Você está na página 1de 202

FINITE ELEMENT ANALYSIS TO SIMULATE

REINFORCED CONCRETE CORROSION

IN BEAMS AND BRIDGE DECKS

by

Diane Wurst

A thesis submitted to the Faculty of the University of Delaware in partial


fulfillment of the requirements for the degree of Master of Civil Engineering

Fall 2013

© 2013 Diane Wurst


All Rights Reserved
FINITE ELEMENT ANALYSIS TO SIMULATE

REINFORCED CONCRETE CORROSION

IN BEAMS AND BRIDGE DECKS

by

Diane Wurst

Approved: __________________________________________________________
Jennifer E. Righman McConnell, Ph.D.
Professor in charge of thesis on behalf of the Advisory Committee

Approved: __________________________________________________________
Harry W. Shenton III, Ph.D.
Chair of the Department of Civil and Environmental Engineering

Approved: __________________________________________________________
Babatunde A. Ogunnaike, Ph.D.
Dean of the College of Engineering

Approved: __________________________________________________________
James G. Richards, Ph.D.
Vice Provost for Graduate and Professional Education
ACKNOWLEDGMENTS

I would like to thank my advisor, Dr. Jennifer Righman McConnell. Without

her, I would not have had the opportunity to perform research and write a thesis. She

also provided invaluable guidance and support throughout the entire process. In

addition, I would like to thank my family and friends. Without their support, I

wouldn’t have had the motivation to apply and attend graduate school.

I would like to thank all of the other structures graduate students at the

University of Delaware for making graduate school a fun and exciting experience, and

for providing a listening ear when my research wasn’t going the way I had hoped.

Finally, I would like to thank the University of Delaware University

Transportation Center (UD-UTC) for providing financial support for my project.

iii
TABLE OF CONTENTS

LIST OF TABLES ....................................................................................................... vii


LIST OF FIGURES ........................................................................................................ x
ABSTRACT ................................................................................................................ xiii

Chapter

1 INTRODUCTION .............................................................................................. 1

1.1 Motivation ................................................................................................. 1


1.2 Objectives and Scope ................................................................................ 2
1.3 Thesis Outline ............................................................................................ 3

2 LITERATURE REVIEW ................................................................................... 5

2.1 Previous Experimental Results .................................................................. 5

2.1.1 Destructive Bridge Testing ............................................................ 5


2.1.2 Corroded Reinforced Concrete ...................................................... 6

2.2 Reinforced Concrete and Corrosion Modeling........................................ 11

3 MODELING APPROACH .............................................................................. 15

3.1 Concrete ................................................................................................... 15

3.1.1 Elastic Behavior........................................................................... 16


3.1.2 Non-Linear Behavior ................................................................... 16

3.2 Rebar........................................................................................................ 23

3.2.1 2-Dimensional Rebar ................................................................... 23


3.2.2 3-Dimensional Rebar ................................................................... 24

3.3 Boundary Conditions ............................................................................... 29


3.4 Analysis Method ...................................................................................... 30

4 MODEL CALIBRATION ................................................................................ 33

4.1 Beam Geometry and Material Properties ................................................ 34

4.1.1 Concrete ....................................................................................... 34


4.1.2 Rebar............................................................................................ 36

iv
4.2 Calibration Metrics .................................................................................. 39
4.3 Strength and Deflection Calculations ...................................................... 42
4.4 2-Dimensional Concrete Model .............................................................. 43

4.4.1 Uncorroded Base Model Input Values ........................................ 44


4.4.2 Mesh Sensitivity Analysis and Uncorroded Base Model Input ... 48
4.4.3 Corroded Model........................................................................... 54

4.5 3-Dimensional Concrete Model .............................................................. 72

4.5.1 2-Dimensional Rebar ................................................................... 73


4.5.2 3-Dimensional Rebar ................................................................... 79

4.6 Conclusions ............................................................................................. 89

5 BRIDGE MODELS .......................................................................................... 92

5.1 7R ............................................................................................................ 92

5.1.1 Bridge Information ...................................................................... 92


5.1.2 Results ......................................................................................... 94

5.2 SR 1 over US 13 .................................................................................... 109

5.2.1 Bridge Information .................................................................... 109


5.2.2 Modeling Results ....................................................................... 111

5.3 SR 299 over SR 1 .................................................................................. 115

5.3.1 Bridge Information .................................................................... 116


5.3.2 Modeling Results ....................................................................... 118

5.4 Conclusions ........................................................................................... 119

6 CONCLUSIONS ............................................................................................ 122

6.1 Summary................................................................................................ 122


6.2 Results ................................................................................................... 125
6.3 Future Work........................................................................................... 129

REFERENCES ........................................................................................................... 132

v
Appendices

A SAMPLE STRENGTH CALCULATIONS ................................................... 136


B COMPLETE RESULTS FOR 2-D BEAM CALIBRATION
VARIATIONS ................................................................................................ 143
C COMPLETE RESULTS FOR 3-D BEAM CALIBRATION
VARIATIONS ................................................................................................ 166
D PERMISSION LETTERS .............................................................................. 183

vi
LIST OF TABLES

Table 2.1. Comparison of strength of DRBA Bridge 7R (Ross 2007). ...................... 6

Table 4.1. Yield and ultimate strengths of SD345 and SD295 rebar found in
literature compared to the minimum specified yield strengths. .............. 37

Table 4.2. True stress and true plastic strain input values for SD295 and SD345
rebar. ........................................................................................................ 39

Table 4.3. Input values for post-cracking commands for CDP model. .................... 47

Table 4.4. Strength results of mesh sensitivity analysis with percent differences
based on the calculated theoretical strength values. ................................ 50

Table 4.5. Deflection results of mesh sensitivity analysis with percent differences
based on the calculated deflection values for the uncracked and cracked
sections. ................................................................................................... 50

Table 4.6. Input and results of final uncorroded 2-D beam model. .......................... 54

Table 4.7. Selected results for initial calibration of 2-D corroded models with values
expressed as a percentage of the base model........................................... 57

Table 4.8. Input and results of calibration of Ec, f’c, and f’t for corroded 2-D beam
model. ...................................................................................................... 59

Table 4.9. Input and results of 2-D beam model after calibrating input
parameters................................................................................................ 67

Table 4.10. Results of optimization of the dilation angle after optimizing Ec, f’c, f’t,
As, and A’s for the corroded 2-D beam model......................................... 69

Table 4.11. Strength results of rebar mesh sensitivity analysis for 3D concrete beam
with 2D rebar including percent differences based on the calculated
theoretical strength values. ...................................................................... 75

Table 4.12. Strength results of concrete mesh sensitivity analysis for 3D concrete
beam with 2D rebar including percent differences based on the calculated
theoretical strength values. ...................................................................... 76

Table 4.13. Base model input for 3-D concrete elements with 2-D rebar elements for
brittle cracking technique. ....................................................................... 77

vii
Table 4.14. Selected results of optimization of corrosion for 3-D beam model with 2-
D rebar. .................................................................................................... 78

Table 4.15. Results of varying friction input for 3-D concrete beam models with 3-D
rebar elements. ......................................................................................... 83

Table 4.16. Comparison of variations in friction and pressure-overclosure of 3-D


concrete beam models with 3-D rebar using brittle cracking. ................. 88

Table 4.17. Comparison of strength values utilizing different analysis techniques of


3-D concrete beams with 3-D rebar using brittle cracking. ..................... 89

Table 5.1. Original capacity results for Bridge 7R. .................................................. 96

Table 5.2. Results of bridge 7R when varying tension stiffening values. .............. 102

Table 5.3. Peak stress values in cross-frames as maximum loading in different 7R


bridge models. ....................................................................................... 105

Table 5.4. Distribution factors of bridge 7R previously determined in research


(McConnell et al. in review). ................................................................. 106

Table 5.5. DF values for finite element models created of Bridge 7R. .................. 107

Table 5.6. Initial results from analyzing bridge US13 including non-linear concrete
and rebar commands. ............................................................................. 112

Table 5.7. Results of Bridge US13 when varying tension stiffening values. ......... 115

Table 5.8. Initial results from analyzing Bridge US299. ........................................ 118

Table 5.9. Results of Bridge US299 when varying tension stiffening values. ....... 119

Table B.1. Results of mesh sensitivity analysis for 2-D beam models. .................. 143

Table B.2. Complete input and results of all concrete damaged plasticity models
tested during 2-D beam calibration. ...................................................... 144

Table C.1. Results of mesh sensitivity analysis for 3-D beam with 2-D rebar models
utilizing the brittle cracking technique………………………………………...... 166

Table C.2. Results of mesh sensitivity analysis for 3-D beam with 2-D rebar models
utilizing the CDP and SC techniques……………………………………………. 167

viii
Table C.3. Results for 3-D beam with 2-D rebar models utilizing the brittle cracking
technique…………………………………………………………………….……………. 169

Table C.4. Results for 3-D beam with 3-D rebar models utilizing the CDP
technique………………………………………..………………………………………….170

Table C.5. Results for 3-D beam with 3-D rebar models utilizing the SC
technique………………………………………..………………………………………….178

Table C.6. Results for the 3-D beam with 3-D rebar models utilizing the brittle
cracking technique…………………………..…………………………………………. 181

ix
LIST OF FIGURES

Figure 2.1. Strength losses reported in literature (McConnell et al. 2012). ................. 9

Figure 2.2. Stiffness losses reported in literature (McConnell et al. 2012). ................ 9

Figure 3.1. Tension stiffening model ABAQUS employs in the smeared crack
technique (Simulia 2011). ....................................................................... 18

Figure 3.2. Response of concrete to uniaxial loading in (a) tension and (b)
compression for CDP model (Simulia 2011). ......................................... 20

Figure 3.3. Exponential decay friction model (Simulia 2011)................................... 25

Figure 3.4. Default “hard” pressure-overclosure relationship (Simulia, 2011). ........ 27

Figure 3.5. “Softened” exponential pressure-overclosure relationship


(Simulia 2011). ........................................................................................ 28

Figure 4.1. Geometry of finite element calibration models, elevation view (top) and
cross-section view (bottom) (dimensions in inches). .............................. 35

Figure 4.2. Example of determination of (a) the last step before failure and (b) the
first step of failure in 3-D models utilizing ABAQUS/Explicit. ............. 41

Figure 4.3. Uncorroded 2-D base model stress contours (units of psi)...................... 53

Figure 4.4. Load and deflection comparison between final uncorroded model and
theoretical calculations for the uncracked and cracked sections. ............ 53

Figure 4.5. Comparison of strength between results from calibration models and
theoretical values caused by varying Ec. ................................................. 60

Figure 4.6. Comparison of deflection between results from calibration models and
theoretical values of the uncracked section caused by varying Ec. ......... 60

Figure 4.7. Strength comparison of results from models and theoretical values while
varying As and using a constant A’s of 100% of the base model value. . 62

Figure 4.8. Strength comparison of results from models and theoretical values while
varying As and using a constant A’s of 90% of the base model value..... 62

Figure 4.9. Strength comparison of results from models and theoretical values while
varying As and using a constant A’s of 80% of the base model value..... 63

x
Figure 4.10. Deflection comparison of results from models and theoretical values
while varying As and using a constant A’s of 100% of the base model
value. ....................................................................................................... 64

Figure 4.11. Deflection comparison of results from models and theoretical values
while varying As and using a constant A’s of 90% of the base model
value. ....................................................................................................... 64

Figure 4.12. Deflection comparison of results from models and theoretical values
while varying As and using a constant A’s of 80% of the base model
value. ....................................................................................................... 65

Figure 4.13. Stress-strain response of element in 2-D beam models reported as percent
decrease in As. ......................................................................................... 66

Figure 4.14. Deflection response of elements in 2-D beam models reported as percent
decrease in As. ......................................................................................... 66

Figure 4.15. Stress contours of (a) final uncorroded and (b) final corroded 2-D beam
models (units of in psi). ........................................................................... 71

Figure 4.16. Load and deflection comparison between uncorroded and corroded finite
element model and theoretical calculations for the uncracked section. .. 72

Figure 4.17. Comparison of compression concrete output to input values for 3-D
beam with 3-D rebar. ............................................................................... 84

Figure 4.18. Comparison of tensile concrete output to input values for 3-D beam with
3-D rebar. ................................................................................................. 85

Figure 4.19. Time comparison for 3-D concrete beam models with 3-D rebar using
brittle cracking. ........................................................................................ 87

Figure 5.1. Map of location of Bridge 7R (Ross 2007). ............................................ 94

Figure 5.2. Finite element model of bridge 7R viewed from (a) the top, (b) the
bottom, and (c) isoparametrically. ........................................................... 95

Figure 5.3. Results of bridge 7R at maximum loading for the (a) elastic version with
no non-linear concrete or rebar commands, (b) the uncorroded version,
and (c) the corroded version. ................................................................... 97

Figure 5.4. Representative example of location of deck element used for stress
distribution analysis in (a) the elastic, (b) uncorroded and (c) corroded
deck.......................................................................................................... 99

xi
Figure 5.5. Tension stiffening values for uncorroded model. .................................. 101

Figure 5.6. Stress-strain response of representative deck element of Bridge 7R at a


stress concentration location using the original tension stiffening input.
............................................................................................................... 103

Figure 5.7. Stress-strain response of element in the deck of Bridge 7R at a stress


concentration location using a tension stiffening value of 0.035. ......... 104

Figure 5.8. Comparison of nonlinear response of concrete for bridge 7R for the (a)
uncorroded and (b) corroded deck. ........................................................ 108

Figure 5.9. Satellite View of SR 1 over US 13 (Ambrose 2012)............................. 110

Figure 5.10. Finite element model of bridge US13 viewed from (a) the top, (b) the
bottom, and (c) in cross-section............................................................. 111

Figure 5.11. Results of bridge US13 for (a) the elastic version at s similar loading as
the maximum for the uncorroded/corroded model, (b) the uncorroded
version at the maximum loading, and (c) the corroded version at the
maximum loading. ................................................................................. 113

Figure 5.12. Compressive stress-strain response of elements in deck of US13 causing


inability to converge. ............................................................................. 114

Figure 5.13. Tensile stress-strain response of elements in deck of US13 causing


inability to converge. ............................................................................. 114

Figure 5.14. Satellite view of SR 299 over SR 1 (Ambrose 2012). .......................... 117

xii
ABSTRACT

Current analysis techniques do not acknowledge the existence of load

redistribution between girders, or system capacity, of bridges due to a lack of

understanding on the redistribution mechanisms. This lack of understanding is the

primary motivation for this research. Specifically, the deck as a load redistribution

mechanism is analyzed. It is thought that including the system capacity of bridges

would help to prioritize repairs and allocate the limited funding available for

infrastructure. For this reason, this research aims to aid in quantifying the system

capacity effects of bridges due to corrosion of reinforcement in the deck.

