxy = 0), which explains why the DSFM-TG plot lines point
to zero at the beggining of the plot. The DSFM-TG results follow a similar pattern as the
experimental results (Figure 5.100(a)). The θε DSFM-TG results shown in Figure 5.100(b)
are also comparable to that of Vector2.

PV10B1−TR−Arch
PV10B2−TR−Arch
PV10B3−TR−Arch
PV10B4−TR−Arch
PV10B5−TR−Arch
Experimental
Vector2

0.5 1 1.5 2 2.5 3 3.5 4 4.5


τ [MPa]
xy
(a)

0.5 1 1.5 2 2.5 3 3.5 4 4.5


τ [MPa]
xy
(b)

Figure 5.100: Panel PV10 θ12c and θε angles, Third group of tests
278

Figure 5.101 shows the concrete response for the third group of tests including concrete
principal strain versus imposed shear stresses and concrete principal stress-strain response,
fc1 vs. εc1 and fc2 vs. εc2 .
Figure 5.101(a) shows an experimental εc1 similar to that of the DSFM-TG test cases,
but with the difference for τxy < 3.9M P a. For higher τxy , the experimental and simulated
results vary greatly.
Figure 5.101(b) shows tensile concrete stresses lower than that of the experimental. The
Vector 2 result completely drop the fc1 stress for εc1 > 0.008, while the DSFM-TG results
maintain the stresses. The case PV10B5-TR-Arch, exhibits considerably lower fc1 stresses,
which is explained by the reduced stiffening response used for this test case.
Figure 5.101(c) shows an experimental εc2 similar to that of the DSFM-TG test cases.
The test DSFM-TG compressive concrete stresses fc2 shown in Figure 5.101(d) are higher
than that of the experimental, while the tensile concrete stresses shown in Figure 5.101(b)
are lower. The higher compressive stresses could in part explain why the RC shear response
is higher than the experimental.
The reduction achieved in the post-cracking tensile stress in test case PV10B5-TR-Arch,
which is considerable lower than that the experimental results generates results close to
that of the experiment. In reality, this artificial reduction of the tensile concrete response
compensating for one of the shortcoming of the the DSFM-TG . The DSFM-TG does not
account for the elongation of the steel at the crack when calculating the total RC strain,
this means that only the steel strains in the average zone can have an effect in the RC total
strains. Reducing the tensile response accelerated the yielding of the steel in the average
region. This earlier yielding softens the system at a lower shear stress, which allows a better
match of the between the analysis and the experimental results.
Vector 2 results are also close to that of the experiment, because the low fc1 stresses (no
residual, Figure 5.101(d)) allow steel yielding in the average zone to occur for a τxy value
closer to the experimental peak.
279

0.02 1.6
1.4
0.015 1.2
1

f [MPa]
0.01 0.8

c1
0.6
0.005 0.4
0.2
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 0.005 0.01 0.015 0.02
τxy [MPa] εc1 [mm/mm]
(a) (b)
−3
x 10 PV10B1−TR−Arch
0 0
PV10B2−TR−Arch
−0.5 PV10B3−TR−Arch
−1 −2 PV10B4−TR−Arch
−1.5 PV10B5−TR−Arch
f [MPa]

−4 Experimental
−2
Vector2
−2.5
c2

−6
−3
−3.5 −8
−4
−4.5 −10
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 −5 −4 −3 −2 −1 0
τxy [MPa] εc2 [mm/mm] −3
x 10
(c) (d)

Figure 5.101: Panel PV10 Concrete fc1 and fc2 stress response, Third group of test
280

Figure 5.102 shows the slip, δslip , versus the shear stress, τxy ; the shear stress at the
crack due to aggregate interlock, vc , versus slip, δslip ; the crack width, wcr , versus the τxy ;
and the crack spacing, §cr , versus τxy .
Vecchio and Collins do not provide these data, as they are not easily measured and
cannot be directly computed from measured quantities. Thus, no experimental data are
included in Figure 5.102. Figure 5.102(a) shows the slip at the crack surface, δslip , increases
with the shear stress up to point of steel yielding at the crack, at which point it decreases.
This is because the concrete shear stress at the crack, vc , is originated by the difference
between the steel stresses at the crack and at the average zone.
The shear stress on the crack, vc , (Figure 5.102(b)) increases with δslip and then returns
to almost zero as the δslip lowers. Vecchio and Collins (1982) noted that the crack width
reached about 0.5 mm near the ultimate shear stress, which is about of the same order
of the results shown in 5.102(c). The experimental observations also note that the crack
spacing Scr was between 50 and 75mm, which is also within the range of the analysis results
(Figure 5.102(d)).
281

0.08 0.8
0.07 0.7
0.06 0.6
0.05 0.5

vc [MPa]
0.04 0.4
0.03 0.3
0.02 0.2
0.01 0.1
0 0
−0.01 −0.1
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 −0.01 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
τxy [MPa] δslip [mm]
(a) (b)

1.4 70

1.2 60

1 50
Scr [mm]

0.8 40 PV10B1−TR−Arch
PV10B2−TR−Arch
0.6 30
PV10B3−TR−Arch
0.4 20 PV10B4−TR−Arch
PV10B5−TR−Arch
0.2 10 Vector2
0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 0 1 2 3 4 5
τxy [MPa] τxy [MPa]
(c) (d)

Figure 5.102: Panel PV10 Cracking Region Response, Third group of tests
282

Figure 5.103 shows the panel PV10 in-plane steel layer 2 (y-direction) stress response.
No experimental data is provided for the steel stresses. The in-plane steel layer 2 in the
average region yields earlier for tests PV10B5-TR-Arch than for the other DSFM-TG anal-
ysis (Figure 5.103(a)) due to the weaker concrete tensile response. There is no significant
difference in the strain εscr2 between the different test cases (Figure 5.103(c)).

300

250

200 PV10B1−TR−Arch
PV10B2−TR−Arch
fs2 [MPa]

150
PV10B3−TR−Arch
100 PV10B4−TR−Arch

50 PV10B5−TR−Arch
Vector2
0

−50
−2 0 2 4 6 8 10 12
2 3 4 5 ε [mm/mm] −3
τ [MPa] s2 x 10
xy (b)
(a)

350 350

300
300
250
250
200
[MPa]
[MPa]

200
150
scr2
scr2

150
100
f
f

100 50

50 0

0 −50
2 3 4 5 0 0.02 0.04 0.06 0.08 0.1 0.12 0 2 4 6
τ [MPa] ε [mm/mm] τ [MPa]
scr2 xy
xy
(c) (d) (e)

Figure 5.103: Panel PV10 In-plane Steel Layer 2 Response, Third group of tests
283

Figure 5.104 shows the panel PV10 in-plane steel layer 1 (x direction) stress response.
No experimental data is provided for the steel stresses. The in-plane steel steel layer 2 in
the average region yields at about the same time for a all the test cases tests as shown in
Figure 5.103(a). It is clear that yielding of the in-plane steel layer 2 in the average region
is what causes the softening of the panel response.

300 0.07

250 0.06

200 0.05

[mm/mm]
[MPa]

150 0.04

100 0.03
s1

scr1
f

ε
50 0.02

0 0.01

−50 0
1 2 3 4 5 −2 0 2 4 6 0 1 2 3 4 5
τ [MPa] ε [mm/mm] x 10−3 τ [MPa]
xy s1 xy
(a) (b) (c)

400 PV10B1−TR−Arch
PV10B2−TR−Arch
300 PV10B3−TR−Arch
PV10B4−TR−Arch
PV10B5−TR−Arch
[MPa]

200 Vector2
scr1

100
f

−100
0.02 0.04 0.06 0.08 0 2 4 6
ε [mm/mm] τ [MPa]
scr1 xy
(d) (e)

Figure 5.104: Panel PV10 In-plane Steel Layer 1 Response, Third group of tests
284

5.4.2 Panel PV16

Panel PV16 was subjected to pure monotonic shear. The panel material properties presented
by Vecchio and Collins (1982) are shown in Table 5.33.

Table 5.33: Panel PV16 Material Properties


0
fc -21.7 Mpa
εco -2.0e-3
fyx 255 MPa
fyy 255 MPa
ρ1 = ρx 0.74%
ρ1 = ρy 0.74%

where x and y-directions are as shown in Figure 5.94. The panel developed crack oriented
45◦ from the x -direction (θ12c = 45◦ ). The reinforcement yielded immediately after cracking,
The initial cracks opened up to about 5 mm.
The amount of reinforcement is equal in each direction; combined with loading in pure
shear, this results in cracking oriented at 45◦ from the x -axis with no slip, δslip = 0, along
the crack.
285

For panel PV16, the first group of tests was intended to asses the the performance of
secant versus tangent material stiffness and fixed versus rotating concrete material stiffness
matrix [Dc ]. Table 5.34 shows the tests included in this group and the analysis times for
each one them.

Table 5.34: Panel PV16, First group of tests

Cases [Dc] type Models Analysis time [s]


PV16-SF-Arch Secant Fixed PopovicHS, LNRBentz, Raynor 214.5
PV16-TF-Arch Tangent Fixed PopovicHS, LNRBentz, Raynor 1.8
PV16-TR-Arch Tangent Rotating PopovicHS, LNRBentz, Raynor 30.5

where the PopovicHS, LNRBentz, and Raynor are the compressive concrete, tensile concrete,
and longitudinal steel constitutive models. The results are presented in Figure 5.105. The
only test case able to complete the analysis is PV16-TR-Arch. Test case PV16-SF-Arch
fails when the steel at the crack yields, and PV16-TF-Arch after cracking. The maximum
shear stress reached by PV16-TR-Arch is higher than that of the experimental and Vector2
results. The PV16-TR-Arch analysis time is considerable lower than that of the test case
PV16-SF-Arch, even though the analysis of the latest fails earlier than PV16-TR-Arch.
Analysis PV16-TF is unabl to continue after crack, which explain the low analysis time
(it completes only a minimal fraction of the analysis). This first group of test probe that
using a tangent material sitffness in combinaiton with a rotating [Dc ] matrix, as test case
PV16-TF-Arch does, is the most stable alternative to perform the analysis of this panel.
286

3.5
PV16−SF−Arch
PV16−TF−Arch
3
PV16−TR−Arch
Experimental
2.5 Vector2

1.5

0.5

0
0 0.005 0.01 0.015 0.02 0.025
γxy [mm/mm]

Figure 5.105: Panel PV16 Shear Response, First group of tests


287

The second group of tests are an extension of PV16-TR-Arch to include additional effect
such as compression softening (CS), dowel action(DA), and Poisson effect (P). The objective
of this group of test is to compare the impact of this additional effects on improving the
results from the analysis. The test cases and their analysis times are listed in Table 5.35.

Table 5.35: PV16 Panel, Second group of tests

Cases Additional Effects Models Analysis time [s]


PV16- TR-Arch - PopovicHS, LNRBentz, 22.8
Vecchio1986
PV16A1-TR-Arch CS Hognestad, LNRBentz, Raynor 25.7
Vecchio1986
PV16B1-TR-Arch CS+DA PopovicHS, Vecchuo1982, 26.7
Vecchio1986, He-Kwan
PV16C1-TR-Arch CS+DA+P PopovicHS, LNRBentz, 28.8
Vechio1986, He-Kwan,
Variable Kupfer

Shear stresses versus strain results are presented in Figure 5.106. No significant difference
is observed in the tests that include compression softening, dowel action and Poisson effect
with respect to the result cases obtained previously from test case PV10-TR-Arch. The
analysis times stay fairly close even despite the incorporation of additional effects.
288

2.5

PV16−TR−Arch
2
PV16A1−TR−Arch
PV16B1−TR−Arch
PV16C1−TR−Arch
1.5 Experimental
Vector2

0.5

0
0 0.005 0.01 0.015 0.02 0.025
γxy [mm/mm]

Figure 5.106: Panel PV16 Shear Response, Second Group of Tests


289

The third group of tests uses the same effects included in the test case PV16B1-TR (com-
pression softening and dowel action), but using different constitutive models. PV16B1-TR-
Arch was selected as the base case for the third group of tests because compression softening
and dowel action could be excepted to affect the results. Thus, a different set of consti-
tutive models could be expected to significantly affect the results. Compression softening
is important because a pure shear stress state crates cracking parallel to the compression
concrete principal direction. Dowel action is important because it helps stabilize the slip at
the crack, δslip . The test cases and the analysis times for each are listed in Table 5.36.

Table 5.36: PV16 Panel, Third group of tests

Cases Additional Effects Models Analysis time [s]


PV16B1-TR-Arch CS+DA PopovicHS, LNRBentz, Raynor 26.7
Vecchio1986
PV16B2-TR-Arch CS+DA Hognestad, LNRBentz, Raynor 26.0
Vecchio1986
PV16B3-TR-Arch CS+DA PopovicHS, Vecchuo1982, 26.9
Vecchio1986
PV16B4-TR-Arch CS+DA PopovicHS, LNRBentz, 21.4
Vechio1986, Reduced Tension Stiff.