Previously, a literature review was executed to determine the effects of

corrosion in reinforced concrete. It was determined through this review of testing that

the change in performance due to corrosion is best estimated as a strength decrease of

50% and an ultimate deflection increase of 82%, mimicking 25 years of corrosion.

These expected performance metrics were used to create finite element models of

uncorroded and corroded reinforced concrete beams. Different concrete material

modeling techniques available within the commercial software ABAQUS were

assessed; these include brittle cracking, smeared crack, and concrete damaged
plasticity techniques. This was done for both 2-dimensional and 3-dimensional

concrete elements, as well as for both 2-dimensional and 3-dimensional rebar elements

within the 3-dimensional concrete elements. In the end, it was determined that using

the concrete damaged plasticity approach with 2-dimensional beam and rebar elements

produced the most accurate results and was also the easiest approach to implement in

existing full-scale bridge models; this approach will also reduce computational effort

in any future full-scale bridge models.

xiii
After the modeling technique was determined, the input was calibrated to

determine the optimal approach to model uncorroded and corroded reinforced

concrete. It was determined which input values to use for the uncorroded concrete,

and how to alter these values to simulate corrosion. Through this optimization

process, it was determined that a 40% decrease in the modulus of elasticity of

concrete, 40% decrease in tensile strength of concrete, 64% decrease in compressive

strength of concrete, 20% decrease in area of compressive steel, and 61.5% decrease

in area of tensile steel resulted in the optimum simulation of reinforced concrete

corrosion.

This uncorroded and corroded input was applied to 3 different full-scale bridge

models which were previously created; these models were created and calibrated

based on actual bridges located in Delaware that had been previously field tested, all

having steel girders. It was found that this modeling approach created convergence

difficulties in some of the bridge models when attempting to load the structures to

their ultimate capacities and subsequently only the results of one of the bridges, for

which convergence was obtained up through a peak loading, was analyzed in depth.

This bridge is referred to as Bridge 7R and served as an exit ramp for Interstate 295

North, just south of the Delaware Memorial Bridge. Initially, convergence with
corroded models was not reached. However, after changing the input parameters

governing tension stiffening, convergence was achieved.

The maximum loading and the distribution factors of the bridge models were

analyzed. It can be seen in these results that the corrosion in the deck caused a more

uniform stress distribution in the deck, and consequently the girders, than with an

uncorroded deck. Contrary to the expected response, the corroded model resulted in

xiv
DF values between those of the elastic and those of the uncorroded models; however,

in all models the DF approached the theoretical inelastic values. The results of these

models also indicated that the corroded models reached higher strengths than their

uncorroded counterparts. It was thought that the cause of the differences in both the

strength and DF values was due to greater load sharing in the corroded model when

compared to the corresponding uncorroded model.

xv
Chapter 1

INTRODUCTION

1.1 Motivation
Current analysis techniques for bridges, as specified by the American

Association of State Highway and Transportation Officials (AASHTO) (2013),

specify that each individual component of a bridge be designed separately to be

capable of carrying the maximum possible loading which may be applied. This

includes utilizing a line girder analysis. However, in reality, bridges act as a system,

redistributing the load when members begin yielding. This produces a higher strength

than through the analysis of individual components. This true strength can then be

used to help better prioritize bridge repair and replacement.

According to the American Society of Civil Engineers (ASCE) report card

(2009), approximately 27% of bridges in the United States are either structurally

deficient or functionally obsolete. ASCE estimates that, over 50 years, $650 billion

would be required to maintain the current level of bridges; that is, leave approximately

27% of bridges structurally deficient or functionally obsolete. To eliminate all of

these deficiencies, $850 billion over 50 years is estimated to be required. However, in

2004 only $10.5 billion was spent on bridge improvements. This indicates a need for

a prioritization method to properly allocate the limited funding.

It was previously shown by McCarthy (2012) that including the load path

redundancy of bridges in the rating process can create a significant savings when

prioritizing bridge repairs. This analysis reviewed 14 steel girder-concrete deck

1
bridges located in Delaware and determined that, when including load path

redundancy in the load carrying capacity, a savings between $2 and $4.7 million from

the estimated $23.4 million could be achieved. This $23.4 million is the approximated

cost of repairing the selected structures based on their current conditions.

One of the current impediments to implementing system capacity analysis of

bridges is the lack of information regarding transverse load distribution mechanisms

and how, in the case of transverse distribution through concrete decks (which tend to

suffer the most accelerated degradation of condition of all structural components due

to factors such as direct traffic loads and applications of deicing agents), these

mechanisms may vary with age and condition. For this reason, the deck is the load

distribution mechanism to be investigated and, due to corrosion of rebar being one of

the most prevalent causes of deterioration facing bridge decks, the influences of this

corrosion on the system capacity of steel girder bridges is the focus of this study.

Previously, a literature review was performed to quantify the effects of corrosion on

the behavior of reinforced concrete specimens and these results were synthesized and

extrapolated to a deck design life of 25 years (McConnell et al. 2012). This research

focuses on these behavior targets and uses them to calibrate a finite element model of

reinforced concrete members, through varying selected input parameters. These


calibrated values of input parameters are then applied to bridge models to estimate the

change in system capacity.

1.2 Objectives and Scope


The two primary objectives of this project were to calibrate a finite element

modeling technique to model reinforced concrete which has experienced corrosion due

to deicing agents and then apply this technique to full-scale bridge models in order to

2
determine how corrosion of reinforcing bars in the deck effects the system capacity of

bridges. Specifically, a reinforced concrete beam was modeled to determine the

material input values required to emulate corrosion using ABAQUS. After

performing a literature review to determine the different modeling approaches for

modeling reinforced concrete within ABAQUS, models were created using three

different approaches. These include the smeared crack, concrete damaged plasticity,

and brittle cracking techniques. The previous strength and deflection results for a 25

year old deck were utilized as the response goals of the corroded concrete. The input

parameters, as well as the different modeling approaches and commands, were then

calibrated to reflect these corrosion goals.

Once the input was calibrated, it was applied to three previously created and

calibrated full-scale bridge models; these models were based off of bridges which are

located in Delaware and previously field tested (Michaud 2011 and Ambrose 2012).

All 3 of these bridges are steel girder bridges with composite reinforced concrete

decks due to steel girder bridges being a common bridge configuration for which the

advantages of system analysis has been previously demonstrated (Michaud 2011).

One of the three bridges is a simple span bridge and the other two are 2-span

continuous structures. These models were loaded in attempts to estimate their ultimate
capacity. The ultimate capacity with the uncorroded input and the corroded input

were compared and this change of strength, caused by the corrosion in the deck, was

analyzed to determine the effects on load redistribution ability.

1.3 Thesis Outline


In the pages that follow is an investigation of how corrosion of reinforced

concrete bridge decks due to deicing agents effects the system capacity of bridges.

3
The primary research tasks were the development of the finite element models as well

as the calibration of these models to demonstrate corrosion, followed by the

application of this calibration to full-scale bridge models. The material has been

divided into the following chapters.

 Chapter 2 presents the background of the project, including a literature review of

how to model reinforced concrete beams and how corrosion effects the material

properties of reinforced concrete.

 Chapter 3 describes the relevant modeling approaches available within ABAQUS

and an associated discussion of how these approaches and corresponding

commands were utilized.

 Chapter 4 presents the results of the reinforced concrete beam models which were

created and the calibration process applied.

 Chapter 5 discusses the application of the calibrated reinforced concrete input to

the full-scale bridge models and the results of this analysis.

 Chapter 6 presents the conclusion of the research; this chapter also provides

recommendations for future work for better calibrating beam models and

processing bridge result data.

4
Chapter 2

LITERATURE REVIEW

2.1 Previous Experimental Results


A literature review was performed focusing on the proof of concept underlying

the motivation for this work and experimental testing performed involving reinforced

concrete that had undergone corrosion. This was done to prove how system capacity

is significantly larger than the strength calculated using standard analysis techniques

and to determine how the strength and deflection of reinforced concrete changes as

reinforced concrete corrodes. Many different tests have been performed to cause

accelerated corrosion; however, many of these studies used an electrical current to

induce corrosion. This method has been questioned by other researchers and therefore

all tests found using this method were disregarded (Li 2000 and Melchers et al. 2006).

Below, the proof of concept and corroded reinforced concrete studies which were

utilized are discussed.

2.1.1 Destructive Bridge Testing


Previous research was performed which destructively tested a full-scale bridge

to determine the ultimate strength (Chajes et al. 2010). For this testing, hydraulic

jacks were placed on a bridge to mimic the presence of an HS-20 truck. However, the

hydraulic jacks and associated load resistance mechanism did not have sufficient

capacity to cause the bridge to fail, although previous predictions using current

evaluation methods (AASHTO 2012) indicated that this should not be the case. A

finite element model was then created of the same bridge. This model was loaded

until failure and the results were compared to the predictions using current evaluation

methods. The results of this comparison can be seen in Table 2.1. Note that the

5
strength reported for the loading system is the maximum force which could be applied

to the bridge, not the force causing failure. It can be seen in these results that the code

predictions are very conservative compared to the actual results and the finite element

predictions.

Table 2.1. Comparison of strength of DRBA Bridge 7R (Ross 2007).


Strength
Quantification Method Number of Equivalent
Load (MN)
HS-20 Trucks
Code-Predicted, with load factors 5 1.6
Code-Predicted, no load factors 12 3.8
Loading System 17 5.4
FEA, Girder Yielding 19 6.1
FEA, System Failure 30 9.6

These results reinforce the concepts that load redistribution in bridges

increases the system capacity and that using finite element analysis can help predict

the true strength. The code-predicted capacity of 12 trucks is based on the same

criteria, load to induce yielding in a bottom flange, as the FEA girder yielding strength

of 19 trucks. This demonstrates conservatism in the existing methods for distributing

load to individual girders. However, these system capacities are based on linear-

elastic concrete deck properties. The approach proposed in this work is to extend this

analysis to include the inelastic response of concrete post-cracking and determine how

this change affects the system capacity of bridges.

2.1.2 Corroded Reinforced Concrete


Li and Zheng (2005), Oyado et al. (2011), and, less significantly, Gu et al.

(2010) were three different studies relied upon in the present work. In these studies,

6
reinforced concrete beams were constructed and accelerated corrosions techniques

were utilized, with the exception of Gu et al. (2010) which analyzed an existing

structure which has undergone natural corrosion. These studies and their relevant

results are described below.

2.1.2.1 Experimental Set-Up


Each of the tests utilized a different test set-up. The first test, Li and Zheng

(2005), used an accelerated salt-spraying technique and then converted the accelerated

time to real time. A marine environment was simulated by alternating wetting and

drying cycles of saltwater spray in a chamber constructed specifically for the testing.

That is, the beams were sprayed for 2 hours then let sit for an hour before being

sprayed for another 2 hours. The relative humidity of the chamber was also

controlled. The test accelerated corrosion by spraying sodium chloride directly on the

cracks and intensifying the drying phases of the wetting and drying cycles. This

accelerated test was calibrated using identical specimens under the natural

environment and the accelerated time was converted to real time. The accelerated test

took place over a seven month period, equivalent to approximately 9 years in real

time. During this testing, both strength and deflection data were recorded initially, at
3 months, 5 months, and 7 months.

In the second test, Oyado et al. (2011), the beams were placed outdoors, in an

urban area away from the coast, for 3 months; after this time, the beams were sprayed

with a saline solution 3 times a day for 17 months in attempts to accelerate corrosion.

After this time, some of the beams were strength tested and the remaining beams were

left outdoors for a total of 12 years, after which time they were also strength tested.

7
The first set of testing, performed at 20 months, provided only strength data. The

second set of testing, performed at 12 years, provided strength and deflection data.

The final test, Gu et al. (2010), only included stiffness data. However, this

report was vague regarding the condition of the concrete which was tested, stating it

was removed from a building which had “gone through decades of natural corrosion”.

For this reason, less emphasis was placed on the results of this testing and it was used

more for comparison than specific quantitative data.

2.1.2.2 Results of Testing and Concrete Performance Goals


The strength losses reported by Li and Zheng (2005) and Oyado et al. (2011)

are summarized in Figure 2.1. In this figure, strength loss is quantified by the

decrease in strength between the corroded and uncorroded specimens normalized by

the strength of the original, uncorroded, beam. The stiffness losses reported by the

previously mentioned literature are summarized in Figure 2.2. In this figure, change

in stiffness is quantified by the increase in ultimate deflection between corroded and

uncorroded specimens normalized by the ultimate deflection of the uncorroded

specimen.

8
Figure 2.1. Strength losses reported in literature (McConnell et al. 2012).

Figure 2.2. Stiffness losses reported in literature (McConnell et al. 2012).

When reviewing the strength loss values, a trend can be observed in that the

data from Li and Zheng (2005) which suggests a linear relationship between chloride

exposure time and strength loss. Comparing this data to Oyado et al. (2011) at 20

9
months, it can be observed that greater strength losses are reported. This is possibly

due to the fact that, for months 3 to 20, the Oyado et al. (2011) specimens were

sprayed with a chloride solution 3 times per day. It can be hypothesized that this

accelerated the corrosion of these specimens over this time period. If the linear trend

using the Li and Zheng (2005) data is extrapolated, it intersects with the 12 year

Oyado et al. (2011) data.

Using this relationship, it can be concluded that the strength loss data from Li

and Zheng (2005) and Oyado et al. (2011) are in general agreement with one another.

For this reason, a linear trend using the data from Li and Zheng (2005) and the 12 year

data from Oyado et al. (2011) was fit. The 20 month data from Oyado et al. (2011)

was not included for the previously stated reasons involving the unexpectedly high

strength loss values. This resulted in the following relationship, shown in Equation

2.1.

𝑠𝑡𝑟𝑒𝑛𝑔𝑡ℎ 𝑙𝑜𝑠𝑠(%) = 2.0002 ∗ (𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑦𝑒𝑎𝑟𝑠) [2.1]

This relationship was extrapolated for a 25 year design life, resulting in a prediction of

50% strength loss. With a lack of better data, this is used as the assumed strength loss

chosen to represent a corroded deck condition in the following analyses.


When reviewing the stiffness data, only the results from Li and Zheng (2005)

includes data from multiple time periods. This data indicates a bi-linear relationship

between ultimate deflection and time where change in deflection increases rather

rapidly for the first 4 years of exposure and then becomes more gradual. Extrapolating

the second linear portion of the trend line results in the Li and Zheng (2005) data

intersecting with the Oyado et al. (2011) data; for this reason, the second portion of the

10
bi-linear curve is fitted with a linear trend line and used to extrapolate the increase in

deflection that would be expected after 25 years of chloride exposure. This trend line

resulted in the following relationship, shown in Equation 2.2.

𝑑𝑒𝑓𝑙𝑒𝑐𝑡𝑖𝑜𝑛 𝑖𝑛𝑐𝑟𝑒𝑎𝑠𝑒(%) = 2.0347 ∗ (𝑛𝑢𝑚𝑏𝑒𝑟 𝑜𝑓 𝑦𝑒𝑎𝑟) + 31.559 [2.2]

This equation indicated an 82% increase in ultimate deflection after 25 years. With a

lack of better data, this is used as the assumed deflection increase chosen to represent

a corroded deck condition in the following analyses.

It should also be noted that information regarding changes in mass, due to

spalling of concrete or corroding of reinforcing rebar, was also evaluated. However, it

was decided that these changes in mass were encompassed by the observed changes in

strength and deflection; therefore, it can be considered a cause of observed changes

versus an effect in the following analyses.

2.2 Reinforced Concrete and Corrosion Modeling


In order to determine the optimal way to model reinforced concrete, a literature

review on available modeling techniques was performed. When modeling rebar and
reinforced concrete, different methods were investigated. The main difference

between these modeling approaches is whether the rebar is modeled as 2-dimensional

(2-D) or 3-dimensional (3-D) elements (Simulia 2011). The 2-D rebar can be

embedded into the concrete, creating a perfect bond between the concrete and rebar.

However, the 3-D rebar allows for more properties discussed in prior research to be

incorporated; these properties included debonding of concrete and rebar which causes

the friction between the concrete and rebar to vary, considered by both Val et al.

11
(2009) and Amleh and Ghosh (2006); increased contact pressure between the rebar

and the surrounding concrete due to the creation of rust during the corrosion process

(Amleh and Ghosh 2006); and associated mass loss of rebar (Amleh and Ghosh 2006).

For the purposes of this research, papers which used ABAQUS were isolated.

These were reviewed and emphasis was placed on these techniques as they would be

easiest to replicate due to models in this file format already existing. Val et al. (2009)

describes a technique to map cracking of corroded reinforced concrete beams. This

approach used 2-D elements for both concrete and rebar, using Rankine criterion to

define crack initiation. Friction between the reinforcing bars and the surrounding

concrete is modeled using Coulomb’s friction model with a constant coefficient of

friction. The corroded rebar was modeled using a penalty contact algorithm for the

interaction between the rebar and concrete, which defines the master and slave

surfaces and how far the slave surface can penetrate into the master surface.

Deformability of corrosion products is neglected as a simplification; however, the

cross-sectional area of the rebar is increased to reflect the creation of corrosion

products. In this model, only corroded reinforced concrete is analyzed.

Amleh and Ghosh (2006) describes a pull-out test which was performed on

corroded concrete then modeled. This approach describes using 3-D elements for both
concrete and rebar, defining the concrete as the master surface and the rebar as the

slave surface. It describes using a pressure-overclosure relationship between the

contact surfaces to include the creation of corrosion by-products. In this study,

friction was defined between the concrete and rebar surfaces as decaying

exponentially from the static value to the kinetic value as the rebar begins to slip

within the concrete. This decay was then related to the mass loss of rebar caused by

12
corrosion. A more comprehensive description of these modeling techniques and how

they apply to ABAQUS are provided in Chapter 3.

In addition to articles which specifically used ABAQUS, other approaches to

modeling reinforced concrete and corrosion were reviewed to determine if they could

be applied to ABAQUS models. Fang et al. (2006) modeled the rebar as 3-D elements

with rust modeled as a granular material between the rebar and concrete elements. A

similar approach to modeling corrosion was used by Dekoster et al. (2003), modeling

rust as a component between the steel and concrete with mechanical properties almost

identical to that of water. Kallias and Rafiq (2010) modeled the rebar as truss

elements, representing the bond between the rebar and concrete using 2-D interface

elements. To model the corrosion, stress was increased in certain locations to mimic

pitting. Chen and Mahadevan (2007) also considered that corrosion by-product takes

up more volume than the structurally-sound amount of rebar lost to create the

corrosion products and therefore creates pressure. Coronelli and Gambarova (2004)

modeled the concrete as 2-D elements, with the rebar modeled as truss elements.

Bond-link elements were used to couple the concrete and rebar elements to one

another. To model corrosion, the cross-sectional area of the rebar was decreased and

the ultimate strains of the steel were decreased. Biondini and Vergani (2012)
specifically focused their model on the difference between uniform corrosion, pitting,

and a combination of the two. Three-dimensional elements were used for both the

concrete and rebar, and equations were created and input into the model to change the

cross-sectional area and shape of the rebar to reflect the effects of the corrosion. The

cross-sectional area was uniformly decreased due to the creation of corrosion by-

product and/or decreased locally due to pitting by altering the cross-sectional shape.

13
All of the previously described models utilized a smeared crack approach to modeling

the cracking initiation and propagation in concrete.

An alternative approach to modeling reinforced concrete is called the

multifiber approach (Richard et al. 2011 and Adelaide et al. 2012). This approach is

primarily used to decrease the global computational costs. The concrete and steel are

represented by fibers and the equilibrium equations provide the nodal displacements

and rotations. However, this technique does not provide accurate local results in terms

of cracking pattern, crack opening, crack spacing etc. and was therefore considered not

able to accurately model small scale beams as required in this research.

14
Chapter 3

MODELING APPROACH

In this chapter, the different ABAQUS commands and techniques which were

utilized in creating a finite element model of reinforced concrete are discussed. This

includes both the mechanics behind each command and the variables which are input

into ABAQUS to quantify the underlying behavior; in addition, the different modeling

techniques available within ABAQUS that were used within this research for the

purposes of modeling corrosion of reinforced concrete due to chlorides are discussed.

Specifically, this includes the different material models of concrete (Section 3.1) and

rebar that are used as well as the different approaches to modeling corrosion of the

rebar (Section 3.2), which differs depending on whether the rebar is modeled as 2-D or

3-D elements. Exact values used in these commands are discussed in Chapter 4. The

boundary conditions (Section 3.3) used in the models are also discussed and this

chapter concludes with a discussion of the different analysis methods available within

ABAQUS as well as the pros and cons of these alternative methods with respect to the

present research objectives (Section 3.4).

3.1 Concrete
ABAQUS offers three different approaches for modeling the non-linear

behavior of concrete; in each of these models, the elastic portion of the material

response is consistent. ABAQUS uses the elastic definition to determine the material

response until the material reaches the defined cracking stress; at this point, the non-

linear behavior of the material governs, including the post-cracking response. The

elastic commands are described in Section 3.1.1 and the non-linear behavior is

described in Section 3.1.2.

15
3.1.1 Elastic Behavior
As stated previously, all of the concrete models utilize the same linear-elastic

behavior. For this behavior, the modulus of elasticity is defined for concrete (Ec), as

well as Poisson’s ratio (ν). A standard relationship between Ec, compressive concrete

strength (f’c), and tensile concrete strength (f’t) was assumed and are expressed in

Equations 3.1 and 3.2, where Ec, f’c, and f’t are expressed in psi.

𝑓 ′ 𝑡 = 7.5√𝑓′𝑐 [3.1]

𝐸𝑐 = 57,000√𝑓′𝑐 [3.2]

These material properties are defined using the “elastic” command within

ABAQUS. For the purposes of these analyses, it was assumed that the material was

isotropic, and this parameter was included in the “elastic” command. In addition to

the “elastic” command, the density was also defined for the concrete. This value was

included using the “density” command within ABAQUS. The exact values which

were used for these commands can be found in Section 4.1.1. These elastic commands

do not directly take into consideration f’c or f’t.

3.1.2 Non-Linear Behavior


As stated previously, three different modeling techniques for modeling non-

linear behavior of concrete are available within ABAQUS: the smeared cracking

model (SC), the concrete damaged plasticity model (CDP), and the brittle cracking

model. These are described in the following Sections 3.1.2.1 to 3.1.2.3, respectively.

16
3.1.2.1 Smeared Crack Model
One modeling approach for post-cracking behavior of concrete which was

explored was the SC model. In this model, ABAQUS employs a smeared cracking

technique; rather than tracking individual cracks, the smeared cracking technique

performs constitutive calculations independently at each integration point and the

presence of cracks enters into these calculations through the stress and material

stiffness associated with the integration point (Simulia 2011). These stress and

material stiffness values are defined by the user in the commands associated with the

SC technique. The SC model is intended as a model of concrete behavior for

relatively monotonic loadings under low confining pressures, with cracking assumed

to be the most important aspect of the behavior.

The SC approach utilizes a Rankine criterion to detect crack initiation; a crack

forms in the direction normal to the maximum principle tensile stress when this stress

reaches a failure surface that is a linear relationship between the equivalent pressure

stress and the Mises equivalent deviatoric stress (Simulia 2011). Once it forms at a

point, the crack orientation is stored for subsequent calculations. A new crack at the

same point can form only in a direction orthogonal to the direction of an existing

crack. These cracks may open and close as the integration point goes into tension and
compression, but remain for all subsequent calculations.

Once a crack has formed, the load transfers across cracks through the rebar,

modeled using tension stiffening; tension stiffening defines the stress-strain response

after cracking and unloads to zero stress at a level of strain defined by the user. This

tension stiffening relationship can be seen in Figure 3.1. ABAQUS also takes into

consideration the change in shear modulus of the concrete, which affects the shear

17
behavior post-cracking. One of the most defining features of the SC model is that it is

only applicable when utilizing ABAQUS/Standard, as described further in Section 3.4.

Figure 3.1. Tension stiffening model ABAQUS employs in the smeared crack
technique (Simulia 2011).

In order to utilize the SC model, certain ABAQUS commands are required.

The first command, called the “concrete” command, defines the stress-strain behavior

of plain concrete in uniaxial compression outside the linear elastic range. The next

command is the “tension stiffening” command. This command defines the fraction of

remaining stress to stress at cracking as a function of the absolute value of the direct

strain minus the direct strain at cracking. The final command used is the “failure

18
ratios” command. This defines the shape of the failure surface for a concrete model

by defining the ratio of the ultimate biaxial compressive stress to the uniaxial

compressive ultimate stress, the absolute value of the ratio of uniaxial tensile stress at

failure to the uniaxial compressive stress at failure, the ratio of the magnitude of a

principal component of plastic strain at ultimate stress in biaxial compression to the

plastic strain at ultimate stress in uniaxial compression, and the ratio of the tensile

principal stress value at cracking in plane stress to the tensile cracking stress under

uniaxial tension. The exact values which were input into the model are described in

Section 4.4.1.3.

3.1.2.2 Concrete Damaged Plasticity Model


Another concrete model investigated was the CDP model. This model is a

continuum, plasticity-based damage model for concrete, assuming that the two main

failure mechanisms are tensile cracking and compressive crushing (Simulia 2011).

Under uniaxial tension, the stress-strain response follows a linear elastic relationship

until the value of the failure stress (σt0), is reached; σt0 is calculated by ABAQUS

when the cracking strain is achieved. This relationship can be seen in Figure 3.2 (a).

The failure stress corresponds to the onset of micro-cracking in the concrete material,
beyond which the formation of micro-cracks is represented macroscopically with a

softening stress-strain response. This softening induces strain localization in the

concrete and under uniaxial loading is linear below the value of initial yield (σc0)

(Simulia 2011). Under multiaxial loading, the stress-strain relations are given by a

scalar damage elasticity equation which utilizes a scalar stiffness degradation variable,

calculated by ABAQUS, that is generalized to the multiaxial stress case to modify the

undamaged elasticity matrix. A similar approach is used to model the compressive

19
behavior, defining the stress-strain behavior of plain concrete in uniaxial compression

outside the elastic range and using a generalized scalar stiffness degradation variable

to model multiaxial behavior, as can be seen in Figure 3.2 (b).

Figure 3.2. Response of concrete to uniaxial loading in (a) tension and (b)
compression for CDP model (Simulia 2011).

20
Tension stiffening is again used to model the stress-strain response between the

concrete and rebar after cracking; that is, tension stiffening defines how the load is

transferred to the rebar from the concrete as it cracks. In addition, damage can be

specified. These variables are treated as non-decreasing material point quantities and

correspond to reductions in stiffness. Should this input not be included, the model

behaves as a plasticity model. There are separate variables to define the tension and

compression damage coefficients. In addition to these other values, flow potential,

yield surface, and viscosity parameters can be defined. As with the smeared crack

model, the CDP model is applicable only when using ABAQUS/Standard, as is

described further in Section 3.4.

To properly define the CDP model using ABAQUS, many different commands

need to be utilized. The first of these is the “concrete damaged plasticity” command.

This command defines the dilation angle, flow potential eccentricity (ϵ), ratio of initial

equibiaxial compressive yield stress to initial uniaxial compressive yield stress

(σb0/σc0), and the ratio of the second stress invariant on the tensile meridian to that of

the compressive meridian at initial yield for any given value of the pressure invariant

such that the maximum principal stress is negative (Kc) (Simulia 2011).The next

command is the “concrete tension stiffening” command. This is used to define post-
cracking properties for concrete that is in tension by using a multi-linear relationship

to define the remaining direct stress after cracking and the associated direct cracking

strain. The “concrete compression hardening” command defines the equivalent post-

cracking properties for concrete in compression by using a multi-linear relationship

and defining the stress after yielding in compression and associated crushing strain.

The “concrete tension damage” and “concrete compression damage” commands define

21
post-cracking damage properties for concrete in tension and compression,

respectively, by defining a damage variable and associated crushing strain to include

stiffness degradation as the cracking proliferates. The exact values which were used

for this input are discussed in Section 4.4.1.2.

3.1.2.3 Brittle Cracking Model


The final concrete constitutive modeling approach to be discussed is the brittle

cracking model. As with the previous SC model, the concrete is modeled using a

smeared crack model; rather than tracking individual cracks, constitutive calculations

are performed independently at each integration point and the presence of cracks

enters into these calculations through the stress and material stiffness associated with

the integration point. This is in contrast to the CDP model where the formation of

micro-cracks is represented macroscopically with a softening stress-strain response.

As with the SC model, the brittle cracking model considers tension stiffening

and shear retention. However, unique to the brittle cracking model, ABAQUS has the

capability of defining brittle failure of a material. That is, when any of the local direct

cracking strain components at a material point reach the input value for failure strain,

the material point fails and all the stress components are set to zero. If all the material
points fail within an element, the element is removed from subsequent calculations.