Shear stress-strain sresults are presented in Figure 5.107. The only test case that differs
significantly from PV16B1-TR-Arch is PV16B4-TR-Arch. This test case uses the reduced
tension stiffening model employed in PV10B5-TR-Arch, the results is earlier yielding o
the in-plane steel layers in the average zone. The weakening of the tension stiffening was
achieved in the LNRBentz model by specifying a bar diameter greater than the actual
diameter, which reduces the steel-concrete bonding parameter used to calculate the tension
stiffening stress.
290

2.5

2
PV16B1−TR−Arch
PV16B2−TR−Arch
PV16B3−TR−Arch
1.5 PV16B4−TR−Arch
Experimental
Vector2
1

0.5

0
0 0.005 0.01 0.015 0.02 0.025
γxy [mm/mm]

Figure 5.107: Panel PV16 Shear Response, Third Group of Tests


291

Figures 5.108 to 5.113 present additional response data for the third group of tests.
Figure 5.108 shows the average RC stress-strain response for the third group of tests.

2.5

[MPa]
1.5

xy
1

τ
0.5

0
0 2 4 6 8 10 12 −2 0 2 4 6 8 10 12
ε [mm/mm] x 10
−3 ε [mm/mm] x 10
−3
x y
(a) (b)

0.01

−0.01
PV16B1−TR−Arch
σ [MPa] PV16B2−TR−Arch
−0.02
PV16B3−TR−Arch
y

PV16B4−TR−Arch
−0.03
Experimental
Vector2
−0.04

−0.05
0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
τ [mm/mm] τ [mm/mm]
xy xy
(c) (d)

Figure 5.108: Panel PV16 RC Response, Third Group of Tests

Figures 5.108(a) and (b) show that the shear stress-strain are accompanied by non-zero
εx and εy . Figures 5.108(c) and (d) show σx and σy versus εx and εy , respectively. Since the
panels are loaded in pure shear, equilibrium requires that these remain zero. The Vector 2
results, however, present some important deviation from the theoretically correct results at
certain point of the analysis, while the DSFM-TG results show much smaller errors.
292

Figure 5.109 shows the concrete stress angle θ12c and the RC average strain angle θε re-
sponse. The angles θ12c and θε are equal because there is no slip δslip (symmetric reinforce-
π
ment about crack), and both are correct θ12c = θε = = 0.79rad. The DSFM-TG assumes
4
that θ12c and θε are zero when the material is unloaded (τxy = 0), which explains why the
DSFM-TG plot line point to zero at the beginning of the plot.
Figure 5.110 shows the concrete response for the third group of tests including concrete
principal strain versus imposed shear stresses and concrete principal stress-strain response,
fc1 vs. εc1 and fc2 vs. εc2 .
Figure 5.110(a) shows an experimental εc1 similar to that of test case PVB5-TR-Arch.
The experimental concrete fc1 curve shown in Figure 5.110(b) seems to lack of the
tension stiffening portion. The reduced tension stiffening of test case PV16B4-TR-Arch
better match the actual tensile response. Again, a weaker tensile response accelerates the
yielding of the reinforcement in the average zone, which is the same behavior observed in
Panel PV16. The tension stiffening was reduced by reducing the concrete - steel bonding.
Figure 5.110(c) shows εc2 strain much larger than for the DSFM-TG and Vector 2 test
cases. This seem to occur because of the way the system unload after the peak shear stress
is reached. The tensile stress fc1 remains fairly constant; thus, the concrete compressive
stress, fc2 , reduction occurs associated to large εc2 strains, as shown in Figure 5.110(d).
293

0.5 1 1.5 2 2.5 3 3.5 4 4.5


τxy [MPa]
(a)

PV10B1−TR−Arch
PV10B2−TR−Arch
PV10B3−TR−Arch
PV10B4−TR−Arch
Vector2

0.5 1 1.5 2 2.5 3 3.5 4 4.5


τ [MPa]
xy
(b)

Figure 5.109: Panel PV16, θ12c and θε angles, Third Group of Tests
294

1.6
1.4
1.2

fc1[MPa]
1
0.8
0.6
0.4
0.2
0
0.5 1 1.5 2 2.5 0 0.005 0.01 0.015 0.02 0.025
τxy [MPa] εc1 [mm/mm]
(a) (b)
PV16B1−TR−Arch
0
PV16B2−TR−Arch
PV16B3−TR−Arch
−2 PV16B4−TR−Arch

f [MPa]
Experimental
c2 Vector2
−4

−6
0.5 1 1.5 2 2.5 −3 −2 −1 0
τxy [MPa] εc2 [mm/mm] x 10
−3

(c) (d)

Figure 5.110: Panel PV16 Concrete fc1 and fc2 stress response, Third Group of Tests
295

Figure 5.111 shows the slip, δslip , versus the shear stress, τxy ; the shear stress at the
crack due to interlock, vc ; the crack width, wcr , versus the τxy ; and the crack spacing, Scr ,
versus τxy .
No experimental curves are provided for the information presented in Figure 5.111.
Figure 5.111(a) and (b) show the values of δslip and vc of nominally zero, which is correct,
since no shear at the crack occurs (reinforcement symmetric about the crack). The cracking
width wcr observed during the experiment was about 5 mm, which is higher than what is
predicted by the DSFM-TG tests.
Figure 5.111(c) shows a crack spacing that reaches a value of about 0.8 mm, which is
considerably lower than the 5 mm for for the widest cracks reported from the experiment.
This a results of the uniformly distributed cracking assumed by the DSFM and DSFM-TG .
Figure 5.111(d) shows a prediction of crack spacing, Scr of about ∼ 47 mm (DSFM-TG test
cases), while the crack spacing observed in the experiment varied between 140, and 50 mm.
296

−16 −16
x 10
6
4
2
0

vc [MPa]
−2
−4
−6
−8
−10
0.5 1 1.5 2 2.5 −2 −1.5 −1 −0.5 0 0.5 1
τ [MPa] δ [mm] −16
x 10
xy slip
(a) (b)

100 PV16B1−TR−Arch
PV16B2−TR−Arch
80 PV16B3−TR−Arch
PV16B4−TR−Arch
Vector2
60
S [mm]
cr

40

20

0
0.5 1 1.5 2 2.5 0 0.5 1 1.5 2 2.5
τ [MPa] τ [MPa]
xy xy
(c) (c)

Figure 5.111: Panel PV16 Cracking Region Response, Third Group of Tests
297

Figure 5.112 shows the in-plane steel layer 2 (y-direction) stress response. The data in
Figure 5.113(a) show the average steel strain as a function of the shear demand. The steel
strain is nearly zero until cracking at which point it increases linearly with τxy until the
steel yields at τxy between 2.25 and 2.5 MPa depending on the test case. The steel yields at
a lower shear stress, τxy , in test PV16B4-TR-Arch as shown in Figure 5.113(a), accelerating
the decay in stress for this case.

300 0.06

250 0.05
200

[mm/mm]
0.04
[MPa]

150
0.03
100
s2

scr2
f

0.02

ε
50

0 0.01

−50 0
0.5 1 1.5 2 2.5 −5 0 5 10 15 0 0.5 1 1.5 2 2.5
τ [MPa] ε [mm/mm] x 10−3 τ [MPa]
xy s2 xy
(a) (b) (c)

400

300 PV10B1−TR−Arch
PV10B2−TR−Arch
PV10B3−TR−Arch
[MPa]

200 PV10B4−TR−Arch
Vector2
scr2

100
f

−100
0.01 0.02 0.03 0.04 0.05 0.06 0 2 4
ε [mm/mm] τ [MPa]
scr2 xy
(d) (e)

Figure 5.112: Panel PV16 In-plane Steel Layer 2 Response, Third Group of Tests
298

Figure 5.113 shows the in-plane steel layer 1 (x -direction) stress response. No experi-
mental data is provided for the steel stress. The data shown in Figure 5.113 show that the
response of the in-plane steel layer 1 is essentially the same as that o the in-plane steel layer
2. This is expected as the uniform steel reinforcement layout and the uniform shear loading
of the panel results in identical demand for both in-plane steel layers.

300 0.06

250 0.05

200

εscr1 [mm/mm]
0.04
fs1 [MPa]

150
0.03
100
0.02
50

0 0.01

−50 0
0.5 1 1.5 2 2.5 −5 0 5 10 15 0 0.5 1 1.5 2 2.5
τxy [MPa] εs1 [mm/mm] x 10
−3 τxy [MPa]
(a) (b) (c)

350

300 PV10B1−TR−Arch
PV10B2−TR−Arch
250 PV10B3−TR−Arch
200 PV10B4−TR−Arch
fscr1 [MPa]

Vector2
150

100

50

−50
0.01 0.02 0.03 0.04 0.05 0.06 0 1 2 3
ε [mm/mm] τ [MPa]
scr1 xy
(d) (e)

Figure 5.113: Panel PV16 In-plane Steel Layer 1 Response, Third Group of Tests
299

5.4.3 Panel PV17

Panel PV17 was subjected to pure monotonic compression. The panel material properties
presented by Vecchio and Collins (1982) are shown in Table 5.37 where x and y-direction are

Table 5.37: Panel PV17 Material Properties


0
fc -18.6 Mpa
εco -2.0e-3
fyl 255 MPa
fyt 255 MPa
ρl 0.74%
ρt 0.74%

as shown in Figure 5.95. Figure 5.114 shows the experimental panel response. The response
was nearly linear up until a compressive stress of about -13.8 MPa. The panel reaches a
peak stress of about -21.4 MPa. Explosive failure, with crushing of concrete completely
across panel, and in-plane steel layer 1 buckled out (compression applied in the x direction).
300

The first group of tests has as a purpose to compare the performance of secant y versus
tangent material stiffness and fixed versus rotating concrete material stiffness matrix [Dc ].
Table 5.38 shows the tests included in this group and the analysis times for each one of
them.

Table 5.38: Panel PV17 First Group of Tests

Cases [Dc] type Models Analysis time [s]


PV17-SF-Arch Secant Fixed PopovicHS, Raynor 8.8
PV17-TF-Arch Tangent Fixed PopovicHS, Raynor 12.4
PV17-TR-Arch Tangent Rotating PopovicHS, Raynor 8.5

The results are presented in Figure 5.114. All the tests were able to complete the
analysis. Analysis PV17-TR-Arch was the one that needed the least amount of time to
complete the analysis. The peak stress are slightly lower than that of the experimental
results. Using a secant material stiffness and a fixed [Dc ] matrix leads to satisfactory
results for this particular example, because no cracking or rotation of the principal concrete
direction (θ12c ) is present. When any of these two phenomena are present the test cases
using those analysis options become unstable and slow as it was shown in the analysis for
panels PV10 and PV17.
301

0
PV17−SF−Arch
PV17−TF−Arch
PV17−TR−Arch
−5 Experimental
Vector2

−10

−15

−20

−25
−9 −8 −7 −6 −5 −4 −3 −2 −1 0
εx [mm/mm] x 10
−3

Figure 5.114: Panel PV17 Compressive Response, First Group of Tests


302

The second group of tests are an extension of PV10-TR-Arch to include additional effects
such as Poisson effect (P) and confinement(CF). The tests are listed in Table 5.39.

Table 5.39: Panel PV17, Second Group of Tests

Cases Additional Effects Models Analysis time [s]


PV17-TR-Arch - PopovicHS 8.5
PV17A1-TR-Arch P PopovicHS 11.8
PV17B1-TR-Arch P + CF PopovicHS 18.2

For this group of tests, stress-strain data for the direction of loading (x -direction) are
shown in Figure 5.115. These data show that the inclusion of the Poisson effect alone
(PV17A1-TR-Arch) does not have a significant impact in the peak compressive strength.
However, the inclusion of confinement and the Poisson effect (PV17B1-TR-Arch) does cap-
tures the confinement of the concrete generated by the in-plane steel layer 2 (y-direction,
perpendicular to the direction of the externally applied compression). It is important to
note that simulation of the confinements effect requires the inclusion of both the confine-
ment and the Poisson effect. The compressive concrete confining stresses are generated
by the in-plane steel layer 2 (y-direction), as it restrains the the lateral expansion due to
Poisson effect in the y-direction. Including confinement without Poisson effect would have
no impact in this particular panel. The PV17B1-TR-Arch peak stress increases to nearly
match the experimental peak.
303

0
PV17−TR−Arch
PV17A1−TR−Arch
−5 PV17B1−TR−Arch
Experimental
Vector2
−10

−15

−20

−25
−0.01 −0.008 −0.006 −0.004 −0.002 0
εx [mm/mm]

Figure 5.115: Panel PV17 Compression Response, Second Group of Tests


304

The third group of tests has as a purpose to measure the impact of using the Hognestad
compressive concrete model instead of the PopovicHS in the test case PV17B1-TR-Arch.
The tests cases and their analysis times are shown in Table 5.40

Table 5.40: Panel PV17, Third Group of Tests

Cases Additional Effects Models Analysis time [s]


PV17B1-TR-Arch P + CF PopovicHS 18.2
PV17B1-TR-Arch P + CF Hognestad 21.3

For this group of tests, stress-strain data for the direction of loading (x -direction)
are shown in Figure 5.116.The difference in strength decay for PV17B2-TR-Arch versus
PV17B2-TR-Arch is a direct results of the difference concrete compressive models used in
the two analysis. PV17B2-TR-Arch uses the Hognestad compressive concrete constitutive
model, which decays much quicker than the PopovicHS model used by PV17B1-TR-Arch.
The Hognestad model is a parabolic curve, symmetric about the peak, while the PopovicHS
decays rather slowly for this particular concrete compressive strength (low strength con-
crete). Analysis PV17B1-TR-Arch matches the experimental results better than PV17B2-
TR-Arch for this particulars case.
305

0
PV17B1−TR−Arch
PV17B2−TR−Arch
−5 Experimental
Vector2

−10

−15

−20

−25
−8 −7 −6 −5 −4 −3 −2 −1 0
εx [mm/mm] −3
x 10

Figure 5.116: Panel PV17 Compression Response, Third Group of Tests


306

Figure 5.117 to 5.121 show additional plots for the third group of tests. The data in
figure 5.117 show that the shear stress, strain, and normal stress in the y-direction are
nominally zero throughout the analysis, as expected due to the pure compressive loading
in the x -direction. The data in Figure 5.117(b) shows a significant error in the Vector2
σy . Figure 5.117(d) shows the expansion in the y-direction, εy . This expansion is due to
the compressive loads in the direction y and the restrain from the in-plane steel layer 2 (y
direction).

0.5
0
σy [mm/mm] −0.5
−1 PV17B1−TR−Arch
−1.5 PV17B2−TR−Arch
−2 Experimental
−2.5 Vector2
−3
−15 −10 −5 0 −25 −20 −15 −10 −5 0
σx [mm/mm] σx [mm/mm]
(a) (b)
−3
x 10
4
3
εy [mm/mm]

2
1
0
−1
−15 −10 −5 0 −30 −20 −10 0
σx [mm/mm] σx [mm/mm]
(c) (d)

Figure 5.117: Panel PV17 RC Response, Third Group of Tests


307

Figure 5.118 shows the concrete principal stress angle, θ12c , and the average RC strain
principal strain angle, θε . Since concrete does not crack there can be no slip, δslip , at
the crack surface (δslip = 0) and the angles θ12c and θε are equal (no cracking, no δslip ).
π
The angles have the value of ± because the principal maximum direction point in the
2
y-direction, as the normal to the compressive external load is applied in x -direction. the
π π
data show that θ12c oscillates between + and − . Both values are correct.
2 2

2
PV17B1−TR−Arch
1 PV17B2−TR−Arch
0 Vector2

−1
−2
−25 −20 −15 −10 −5 0
σx [MPa]
(a)
0
−0.2
−0.4
−0.6
−0.8
−1
−1.2
−1.4
−1.6
−25 −20 −15 −10 −5 0
σx [MPa]
(b)

Figure 5.118: Panel PV17 θ12c and θε angles, Third Group of Tests
308

Figure 5.119 shows the concrete fc2 stress. The Hognestad constitutive model decays
quicker than the PopovicHS model as can be seen in Figure 5.119. The peak strain and
strength are equal for both models.

−3
x 10
0
−5

fc2[MPa]
−10
−15
−20
−25 −20 −15 −10 −5 0 −8 −6 −4 −2 0
σx [MPa] εc2 [mm/mm] −3
x 10 PV17B1−TR−Arch
(a) (b) PV17B2−TR−Arch
−3
x 10 0 Experimental
0
−1 −2 Vector2
−2 −4
f [MPa]

−6
−3 −8
−4 −10
−5 −12
c2

−6 −14
−7 −16
−18
−8 −20
−25 −20 −15 −10 −5 0 −8−7−6−5−4−3−2−1 0
σx [MPa] σx [MPa]x 10−3
(c) (d)

Figure 5.119: Panel PV17 Concrete fc2 Stress Response, Third Group of Tests
309

Figure 5.120 shows the concrete fc1 stress. The fc1 stress matches the stress response
of the steel in that direction, since the total stress in the y direction has to be zero for
equilibrium to hold. The εc1 strain is positive as the concrete expand in that direction. If
that strain εc1 (> 0) was the one used to calculate the concrete stress fc1 , then fc1 would
be in tension. This is not valid, since the in-plane steel layer 2 (y direction) restrains this
expansion by generating compressive stresses in the concrete. The load strain εload
c1 (< 0)
is the one used to calculate fc1 which leads to compressive stresses in concrete in the y
direction as shown in Figure 5.120(d).

PV17B1−TR−Arch
0
PV17B2−TR−Arch
Experimental −0.5

f [MPa]
Vector2 −1
−1.5
c1 −2
−2.5
−20 −10 0 0 0.5 1 1.5 2 2.5 3 3.5 4
σx [MPa] εc1 [mm/mm]x 10
−3

(a) (b)
0
−0.2
−0.4
−0.6
fc1[MPa]

−0.8
−1
−1.2
−1.4
−1.6
−1.8
−2
−15 −10 −5 0 −1.2−1−0.8−0.6−0.4−0.2 0
σx [MPa] −4
εLoad[mm/mm] x 10
c1
(c)
(d)

Figure 5.120: Panel PV17 Concrete fc1 Stress Response, Third Group of Tests
310

Figure 5.121 shows the in-plane steel layer 1 (x direction) and 2 (y direction) stress
response. The in-plane steel layer is in compression because is oriented in the same direc-
tion as the externally applied compressive stress (5.121(a)). The in-plane steel layer 2 is
in tension because of the lateral concrete expansion. Notice that both steel layers reach
yielding.

0
PV17B1−TR−Arch −50

fs1[MPa]
PV17B2−TR−Arch −100
−150
−200
−250
−300
−20 −10 0 −8 −7 −6 −5 −4 −3 −2 −1 0
σx [MPa] εs1 [mm/mm] x 10
−3

(a) (b)
300
250
fs2[MPa]

200
150
100
50
0
−15 −10 −5 0 0 0.5 1 1.5 2 2.5 3 3.5 4
σx [MPa] εs2[mm/mm] −3
x 10
(c) (d)

Figure 5.121: Panel PV17 Steel Stress Response, Third Group of Tests

5.5 Summary and Conclusions

The results of a series of analyzes conducted to evaluate the DSFM-TG formulation and
implementation are presneted in this Chapter. These analyzes include applications of the
1D response model (e.g. 1D concrete compression model response) and 2D RC material
to simulate response under basic loading scenarios, as well as applications of the 2D RC
material model to simulate the response of a set of panels tested by Vecchio and Collins
(1982).
The 1D models tests were designed to detect problems in the derivation of the equations
used for the basic models. During this process the errors found in the derivation and coding
311

of the equations were corrected, and the implementation of these 1D models is believed to
be working correctly.
Several strain and stress-driven tests were conducted on an RC material using the DSFM-
TG . The tresults of these simulations show that the use of the tangent rotating [Dc ] matrix
leads to much faster and more stable convergence than observed with the matrices secant
fixed and tangent fixed. The improved performance observed with the [Dc ] matrix becomes
more evident as the changes in the principal concrete stress direction angle, θ12c , become
more significant. The secant fixed [Dc ] matrix was observed to be more stable than the
tangent fixed. This is likely due to the fact that a secant stiffness never reaches extreme
values as a tangent does. However, both types of concrete material stiffness matrices [Dc ]
lead in general the analysis to failure when phenomena such as cracking or steel yielding
occurs.
Three panels were tested, and the results obtained match the experimental results with a
level of accuracy similar to that of Vector2 for the same panels. Problems such as the delayed
yielding of the steel (Panels PV10 and PV16 )in the average region caused the response to
soften at a higher shear stress than that of the experimental. However, reducing the post-
cracking concrete tensile response helped improve the results. The confinement present
in panel PV17 (pure compression), due to the presence of transverse steel, was effectively
captured by using the confinement and Poisson effect simultaneously.
The main conclusions obtained from this Chapter are

• The results obtained match the experimental results with a level of accuracy similar
to that of Vector2 for the same panels.

• Delayed yielding of the steel (Panels PV16 and PV17 )in the average region caused
the response to soften at a higher shear stress than that of the experimental. However,
reducing the post-cracking concrete tensile response improved the results.

• The consistent or tangent rotating [Dc ] matrix leads to a much quicker and more
stable convergence than the secant fixed and tangent fixed options.

• The difference in the performance of the analysis between the tangent rotating [Dc ]
matrix and the secant fixed and tangent fixed becomes more evident as the the changes
312

in the angle θ12c become more significant.


313

Chapter 6

SUMMARY, CONCLUSIONS AND RECOMMENDATIONS

6.1 Summary

The main goal of this research was to enable the use of the Disturbed Stress Field Model
(DSFM) in a traditional tangent-based FEM analysis. Accomplishing this goal required 1)
the reformulation of the stress strain model, including the compatibility, constitutive, and
equilibrium equations that govern the model to be fully implicit; 2) the formulation of a
consistent tangent for this new model; and 3) the development of a set of procedures to
implement the new formulation.
The original DSFM model and the DSFM-TG developed here are presented in Chapter
2. The DSFM compatibility, constitutive, and equilibrium equations are an extension of
the Modified Compression Field Theory (MCFT), which was developed using data from
panel tests conducted by Vecchio and Collins (1982) at the University of Toronto. The
DSFM-TG incorporates additional features with respect to the DSFM such as the dowel
effect DSFM-TG , concrete confinement, Poisson effect (lateral expansion of concrete), and
out-of-plane steel. Information provided in the Vector2 user’s manual (2002) was used to
incorporate these effects.
Chapter 3 uses the DSFM-TG compatibility, constitutive, and equilibrium governing
equations to develop a set of mathematical expressions that define the internal RC material
state. The internal RC material is uniquely defined by 15 internal variables that include
parameters related to the multiaxial loading effects such as Poisson, confinement, compres-
sion softening and out-of-plane reinforcement. The solution for the internal RC material
state requires the solution of a 15x15 nonlinear system of equations. Chapter 3also presents
an expression for the consistent tangent material stiffness matrix which replaces the secant
stiffness matrix used in the original DSFM.
Chapter 4 presents the implementation of the theory developed in Chapter 3. The
314

implementation of the DSFM-TG formulation consists of Matlab code that accomplishes


two main tasks. The first task is to solve the internal RC material state problem for a
given strain state (Process 1). The second task is to assemble the the RC stress vector and
material stiffness matrix (Processes 2 and 3).
Testing of the new DSFM-TG formulation and implementation is presented in Chapter
5. Tests evaluated the numerical efficiency of the DSFM-TG as well as the accuracy with
which the new model simulates experimentally observed responses.

6.2 Conclusions

This section highlights the main conclusions derived from this research organized by Chap-
ters.
Chapter 2 presents a description of the DSFM and DSFM-TG . The main conclusions
that were obtained from the information presented in Chapter 2 are

• The concrete constitutive, Poisson effect and confinement equations are the results
of the independent application of experimentally derived 1D expressions in the three
principal concrete stress directions.

• The crack opening, wcr , and the steel strains at the crack, εscr , are not accounted for
when calculating the average RC strains.

• The perfect steel concrete bonding is not compatible with the DSFM governing equa-
tions when the concrete is cracked.

Chapter 3 presents the mathematical framework used for the implementation of DSFM-
TG . The main conclusions obtained from Chapter 3 are

• The internal RC material state is defined by a set of equations and parameters con-
densed in a 15 × 15 nonlinear system of equations.

• The solution of the in-plane stresses requires the consideration of the out-of-plane
stresses.
315

• The original DSFM concrete material stiffness matrix (secant fixed), [Dc ], written in
the tangent form (tangent fixed), does not include the effect of the principal concrete
stress rotation

• The DSFM-TG concrete material stiffness matrix, [Dc ], includes the effect of the
principal concrete stress rotation and is calculated using the Jacobian matrix, [J],
from the internal RC material problem.

Chapter 5 presents the implementation of the DSFM-TG . The main conclusions ob-
tained 5 are

• The implementation is divided into three processes: Process 1 finds the internal RC
material process; Process 2 finds the concrete and steel stress vectors and material
stiffness matrices; and Process 3 assembles and obtains the RC stress vector and
stiffness matrices by assembling the concrete and steel contributions.

• Process 1 is the most time consuming since it is the only one that requires an iterative
algorithm (Newton) to generate its output.

• Processes 2 and 3 use explicit expressions to generate their output, which makes them
faster than Process 1 to execute.

• The internal RC problem cannot be solved when the Jacobian matrix, [J], used in the
system of equations becomes singular. This singularity of the Jacobian is associated
with the loss of stiffness due to steel failure at the crack.

Chapter 3 presents the evaluation and validation of the formulation and implementation
of the DSFM-TG . The main conclusions obtained 5 are

• The results obtained match the experimental results with a level of accuracy similar
to that of Vector2 for the same panels.