The primary difference between the brittle cracking model and the SC model is that,

rather than using ABAQUS/Standard, the brittle cracking model is only applicable

when ABAQUS/Explicit is utilized, as is explained further Section 3.4.

To implement the brittle cracking model in ABAQUS, the first command that

is specified is the “brittle cracking” command. This command defines the stress

which causes a crack to form, as well as the constitutive relationship after the element

22
fails, which is referred to as tension stiffening. The next command is the “brittle

shear” command. A retention factor is employed to specify the post-cracking shear

behavior by entering the ratio of shear strength to original shear strength as a function

of the crack opening strain. The final command controlling the post-cracking behavior

of concrete used is the “brittle failure” command. This is the command which

specifies the strain at which the material points should fail and the element is removed

from subsequent calculations. The exact input values which are used for each of these

commands is discussed in Section 4.5.1.1.

3.2 Rebar
In order to determine the most accurate way to model corroded reinforced

concrete using ABAQUS, previous research was reviewed. Specifically, research

using ABAQUS was isolated and evaluated. Two different techniques were selected;

one which models rebar as 2-D elements and embeds them within the concrete, and

another which models the rebar as 3-D elements, defining how the surfaces of the

rebar and concrete elements interact with one another. These approaches are detailed

in the following Sections 3.2.1 and 3.2.2.

3.2.1 2-Dimensional Rebar


For the 2-D rebar model, the rebar are modeled as 2-D rod elements. These

elements and nodes are defined separately from those defining the concrete portion of

the model. The elements are then embedded in the concrete, constraining the response

of the rebar’s nodal translational degrees of freedom to that of the concrete, using the

“embedded element” command and defining the host and embedded element sets.

This approach is specifically designed by ABAQUS to model rebar in reinforced

23
concrete, although it can be used for other purposes. This approach is the most simple

and straightforward, but provides few opportunities for modeling the effects of

corrosion. The only properties which can be changed for the rebar are the cross-

sectional area and strength; no changes can be made to affect the bond between the

rebar and concrete.

3.2.2 3-Dimensional Rebar


When modeling the rebar as 3-D elements, the concrete and rebar elements are

modeled with separate nodes and elements from one another. The literature then

suggests a surface interaction between the concrete and rebar surfaces be defined (Val

et al. 2009). Once the surface interaction has been defined, many different properties

can be included. Most notable of these properties are friction and the pressure applied

by the creation of corrosion by-products around the rebar.

ABAQUS has many different options when defining the friction between two

surfaces. The most simple of these is to define a constant friction coefficient (µ). This

coefficient is defined independent of the slip rate, although a dependency can be

manually defined. Another approach is to define an exponential decay curve, where

the friction value starts at the static friction value (μs) when the slip rate is zero, then
exponentially decays to the kinetic friction value (μk) based on increasing slip rate
(γ̇ eq ) and a decay coefficient (dc), as shown in Figure 3.3 (Amleh and Ghosh 2006).

In this approach, the slip rate is calculated by ABAQUS at each loading increment and

is applicable to both ABAQUS/Standard and ABAQUS/Explicit. It is suggested that

this approach is more appropriate when modeling corrosion of rebar as the effects of

corrosion can be more directly incorporated. The exponential decay function is

defined by Equation 3.3.

24
𝜇 = 𝜇𝑘 + (𝜇𝑠 − 𝜇𝑘 )𝑒 −𝑑𝑐𝛾̇ 𝑒𝑞 [3.3]

The decay coefficient is defined by the user. Amleh and Ghosh (2006) used previous

test results and compared them to their model to determine the most accurate dc value,

which was also used for this analysis. This value, along with µs and µk, are detailed in

Section 4.5.2.2.

Figure 3.3. Exponential decay friction model (Simulia 2011).

In order to define the friction, the “friction” command is utilized. When

defining a constant friction coefficient, no other information is required besides the

value of µ. However, the parameter “exponential decay” must be included to use the

exponential decay curve. After including this parameter, μs, μk, and dc are defined by
the user and γ̇ eq is calculated at each loading increment automatically by ABAQUS,

then applied using Equation 3.3. In addition to the “friction command,” the “surface

interaction” and “contact pair” commands are required. The “surface interaction”

25
command creates a surface interaction property definition; this command defines a

label that will be used to reference the surface interaction property in the “contact

pair” command and also allows for the inclusion of an interfacial layer between the

contact surfaces. The “contact pair” command defines pairs of node sets or surfaces

that may contact or interact with each other during the analysis. This is where the

surfaces of the concrete and rebar which are in contact are directly defined.

To model the pressure caused by the creation of rust, literature suggests the use

of a pressure-overclosure relationship (Amleh and Ghosh 2006). The most basic of

this approach is to use a “hard” contact relationship where the surfaces transmit no

contact pressure unless nodes of the slave surface contact the master surface. In the

case of reinforced concrete, the slave surface is the rebar and the master surface is the

concrete. This approach does not allow penetration at each constraint location and

there is no limit to the magnitude of contact pressure that can be transmitted when the

surfaces are in contact. A graph of this relationship can be seen in Figure 3.4. A

contact clearance at which the contact pressure is zero (c0) can be defined, allowing

for space between surfaces to be defined before contact is made. In addition, a linear

penalty stiffness value can be defined; typically this value is calculated by ABAQUS

and assumed to be 10 times a representative underlying element stiffness and defines


the relationship between contact pressure and overclosure.

26
Figure 3.4. Default “hard” pressure-overclosure relationship (Simulia, 2011).

In contrast to this “hard” contact relationship, a “softened” contact relationship

is available. This is used to model a soft, thin layer on one or both surfaces and can be

better numerically because it can be easier to resolve the contact conditions (Simulia

2011). In order to define a “softened” contact relationship, a type has to be chosen.

The available types include using a linear law, a tabular piecewise-linear law, or an

exponential law. The optimal version to model corrosion was found to be the

exponential law (Amleh and Ghosh 2006), which is shown in Figure 3.5. This

relationship takes into consideration the increase in pressure as the surfaces get closer,

and allows for the pressure to become zero should the surfaces no longer be in contact.

This exponential relationship is based on c0, the pressure at zero clearance (p0), and,

when employing ABAQUS/Explicit, the maximum stiffness value (kmax). The kmax

value is a required parameter when utilizing ABAQUS/Explicit and is not available in

ABAQUS/Standard; limiting this value can be useful for penalty contact to mitigate

the effect that large stiffnesses have on reducing the stable time increment. Exact

values which were used and input into the model are discussed in Section 4.5.2.2.

27
Figure 3.5. “Softened” exponential pressure-overclosure relationship (Simulia 2011).

In order to use a pressure-overclosure relationship, the “surface behavior”

command must be utilized. The default for this command is to apply the “hard”

relationship. In this approach, c0 is defined, along with a linear penalty stiffness. A

parameter can be added to utilize the exponential pressure-overclosure relationship

where c0 and p0 are defined. When employing ABAQUS/Explicit, a kmax value is also

defined.

One of the benefits of using the exponential pressure-overclosure relationship

over the default “hard” relationship is the ability to include corrosion effects. Using
the results from pull out tests, the pressure and friction were related to the concrete

cover thickness (C) and the mass loss as a percentage (M) caused by corrosion in prior

work by Amleh and Ghosh (2006). The pressure at zero clearance for uncorroded

concrete is defined by Equation 3.4, where p0 is expressed in MPa and C is expressed

in mm.

𝑝0 = 0.128𝐶 + 1.5 [3.4]

28
This p0 changes as rebar corrodes and M increases. The percentage loss of contact

pressure (L) is related to M and f’c by Equation 3.5, which is then multiplied by p0 to

determine a new pressure at zero clearance for corroded rebar (Amleh and Ghosh

2006).

𝐿 = [(−0.00024𝑓 ′ 𝑐 − 0.0028)𝐶 + 4.3]𝑀 [3.5]

In this empirical equation, L and M are expressed as a percentage, f’c is in MPa, and C

is in mm.

In addition to the pressure changing, μs and dc also change due to mass loss.

These changes can be calculated using the following Equations 3.6 and 3.7,

respectively (Amleh and Ghosh 2006), where M is expressed as a percentage.

𝜇𝑠 = 𝑒 −0.035𝑀 [3.6]

𝑑𝑐 = 0.0261𝑀 + 0.45 [3.7]

3.3 Boundary Conditions


The original beam test set-up utilized by Oyado et al. (2010) loaded the beam
under simple support conditions. It was found during initial modeling that using one

row of nodes as a pin and another row of nodes as a roller caused high bearing forces

and deformations at the supports. For this reason, each of the supports were modeled

across multiple rows of nodes. The number of rows of nodes was directly related to

the size of the mesh; the supports were modeled over 1.57 in (40 mm). This was

chosen based on the 0.7874 in (20 mm) mesh size using 3 rows of nodes. This model

was the one used to analyze the support condition, and the size of the support was kept

29
consistent between mesh sizes. All of the supports nodes were modeled as rollers,

only limiting vertical displacement, with the exception of one node on the pinned end.

This node, located in the center of the defined support, was fixed in all directions with

the exception of rotation about the length of the beam. It was found that this provided

adequate restraint while still accurately modeling the fixity of the support. The

bearing forces were no longer considered high when the nodes defined for the support

and the elements attached to those nodes displayed no deformation before the model

failed.

3.4 Analysis Method


ABAQUS offers two different techniques for performing analyses:

ABAQUS/Explicit and ABAQUS/Standard. ABAQUS/Explicit is an explicit

dynamic analysis. It is more computationally efficient for large models with relatively

short dynamic response times and allows for the definition of general contact

conditions (Simulia 2011). This approach uses a consistent, large-deformation theory

where models can undergo large rotations and large deformations. It can also use a

geometrically linear deformation theory where strains and rotations are assumed to be

small. It allows for either automatic or fixed time incrementation to be used and can
be used to perform quasi-static analyses with complicated contact conditions. To

implement this type of analysis, the “dynamic” command is used, specifying the

optional parameter of “explicit.”

As opposed to the explicit dynamic analysis of ABAQUS/Explicit,

ABAQUS/Standard is a static stress analysis. This is used when inertia effects can be

neglected and can be linear or nonlinear (Simulia 2011). This analysis ignores time-

dependent material effects such as creep, swelling, and viscoelasticity; however, it

30
takes rate-dependent plasticity and hysteretic behavior for hyperelastic materials into

account. To implement this type of analysis, the “static” command is used.

With each of these different analysis techniques come different loading

definitions. For ABAQUS/Explicit, a loading amplitude is defined using the

“amplitude” command. With this command, the time and load proportion are defined

by the user and can be applied as a ramp or sustained. With the loading defined, the

analysis then performs the applicable calculations to determine stress and

displacements.

For ABAQUS/Standard, the “riks” command was utilized. This method is

generally used to predict unstable, geometrically nonlinear collapse of a structure and

can include nonlinear materials and boundary conditions (Simulia 2011). This method

uses the load magnitude as an additional unknown and solves simultaneously for loads

and displacements. This approach provides solutions regardless of whether the

response is stable or unstable and is only applicable to ABAQUS/Standard.

There are different pros and cons associated with these different analysis

techniques. ABAQUS/Standard allows for calculating a static loading corresponding

to the equilibrium condition of the deformed structure and clearly indicates failure by

reaching a peak loading then decreasing; this gives a direct quantitative definition of
failure. Conversely, ABAQUS/Explicit utilizes a dynamic loading approach that may

over-estimate realistic loads, requiring judgment to assess failure; this approach also

includes a time component, adding another parameter needing to be analyzed and

calibrated. These loading differences cause the two methods to require differing levels

of judgment to determine when failure has occurred, with ABAQUS/Standard being

more straightforward. The biggest con with utilizing ABAQUS/Standard is that,

31
ideally, the concrete input should be taken from actual concrete testing results; that is,

the “concrete compression hardening” and “concrete tension stiffening” commands

require multiple input values which should be based on actual testing results that are

not widely documented. However, applying the results of concrete testing similar to

the problem of interest, the approach which is used in this research, can also result in

calibrated input. This is in contrast to ABAQUS/Explicit modeling techniques with

input values which can be more readily estimated based on common concrete

properties.

32
Chapter 4

MODEL CALIBRATION

The ABAQUS variables and commands described in Chapter 3 were utilized to

model corrosion within a reinforced concrete beam. An uncorroded base model was

created using the original material property input and modeling commands as

described in Sections 3.1 and 3.2. Once the input and modeling approach for the final

uncorroded model were determined, these input variables were altered in attempts to

simulate corroded reinforced concrete and obtain the targeted amounts of strength

decrease and deflection increase associated with corrosion, as previously explained in

Section 2.1.2.2. The geometry and material properties which were modeled are

described in Section 4.1, with the concrete modeling described in Section 4.1.1 and the

rebar modeling described in Section 4.1.2. Section 4.2 defines how the results are

evaluated and standardized in order to easily compare different modeling techniques.

Hand calculations were performed to compare to the results of the finite element

model; this was done for both strength and deflection. These calculations are

described in Section 4.3.

The 2-D beam model is described in Section 4.4. Within this section, the

uncorroded base model input to simulate an uncorroded reinforced concrete beam is

described in Section 4.4.1, the mesh sensitivity analysis and resulting final uncorroded

model input are described in Section 4.4.2, and the calibration and determination of

the input for the corroded model are described in Section 4.4.3.

In addition to the 2-D beam model, a 3-D beam model was created. This

model is described in Section 4.5. Initially, this modeling was performed with 2-D

rebar elements, as described in Section 4.5.1; the model was then refined by using 3-D

33
rebar elements, as described in Section 4.5.2. The conclusions drawn from these

different modeling techniques can be found in Section 4.6.

4.1 Beam Geometry and Material Properties


Certain input was constant throughout all different types of beam modeling.

This includes the strength and modulus of elasticity of the concrete as well as the

elastic and plastic parameters of the rebar. This concrete and rebar input is referred to

as the uncorroded base model and is described in Sections 4.1.1 and 4.1.2,

respectively.

4.1.1 Concrete
For the purposes of this study, a beam design which was previously corroded

and tested (Oyado et al. 2010) was used to calibrate modeling techniques and inputs.

This beam was chosen based on the review of experiments of this type conducted by

McConnell et al. (2012), which is described in Section 2.1.2. The geometry of this

beam can be seen in Figure 4.1. For ease of creating the model, the hooks that were

located on the ends of the tension and compression reinforcement as well as the

stirrups in the physical specimen were ignored. The rebar which were used for

compression, tension, and stirrups had varying sizes and strengths; the #2 bars used for

compression reinforcement and stirrups were SD295 and the #4 bars used for tensile

reinforcement were SD345. These material designations correspond to a minimum

yield strength of 42,786 psi (295 MPa) and 50,038 psi (345 MPa), respectively.