• Delayed yielding of the steel (Panels PV16 and PV17 )in the average region caused
the response to soften at a higher shear stress than that of the experimental. However,
reducing the post-cracking concrete tensile response improved the results.
316

• The consistent or tangent rotating [Dc ] matrix leads to a much quicker and more
stable convergence than the secant fixed and tangent fixed options.

• The difference in the performance of the analysis between the tangent rotating [Dc ]
matrix and the secant fixed and tangent fixed becomes more evident as the the changes
in the angle θ12c become more significant.

6.3 Recommendations

For future improvements of the DSFM-TG the following is recommended:

• Define an uncracked region instead of an average region

• Modify the 1D compressive concrete models to add ductility due to confinement

• Account for the longitudinal and dowel stresses in the steel at the crack

• Include cyclical loading

The DSFM and DSFM-TG define two regions for the RC material: average region and
cracking region. Both regions are related by equilibrium and compatibility equations. The
strain compatibility equation that relates both regions includes only the effect of the slip
at the crack, δslip , and ignores the effect of the crack opening, wcr . This simplification
is associated with the definition of the average region. This region is physically located
between cracks, and even though it is defined as a zone where the strain and stresses are
average, its strains are not average. The strains are not average because the average strains
need to account for the localized displacements at the crack, wcr and δslip , and they do not.
Instead of dividing the RC material into a cracking and average regions, it would be
better to define a cracking and uncracked regions. This would eliminate the confusion
associated with the strain in the average region, because it would be clear that the strains
in the uncracked region are in fact the stresses in that region. The equilibrium equation
between the cracking and uncracked region is identical to that of the current cracking and
average region. The compatibility equation, however, requires the definition of an area for
the cracking region. This area could be defined by an equivalent unbonded length of the
steel at the crack, where the concrete tensile stresses are zero. The task of finding a simple
317

expression for an equivalent unbonded length could be based on existing research on the
subject such as the work developed by Raynor (2000).
The current implementation of confinement focuses on increasing the peak concrete stress
of concrete. The increase in ductility observed in numerous confined concrete tests is not
captured by the current implementation of the model. The currently available 1D compres-
sive concrete models, Hognestad and PopovicHS, do not include a parameter that modifies
the post peak response; they only account for the peak concrete response parameters, εp
and fp . Thus, a new parameter would have to be added to these models to account for
added ductility in concrete due to confinement. This parameter would flatten the concrete
post-peak response as a function of the confinement modification factor, βl .
The dowel steel stresses at the crack are calculated independently of the longitudinal
steel stresses at the crack. This overestimates the steel stresses at the crack when yielding
is present. An interaction relationship that relates the bending stresses due to dowel action
and tension due to the longitudinal stresses is necessary to capture this stress interaction.
Cyclical loading is particularly important for seismic applications where the damage
caused by successive loading and unloading of structures needs to be estimated. However,
it has not yet been included in the model. In order to include the cyclical loading capabil-
ities into DSFM-TG , it is necessary to include history variables and constitutive models
able to consider unloading. This task is quite complex, considering the multiaxial effects
such as Poisson effect, Confinement, and Compression softening. Regardless of the partic-
ularities of these future advancements, the research presented in this document has laid a
framework flexible enough to accommodate changes in the basic constitutive models and
internal variables without needing to change the expression for the material stiffness matrix
and stress vector.
Despite all the possible improvements to the DSFM, it still predicts the behavior of RC
relatively well as shown in Chapter 5. Most of the problems identified could be solved by
including additional internal variables and equations, leading to an even more complicated
model. One of the good qualities of the DSFM is that it is able to give reasonable good
prediction for the behavior of RC using a relatively simple material model. Improvements
to the DSFM come with the cost of losing some of its original simplicity.
318

BIBLIOGRAPHY

He, X. and A. Kwan (2001). Modeling dowel action of reinforcement bars for finite element
analysis of concrete structures. Computers and Structures 79 (6), 595–604.
Kupfer, H. and K. Gerstle (1973). Behavior of concrete under biaxial stress. ASCE Journal
of Engineering Mechanics 99, 853–866.
Kupfer, H., H. H. R. H. (1969). Behavior of concrete under biaxial stress. ACI Jour-
nal 87 (2), 656–666.
Popovic, S. (1973). A numerical approach to the complete stress-strain curve of concrete.
Cement and Concrete Research 3 (5), 583–599.
Raynor, D. (2000). Bond assesment of hybrid frame continuity reinforcement. Master’s
thesis, University of Washington, Seattle, Washington.
Richart, F.E., B. A. and R. Brown (1928). A study of the failure of concrete under com-
bined compressive stresses. University of Illinois Engineering Experimental Station.
Vecchio, F. J. (2000). Disturbed stress field model for reinforced concrete: Formulation.
ASCE Journal of the Structural Engineering 126 (9), 1070–1077.
Vecchio, F. J. (2001). Disturbed stress field model for reinforced concrete: Implementa-
tion. ASCE Journal of the Structural Engineering 127 (1), 12–20.
Vecchio, F. J. and M. P. Collins (1982, March). The response of reinforced concrete to
in-plane shear and normal stresses. Technical report, University of Toronto, Toronto,
Canada.
Vecchio, F. J. and M. P. Collins (1986). The modified compression-field theory for rein-
forced concrete elements subjected to shear. ACI Journal 83 (2), 219–231.
Walraven, J. (1981). Fundamental analysis of aggregate interlock. ASCE Journal of the
Structural Division 107 (11), 2245–2270.
Wong, P. and F. Vecchio (2002, August). Vector2 and formoworks users’s manual. Tech-
nical report, University of Toronto, Toronto, Canada.
319

Appendix A

MATLAB CODE

A.1 averageRegionModels.m

function[fs,Es,fscr,Escr,fsz,Esz,sigmac3PD,Dc3PDLoad,Dcep,Dcfp,b12c,RCstate]=...
averageRegionModels(epss, epss_cr, epssz, epsc3PDLoad, epsp, fp, ...
phi, S_cr, dS_cr_dtheta12c, RCstate, RCstate0, MatProp, AnlPar);
%Subprocess 1AR: Average Region Models

%Subprocess 1AR-MS: Multiple Steel Longitudinal Response


[fs, Es, fscr, Escr, fsz, Esz, RCstate] = multipleSteel(epss,
epss_cr, epssz, RCstate, RCstate0, MatProp, AnlPar);

%Subprocess 1AR-MC: Multiple Concrete Direction Response


[sigmac3PD, Dc3PDLoad, Dcep, Dcfp, b12c, RCstate] =
multipleConcrete(epsc3PDLoad, epsp, fp, phi, S_cr, dS_cr_dtheta12c,
RCstate, RCstate0, MatProp, AnlPar);

A.2 bilinear.m

function [fs, Es] = Bilinear(epss, SteelProp, AnlPar)


%Bilinear steel model
%1. Steel properties
epssy = SteelProp.epssy(SteelProp.layer); fy =
SteelProp.fy(SteelProp.layer);

%2. Derived properties


Eso = fy/epssy;

%2. Derived properties

%3. Stress and stiffness


if abs(epss)<=epssy
fs = Eso*epss;
Es = Eso;

else
fs = fy*sign(epss);
320

switch AnlPar.E_slope_type;
case ’E_tangent’
Es = 0;
case ’E_secant’
Es = fs/epss;
end
end

A.3 confinement kupfer richart.m

function [betal_eps, betal_sigma, dbetal_eps_dfcmin, dbetal_sigma_dfcmin, ...


dbetal_eps_dfcmax, dbetal_sigma_dfcmax ] = confinement_kupfer_richart(fc_min,fc_max,MatProp);
%Kupfer-Richart confinement model
fpc = MatProp.concrete.fpc;
%1. Calculation of uniform and additional lateral pressure
%pu: uniform lateral pressure
%pa: additional lateral pressure

if fc_min<0 & fc_max > 0 %No pu


pu = 0;
pa = fc_min/fpc;
dpu_dfc_max = 0;
dpu_dfc_min = 0;
dpa_dfc_max = 0;
dpa_dfc_min = 1/fpc;

elseif fc_min >= 0 %No pu, no pa


pu = 0;
pa = 0;
dpu_dfc_max = 0;
dpu_dfc_min = 0;
dpa_dfc_max = 0;
dpa_dfc_min = 0;

else
pu = fc_max/fpc;
pa = (fc_min-fc_max) / fpc;
dpa_dfc_max = -1/fpc;
dpa_dfc_min = 1/fpc;
dpu_dfc_max = 1/fpc;
dpu_dfc_min = 0;
end
321

%2. Peak modification factor


betal = 1 + max((0.92*pa - 0.76*pa^2),0) + 4.1*pu;

%3. Peak modification factor derivatives


if 0.92*pa - 0.76*pa^2 >= 0
dbetal_dpa = 0.92-1.52*pa;
else
dbetal_dpa = 0;
end

dbetal_dpu = 4.1;

dbetal_dfc_min = dbetal_dpu * dpu_dfc_min + dbetal_dpa *


dpa_dfc_min; dbetal_dfc_max = dbetal_dpu * dpu_dfc_max +
dbetal_dpa * dpa_dfc_max;

%4. PEak modifications


betal_eps = betal; dbetal_eps_dfcmin = dbetal_dfc_min;
dbetal_eps_dfcmax = dbetal_dfc_max;

betal_sigma = betal; dbetal_sigma_dfcmin = dbetal_dfc_min;


dbetal_sigma_dfcmax = dbetal_dfc_max;

A.4 cracking spacing basic.m

function [S_cr,dS_cr_dtheta_sc] = cracking_spacing_basic(theta_sc,MatProp);%


%Basic cracking spacing model

Smx = MatProp.concrete.Smx;
Smy = MatProp.concrete.Smy;
ct=cos(theta_sc);
st = sin(theta_sc);

S_cr = (abs(ct)/Smx + abs(st)/Smy ) ^-1;


dS_cr_dtheta_sc = -1*(S_cr^2)*(-st*sign_mod(ct)/Smx+ct*sign_mod(st)/Smy );

A.5 crackingRegionModels.m

function [S_cr,dS_cr_dtheta12c,...
w_cr, dw_cr_depsc1, dw_cr_dS_cr,...
vc, dvc_dw_cr, dvc_ddelta_slip,...
fdcr, Edcr, RCstate] = ...
322

crackingRegionModels(theta12c, epsc1, delta_slip, Deltad_cr,...


RCstate, RCstate0, MatProp, AnlPar)

%I.1.1 Cracking spacing model


cracking_spacing_model = AnlPar.cracking_spacing_model;
[S_cr,dS_cr_dtheta12c] = cracking_spacing_model(theta12c,MatProp);

%I.1.2 Cracking width model


cracking_width_model = AnlPar.cracking_width_model;
[w_cr, dw_cr_depsc1, dw_cr_dS_cr] =cracking_width_model(epsc1,S_cr);

%I.1.3 Aggregate Interlock model


aggregate_interlock_model = AnlPar.aggregate_interlock_model;
[vc, dvc_dw_cr, dvc_ddelta_slip] = aggregate_interlock_model(w_cr,
delta_slip, MatProp);

%I.1.4 Dowel Action Model


dowel_action_model = AnlPar.dowel_action_model; np
= MatProp.steel.inplane.no_layers; switch AnlPar.dowel_action_effect
case ’On’
for layer=1:np
MatProp.steel.inplane.layer = layer;
[fdcr(layer,1),Edcr(layer,1)] = dowel_action_model(Deltad_cr(layer), MatProp);
end
case ’Off’
fdcr = zeros(np,1);
Edcr = zeros(np,1);
end
end

A.6 findRCstate.m

function [J, fs, Es, sigmac3PD, Dc3PDLoad, Dcep, Dcfp,b12c,theta12c, RCstate] =...%
findRCstate(eps, RCstate0, MatProp,AnlPar);%
%Process 1: Internal RC State

%Subprocess 1NLS-Xinitial: Initial x


[eps1,eps2,theta_eps]=principal_alt(eps); x = RCstate0.x; %
if x(1)==0 & x(2) ==0
x(1) = eps1;
x(2) = eps2;
x(13) = theta_eps;
end
323

%Call Subprocess 1RJ for the initial x


[R, J, fs, Es, sigmac3PD, Dc3PDLoad, Dcep, Dcfp,b12c,theta12c,RCstate] = getRJ(eps, x, RCstate0,...
MatProp, AnlPar);%

%Call Subprocess 1NLS-Convergence for the intial x


Rnt_tol = AnlPar.Rnt_tol; Reps_tol = AnlPar.Reps_tol;
[convergenceStatus,errorStress,errorStrain] = isItConverged(R,Rnt_tol,Reps_tol,RCstate.crackStatus);%
n_iterMax = AnlPar.max_iterX; n_iter = 1;

%Subprocess 1NLS-Xupdate
updateStatus = isUpdateNeeded(convergenceStatus, n_iter, n_iterMax);
solutionStatus = 1;
while updateStatus & solutionStatus %solutionStatus = 1 means OK
[x, solutionStatus] = xUpdate(x, J, R, RCstate.crackStatus);
if solutionStatus == 1
[R, J, fs, Es, sigmac3PD, Dc3PDLoad, Dcep, Dcfp, b12c, theta12c, RCstate] = getRJ(eps, x, ...
RCstate0, MatProp, AnlPar);
[convergenceStatus,errorStress(end+1),errorStrain(end+1)] = isItConverged(R,Rnt_tol,Reps_tol,...
RCstate.crackStatus);
n_iter = n_iter + 1;
updateStatus = isUpdateNeeded(convergenceStatus, n_iter, n_iterMax);
end
RCstate.solutionStatus = solutionStatus;
end

%Record information
fields = {’eps1’,’eps2’,’theta_eps’,’n_iter’,’errorStress’,’errorStrain’, ’x’};%
values = { eps1,eps2,theta_eps, n_iter, errorStress, errorStrain , x };%
[RCstate] = record( values, fields, RCstate);

A.7 get steel layer strain.m

function [epss]= get_steel_layer_strain(eps,alpha)


%This function calculates the steel strain for a given total strain vector

for layer=1:length(alpha)
eps_alpha = transformation(eps,-alpha(layer),’strain’);
epss(layer,1) = eps_alpha(1);
end
324

A.8 getRCStressStiff.m

function [sigma, D] = getRCStressStiff (sigmas, Ds, sigmac, Dc);

%Assemble
sigma = sigmas + sigmac; D = Ds + Dc;

A.9 getRJ.m

function[R,J,fs,Es,sigmac3PD,Dc3PDLoad,Dcep,Dcfp,b12c,theta12c,RCstate]=...
getRJ(eps, x, RCstate0, MatProp, AnlPar);
%DESCRIPTION:
% Finds the residual and jacobian to the internal RC problem.
%INPUT:
%
% 1. x: internal RC vector containing the following components
% x = [ epsc_3pd; epsc_load_3pd; epsp; fp; p];
% epsc_3pd = [ epsc1 ; epsc2 ; epscz ]; principal concrete strain vector
% epsc_load_3pd = [ epsc1_load ; epsc2_load ; epscz_load ]; principal load concrete strain vector
% epsp = [ epsp1 ; epsp2 ; epspz ]; modified strain at peak
% concrete strength
% fp = [ fp1 ; fp2 ; fpz ]; modified peak concrete strength
% p = [ theta_sc ; delta_eps1_cr; delta_slip ]; cracking parameter vector
% theta_sc : maximum principal concrete direction w/r to X
% delta_eps1_cr: strain increase at crack parameter
% delta_slip : slip at crack
%
% 2. eps: average RC strain vector
% eps = [epsx; epsy; gammaxy];
%
% 3. RCstate0: structure variable with a record of all the internal
% variables from the previous converged stage.
%
%
% 4. MatProp: material property struct variable.
%
% 5. AnlPar: analysis properties structure
%
% OUTPUT:
% 1. R: residual vector
% R = [R_eps; R_epsc_load_3pd; R_ep;R_fp;R_eq] (15x1) vector
%
% 2. J: Jacobian
325

% J = dR/dx
%
% 3. sigmac3PD, Dc3PDLoad, Dcep, Dcfp, b12c,theta12c: concrete variables.
%
% 4. RCstate: structure variable with a record of all the internal
% variables of the current stage.
%
%Unbundling x
epsc3PD = x(1:3); epsc3PDLoad = x(4:6); epsp = x(7:9);
fp = x(10:12); p = x(13:15);

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% I. CALLING MODELS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

%%%%%%%%%%%%%%%I.1 Subprocess 1RJ-CR: Cracking Region %%%%%%%%%%%%%%%%%%


alpha = MatProp.steel.inplane.alpha; %
np = MatProp.steel.inplane.no_layers;%
epsc_cr = MatProp.concrete.epsc_cr; %
theta12c = p(1); %
delta_slip = p(2); %
phi = theta12c - alpha;%

if epsc3PDLoad(1) > epsc_cr


RCstate.crackStatus = 1; %cracked
epsc1 = epsc3PD(1);
Deltad_cr = delta_slip*cos(phi);
[S_cr,dS_cr_dtheta12c, w_cr,dw_cr_depsc1,dw_cr_dS_cr, vc,dvc_dw_cr,dvc_ddelta_slip, ...
fd_cr, Ed_cr, RCstate] = ...
crackingRegionModels(theta12c, epsc1, delta_slip, Deltad_cr, RCstate, RCstate0, MatProp, AnlPar);
else %default values for uncracked case
RCstate.crackStatus = 0; %uncracked
S_cr = 1;
dS_cr_dtheta12c = 0;
w_cr = 0;
dw_cr_depsc1 = 0;
dw_cr_dS_cr = 0;
vc = 0;
dvc_dw_cr = 0;
dvc_ddelta_slip = 0;
Deltad_cr = zeros(np,1);
326

fd_cr = zeros(np,1);
Ed_cr = zeros(np,1);
end

%%%%%%%%%%%%%%%I.2 Subprocess 1RJ-AR: Average Region %%%%%%%%%%%


epssz = epsc3PD(3);%
delta_eps1_cr = p(3); %
epss = get_steel_layer_strain(eps, alpha);

if RCstate.crackStatus == 0
epss_cr = zeros(length(epss),1);
else
epss_cr = epss + delta_eps1_cr*cos(phi).^2;
end

[fs,Es,fs_cr,Es_cr,fsz,Esz,sigmac3PD,Dc3PDLoad,Dcep,Dcfp,b12c,RCstate]=...
averageRegionModels(epss, epss_cr, epssz, epsc3PDLoad, epsp, fp, phi, S_cr, ...
dS_cr_dtheta12c, RCstate, RCstate0, MatProp, AnlPar);

%%%%%%%%%%%%%%%I.3 Subprocess 1RJ-ME %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%


[beta_eps, beta_sigma, S1, S2, S3, S4, S5, S6, S7, S8, S9, nu, P,...
dnu_depsc3PDLoad, dnu_depsp, dP_dnu1, dP_dnu2, dP_dnuz, RCstate]=...
multiaxialLoadingEffectModels (epsc3PD, sigmac3PD, epsc3PDLoad, epsp, RCstate, ...
RCstate0, MatProp, AnlPar );

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% III.RESIDUAL %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%


R2D = [1, 0, 0;
0, 1, 0;
0, 0, 0];
epsc_12c = R2D*epsc3PD; %
eps_slip_12c = [0;0;delta_slip/S_cr];
fc1 = sigmac3PD(1); %
fcz = sigmac3PD(3);
[dummy,T_strainNEG,T_Der_strainNEG] = transformation(zeros(3,1),-theta12c,’strain’); %
rhoz = MatProp.steel.outofplane.rho; %
rho = MatProp.steel.inplane.rho;
epsc0 = MatProp.concrete.epsc0;%
fpc = MatProp.concrete.fpc;

%III.1 Total strain residual


R_eps = eps - T_strainNEG*(epsc_12c + eps_slip_12c);
327

%III.2 Load concrete strain residual


R_epsc_load_3pd = epsc3PDLoad - P*epsc3PD;

%III.3 Strain at peak concrete compressive response residuals


R_epsp = epsp - beta_eps * epsc0;

%III.4 Peak concrete compressive stress response residuals


R_fp = fp - beta_sigma * fpc;

%III.5 Equilibrium residual


R_normal = fc1 - sum(rho.*(fs_cr-fs).*cos(phi).^2) -...
sum(rho.*fd_cr.*cos(phi).*sin(phi)) ; %
R_tangent = vc - sum(rho.*(fs_cr-fs).*sin(phi).*cos(phi))+...
sum(rho.*fd_cr.*cos(phi).^2);
R_outofplane = fcz + rhoz*fsz;%
R_eq = [R_normal; R_tangent;R_outofplane];

%III.6 Residual
R = [R_eps; R_epsc_load_3pd; R_epsp; R_fp; R_eq];

%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% IV. JACOBIAN %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%

e1 = [1;0;0]; %
e2 = [0;1;0]; %
e3 = [0;0;1]; %
I = eye(3);

%IV.1 Construction of Jacobian submatrices


J=zeros(15);

%IV.2 Total strain matrices, dR_eps_dx


dR_eps_depsc3PD = -T_strainNEG*R2D;
dR_eps_depsc3PDLoad = zeros(3);%
dR_eps_depsp = zeros(3);
dR_eps_dfp = zeros(3);

H1 = (1/S_cr)*e3*e2’ - delta_slip/(S_cr^2)*dS_cr_dtheta12c*e3*e1’;%
dR_eps_dp = T_Der_strainNEG*( epsc_12c + eps_slip_12c)*e1’ - T_strainNEG*H1;%
328

%IV.3 Load strains matrices, dR_epsc_load_3pd_dx


Q1 = (dP_dnu1*epsc3PD*e1’ + dP_dnu2*epsc3PD*e2’ + dP_dnuz*epsc3PD*e3’)*dnu_depsc3PDLoad; %
Q2 = (dP_dnu1*epsc3PD*e1’ + dP_dnu2*epsc3PD*e2’ + dP_dnuz*epsc3PD*e3’)*dnu_depsp;%

dR_epsc3PDLoad_depsc3PD = -P;
dR_epsc3PDLoad_depsc3PDLoad = I- Q1; %
dR_epsc3PDLoad_depsp = -Q2;
dR_epsc3PDLoad_dfp = zeros(3); %
dR_epsc3PDLoad_dp = zeros(3);

%IV.4 Peak strains matrices, dR_epsp_dx


dR_epsp_depsc3PD = - S1*S4*epsc0; %
dR_epsp_depsc3PDLoad = -(S2*S5*Dc3PDLoad + S3)*epsc0;%
dR_epsp_depsp = I - S2*S5*Dcep*epsc0; %
dR_epsp_dfp = - S2*S5*Dcfp*epsc0;
dR_epsp_dp = - S2*S5*b12c*e1’*epsc0;

%IV.5 Peak strength matrices, dR_fp_dx


dR_fp_depsc3PD = - S6*S8*fpc; %
dR_fp_depsc3PDLoad = - S7*S9*Dc3PDLoad*fpc; %
dR_fp_depsp = - S7*S9*Dcep*fpc;
dR_fp_dfp = I - S7*S9*Dcfp*fpc;
dR_fp_dp = - S7*S9*b12c*e1’*fpc;%

%IV.6 Crack equilibrium matrices, dR_eq_dx


E1 = [1,0,0; 0,0,0; 0,0,0];
E3 = [0,0,0; 0,0,0; 0,0,1];
a1 = dvc_dw_cr*dw_cr_depsc1*e1’; %
a2 = dvc_ddelta_slip*e2’ + dvc_dw_cr*dw_cr_dS_cr*dS_cr_dtheta12c*e1’; %
gnormal = 0;
gtangent =0;

for i=1:np
c1i = cos(phi(i))^2*e3’ - 2*delta_eps1_cr*cos(phi(i))*sin(phi(i))*e1’;
c2i = cos(phi(i))*e2’ - delta_slip*sin(phi(i))*e1’;
c3i = -2*cos(phi(i))*sin(phi(i))*e1’;
c4i = (cos(phi(i))^2 - sin(phi(i))^2)*e1’;
gnormal = gnormal - rho(i)*( Es_cr(i)*cos(phi(i))^2*c1i + (fs_cr(i)-fs(i))*c3i ...
+ Ed_cr(i)*cos(phi(i)).*sin(phi(i))*c2i + fd_cr(i)*c4i );
gtangent = gtangent - rho(i)*( Es_cr(i)*sin(phi(i))*cos(phi(i))*c1i + (fs_cr(i)-fs(i))*c4i ...
- Ed_cr(i)*cos(phi(i))^2*c2i - fd_cr(i)*c3i );
329

end

dR_eq_depsc3PD = rhoz*Esz*E3 + e2*a1; %Out of plane equilibrium


dR_eq_depsc3PDLoad = (E1+E3)*Dc3PDLoad; dR_eq_depsp =
(E1+E3)*Dcep; dR_eq_dfp = (E1+E3)*Dcfp; dR_eq_dp
= (E1+E3)*b12c*e1’ + e1*gnormal + e2*(a2+gtangent);

%IV.7 Jacobian
J = [ dR_eps_depsc3PD, dR_eps_depsc3PDLoad,dR_eps_depsp, dR_eps_dfp, dR_eps_dp;%
dR_epsc3PDLoad_depsc3PD, dR_epsc3PDLoad_depsc3PDLoad,...
dR_epsc3PDLoad_depsp,dR_epsc3PDLoad_dfp,dR_epsc3PDLoad_dp;
dR_epsp_depsc3PD, dR_epsp_depsc3PDLoad, dR_epsp_depsp, dR_epsp_dfp, dR_epsp_dp;
dR_fp_depsc3PD, dR_fp_depsc3PDLoad, dR_fp_depsp, dR_fp_dfp, dR_fp_dp;
dR_eq_depsc3PD, dR_eq_depsc3PDLoad, dR_eq_depsp, dR_eq_dfp, dR_eq_dp];

% V. RECORDING %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
%V.1 Results from Cracking models
fields = {’phi’,’S_cr’, ’dS_cr_dtheta12c’,...
’w_cr’,’dw_cr_depsc1’, ’dw_cr_dS_cr’, ’vc’, ’dvc_dw_cr’,...
’dvc_ddelta_slip’,’Deltad_cr’,’fd_cr’, ’Ed_cr’ };%
values = {phi, S_cr , dS_cr_dtheta12c , w_cr , dw_cr_depsc1,...
dw_cr_dS_cr , vc , dvc_dw_cr , dvc_ddelta_slip, Deltad_cr, fd_cr, Ed_cr };
[RCstate] = record( values, fields, RCstate);

%V.2 Results from Average models


fields = {’epss’, ’epss_cr’, ’epssz’,’fs_cr’, ’Es_cr’, ’fsz’,’Esz’};
values = { epss, epss_cr, epssz, fs_cr, Es_cr, fsz, Esz }; %
[RCstate] = record( values, fields, RCstate);

%V.3 Results from Multiaxial Loading Effects models


fields = {’beta_eps’, ’beta_sigma’, ’S1’, ’S2’, ’S3’, ’S4’,...
’S5’, ’S6’, ’S7’, ’S8’, ’S9’, ’nu’, ’P’, ’dnu_depsc3PDLoad’,...
’dnu_depsp’, ’dP_dnu1’, ’dP_dnu2’, ’dP_dnuz’};

values = { beta_eps, beta_sigma, S1, S2, S3, S4, S5,...