34
Figure 4.1. Geometry of finite element calibration models, elevation view (top) and
cross-section view (bottom) (dimensions in inches).

The material property inputs selected for the uncorroded base model (using the

material properties of the tested specimen and the input values suggested in literature)

are based on those from Oyado et al.’s (2010) uncorroded specimen S-0N, their

specimen which carried the highest load. The tested material properties of this

specimen were reported as a compressive strength of 3,147.3 psi (Oyado et al. 2010),

resulting in a modulus of elasticity of 3,197,746 psi and a tensile strength of 420.8 psi,

when calculated using Equations 3.2 and 3.1, respectively. A standard value of

Poisson’s ratio of 0.2 was also assumed. A failure crack strain value of 0.0027 was

assumed; this is the amount of additional strain which can be carried by the concrete

after an initial crack forms and before complete failure. This value was based on the

difference between the ultimate strain and the strain at first cracking; an ultimate strain

(the strain at which the concrete fails completely and can no longer carry load) of

35
0.003 was utilized and it was assumed that the first crack occurs at a strain of 10% of

the ultimate strain (i.e., 0.0003).

4.1.2 Rebar
Standard elastic steel material property inputs were specified for the rebar,

which included the following assumptions: the modulus of elasticity of the rebar (Es)

was assumed to be 29,000 ksi, the Poisson’s ratio was assumed to be 0.3, and the

density was assumed to be 0.000734 lb/in3 (1.27 lb/ft3). The input into ABAQUS is

based on inputting all values in consistent units, where pounds and inches were used in

these models.

In addition to this value, the plastic properties of the rebar were also included.

For these, the yield and ultimate strengths of the SD295 and SD345 rebar, used by

Oyado et al. (2010) and serving as the calibration specimen, were researched. As

stated in Section 4.1.1, the stirrups and compressive reinforcement are comprised of

#2 bars which were SD295 and the tensile reinforcement are comprised of #4 bars

which were SD345. The minimum yield strengths of SD295 and SD345 are 42,786

psi (295 MPa) and 50,038 psi (345 MPa), respectively. However, two different

research papers were found which included actual yield and ultimate strength values
of both rebar types: Takahashi (2008) and Shirai et al. (2002). These values can be

found in Table 4.1, where Takahashi (2008) provided 1 set of values for the rebar

strengths and Shirai et al. (2002) provided 2 sets of values for the rebar strength,

which were averaged before being included in Table 4.1. In this table, the percent

difference is compared to the minimum specified strengths of 295MPa and 345MPa

for SD295 and SD345, respectively. These values are consistent with a trend which

exists for rebar grades typically used in the US, where a factor of 1.1 is applied in

36
some situations to approximate the actual rebar strength relative to the minimum

specified strength (Morales n.d.). Thus, this factor was multiplied by the minimum

specified yield strength to obtain the yield stress in the models. For the ultimate

strength, the individual values provided by Takahashi (2008) and Shirai et al. (2002)

were averaged and this average was used.

Table 4.1. Yield and ultimate strengths of SD345 and SD295 rebar found in literature
compared to the minimum specified yield strengths.
Takahashi (2008) Shirai et al. (2002)
Fy (psi) % Difference Fu (psi) Fy (psi) % Difference Fu (psi)
SD345 53,809 8% 82,671 55,694 11% 78,465
SD295 54,679 28% 77,885 48,588 14% 91,519

For both yield and ultimate strengths, these engineering stresses were

converted to true stresses and plastic logarithmic strains for input into ABAQUS. This

was done by first calculating the engineering strain from the engineering yield stress

using Es by utilizing Equation 4.1, where σy is the yield stress value and εy is the

corresponding yield strain value.

𝜎𝑦
𝐸𝑠 = [4.1]
𝜀𝑦

In addition to the yield and ultimate stresses, two more values are calculated to

make a more complete plastic response. Together, these data points describe a linear

elastic regime, followed by a yield plateau, followed by strain hardening, followed by

a second plateau after strain hardening terminates. This rationale, as well as the

specific stiffnesses and strain values associated with this multi-linear response, are

37
based on the steel material modeling discussed in Barth et al. (2005). The end of the

yield plateau corresponds to a strain of 0.011. The calculation of this strain, σ1, was

done by using Equation 4.2, where σ1 is in psi.

𝜎1 = 𝜎𝑦 + 145,000(0.011 − 𝜀𝑦 ) [4.2]

The strain corresponding to the ultimate stress was calculated using Equation 4.3,

where εu is the strain corresponding to the ultimate stress and σu is the ultimate stress

in psi.

𝜎 −𝜎
𝑢 1
𝜀𝑢 = 0.011 + 720,000 [4.3]

The final stress which was utilized was a value larger than the ultimate stress and

corresponding to a strain of 0.3. This value, σ2, was calculated using Equation 4.4,

where σ2 is in psi.

𝜎2 = 𝜎𝑢 + 145,000(0.3 − 𝜀𝑢 ) [4.4]

Once the 4 pairs of stress-strain input were determined, the true stress was then

calculated by multiplying the engineering stress by the engineering strain using

Equation 4.5, where σtrue is the true stress value, σeng is the engineering stress value,

and εeng is the corresponding engineering strain.

𝜎𝑡𝑟𝑢𝑒 = 𝜎𝑒𝑛𝑔 (1 + 𝜀𝑒𝑛𝑔 ) [4.5]

38
Lastly, the plastic logarithmic strain is calculated by using Equation 4.6, where εlnplastic

is the logarithmic plastic strain.

𝑝𝑙𝑎𝑠𝑡𝑖𝑐 𝜎𝑡𝑟𝑢𝑒
𝜀𝑙𝑛 = ln(1 + 𝜀𝑒𝑛𝑔 ) − [4.6]
𝐸𝑠

The plastic input values for SD295 and SD345 rebar resulting from these equations

can be found in Table 4.2.

Table 4.2. True stress and true plastic strain input values for SD295 and SD345 rebar.
SD295 SD345
σtrue (psi) εlnplastic σtrue (psi) εlnplastic
47,076 0.0000 55,104 0.0000
48,892 0.0093 56,940 0.0090
85,563 0.0519 84,668 0.0414
109,893 0.2586 110,102 0.2586

4.2 Calibration Metrics


In order to directly compare the different modeling techniques discussed in

Section 3.1, the strength and deflection values from various models were compared.

The deflection results were obtained directly; a node in the center of the bottom of the

beam was chosen, as this location should provide the highest deflection result, and the

deflection at this node at the maximum loading is the value reported.

The load proportionality factor (LPF) was used to easily compare the results

obtained utilizing ABAQUS/Standard and ABAQUS/Explicit. This value is the

proportion of the loading at a given step time relative to the total load specified in the

input file of the model. In ABAQUS/Standard, this value is printed directly to the

39
output file, as described below. In ABAQUS/Explicit, this value is hand calculated

based on the step time and the total loading applied, as specified in the input file. This

is done by calculating a loading rate and multiplying this by the step time. The

loading rate is calculated based on the linear input defined in the “amplitude”

command, using the total load input divided by the time over which the load is

applied.

Different approaches were used to determine the maximum loading depending

on the analysis technique used. ABAQUS/Standard was straightforward; an output

file is created while an analysis is running which includes the LPF applied during each

step time. The output file containing the LPF values was analyzed to find the highest

LPF, which denotes the maximum loading of the model. In some versions of the

models, the LPF would reach a peak value before the beam would begin to unload. In

these cases, the first peak value was considered the maximum loading. Ideally, the

model would reach a peak LPF value, followed by a decrease in load. This would

indicate that the model did not experience any problems reaching convergence prior to

achieving its maximum capacity.

When utilizing ABAQUS/Explicit, an LPF is not directly printed. The loading

is defined using the “amplitude” command, as described in Section 3.4; the load can
then be calculated directly based on the step. The step is indicated in the results files

created by ABAQUS during analysis. In contrast to ABAQUS/Standard,

ABAQUS/Explicit does not reach a maximum LPF followed by a decrease, but rather

increases until reaching the maximum load input using the “amplitude” command or

the model terminates due to excessive distortion, which may be well past the point of

40
realistic behavior. To determine the maximum realistic load, the visual output (.odb)

file was analyzed as follows.

Two different methods were used to determine the loading at which the model

is classified as having failed. Initially, visual inspection was used to determine the

step time at the point where the beam displayed an abrupt and obvious non-linear

change in deflection and/or element distortion from one step to the next. An example

of this for the 3-D beam can be seen in Figure 4.2, where Figure 4.2 (a) shows the last

step before failure and Figure 4.2 (b) shows the step where failure occurs. Every

version of this model had similarly clear indications of the failure; although the

excessive deformation was not always located in the center, the distortion of elements

was always evident.

(a) (b)
Figure 4.2. Example of determination of (a) the last step before failure and (b) the first
step of failure in 3-D models utilizing ABAQUS/Explicit.

To determine whether this visual method for determining the failure of the

beam was too qualitative, a second method of assessing maximum load was

formulated. In this method, a row of elements along the bottom of the beam, halfway

between the center of the beam and the loading, was used. It was thought that this

location would provide a more accurate indication of if the beam was globally

41
unloading as the elements are not directly under the load nor experiencing the

maximum shear and moment, such that the results are not sensitive to localized

unloading as individual elements fail; in addition, the center of the beam was the most

common location of deformation in the model. The maximum principal stress at the

integration point of each of these elements was determined using the output file, and

this value for each element was added together for each step. The step where the

highest stresses were seen was then considered to be the failure step.

4.3 Strength and Deflection Calculations


The study that the finite element beam calibration model was based on

reported the tested strength of an uncorroded specimen by indicating the total load

which was applied at failure (Oyado et al. 2010). However, this value of 11,802.5 lbs

(52.5 kN) was thought to be high; therefore, strength calculations based on traditional

reinforced concrete design (Wight and MacGregor 2009) were computed. In order to

easily account for changes in material properties which were performed to model

corrosion, a spreadsheet was created, based on geometry and material properties of the

beam, to calculate the expected strength and deflection for all models. With these

strength and deflection values, the accuracy of each beam modeling technique could
be estimated by comparing the model results to these theoretical values. The

equations and an example of the calculations put into these spreadsheets can be seen in

Appendix A.

Using these calculations, the expected strength was calculated for both the

uncorroded and corroded models. For the uncorroded model, the strength was

calculated to be 5,392 lbs based on the loading configuration shown above in Figure

4.1. This corresponds to an LPF of 0.914 based on a total applied load of 5901.4 lbs

42
in the model, which is half of the experimental load. The deflection calculations were

performed based on both the uncracked and cracked sections, as is done in typical

design calculations for reinforced concrete members. These deflections were

calculated, based on the loading which is associated with the calculated strength, to be

0.0181 in for the uncracked section and 0.1245 in for the cracked section. For the

corroded model, these values varied depending on the material properties which were

defined. The exact values as the input parameters are varied are reported in the

discussion of the calibration process, and can be found in Section 4.4.3.

These theoretical values are later used for comparison with the results of the

models to estimate accuracy. It was thought that these values, based on similar

theoretical equations that ABAQUS utilizes when analyzing models, were a more

accurate representation of the model accuracy than the experimental values; there is a

large amount of deviation in concrete and rebar properties that are possible and for

which specific values are not reported for the experimental specimen. These

deviations could account for the higher failure load observed experimentally when

comparing to the theoretical strength of the concrete beam. The experimental strength

was more than twice what the theoretical calculations indicated. While some

discrepancy between the theoretical and experimental values are expected, as the
theoretical equations are relatively conservative when predicting the response of a

reinforced concrete beam, this large deviation suggests that the experimental specimen

displayed unusual characteristics which were not attempted to be simulated.

4.4 2-Dimensional Concrete Model


In attempts to keep computational effort and time to a minimum, 2-D concrete

beams were analyzed. In using this approach for modeling reinforced concrete,

43
previously created bridge models could be utilized with minimal modeling efforts,

which involved only incorporating post-cracking behavior into the decks. These

existing models used 2-D linear-elastic concrete elements for the deck and the “rebar”

command for the rebar. By modeling the reinforced concrete beam the same way, the

only changes to be made were those related to the material input variables. The

definition of the material property input values used in the uncorroded base model,

including the rebar properties applied to all models, the concrete damaged plasticity

input, and the smeared crack input applied when evaluating different modeling

techniques, are described in Section 4.4.1. A mesh sensitivity analysis was performed,

and the results of this, along with the modeling technique determined to most

accurately model uncorroded reinforced concrete, are described in Section 4.4.2.

Once the uncorroded modeling technique and input were determined, a calibration

process was performed to determine the optimal approach for modeling corrosion to

achieve the target changes in deflection and strength. This calibration process and

final corroded model input are described in Section 4.4.3.

4.4.1 Uncorroded Base Model Input Values


To compare how the different modeling approaches affect the accuracy of the
results, both the concrete damaged plasticity (CDP) and smeared crack (SC)

approaches were analyzed. These are both approaches which utilize

ABAQUS/Standard and therefore do not require interpretation for determining

maximum realistic loadings as well as directly simulating static loadings such that the

rate of load application is not applicable and does not need to be included. The

material properties which were used in the input commands for the uncorroded base

model are defined below in Sections 4.4.1.1 through 4.4.1.3.

44
4.4.1.1 Rebar
In order to accurately create the beam models, the values for the inputs

described in Chapter 3 needed to be determined. The “rebar” command which was

utilized involves defining the cover of the rebar and an equal rebar spacing. Due to

there being only 2 rows of rebar present, which are not constantly spaced between

each other and the edge (see Figure 4.1), it was assumed for this command that the

cover distance was equal to a horizontal rebar spacing of 1.2992 in (33 mm); that is,

rather than defining a 0.7874 in (20 mm) cover to the outer edge of the beam with a

2.3622 in (80 mm) spacing between the bars, a constant spacing of 1.2992 in (33 mm)

between the outer edges and between the bars was defined. It was thought that this

slight spacing change should have no significant effect on the strength results of the

beam, as theoretically the horizontal position of the rebar is not influential. The

vertical spacing was consistent with the spacing utilized in the experimental beam, as

described in Section 4.1. The area of the rebar was also used in the “rebar” command,

defining values of 0.0438 in2 (6 mm2) and 0.20563 in2 (13 mm2) for #2 (SD295) and

#4 (SD345) bars, respectively. These are the actual cross-sectional areas of the #2 and

#4 bars.