S6, S7, S8, S9, nu, P, dnu_depsc3PDLoad, dnu_depsp,...
dP_dnu1, dP_dnu2, dP_dnuz };

[RCstate] = record( values, fields, RCstate);


330

A.10 getSteelConcStressStiff.m

function [sigmas, Ds, sigmac, Dc] = ...


getSteelConcStressStiff (J,fs, Es, sigmac3PD, Dc3PDLoad, Dcep, Dcfp, b12c,theta12c, RCstate,...
MatProp,AnlPar);

%Subprocess 2C: Concrete stress vector and stiffness matrix


% Concrete stress vector
R2D = [1 0 0; 0 1 0; 0 0 0];
[dummy,TstressNeg, TDerstressNeg] = transformation(zeros(3,1),-theta12c,’stress’); %
sigmac12c = R2D*sigmac3PD;%
sigmac = TstressNeg*sigmac12c;

% Concrete material stiffness matrix


ZeroMatrix = zeros(3);%
I = eye(3); %
Eeps = [I, ZeroMatrix, ZeroMatrix, ZeroMatrix, ZeroMatrix]’;
E3PDLoad = [ZeroMatrix , I, ZeroMatrix, ZeroMatrix, ZeroMatrix]; %
Eep = [ZeroMatrix , ZeroMatrix, I, ZeroMatrix, ZeroMatrix]; %
Efp = [ZeroMatrix , ZeroMatrix, ZeroMatrix, I, ZeroMatrix];
Ep = [ZeroMatrix , ZeroMatrix, ZeroMatrix, ZeroMatrix, I];
e1 = [1;0;0];

if RCstate.crackStatus == 0 %uncracked
vardel = [14 15]; %eliminate delta_slip, and Deltaeps1cr
eqdel = [13 14]; %eliminate normal and tangent crack equilibrium equations
Eeps(eqdel,:) = []; %15x3 to 13x3
E3PDLoad(:, vardel)= []; %3x15 to 3x13
Eep(:, vardel) = []; %3x15 to 3x13
Efp(:, vardel) = []; %3x15 to 3x13
Ep(:, vardel) = []; %3x15 to 3x13
J(eqdel,:) = [];
J(:,vardel) = [];
end

A = R2D*( Dc3PDLoad*E3PDLoad + Dcep*Eep + Dcfp*Efp + b12c*e1’*Ep); %


B = - TDerstressNeg*sigmac12c*e1’*Ep + TstressNeg*A;

if isequal(AnlPar.Dc_type,’rotating’) & rank(J)==size(J,1)


Dc = -B*inv(J)*Eeps;
else
Dco12c = [Dc3PDLoad(1,1),0,0;
331

0,Dc3PDLoad(2,2),0;
0,0, (Dc3PDLoad(1,1)*Dc3PDLoad(2,2))/ (Dc3PDLoad(1,1)+Dc3PDLoad(2,2))];
[Dc,dummy, dummy] = transformation(Dco12c,-theta12c,’stiffness’);
end

A.11 Hognestad.m

function [fc, Ec, dfc_depsp, dfc_dfp] = Hognestad(epsc,epsp,fp,AnlPar)%


%Hognestad compression model

%1. Concrete Stress and stiffness


if epsc/epsp<=2
fc = fp*(2*(epsc/epsp) - (epsc/epsp)^2);
switch AnlPar.E_slope_type;
case ’E_tangent’
Ec = fp*(2/epsp - 2*(epsc/epsp)*(1/epsp));
dfc_depsp = fp*(-2*epsc/(epsp^2) - 2*(epsc/epsp)*(-epsc/epsp^2));
dfc_dfp = 2*(epsc/epsp) - (epsc/epsp)^2;
case ’E_secant’
Ec = fc/epsc;
dfc_depsp = fc/epsp;
dfc_dfp = fc/fp;
end
else
fc = 0;
Ec = 0;
dfc_depsp = 0;
dfc_dfp = 0;
end

A.12 IsItConverged.m

function [convergeStatus,errorXnt,errorXeps]=IsItConverged(R,Rnt_tol,Reps_tol,crackStatus);%
%Subprocess 1NLS-Convergence

if crackStatus == 0
errorXnt = norm( R([10:12 15]));
else
errorXnt = norm( R([10:15]));
end errorXeps = norm( R([1:9]));

if errorXnt<Rnt_tol & errorXeps<Reps_tol


332

convergeStatus = 1;
else
convergeStatus = 0;
end

A.13 isUpdateNeeded.m

function updateStatus = isUpdateNeeded(convergenceStatus, n_iter,n_iterMax)%

updateStatus = ~convergenceStatus & (n_iter < n_iterMax);

A.14 Linear no residualSOFT Bentz 2003STIFF.m

function [fc ,Ec, dfc_dphi, dfc_dS_cr] =


Linear_no_residualSOFT_Bentz_2003STIFF(epsc, phi,S_cr, MatProp,
AnlPar)
% Linear with no residual/softening, Bentz 2003/stiffening (used by DSFM)

epsc_cr = MatProp.concrete.epsc_cr;
fpt = MatProp.concrete.fpt;

%1. Stress and stiffness


if epsc <= epsc_cr %Uncracked
Ecto = fpt/epsc_cr;
fc = Ecto*epsc; %Stress
Ec = Ecto;
dfc_dphi = 0;
dfc_dS_cr = 0;

else %Cracked
[fc, Ec, dfc_dphi, dfc_dS_cr] = model_response(epsc, phi, S_cr, MatProp, AnlPar);
switch AnlPar.E_slope_type
case ’E_tangent’
%do nothing
case ’E_secant’
Ec = fc/epsc;
np = MatProp.steel.inplane.no_layers;
dfc_dphi = fc*ones(np,1)./phi;
dfc_dS_cr = fc/S_cr;
end
end

%2. Sub-function - Call and select tension softening or stiffning


333

function [fc, Ec, dfc_dphi, dfc_dS_cr] = model_response(epsc, phi,S_cr, MatProp, AnlPar)%

%2.1 Get tension softening response


[fca,Eca,dfca_dphi, dfca_dS_cr] = tension_softening (epsc,phi,S_cr, MatProp, AnlPar);%

%2.2 Get tension stiffening response


[fcb,Ecb,dfcb_dphi, dfcb_dS_cr] = tension_stiffening(epsc, phi,S_cr, MatProp, AnlPar);%

%2.3 Select the response that gives greater stress


if fca >= fcb
fc = fca;
Ec = Eca;
dfc_dphi = dfca_dphi;
dfc_dS_cr = dfca_dS_cr;
else
fc = fcb;
Ec = Ecb;
dfc_dphi = dfcb_dphi;
dfc_dS_cr = dfcb_dS_cr;
end

%3. Tension softening


function [fc,Ec,dfc_dphi, dfc_dS_cr] = tension_softening(epsc, phi,S_cr, MatProp, AnlPar)%
%Tension SOFTENING

%3.1 Fracture energy constant


Gf = MatProp.concrete.Gf;%75e-3; %75N/m =75e-3 N/mm =75 MPa*mm;

%3.2 Crack spacing


Lr = S_cr/2; %uniform concrete strain length

%3.3 Strain at which softening stress becomes zero


fpt = MatProp.concrete.fpt; %
epsc_cr = MatProp.concrete.epsc_cr;%
epsts = (2*Gf/fpt)/Lr; %
epsts = max(1.1*epsc_cr,epsts); %
epsts = min(10*epsc_cr,epsts);

%3.4 Stress and stiffness


epsc_cr = MatProp.concrete.epsc_cr; %
fc = fpt*(1-(epsc-epsc_cr)/(epsts-epsc_cr)); %
334

Ec = -fpt/(epsts-epsc_cr);

%3.5 dfc_dphi and dfc_dS_cr


np = MatProp.steel.inplane.no_layers; dfc_dphi = zeros(1,np) ;
dLr_dS_cr = 1/2; depsts_dLr = -(2*Gf/fpt)/Lr^2;

if epsts<10*epsc_cr & epsts>1.1*epsc_cr


dfc_depsts = + fpt*(epsc-epsc_cr)/(epsts-epsc_cr)^2;
else
dfc_depsts = 0;
end

dfc_dS_cr = dfc_depsts*depsts_dLr*dLr_dS_cr;

%4. Tension stiffening


function [fc, Ec, dfc_dphi, dfc_dS_cr] = tension_stiffening(epsc, phi, S_cr, MatProp, AnlPar)%
%Tension STIFFENING
%Calculate parameter "Ct"
rho = MatProp.steel.inplane.rho;
db = MatProp.steel.inplane.db;
m = 1/sum(4*rho.*abs(cos(phi))./db);
Ct = 2.2*m;

%Stress and stiffness


fpt = MatProp.concrete.fpt;
fc = fpt/(1+sqrt(Ct*epsc));
Ec = -fpt*(1+sqrt(Ct*epsc))^(-2)*(0.5)*(Ct*epsc)^(-0.5)*Ct;

%3.5 dfc_dphi and dfc_dS_cr


dfc_dCt = -fpt*(1+sqrt(Ct*epsc))^(-2)*(0.5)*(Ct*epsc)^(-0.5)*epsc;
dCt_dm = 2.2;
t = [-4*rho./db.*sin(phi).*sign(cos(phi))]’ ; %Row vector
dm_dphi = -1*m^2 *t ; dfc_dphi = dfc_dCt*dCt_dm*dm_dphi;
dfc_dS_cr = 0;

A.15 multipleCompressionSoftening.m

function [betad_eps,betad_sigma, S4, S8, RCstate] =...


multipleCompressionSoftening(epsc3PD,RCstate, RCstate0, MatProp,AnlPar)%
%Subprocess 1ME-MCS: Multiple Compression Softening Direction Effects

if RCstate.crackStatus %cracked
335

compression_softening_model = AnlPar.compression_softening_model;
perp = [2;1;1];
S4 = zeros(3);
S8 = zeros(3);
for j = 1:3
if epsc3PD(j)<0 & epsc3PD(perp(j))>0
[betad_eps(j,1) , betad_sigma(j,1), S4(j,j), S8(j,j) , S4(j,perp(j)), S8(j,perp(j))] =...
compression_softening_model(epsc3PD(j),epsc3PD(perp(j)),MatProp);
else
betad_eps(j,1) = 1;
betad_sigma(j,1) = 1;
S4(j,j) = 0;
S8(j,j) = 0;
S4(j,perp(j)) = 0;
S8(j,perp(j)) = 0;
end
end
else
betad_eps = [1;1;1];
betad_sigma = [1;1;1];
S4 = zeros(3); %dbetaddepsc3PD
S8 = zeros(3); %dbetaddepsc3PD
end

A.16 multipleConcrete.m

function [sigmac3PD, Dc3PDLoad, Dcep, Dcfp, b12c, RCstate ] = ...


multipleConcrete(epsc3PDLoad,epsp,fp,phi,S_cr,dS_cr_dtheta12c,RCstate,RCstate0,MatProp,AnlPar)
%Subprocess 1AR: Average Region models

%2. Concrete stress models


conc_model_tens = AnlPar.conc_model_tens; %
conc_model_comp = AnlPar.conc_model_comp;

for j=1:3
if epsc3PDLoad(j)<0 %Compression
[sigmac3PD(j,1),Dc3PDLoad(j,j),Dcep(j,j), Dcfp(j,j)] =...
conc_model_comp(epsc3PDLoad(j), epsp(j), fp(j), AnlPar );
b12c(j,1) = 0;
else %Tension
[sigmac3PD(j,1),Dc3PDLoad(j,j) , dfc_dphi, dfc_dS_cr] = ...
conc_model_tens(epsc3PDLoad(j), phi, S_cr, MatProp, AnlPar);
336

b12c(j,1) = dfc_dS_cr*dS_cr_dtheta12c + sum(dfc_dphi);