4.4.1.2 Concrete Damaged Plasticity


The elastic input values previously described in Section 3.1.1 were

used in combination with CDP inputs as one means to model the concrete constitutive

response. These elastic properties included Ec and ν values of 3,197,746 psi and 0.2,

respectively. In addition to these elastic values, the post-cracking values of concrete

needed to be determined for use in the “concrete damaged plasticity”, “concrete

tension stiffening”, “concrete compression hardening”, “concrete tension damage”,

45
and “concrete compression damage” commands, as described in Section 3.1.2.2. A

study was found which used experimental tests to calibrate the input for these CDP

commands based on actual reinforced concrete beams (Jankowiak and Lodygowski

2005). This study provided exact tabulated input for each command. Jankowiak and

Lodygowski’s input (2005) for the “concrete tension stiffening” and “concrete

compression hardening” commands was scaled based on the strength of the concrete

and used directly. This scaling involved taking the proportion of each stress value in

the input relative to the maximum stress value reported in the input for the

corresponding command and multiplying this by the experimental strength, f’c. The

“concrete tension damage” and “concrete compression damage” command input

values were used directly. These input values can be seen in Table 4.3. It should be

noted that, for the “concrete tension damage” and “concrete compression damage”

commands, multiple rows of 0.0 coefficients are required as the strain for these

commands must be the same as the corresponding tension stiffening and compression

hardening commands, respectively, even if no coefficient value is used. These

coefficients define post-cracking damage properties for concrete in tension and

compression by defining a damage variable and associated crushing strain to include

stiffness degradation as the cracking proliferates.

46
Table 4.3. Input values for post-cracking commands for CDP model.
Compression Hardening Compression Damage
σ (psi) ε Coefficient ε
944 0.0 0.0 0.0
1271 0.0000747 0.0 0.0000747
1881 0.0000988 0.0 0.0000988
2537 0.000154 0.0 0.000154
3147 0.000762 0.0 0.000762
2532 0.00256 0.195 0.00256
1274 0.00568 0.596 0.00568
331 0.0117 0.895 0.01173
Tension Stiffening Tension Damage
σ (psi) ε Coefficient ε
269 0.0 0.0 0.0
421 0.0000333 0.0 0.0000333
277 0.000160 0.406 0.000160
128 0.000280 0.696 0.000280
34 0.000685 0.920 0.000685
8 0.00109 0.980 0.00109

The final command which is used for the CDP model is the “concrete damaged

plasticity” command. The dilation angle which was used was 38, as determined by

Jankowiak and Lodygowski (2005). The flow potential eccentricity (ϵ) and ratio of

initial equibiaxial compressive yield stress to initial uniaxial compressive yield stress

(σb0/σc0) (Simulia 2011) were also determined by Jankowiak and Lodygowski (2005)
to be 1.0 and 1.12, respectively; these values were also used directly. The final input

values required is Kc, and a default value of 2/3 was used.

47
4.4.1.3 Smeared Crack
As with the CDP model, the elastic commands and corresponding input values

used are as described in Section 3.1.1. A description of each command and how it

related to the SC model is detailed in Section 3.1.2.1. The first command for post-

cracking behavior for the SC model is the “concrete” command. The strength of 3,147

psi, which is the experimental strength as determined by Oyado et al. (2010), was

used, along with an absolute value of plastic strain of 0.0. This point defines the yield

stress of the concrete. A second data point of 5,500 psi stress and 0.0015 absolute

plastic strain was used. This point defines the maximum stress of the concrete before

crushing and the associated plastic strain. The “tension stiffening” command specified

the relationship between the fraction of remaining stress to stress at initial cracking

and the cracking strain, which is the absolute value of direct strain minus direct strain

at initial cracking. Two data points were used to define this relationship. The first

data point was 1 and 0, where the remaining stress is all of the stress and cracking has

not yet initiated, thus the cracking strain is zero. The second data point was 0 and

0.0027, which specifies the strain at which the stress has been entirely relieved; this

strain value assumes a compressive strain limit of 0.003 and a strain at cracking of

0.0003. For the “failure ratios” command, all of the ABAQUS default values were
used in the absence of any reason for altering these defaults.

4.4.2 Mesh Sensitivity Analysis and Uncorroded Base Model Input


The creation of the 2-D geometry was relatively straightforward for the

concrete elements; the beam was broken up into equally sized square elements. The

exact size of these elements was varied to determine the most computationally

efficient and accurate size. A mesh sensitivity analysis was performed and the mesh

48
sizes which were tested were 3.937 in (100 mm), 1.9685 in (50mm), 0.7874 in (20

mm), 0.3937 in (10 mm), and 0.19685 in (5 mm) squares. For each of these mesh

sizes, the CDP and SC approaches were utilized. For these initial models, the

uncorroded base model values previously described in Section 4.4.1 were used.

Ideally, as the mesh sizes decrease, the results should asymptote to a constant value.

The load which was applied was 5901.4 lbs, half of the experimental load at failure as

reported by Oyado et al. (2010). The strength and deflection results of this sensitivity

analysis can be seen in Table 4.4 and Table 4.5, respectively. The percent difference

values are based on a strength LPF of 0.914 and deflection values of 0.0181 in for the

uncracked section and 0.1245 in for the cracked section, as determined from the hand

calculations described in Section 4.3, where positive changes indicate results larger

than the calculated values and negative changes indicate lesser values. Since these

models utilize ABAQUS/Standard, the loading would ideally reach a maximum peak

before decreasing. Should this not occur, convergence issues are the cause. These

results indicate that, with the exception of the 5 mm mesh, the values begin to become

relatively constant as the mesh sizes decrease. It was thought that the reasoning for

the 5 mm mesh results being significantly higher than the previous models with larger

meshes is that, when tracking cracks, the elements are smaller and therefore have a
smaller area effected by each crack, allowing the elements to reach higher strengths.

It should also be noted that, although the models did not converge, the results of the 10

mm, 20 mm, and 50 mm models utilizing the SC approach were relatively close to the

experimental load of an LPF of 1.83 (11,802 lbs).

49
Table 4.4. Strength results of mesh sensitivity analysis with percent differences based
on the calculated theoretical strength values.
Mesh Size Concrete Highest Load % Difference from
(mm) Model (LPF) Expected Load
CDP1 1.75 92%
100 1
SC 4.72 417%
CDP 2.55 179%
50 1
SC 3.02 231%
CDP1 0.89 -3%
25
SC 4.10 349%
CDP 1.10 20%
20
SC1 2.98 226%
CDP 1.08 18%
10
SC1 2.39 162%
CDP 3.69 304%
5
SC1 10.2 1017%
1
Model terminated at peak load, indicating inability to converge.

Table 4.5. Deflection results of mesh sensitivity analysis with percent differences
based on the calculated deflection values for the uncracked and cracked
sections.
Mesh Size Concrete Highest % Difference, % Difference,
(mm) Model Deflection (in) Uncracked Cracked
1
CDP 0.0533 196% -57%
100
SC1 0.3831 2022% 208%
CDP 0.2820 1462% 127%
50
SC1 1.4130 7728% 1035%
CDP1 0.0233 29% -81%
25
SC 0.2109 1068% 69%
CDP 0.0435 141% -65%
20
SC1 0.1361 654% 9%
CDP 0.0425 135% -66%
10
SC1 0.0993 450% -20%
CDP 0.0024 -87% -98%
5 1
SC 0.0069 -62% -94%
1
Model terminated at peak load, indicating inability to converge.

50
When only taking into consideration the models which converged, it was seen

that only 1 SC model and 4 CDP models converged. Of these CDP models, 2 had

relatively consistent results; 1.1 and 1.08. Due to the lack of information regarding

converged SC models, and the inconsistency of the available results, this technique

was not applied to the final uncorroded model input. Ultimately, the CDP approach

was chosen with a mesh size of 20 mm rather than 10 mm (of the two models that

converged and gave consistent values) to save on computational time.

After performing the mesh sensitivity analysis, the results of the analysis were

compared to the calculated strength and deflection values to determine accuracy. It

should be noted that the models include post-cracking behavior which is not included

in the corresponding strength and deflection equations. This is likely the explanation

for the increased strength values seen in Table 4.4. Thus, the 20% increase in strength

associated with the 20 mm CDP model was considered to be within reason. As for the

deflection, these values were consistently between the cracked and uncracked

expected values based on the theoretically expected strengths, with the exception of

the 50 mm and 5 mm mesh size. Although the values are not the same as the expected

values, it was thought that ABAQUS can take into consideration the post-cracking

behavior more accurately than the theoretical equations and therefore the 20 mm mesh
size was used as the final uncorroded model for comparison with the calibration of the

corroded model. That is, the values described above in Section 4.4.1.2 were used as

the uncorroded material input and the results, outlined in Tables 4.4 and 4.5, were

used as a baseline for comparison to determine if a 50% decrease in strength and 82%

increase in deflection, for reasons described in Section 2.1.2.2, were achieved during

the calibration process. The stress contours of this uncorroded model can be seen in

51
Figure 4.3, where darker colors indicate a higher stress and lighter colors indicate a

lower stress and the stress which is reported is the Mises stress at the integration point

of each element; subsequent stress data in 2-D elements is also reported at integration

points, which means that peak tensile and compressive stresses at the extreme fiber of

the elements is not considered. Similarly, the relative tensile and compressive stress

on the top and bottom cannot be discerned from one another visually in Figure 4.3.

The maximum and minimum principal stress values can be used to determine the

compressive and tensile stresses at the integration points, respectively. Furthermore,

the deformation of the beam cannot be visually seen in Figure 4.3. Thus, this

deflection was plotted versus load in Figure 4.4, where the FEA data includes the

loading response and the unloading response, which nearly traces the loading

response. To determine the validity of the model, this data is compared to the

deflection results calculated for the uncracked and cracked sections using the same

loading applied in the model, where the theoretical deflection results are limited to the

elastic regime, which is assumed to occur up to the predicted strength of the beams.

These results indicate that the finite element model follows the expected deflection

results relatively well. It should be noted, however, that these results were calculated

using the actual loading applied in the finite element model and therefore the resulting
deflections are not the same as those compared to in Table 4.5, where the load was

based on the loading associated with the theoretically expected strength and thus

indicated the model deflection results were between the uncracked and cracked

sections deflections. It can be seen in Figure 4.4 that this is not the case. A summary

of the input used in the final uncorroded model for future comparison during the

calibration process of determining the corroded model input can be found in Table 4.6.

52
Figure 4.3. Uncorroded 2-D base model stress contours (units of psi).

1.2

0.8
Load (LPF)

0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5
Deflection (in)

FEA Uncracked Cracked

Figure 4.4. Load and deflection comparison between final uncorroded model and
theoretical calculations for the uncracked and cracked sections.

53
Table 4.6. Input and results of final uncorroded 2-D beam model.
Dilation Angle 38 Es (ksi) 29000
ϵ 1 f'y (psi) 47000
σb0/σc0 1.12 fy (psi) 55000

Input
Kc 2/3 f'c (psi) 3147.3
A's (in2) 0.0438 f't (psi) 420.8
As (in2) 0.20563 Ec (psi) 3197746
E's (ksi) 29000
Strength (LPF) 1.1
Results

Deflection (in) 0.0435

4.4.3 Corroded Model


While calibrating the optimal method to simulate the effects of corrosion in the

2-D models, different input values and the use of numerous commands were varied.

These included the inclusion of the “concrete compression damage” and “concrete

tension damage” command, as well as variation of the dilation angle, flow potential

eccentricity (ϵ), σb0/σc0, Kc, compression rebar area (A’s), compression rebar strength

(f’y), compression rebar modulus of elasticity (E’s), tensile rebar area (As), tensile

rebar strength (fy), tensile rebar modulus of elasticity (Es), f’c, f’t, and Ec. The

inclusion of the “concrete compression damage” and “concrete tension damage”

commands, the dilation angle, ϵ, σb0/σc0, and Kc were varied as their effects on the

model were not well understood and the sensitivity of the models to these values

needed to be determined. Changing the A’s and As was used to reflect the decrease in

area which occurs during corrosion and varying f’y, fy, E’s, and Es was used as an

indirect means to produce the desired changes in strength and deflection. It was

thought that decreasing f’c, f’t, and Ec would be a valid approach for simulating the

54
decrease in strength associated with corrosion and the theoretical change in concrete

stiffness; although the theoretical change in stiffness occurs in the non-linear range,

modeling the change in this region was an indirect approach.

Initially, the parameters were varied individually to assess the effects of each

variable on strength and deflection of the beam. A discussion of this calibration can

be seen in Section 4.4.3.1. After this was performed, the model was then optimized

using different variables individually. The optimization of Ec, f’c, and f’t is described

in Section 4.4.3.2, the optimization of A’s and As is described in Section 4.4.3.4,

followed by the calibration of the dilation angle, described in Section 4.4.3.4. During

the optimization process, the results of the models were compared to the strength and

deflection results expected through the theoretical calculations described in Section

4.3, where the basis for this comparison is also described.

4.4.3.1 Individual Parameter Variation


Initially, the variations of parameters were performed with each variable

independently. One of the first variables tested was the removal of the commands

“concrete compression damage” and “concrete tension damage.” The values used in

these commands were taken directly from the literature (Jankowiak and Lodygowski
2005); consequently, their importance to the model was not known. It was found that

removing these commands did not affect the results of the uncorroded base model and

therefore subsequent models include these commands.

Next, the remaining variables were changed independently. A complete list of

all models tested, both the input in the model and the resulting strength and deflection

values, can be found in Appendix B. For brevity, only the models which affected the

strength and deflection in the targeted manner (i.e., helped towards achieving the 50%

55
decreased strength and 82% increase in ultimate deflection goals) are included; these

can be found in Table 4.7. Thus, the implicit goal in this work is that the generalized

effects of corrosion are intended to be represented rather than calibrating the model to

the specific performance of a given specimen given the ambiguities regarding the

exact properties of the relevant experimental data available in the literature. For the

sake of comparison, the base model results are also included. The input and output

values of the different corroded models are expressed as a percentage of the base

model values. For the deflection results, any value with increased deflection from the

base model is indicated in green and any value with decreased deflection is indicated

in red. For strength, any value with decreased strength from the base model is

indicated in green, and any model with increased or unchanged strength is indicated in

red. This is done to illustrate which input variables help achieve the goals for the

corroded model.