Dcep(j,j) = 0;
Dcfp(j,j) = 0;
end
end

A.17 multipleConfinement.m

function [betal_eps,betal_sigma, S5, S9, RCstate] =...


multipleConfinement(sigmac3PD,RCstate, RCstate0, MatProp, AnlPar);
%Subprocess 1ME-MCF: Multiple Confinement Directions

confinement_model = AnlPar.confinement_model;
S5 = zeros(3);%
S9 = zeros(3);

for j = 1:3
confStresses = sigmac3PD;
confStresses(j) = [];
jmin = find(min(confStresses)==sigmac3PD);
jmax = find(max(confStresses)==sigmac3PD);
if length(jmin)>1
jmin = jmin(1);
end
if length(jmax)>1
jmax = jmax(1);
end
[betalj_eps, betalj_sigma, dbetalj_eps_dfcmin, dbetalj_sigma_dfcmin, dbetalj_eps_dfcmax, ...
dbetalj_sigma_dfcmax] = confinement_model(sigmac3PD(jmin),sigmac3PD(jmax),MatProp);
betal_eps(j,1) = betalj_eps;
betal_sigma(j,1) = betalj_sigma;
S5(j,jmin) = dbetalj_eps_dfcmin; %dbetal_eps_dsigmac3PD
S5(j,jmax) = dbetalj_eps_dfcmax;
S9(j,jmin) = dbetalj_sigma_dfcmin;
S9(j,jmax) = dbetalj_sigma_dfcmax; %dbetal_sigma_dsigmac3PD
end

A.18 multipleLoadingEffectModels.m

function [beta_eps, beta_sigma, nu, P, S1, S2, S3, S4, S5, S6, S7,
S8, S9, dnudepsc3PDLoad, dnudepsp, dPdnu1,dPdnu2,dPdnuz,RCstate]=...
multipleLoadingEffectModels(epsc3PD, sigmac3PD, epsc3PDLoad, epsp, RCstate0, MatProp )
337

%Subprocess 1ME

%Subprocess 1ME-MCSc: Multiplae Compression Softening Direction Effects


switch AnlProp.compression_softening_effect
case ’on’
[betad_eps,betad_sigma, S4, S8, RCstate] = ...
multipleCompressionSoftening(epsc3PD,RCstate, RCstate0, MatProp, AnlPar);
case ’off’
betad_eps = ones(3,1);
betad_sigma = ones(3,1);
S4 = zeros(3);
S8 = zeros(3);
end

%Subprocess 1ME-MCF: Multiple Confinement Directions


switch AnlProp.confinement_effect
case ’on’
[betal_eps,betal_sigma, S5, S9, RCstate] = ...
multipleConfinement(sigmac3PD,RCstate, RCstate0, MatProp, AnlPar);
case ’off’
betal_eps = ones(3,1);
betal_sigma = ones(3,1);
S5 = zeros(3);
S9 = zeros(3);
end

%Subprocess 1ME-COMB: Combine Compression softening and Confinement effects


switch AnlProp.compression_softening_effect
case ’On’
[beta_eps,beta_sigma, S1, S2, S3, S6, S7, RCstate] = ...
multipleCSandCFcombination(epsc3PDLoad, betad_eps, betad_sigma, betal_eps, betal_sigma, ...
RCstate,RCstate0, MatProp, AnlPar);
case ’Off’
beta_eps = betal_eps;
beta_sigma = betal_sigma;
S1 = zeros(3);
S2 = zeros(3);
S3 = zeros(3);
S6 = zeros(3);
S7 = zeros(3);
end
338

%Subprocess 1ME-MP: Multiple Poisson Effect Directions


switch AnlProp.poisson_effect
case ’on’
[nu, dnudepsc3PDLoad, dnudepsp, P, dPdnu1, dPdnu2, dPdnuz, RCstate] = ...
multiplePoisson(epsc3PDLoad, epsp,RCstate, RCstate0, MatProp, AnlPar);
case ’off’
nu = zeros(3,1);
dnudepsc3PDLoad = zeros(3);
dnudepsp = zeros(3);
P = zeros(3);
dPdnu1 = zeros(3);
dPdnu2 = zeros(3);
dPdnuz = zeros(3);
end

A.19 multiplePoisson.m

function[nu,dnudepsc3PDLoad,dnudepsp,P,dPdnu1,dPdnu2,dPdnuz,RCstate]= ...
multiplePoisson(epsc3PDLoad, epsp,RCstate, RCstate0, MatProp, AnlPar);

poisson_tension_model = AnlPar.poisson_tension_model;
poisson_comp_model = AnlPar.poisson_compression_model;

for j=1:3
if epsc3PDLoad(j)<0 %Compression
[nuj,dnuj_depscjLoad,dnuj_depspj] = poisson_comp_model(epsc3PDLoad(j),epsp(j), MatProp);
else %Tension
[nuj,dnuj_depscjLoad] = poisson_tension_model(epsc3PDLoad(j),MatProp);
dnuj_depspj = 0;
end
nu(j) = nuj;
dnudepsc3PDLoad(j,j) = dnuj_depscjLoad;
dnudepsp(j,j) = dnuj_depspj;
end

%P matrices
nu1 = nu(1); nu2 = nu(2); nuz = nu(3);

%1. P matrix
a = 1-nuz*nu2-nu1*nu2-2*nu1*nuz*nu2-nu1*nuz;
P = 1/a* [1-nuz*nu2,nu2*(1+nuz), nuz*(nu2+1);
nu1*(1+nuz), 1-nu1*nuz, nuz*(1+nu1);
339

nu1*(nu2+1), nu2*(1+nu1), 1-nu1*nu2];

%2. dP_dnu1 matrix


b = (-1+nuz*nu2+nu1*nu2+2*nu1*nuz*nu2+nu1*nuz)^2;
dPdnu1 =...
1/b*[ -(-1+nuz*nu2)*(nu2+2*nuz*nu2+nuz),nu2*(1+nuz)*(nu2+2*nuz*nu2+nuz), nuz*(nu2+1)*(nu2+2*nuz*nu2+nuz);
-(1+nuz)*(-1+nuz*nu2), nu2*(nuz^2+1+2*nuz), nuz*(nu2+nuz*nu2+nuz+1);
-(nu2+1)*(-1+nuz*nu2), nu2*(nu2+nuz*nu2+nuz+1), nuz*(nu2^2+2*nu2+1)];

%3. dP_dnu2 matrix


c = (-1+nuz*nu2+nu1*nu2+2*nu1*nuz*nu2+nu1*nuz)^2;
dPdnu2 =...%
1/c*[ nu1*(nuz^2+1+2*nuz),-(1+nuz)*(-1+nu1*nuz),nuz*(nuz+nu1+nu1*nuz+1);...%
nu1*(1+nuz)*(nuz+nu1+2*nu1*nuz), -(-1+nu1*nuz)*(nuz+nu1+2*nu1*nuz), nuz*(1+nu1)*(nuz+nu1+2*nu1*nuz);...
nu1*(nuz+nu1+nu1*nuz+1),-(1+nu1)*(-1+nu1*nuz), nuz*(nu1^2+1+2*nu1)];

%4. dP_dnuz matrix


d = (-1+nuz*nu2+nu1*nu2+2*nu1*nuz*nu2+nu1*nuz)^2;
dPdnuz =...
1/d*[ nu1*(nu2^2+2*nu2+1), nu2*(nu2+nu1*nu2+nu1+1),-(nu2+1)*(-1+nu1*nu2);...
nu1*(nu2+nu1*nu2+nu1+1), nu2*(nu1^2+1+2*nu1), -(1+nu1)*(-1+nu1*nu2);...
nu1*(nu2+1)*(nu2+2*nu1*nu2+nu1), nu2*(1+nu1)*(nu2+2*nu1*nu2+nu1),
-(-1+nu1*nu2)*(nu2+2*nu1*nu2+nu1)]

A.20 multipleSteel.m

function [fs, Es, fs_cr, Es_cr, fsz, Esz, RCstate] =...


multipleSteel(epss, epss_cr, epssz, RCstate, RCstate0, MatProp, AnlPar);%
%Subprocess 1AR-MS: Multiple Steel Longitudinal Response

%Constitutive models
steel_longitudinal_model = AnlPar.steel_longitudinal_model;

%1. In-plane layers stresses and stiffness


np = MatProp.steel.inplane.no_layers;
for layer = 1:np
MatProp.steel.inplane.layer = layer;
%Average
[fs(layer,1) , Es(layer,1)] = steel_longitudinal_model(epss(layer), MatProp.steel.inplane, AnlPar);
%At crack
[fs_cr(layer,1), Es_cr(layer,1)] = ...
steel_longitudinal_model(epss_cr(layer),MatProp.steel.inplane, AnlPar);
end
340

%2. Out of plane layer stress and stiffness


switch AnlPar.out_of_plane_effect
case ’On’
MatProp.steel.outofplane.layer = 1;
[fsz, Esz] = steel_longitudinal_model(epssz, MatProp.steel.outofplane, AnlPar);
case ’Off’
fsz = 0;
Esz = 0;
end
end

A.21 poisson kupfer comp.m

function [nu,dnu_depsc,dnu_depsp] = poisson_kupfer_comp(epsc,epsp,MatProp)


%This function calculates the poisson’s modulus using the kupfer model

nu0 = MatProp.concrete.nu0;
%1. Poisson modulus and its derivatives
if epsp/2 < epsc
nu = nu0;
dnu_depsc = 0;
dnu_depsp = 0;
else
nu = min( 0.5, nu0*(1+1.5*(2*epsc/epsp-1)^2) );
if nu < 0.5
dnu_depsc = 3*nu0*(2*epsc/epsp-1)*( 2/epsp);
dnu_depsp = 3*nu0*(2*epsc/epsp-1)*(-2*epsc/epsp^2);
else
dnu_depsc = 0;
dnu_depsp = 0;
end

end

A.22 poisson V2 tension.m

function [nu,dnu_depsc] = poisson_V2_tension(epsc, MatProp)


%This function returns the Poisson modulus for concrete under tension

epsc_cr = MatProp.concrete.epsc_cr; nu0 = MatProp.concrete.nu0;

%1. Poisson modulus and its derivatives


341

if epsc <= epsc_cr


nu = nu0;
dnu_depsc = 0;
else
nu = max(nu0*(2 - epsc/epsc_cr),0);
if nu>0
dnu_depsc = -nu0/epsc_cr;
else
dnu_depsc = 0;
end
end

A.23 PopovicHS

function [fc,Ec,dfc_depsp,dfc_dfp]= PopovicHS(epsc,epsp,fp, AnlPar)


%Popovic High Strength compression model(used by DSFM)

%1. Model factors


%K factor
if epsp <= epsc
k = 1;
dk_dfp = 0;
else
k = max(0.67 - fp/62,1);
if k>1
dk_dfp = -1/62;
else
dk_dfp = 0;
end
end

%n factor
n = 0.8 - fp/17; dn_dfp = -1/17;

%2. Concrete Stress and stiffness


fc = fp*n*(epsc/epsp) / ((n-1)+(epsc/epsp)^(n*k));

switch AnlPar.E_slope_type
case ’E_tangent’
Ec = fp*n/epsp /(n-1+(epsc/epsp)^(n*k))-...
fp*n^2/epsp/(n-1+(epsc/epsp)^(n*k))^2*(epsc/epsp)^(n*k)*k;
342

%dfc_depsp
dfc_depsp = -fp*n*(epsc/epsp^2) / ((n-1)+(epsc/epsp)^(n*k)) - ...
fp*n*(epsc/epsp) / ((n-1)+(epsc/epsp)^(n*k))^2 * ...
(n*k*(epsc/epsp)^(n*k-1)*(-epsc/epsp^2));

%dfc_dfp
if epsc == 0 %term = d ((epsc/epsp)^(n*k))_dfp
term = 0;
else
term = (epsc/epsp)^(n*k) * log(epsc/epsp) * (dn_dfp*k + n*dk_dfp);
end

dfc_dfp = (epsc/epsp)*(n+fp*dn_dfp)/(n-1+(epsc/epsp)^(n*k)) - ...


fp*n*(epsc/epsp) / ((n-1)+(epsc/epsp)^(n*k))^2 * ( dn_dfp + term);

case ’E_secant’
Ec = fc/epsc;
dfc_depsp = fc/epsp;
dfc_dfp = fc/fp;
end

A.24 principal alt.m

function [eps1,eps2,theta]=principal_alt(eps)
%This function calculate the principal direcitons of the strain state eps
%Strain
epsx = eps(1);
epsy = eps(2);
gammaxy = eps(3);
R = sqrt(((epsx-epsy)/2)^2+(gammaxy/2)^2);%

if R==0
theta = 0;
else
theta = acos((epsx-epsy)/(2*R))/2;
end

if gammaxy<0
theta = - theta;
end

aver = (epsx+epsy)/2;
343

eps1 = aver + R;
eps2 = aver - R;