56
Table 4.7. Selected results for initial calibration of 2-D corroded models with values
expressed as a percentage of the base model.
Model # Base 1 2 3 4 5
Dilation Angle 38 100% 100% 105% 100% 100%
ϵ 1 100% 100% 100% 100% 100%
σb0/σc0 1.12 100% 100% 100% 100% 100%
Kc 2/3 100% 100% 100% 100% 100%
A's (in2) 0.0438 100% 75% 100% 100% 100%
As (in2) 0.20563 75% 75% 100% 100% 100%
Input

E's (ksi) 29000 100% 100% 100% 100% 100%


Es (ksi) 29000 100% 100% 100% 100% 90%
f'y (psi) 47000 100% 100% 100% 100% 100%
fy (psi) 55000 100% 100% 100% 100% 100%
f'c (psi) 3147.3 100% 100% 100% 100% 100%
f't (psi) 420.8 100% 100% 100% 100% 100%
Ec (psi) 3197746 100% 100% 100% 75% 100%
Strength (LPF) 1.10 1.06 1.01 1.10 1.06 1.09
Deflection (in) 0.0435 0.0451 0.0430 0.0452 0.0501 0.0441
Results

Change in
N/A -3.64% -8.18% 0.00% -3.64% -0.91%
Strength
Change in
N/A 3.84% -1.05% 4.03% 15.14% 1.40%
Deflection

It can be seen by the results in Table 4.7 that varying As, A’s, Ec, the dilation
angle, and Es have positive effects on the results. However, changing Es was

considered to be unrealistic and was therefore not considered in the final model. This

was not decided until the final calibration steps were being performed and therefore

most subsequent models will include a decrease in this value. Due to f’c, f’t, and Ec

being related through Equations 3.1 and 3.2, f’c and f’t were modified in conjunction

with the change in Ec for subsequent models according to this relationship.

57
Once the variables which positively affected the results were identified, they

were calibrated. The order of this calibration was based on how significantly they

affected the strength and deflection results of the model; variables which changed

these values more drastically were calibrated first. Thus, the values of Ec, f’c, and f’t

were varied to begin with, as their effect was the greatest and their values related. The

values of A’s and As were varied next, followed by the dilation angle. These

calibrations are described in Sections 4.4.3.2, 4.4.3.3, and 4.4.3.4, respectively.

4.4.3.2 Calibration of Ec, f’c, and f’t


The results of the calibration of Ec, f’c, and f’t can be seen in Table 4.8. In this

calibration process, Ec was the value which was changed and the respective f’c and f’t

values were calculated based on this. The input and output values of the different

corroded models are expressed as a percentage of the base model values. For the

deflection results, any value with increased deflection from the base model is indicated

in green and any value with decreased deflection is indicated in red. For strength, any

value with decreased strength from the base model is indicated in green, and any

model with increased or unchanged strength is indicated in red. This is done to

illustrate which input variables help achieve the goals for the corroded model. A
strength comparison between the model results and the theoretical results can be seen

in Figure 4.5. In this graph, the black line represents a 50% decrease in strength from

the base model, an LPF value of 0.55, which is the targeted strength value. A

deflection comparison between the model results and the theoretical results of can be

seen in Figure 4.6. In this graph, the black line represents the targeted 82% increase in

ultimate deflection from the base model, represented by a value of 0.0791 in.

58
Table 4.8. Input and results of calibration of Ec, f’c, and f’t for corroded 2-D beam
model.
Model # Base 6 7 8 9 10
Dilation Angle 38 100% 100% 100% 100% 100%
ϵ 1 100% 100% 100% 100% 100%
σb0/σc0 1.12 100% 100% 100% 100% 100%
Kc 2/3 100% 100% 100% 100% 100%
A's (in2) 0.0438 100% 100% 100% 100% 100%
As (in2) 0.20563 100% 100% 100% 100% 100%
Input

E's (ksi) 29000 100% 100% 100% 100% 100%


Es (ksi) 29000 100% 100% 100% 100% 100%
f'y (psi) 47000 100% 100% 100% 100% 100%
fy (psi) 55000 100% 100% 100% 100% 100%
f'c (psi) 3147.3 90% 81% 56% 49% 36%
f't (psi) 420.8 95% 90% 75% 70% 60%
Ec (psi) 3197746 95% 90% 75% 70% 60%
Strength (LPF) 1.10 1.06 1.02 0.852 0.798 0.689
Deflection (in) 0.0435 0.0435 0.0435 0.0412 0.0406 0.0394
Results

Change in
N/A -3.64% -7.27% -22.5% -27.5% -37.4%
Strength

Change in
N/A 0.04% 0.03% -5.32% -6.71% -9.26%
Deflection

59
1.20

1.00
Strength (LPF)
0.80

0.60

0.40

0.20

0.00
0 5 10 15 20 25 30 35 40
% Decrease in Ec

Model Theoretical
Figure 4.5. Comparison of strength between results from calibration models and
theoretical values caused by varying Ec.

0.09
0.08
0.07
Deflection (in)

0.06
0.05
0.04
0.03
0.02
0.01
0.00
0 5 10 15 20 25 30 35 40
% Decrease in Ec

Model Theoretical Uncracked


Figure 4.6. Comparison of deflection between results from calibration models and
theoretical values of the uncracked section caused by varying Ec.

60
Using a decrease in Ec of 40%, as seen in Model 10 of Table 4.8, was chosen

as the calibrated model because this produced the largest decrease in strength.

Although this produces a decrease in deflection, it was anticipated to counteract this

change by varying the remaining parameters. A decrease larger than 40% wasn’t

utilized as it was thought that this would be unrealistic as it caused too large a change

in f’c, as a decrease in Ec of 40% causes an associated decrease in f’c of 67%. This

40% decrease in Ec and the associated decreases of 67% and 40% in f’c and f’t,

respectively, were maintained for all subsequent calibration models.

4.4.3.3 Calibration of A’s and As


The next input parameters which were varied were A’s and As; these

parameters were varied concurrently to one another. For ease of comparison, the

strength results of the variation of As and A’s are presented in Figure 4.7, Figure 4.8,

and Figure 4.9 corresponding to an A’s of 100%, 90%, and 80% of the uncorroded

model value, respectively. The remaining input variables are consistent with those

used in Model 10, as shown in Table 4.8. Again, the black line represents a 50%

strength decrease from the uncorroded base model.

61
0.8

Strength (LPF) 0.7

0.6

0.5

0.4

0.3

0.2
25 35 45 55 65 75
% Decrease As
Model Theoretical
Figure 4.7. Strength comparison of results from models and theoretical values while
varying As and using a constant A’s of 100% of the base model value.

0.7
0.65
0.6
Strength (LPF)

0.55
0.5
0.45
0.4
0.35
0.3
45 50 55 60 65 70
% Decrease As
Model Theoretical
Figure 4.8. Strength comparison of results from models and theoretical values while
varying As and using a constant A’s of 90% of the base model value.

62
0.75

0.65
Strength (LPF)

0.55

0.45

0.35

0.25
48 53 58 63
% Decrease As
Model Theoretical
Figure 4.9. Strength comparison of results from models and theoretical values while
varying As and using a constant A’s of 80% of the base model value.

The model results were also evaluated in terms of the deflection values for

each model. For ease of comparison, the deflection results of the variation of As and

A’s are presented in Figure 4.10, Figure 4.11, and Figure 4.12 corresponding to an A’s

of 100%, 90%, and 80% of the base model value, respectively. The black line

represents an 82% increase in deflection from the base model. As with the strength
values, there is a convergence problem at approximately 63% decrease in As. It can

also be seen that, with all the variations of As and A’s, the deflection values were not

reaching the goal 82% increase.

63
0.083

0.073
Deflection (in)
0.063

0.053

0.043

0.033

0.023
40 45 50 55 60 65 70
% Decrease As
Model Theoretical
Figure 4.10. Deflection comparison of results from models and theoretical values
while varying As and using a constant A’s of 100% of the base model
value.

0.083

0.073
Deflection (in)

0.063

0.053

0.043

0.033

0.023
50 55 60 65
% Decrease As
Model Theoretical
Figure 4.11. Deflection comparison of results from models and theoretical values
while varying As and using a constant A’s of 90% of the base model
value.

64
0.083

0.073
Deflection (in)
0.063

0.053

0.043

0.033

0.023
50 55 60 65
% Decrease As
Model Theoretical
Figure 4.12. Deflection comparison of results from models and theoretical values
while varying As and using a constant A’s of 80% of the base model
value.

As can be seen in the results of Figures 4.7 through 4.12, there is a

convergence problem with the resulting values at approximately 63% decrease in As.

This convergence issue is centered on the 50% strength loss goal and the cause of this

is unknown. The stress-strain responses of the different models were analyzed in

hopes of determining this cause. This stress-strain response can be seen in Figure
4.13, where the A’s value is constant at 80% of the original value, the percentages

reported are the percent decrease of As from the original, and the stresses are the

values reported at the integration point. In addition, Figure 4.14 shows the deflection

response compared to the loading, where A’s is again constant at 80% of the original

value and the percentages reported are the percent decrease of As from the original

value. These results indicate that the stress-strain and deflection responses of all

models were consistent. Furthermore, it is observed from Figure 4.14 that the erratic

65
deflection results are caused by the erratic strengths rather than both parameters

varying unpredictably.

800
700
50%
600
60%
Stress (psi)

500
62%
400
62.25%
300
62.50%
200
62.75%
100
63%
0
0 10 20 30 40 50 65%
Strain

Figure 4.13. Stress-strain response of element in 2-D beam models reported as percent
decrease in As.

0.7
0.6
50%
0.5 60%
Load (LPF)

0.4 62%
0.3 62.25%
0.2 62.50%
62.75%
0.1
63%
0
0 0.01 0.02 0.03 0.04 0.05 0.06 65%
Deflection (in)

Figure 4.14. Deflection response of elements in 2-D beam models reported as percent
decrease in As.

66
When choosing which model input to proceed with, it was decided that a

model before the convergence problems would be best; prior to the erratic spikes in

values, the model results followed a relatively linear trend. This trend is expected, as

can be seen in the theoretical results. By choosing a value prior to the convergence

problems, it was hoped that the convergence issues would be avoided when applying

the input to the full-scale bridge tests. This limited the results to not reducing As by

more than 63%. When considering the strength and deflection results of the different

models, it was found that using an A’s decrease of 20% in combination with an As

decrease of 62.5% produced the strength and deflection values closest to the goal

values. The exact input and results of this model can be seen in Table 4.9.

Table 4.9. Input and results of 2-D beam model after calibrating input parameters.
Dilation Angle 38
2
A's (in ) 0.03504
2
As (in ) 0.07711
Input

f'c (psi) 1133


f't (psi) 252.4
Ec (psi) 1,918,648
Strength (LPF) 0.603
Results

Deflection (in) 0.0485


Change in Strength -45.2%
Change in Deflection 11.6%

4.4.3.4 Calibration of Dilation Angle


The final value which was optimized was that of the dilation angle. The exact

effects of varying this value were unknown; however, the initial variations done in

Section 4.4.3.1 indicated that changing the dilation angle increased the deflection.

Since this is the goal which was most difficult to achieve, it was thought that

67
calibrating this value may allow the model to achieve the strength and deflection

goals. The dilation angle was slowly increased to determine the optimum value. The

results of this variation can be seen in Table 4.10. The input and output values of the

different corroded models are expressed as a percentage of the base model values. For

the deflection results, any value with increased deflection from the base model is

indicated in green and any value with decreased deflection is indicated in red. For

strength, any value with decreased strength from the base model is indicated in green,

and any model with increased or unchanged strength is indicated in red. This is done

to illustrate which input variables help achieve the goals for the corroded model. It

can be seen that there is little variation in the strength and deflection results between

models until a dilation angle of 44 or above is used. The largest strength in

conjunction with deflection change were seen when using a dilation angle of 38, the

uncorroded model value; therefore, a dilation angle of 38 was chosen as the optimal

value for the corroded model.

68
Table 4.10. Results of optimization of the dilation angle after optimizing Ec, f’c, f’t,
As, and A’s for the corroded 2-D beam model.
Dilation
38 40 41 42 43 44 45
Angle
A's (in2) 80% 80% 80% 80% 80% 80% 80%
Input

As (in2) 37.5% 37.5% 37.5% 37.5% 37.5% 37.5% 37.5%


f'c (psi) 36% 36% 36% 36% 36% 36% 36%
f't (psi) 60% 60% 60% 60% 60% 60% 60%
Ec (psi) 60% 60% 60% 60% 60% 60% 60%
Strength
0.603 0.624 0.624 0.624 0.625 0.482 0.484
(LPF)
Deflection
0.0485 0.0481 0.0480 0.0480 0.0480 0.0328 0.0330
Results

(in)
Change in
-45.2% -43.3% -43.3% -43.3% -43.2% -56.2% -56.0%
Strength
Change in
11.6% 10.6% 10.5% 10.4% 10.4% -24.7% -24.1%
Deflection

The exact final input which were used for the corroded model were unchanged

by this final calibration and remain the same as the values previously shown in Table

4.9. Although the final strength and deflection values weren’t exactly the goal values

of 50% decrease in strength and 82% increase in ultimate deflection, it was decided

that the input in this final version was physically possible and well-achieved the

targeted strength with a 45.2% decrease, while also being accompanied by an increase

in deflection; many models tended to fail at a lower deflection in proportion to the

lower strength. It was determined that it would be near impossible to create input for a

model which would produce both the desired strength and deflection changes with the

2-D modeling technique used here without utilizing input values which were outside

physical possibilities. The stress contours of this model, along with the contours of

the uncorroded model for comparison, can be seen in Figure 4.15 (b) and (a),

69
respectively. To help determine the validity of the model, this data is compared to the

deflection results calculated for the uncracked and cracked sections using the same

loading applied in the model, as shown in Figure 4.16, where the FEA data includes

the loading response and the unloading response. These results were also compared to

those of the uncorroded model. It should be noted that the theoretical deflection

results are limited to the elastic regime, which is assumed to occur up to the predicted

strength of the beams; in addition, these results were calculated using the actual

loading applied in the finite element model and therefore the resulting deflections are

not the same as those compared to in Table 4.10, where the load was based on the

theoretically expected strength and thus indicated the model deflection results were

between the uncracked and cracked sections deflections. It can be seen in Figure 4.16

(which plots the loading and unloading responses of the specimens, although the

unloading responses trace the loading response) that this is not the case. The results of

the corroded comparison to theoretical indicate that the finite element model follows

the expected deflection results relatively well when compared to the uncracked

section, diverging more as the load increases. The results of the uncorroded

comparison to the corroded follow the expected trend, with the corroded model

reaching higher deflections for the same loading values.

70
(a)

(b)

Figure 4.15. Stress contours of (a) final uncorroded and (b) final corroded 2-D beam
models (units of in psi).

71
1.2

0.8
Load (LPF)

0.6

0.4

0.2

0
0 0.1 0.2 0.3 0.4 0.5
Deflection (in)
Uncorroded FEA Uncorroded Uncracked Uncorroded Cracked
Corroded FEA Corroded Uncracked Corroded Cracked

Figure 4.16. Load and deflection comparison between uncorroded and corroded finite
element model and theoretical calculations for the uncracked section.