A.25 Raynor.m

function [fs, Es] = Raynor(epss, SteelProp,AnlPar)


%Raynor steel model

%1. Steel properties


n_sh = SteelProp.n_sh(SteelProp.layer);
epssy = SteelProp.epssy(SteelProp.layer);%
epssh = SteelProp.epssh(SteelProp.layer);%
epssu = SteelProp.epssu(SteelProp.layer);%
fy = SteelProp.fy(SteelProp.layer);%
fsh = SteelProp.fsh(SteelProp.layer);%
fu = SteelProp.fu(SteelProp.layer);

%2. Derived properties


Eso = fy/epssy; Esy = (fsh-fy)/(epssh - epssy);

%3. Stress and stiffness


if abs(epss) <= epssy %Pre-yielding
fs = Eso*epss;
Es = Eso;

elseif abs(epss) <= epssh %Yielding


fs = (fy+(abs(epss)-epssy)*Esy)*sign(epss);
switch AnlPar.E_slope_type
case ’E_tangent’
Es = Esy;
case ’E_secant’
Es = fs/epss;
end

elseif abs(epss) <= epssu %Hardening


fs = ( fu - (fu-fsh) * ( (epssu-abs(epss)) / (epssu-epssh) )^n_sh )*sign(epss);
switch AnlPar.E_slope_type
case ’E_tangent’
Es = n_sh * (fu-fsh) * (epssu-abs(epss))^(n_sh-1) / (epssu-epssh)^(n_sh);
case ’E_secant’
Es = fs/epss;
end
344

else %Failure
fs = 0;
Es = 0;
end

A.26 RCmaterial.m

function [sigma, D, RCstate] = RCmaterial(eps, RCstate0, MatProp,AnlPar)%


%It finds the material stiffness matrix (D) and strain vector (sigma) for a given strain vector

%Process 1: Find internal RC state for a given eps


[J, fs, Es, sigmac3PD,Dc3PDLoad,Dcep,Dcfp,b12c,theta12c,RCstate]=...
findRCstate(eps, RCstate0, MatProp, AnlPar);

%Process 2: Concrete and Steel Material Stiffness Matrices and Stress Vectors
[sigmas, Ds, sigmac, Dc] = ...
getSteelConcStressStiff(J,fs,Es,sigmac3PD, Dc3PDLoad, Dcep,...
Dcfp,b12c,theta12c, RCstate, MatProp, AnlPar);

%Process 3: RC Material Stiffness Matrices and Stress Vectors


[sigma, D] = getRCStressStiff (sigmas, Ds, sigmac, Dc);

% Record
fields = { ’J’, ’fs’, ’Es’, ’sigmac3PD’, ’Dc3PDLoad’, ’Dcep’,...
’Dcfp’, ’b12c’,’theta12c’, ’sigmas’, ’Ds’, ’sigmas’, ’Dc’ };
values ={ J, fs, Es, sigmac3PD,Dc3PDLoad,Dcep,Dcfp,b12c,...
theta12c, sigmas, Ds, sigmas , Dc };
[RCstate] =record( values, fields, RCstate );

A.27 record.m

function [structvar] = record(values, fields, structvar)

for i=1:length(fields)
structvar.(char(fields(i)))= cell2mat(values(i));
end

A.28 steel dowel he kwan.m

function [fd,Ed] = steel_dowel_he_kwan(Deltad,MatProp)


%Dowel action based on He-Kwan model
%1. Steel properties
345

layer = MatProp.steel.inplane.layer;
epssy = MatProp.steel.inplane.epssy(layer);%
fy = MatProp.steel.inplane.fy(layer);%
Es = fy/epssy;

fpc = MatProp.concrete.fpc;
c1 = MatProp.concrete.c1; %should be a value between 0.6 and 1
db = MatProp.steel.inplane.db(layer);
Is = pi*(db)^4/64;%
As =pi*(db/2)^2;
kc = 127*c1*sqrt(abs(fpc))/(db^(2/3)); %should be a value between 75 and 450
lambda = (kc*db/(4*Es*Is))^(1/4);

Vd = Es*Is*lambda^3*Deltad; Vdu = 1.27*db^2*sqrt(abs(fpc)*fy);

if Vd < Vdu
fd = Vd/As;
Ed = Es*Is*lambda^3/As;
else
fd = Vdu/As;
Ed = 0;
end

A.29 transformation.m

function [A,T,dT_dangle]=transformation(Ao,angle,var_type)
%It transforms a stiffness material matrix, strain,
%or stress vector from one coordinate system to another.
%clockwise rotation is defined as negative
%var_type=strain,stress,stiffness
cos_angle = cos(angle);
sin_angle = sin(angle);

switch var_type
case ’strain’ %Strain transformation
T = [ cos_angle^2, sin_angle^2 , cos_angle*sin_angle;
sin_angle^2, cos_angle^2 , -cos_angle*sin_angle;
-2*cos_angle*sin_angle, 2*cos_angle*sin_angle, cos_angle^2-sin_angle^2];

dT_dangle = [ -2*cos_angle*sin_angle, 2*cos_angle*sin_angle, cos_angle^2-sin_angle^2;


2*cos_angle*sin_angle, -2*cos_angle*sin_angle, -cos_angle^2+sin_angle^2;
-2*cos_angle^2+2*sin_angle^2, 2*cos_angle^2-2*sin_angle^2, -4*cos_angle*sin_angle];
346

A = T*Ao;

case ’stress’ %Stress transformation


T = [ cos_angle^2, sin_angle^2 , 2*cos_angle*sin_angle;
sin_angle^2, cos_angle^2 , -2*cos_angle*sin_angle;
-sin_angle*cos_angle, sin_angle*cos_angle, cos_angle^2-sin_angle^2];

dT_dangle = [ -2*cos_angle*sin_angle, 2*cos_angle*sin_angle, 2*cos_angle^2-2*sin_angle^2;


2*cos_angle*sin_angle, -2*cos_angle*sin_angle, -2*cos_angle^2+2*sin_angle^2;
-cos_angle^2+sin_angle^2, cos_angle^2-sin_angle^2, -4*cos_angle*sin_angle];

A = T*Ao;

case ’stiffness’ %Stiffness transformation


T = [ cos_angle^2, sin_angle^2 , -cos_angle*sin_angle;
sin_angle^2, cos_angle^2 , cos_angle*sin_angle;
2*cos_angle*sin_angle, - 2*cos_angle*sin_angle, cos_angle^2-sin_angle^2 ];

dT_dangle = [-2*cos_angle*sin_angle, 2*cos_angle*sin_angle, -cos_angle^2+sin_angle^2;...


2*cos_angle*sin_angle, -2*cos_angle*sin_angle, cos_angle^2-sin_angle^2;...
2*cos_angle^2-2*sin_angle^2, -2*cos_angle^2+2*sin_angle^2, -4*cos_angle*sin_angle];

A = T’*Ao*T;
end

A.30 Trilinear.m

function [fs, Es] = Trilinear(epss, SteelProp,AnlPar)


%Trilinear steel model

%1. Steel properties


epssy = SteelProp.epssy(SteelProp.layer); %
epssh = SteelProp.epssh(SteelProp.layer); %
epssu = SteelProp.epssu(SteelProp.layer); %
fy = SteelProp.fy(SteelProp.layer); %
fu = SteelProp.fu(SteelProp.layer);

%2. Derived properties


Eso = fy/epssy; %
Esh = (fu-fy)/(epssu-epssh);

%3. Stress and stiffness


if abs(epss) <= epssy
347

fs=Eso*epss;
Es=Eso;

elseif abs(epss) <= epssh


fs=fy*sign(epss);
switch AnlPar.E_slope_type
case ’E_tangent’
Es=0;
case ’E_secant’
Es = fs/epss;
end

elseif abs(epss) <= epssu


fs = (fy+Esh*(abs(epss)-epssh))*sign(epss);
switch AnlPar.E_slope_type;
case ’E_tangent’
Es = Esh;
case ’E_secant’
Es = fs/epss;
end

else
fs = 0;
Es = 0;
end

A.31 Vecchio 1982

function [fc ,Ec, dfc_dphi, dfc_dS_cr] = Vecchio_1982(epsc,phi,S_cr, MatProp, AnlPar)%


%Vecchio 1982 concrete tension models (Used by MCFT)

epsc_cr = MatProp.concrete.epsc_cr;
fpt = MatProp.concrete.fpt;

%1. Stress and stiffness


if epsc <= epsc_cr %Uncracked concrete
Ecto = fpt/epsc_cr;
fc = Ecto*epsc;
Ec = Ecto;

else %Cracked concrete


fc = fpt/(1 + sqrt(200*epsc));
348

switch AnlPar.E_slope_type
case ’E_tangent’
Ec = -fpt*(1+sqrt(200*epsc))^(-2)*0.5*(200*epsc)^(-0.5)*200;
case ’E_secant’
Ec = fc/epsc;
end
end

%2. dfc_dphi and dfc_dS_cr


np = MatProp.steel.inplane.no_layers; dfc_dphi =
zeros(1,np); dfc_dS_cr = 0;

A.32 Vecchio 1986 comp soft.m

function [betad_eps, betad_sigma,


dbetad_eps_depsc,dbetad_sigma_depsc, dbetad_eps_depsc_perp,...
dbetad_sigma_depsc_perp] = Vecchio_1986_comp_soft(epsc, epsc_perp, MatProp);
%Vecchio 1986, Compression softening model
%This is STRESS-only model

%1. Modification factor


epsc0 = MatProp.concrete.epsc0;
r = -epsc_perp/epsc0;%
betad =min([1,1/(0.8+0.34*r)]);

%2. Modification factor derivatives


if betad < 1
dr_depsc_perp = -1/epsc0;
dbetad_depsc_perp = -betad^2 * 0.34* dr_depsc_perp;
else
dbetad_depsc_perp = 0;
end

dbetad_depsc = 0;

%3. This is a strength-reduction only model


betad_eps = 1; %
dbetad_eps_depsc = 0;
dbetad_eps_depsc_perp = 0;%
betad_sigma = betad; %
dbetad_sigma_depsc = dbetad_depsc; %
dbetad_sigma_depsc_perp =dbetad_depsc_perp;
349

A.33 Vecchio 1992 A comp soft.m

function[betad_eps,betad_sigma,dbetad_eps_depsc,dbetad_sigma_depsc,dbetad_eps_depsc_perp,...
dbetad_sigma_depsc_perp]=...
Vecchio_1992_A_comp_soft(epsc,epsc_perp, MatProp)
%Compression softening model based on Vecchio 1992-A model (epsc1/epsc2)
%This is a STRESS-STRAIN reduction model

%1. Modification factor parameters


Cs = 0.55;%DSFM, slip considered

r = min(-epsc_perp/epsc,400); if r<400
dr_depsc = epsc_perp/epsc^2;
dr_depsc_perp = -1/epsc;
else
dr_depsc = 0;
dr_depsc_perp = 0;
end

%parameter Cd
if r>0.28
Cd = 0.35*(r-0.28)^0.8;
else
Cd = 0;
end

%2. Modifcation factor


betad = min(1,1/(1+Cs*Cd));

%3. Modification factor derivatives


if betad < 1
dCd_dr = 0.8*0.35*(r-0.28)^(-0.2);
dbetad_depsc = - betad^2 * Cs * dCd_dr * dr_depsc;
dbetad_depsc_perp = - betad^2 * Cs * dCd_dr * dr_depsc_perp;
else
dbetad_depsc = 0;
dbetad_depsc_perp = 0;
end

%4. This is a strength/strain-reduction model


betad_eps = betad;
dbetad_eps_depsc = dbetad_depsc;
350

dbetad_eps_depsc_perp = dbetad_depsc_perp;

betad_sigma = betad;
dbetad_sigma_depsc = dbetad_depsc; %
dbetad_sigma_depsc_perp = dbetad_depsc_perp;

A.34 Walraven slip resistance.m

function [vc, dvc_dw_cr, dvc_d_delta_slip]=...


Walraven_slip_resistance(w_cr, delta_slip, MatProp);
%Aggregate interlock model based on Walraven’s equation

fcc = MatProp.concrete.fcc;

if w_cr ==0
vc = 0;
dvc_dw_cr = 0;
dvc_d_delta_slip = 0;
else
value = ( 0.234 * w_cr ^ -0.707 - 0.2);
if value < 0;
value = 0;
dvalue_dw_cr = 0;
else
dvalue_dw_cr = ( 0.234 * (-0.707) * w_cr ^(-1.707) );
end

f_w_cr = 1.8 *w_cr ^ -0.8 + value*abs(fcc);


df_dw_cr = 1.8*(-0.8)*w_cr ^ (-1.8) + dvalue_dw_cr *abs(fcc);
vc = delta_slip * f_w_cr;
dvc_dw_cr = delta_slip * df_dw_cr;
dvc_d_delta_slip = f_w_cr;
end

Das könnte Ihnen auch gefallen