4.5 3-Dimensional Concrete Model


In order to model corrosion with more input variables, 3-D concrete models

were analyzed. Initially, these models were created using 2-D rebar elements.

However, previous research offers techniques for using 3-D rebar elements (Val et al.

2009; Amleh and Ghosh 2006; Fang et al. 2006) to model corrosion which can more

completely simulate all of the effects of corrosion, whereas there are limited modeling

approaches for 2-D rebar elements (Coronelli and Gambarova 2004; Dekoster et al.

2003; Kallias and Rafiq 2010). The 2-D and 3-D rebar element modeling are

described in Sections 4.5.1 and 4.5.2, respectively.

72
4.5.1 2-Dimensional Rebar
Initially, 2-D elements were used to define the rebar. These elements were

then embedded within the concrete using the “embedded element” command, as

described in Section 3.2.1. The compressive and tensile rebar, as well as the stirrups,

were modeled as 2-D rod elements. This process is described in the following

Sections 4.5.1.1 through 4.5.1.3, where Section 4.5.1.1 describes input values

associated with the use of the brittle cracking technique, Section 4.5.1.2 details the

results of the mesh sensitivity analysis which was performed, and Section 4.5.1.3

outlines the attempted calibration of the corroded model input.

4.5.1.1 Brittle Cracking Base Model Input


Initially, the brittle cracking technique was used to model the non-linear

response of the concrete. This technique was chosen as it allowed for visually

tracking cracks within the output file as elements fail and are removed from the mesh.

However, if applied to 2-D models, the elements would not reach the required failure

criteria to be removed, as all of the elements are experiencing both tension and

compression (at opposing section points) and therefore cannot reach the defined

failure criteria at all material points. For this reason, this technique was not applied to

the previous 2-D models.

The elastic and geometric input values described in Section 4.1 were used as

the uncorroded base model for the brittle cracking technique. The nonlinear input

commands were previously described in Section 3.1.2.3. For the “brittle cracking”

command, the values were defined through tabular input. An f’t value of 420.8 psi

was used, based on applying Equations 3.1 and 3.2 to the elastic Ec value, with a value

of 0.0 direct cracking strain; a value of 0.0 stress was then used with a value of 0.0027

73
direct cracking strain using the same logic as previously explained in Section 4.1.1.

The “brittle shear” command was also defined through tabular input; a shear retention

factor of 1 was used with a crack opening strain of 0.0, and a shear retention factor of

0.0 used with a crack opening strain of 0.0027. The “brittle failure” direct cracking

failure strain was defined as 0.0027.

4.5.1.2 Mesh Sensitivity Analysis


A mesh sensitivity analysis was performed for 2-D rebar elements. This brittle

cracking technique using the input described in Section 4.5.1.1 was utilized, along

with the CDP and SC techniques, described in Sections 4.4.1.2 and 4.4.1.3,

respectively. Both the concrete and rebar element sizes were varied; the mesh sizes

for both the concrete and rebar elements ranged from 3.937 in (100 mm) to 0.19685 in

(5 mm) cubes and rods, respectively. Every permutation of the concrete and rebar

sizes was analyzed.

Table 4.11 shows representative results of the rebar sensitivity in this analysis

and Table 4.12 shows representative results of the concrete element sensitivity done in

this analysis. In these tables, the percent difference which is reported is compared to

the theoretical beam strength LPF of 0.914, with negative values being below this
strength and positive values being above. It should be noted that, for the brittle

cracking technique, results were only available in LPF increments of 0.05. It was

found that the 2-D rebar element size was negligible, both in results and in

computational time; the CDP and SC results remained relatively constant and the

brittle cracking varied between 0.35 and 0.45. Due to the limited availability of LPF

increments, it was thought that this difference was minimal. In addition, the CDP and

SC models failed to converge for all rebar sizes with the exception of 100 mm. For

74
these reasons 3.937 in (100 mm) rods for the rebar were chosen. The results of the

concrete elements began to converge to a constant value with the brittle cracking

approach; however, the CDP and SC models failed to converge on almost all of the

mesh sizes. For this reason, more emphasis was placed on the results of the brittle

cracking models. It was found that a 0.7874 in (20 mm) concrete mesh provided

consistent results and limited computational efforts; these (0.7874 in, 20 mm) mesh

results were also closest to the theoretical results while also providing the largest mesh

size to reduce computational effort.

Table 4.11. Strength results of rebar mesh sensitivity analysis for 3D concrete beam
with 2D rebar including percent differences based on the calculated
theoretical strength values.
Rebar Size (mm) Model Type Highest Load (LPF) % Difference
CDP1 0.334 -63.4%
1
5 SC 0.287 -68.6%
Brittle Cracking 0.450 -50.7%
1
CDP 0.341 -62.7%
10 SC1 0.287 -68.6%
Brittle Cracking 0.450 -50.7%
1
CDP 0.337 -63.1%
20 SC 0.230 -74.8%
Brittle Cracking 0.450 -50.7%
CDP1 0.339 -62.9%
1
50 SC 0.234 -74.4%
Brittle Cracking 0.400 -56.2%
CDP 0.333 -63.5%
100 SC 0.234 -74.4%
Brittle Cracking 0.350 -61.7%
1
Model terminated at peak load, indicating inability to converge.

75
Table 4.12. Strength results of concrete mesh sensitivity analysis for 3D concrete
beam with 2D rebar including percent differences based on the calculated
theoretical strength values.
Concrete Size (mm) Model Type Highest Load (LPF) % Difference
CDP1 0.321 -64.8%
1
5 SC 0.321 -64.8%
Brittle Cracking 0.400 -56.2%
1
CDP 0.333 -63.5%
10 SC 0.234 -74.4%
Brittle Cracking 0.350 -61.7%
CDP 0.354 -61.2%
20 SC1 0.395 -56.7%
Brittle Cracking 0.600 -34.3%
CDP1 0.323 -64.6%
1
25 SC 0.434 -52.5%
Brittle Cracking 0.450 -50.7%
1
CDP 0.650 -28.8%
1
50 SC 1.56 70.8%
Brittle Cracking 1.50 64.3%
1
CDP 1.25 36.9%
100 SC1 1.67 82.9%
Brittle Cracking 2.25 146.4%
1
Model terminated at peak load, indicating inability to converge.

A summary of this input and the results of the base model can be seen in Table

4.13. It should be noted that the expected strength calculated in Section 4.3 was an

LPF of 0.914; the base model was only 7% below this value. However, the expected

deflection is 0.2593 in. for the cracked section; the base model is 138% higher than

this value. Although this deflection value is higher than the expected, this higher

deflection is similar to that seen in the 2-D beam models described in Section 4.4.3;

therefore, these input values were still used for the base model. These results are the

ones which are used for comparison of the corroded model to determine 50% decrease

in strength and 82% increase in deflection.

76
Table 4.13. Base model input for 3-D concrete elements with 2-D rebar elements for
brittle cracking technique.
f't (psi) 420.8
Direct Cracking Strain 0.0027
Direct Cracking

Input
0.0027
Failure Strain
Crack Opening Strain 0.0027
A’s (in2) 0.0438
Results As (in2) 0.20563
Strength (LPF) 0.850

Deflection (in) 0.616

A factor not taken into consideration during this initial modeling using 2-D

rebar is the time associated with the loading command using the brittle cracking

technique. During the time this modeling was performed, it was not considered that

this would significantly affect the results. For this reason, a time of 1 second was used

for all of the models involving 2-D rebar. However, time was taken into consideration

when using 3-D rebar, and this results of this analysis are discussed in Section

4.5.2.2.2.

4.5.1.3 Corroded Model


Due to the brittle cracking technique having few input variables, modeling

corrosion was relatively straight forward. The only variables which were changed

were f’t, strain at cracking, failure strain, shear retention strain, A’s, and As. As with

the 2-D model, these values were varied independently, the effects were analyzed, and

the variations were combined to attempt to produce the targeted results. For brevity,

select results can be found in Table 4.14. It should be noted that in these models the

77
maximum strength is determined by visually analyzing the result files. These models

represent the results which most closely achieved the 50% strength decrease goal. For

the sake of comparison, the base model results are also included. The input and output

values of the different corroded models are expressed as a percentage of the base

model values. For the deflection results, any value with increased deflection from the

base model is indicated in green and any value with decreased deflection is indicated

in red. For strength, any value with decreased strength from the base model is

indicated in green, and any model with increased or unchanged strength is indicated in

red. This is done to illustrate which input variables help achieve the goals for the

corroded model. The results from all of the calibration models can be found in

Appendix C.

Table 4.14. Selected results of optimization of corrosion for 3-D beam model with 2-
D rebar.
Model # Base 1 2 3 4
f't (psi) 420.8 50% 25% 50% 100%
Direct Cracking
0.0027 50% 200% 50% 100%
Strain
Direct Cracking
0.0027 50% 200% 50% 100%
Input

Failure Strain
Crack Opening
0.0027 50% 200% 200% 100%
Strain
A's (in2) 0.04235 100% 100% 100% 50%
As (in2) 0.20574 100% 100% 100% 50%
Strength (LPF) 0.425 0.375 0.425 0.4 0.225
Results

Deflection (in) 0.616 0.34674 0.5629 0.41765 0.11086


Change in Strength N/A -11.8% 0.0% -5.9% -47.1%
Change in
N/A -43.7% -8.6% -32.2% -82.0%
Deflection

78
It can be seen from Table 4.10 that this particular modeling technique was not

producing the desired deflection increases, but rather was producing decreases in

deflection. Although close strengths were reached, as close as within 5% of a 50%

strength decrease from the base model, it was thought that using a different modeling

approach would produce the desired deflection changes. For this reason, this

modeling technique was not applied to full-scale bridge models and 3-D rebar

elements were analyzed, as described in Section 4.5.2.

4.5.2 3-Dimensional Rebar


After determining the use of 2-D rebar elements within 3-D concrete elements

was not producing the desired results, the model was refined using 3-D elements for

rebar. This allowed for consideration of different corrosion factors, including the

effects of the creation of corrosion by-products; these by-products effect the contact

properties between the rebar and the concrete. The creation and optimization of this

model is described in Sections 4.5.2.1 and 4.5.2.2.

4.5.2.1 Mesh Sensitivity Analysis


In the mesh of 3-D concrete with 3-D rebar elements, the cross-sectional area

of the rebar limited the available sizes of the mesh. The initial mesh size used

measured 0.3937 in x 0.5249 in x 0.3937 in (10 mm x 13.33 mm x 10 mm). This size

was chosen as it provided a concrete cross-section height decrease of 0.03%, no

difference in concrete cross-section width, and a rebar area decrease of 0.5%, which

were deemed negligible, while providing a uniform mesh size throughout the model

for both the rebar and concrete elements to simplify the modeling effort. Using this

mesh size placed the centroid of the tensile rebar at approximately the same location

79
as if the rebar within the specimen created by Oyado et al. (2010). The vertical

distance to the centroid was almost exactly the same as the physical specimen, located

at 0.78735 in (19.995 mm) rather than 0.7874 in (20 mm) from the outer edge. The

horizontal location of the centroid was 0.59055 in (15 mm) from the edge, compared

to the 0.7874 in (20 mm) in the physical specimen. These differences were both

considered negligible, especially when considering that the strength and deflection

characteristics of the beam are theoretically not influenced by the horizontal position

of the rebar. Also, when comparing to the 2-D beam models, whose horizontal rebar

spacing was standardized due to input limitations, these 3-D rebar locations were more

accurate when compared to the original specimen. Considering this mesh size was

smaller than those determined through previous sensitivity analyses, and due to

computational limitations when creating smaller mesh sizes, no other mesh sizes were

analyzed.

4.5.2.2 Uncorroded Base Models Using Brittle Cracking


The commands and base values for the different input described by Amleh and

Ghosh (2006) and Val et al. (2009) were used initially. These commands are “surface

interaction”, “surface behavior”, “friction”, and “contact pair”, as described in Section


3.1.2.3. For the “surface interaction” command, a pad thickness of 0 was chosen; this

quantity represents the thickness of an interfacial layer between the contact surfaces.

For the “friction” command, exponential decay was specified, µs was set to 1, µk was

set to 0.4, and dc was set to 0.45, all of which were suggested by Amleh and Ghost

(2006). When utilizing the “surface behavior” command, exponential pressure-

overclosure was specified, with c0 set to 0.01 inches because the surfaces should

constantly be in contact, and p0 set to 589 psi, as calculated by Equation 3.4. The

80
“contact pair” command specified the rebar as the slave surface and the concrete as the

master surface, as suggested by Amleh and Ghosh (2006). Initially, all 4 surfaces of

the rebar in contact with the concrete were used in the “contact pair” command;

however, errors were found involving the surface normals of these surfaces. It was

found that only the top and bottom surfaces of the rebar and concrete had surface

normals pointing in the correct directions and were therefore the only surfaces used; it

was unclear in the ABAQUS documentation how to redefine the surface normal

direction (Simulia 2011). It was thought that only using the top and bottom surfaces

would accurately model the rebar as the only loading applied to the model was in the

vertical direction, and presence of the top and bottom contact surface between the

rebar and concrete should induce frictional forces which would limit any horizontal

movement during analysis. During analysis, the horizontal displacement of the rebar

was determined to validate this theory. It was found that the horizontal displacement

was less than 0.004 inches, which is less than 10% of the vertical displacement of 0.05

inches, and thus considered to be negligible.

4.5.2.2.1 Smeared Crack and Concrete Damaged Plasticity Models


The uncorroded base model input values for the reinforced beam using the SC
and CDP techniques were described in Sections 4.4.1.3 and 4.4.1.2, respectively.

These were applied to the 3-D beam with 3-D rebar model. It was found that, when

using ABAQUS/Standard, errors regarding excessive distortion of elements during the

first loading increment resulted in the model terminating. When analyzing the

effected elements, it was found that they were all rebar elements. In order to

accommodate this problem, commands were added which linked the rebar nodes to

concrete nodes. This was done using the ABAQUS command “MPC”, which stands

81
for multi-point constraint. Initially, the MPC type slider was used. This command

keeps a node on a straight line defined by two other nodes, but allows the possibility

of moving along the line and allows the line to change length (Simulia 2011). The line

was defined using the concrete nodes on the elements which were in contact with the

rebar and the rebar node was confined to move along this line. In the model, the

concrete and rebar elements did not share nodes; however, the nodes were located at

coincident locations. The concrete nodes on either side of a given rebar node were the

2 used to define the line and the rebar node was limited to moving along this. This

was done for varying numbers of nodes per cross-section of rebar, ranging from 1

corner of the rebar to all 4. Eventually, it was determined that doing this to 2 opposite

corners provided sufficient restraint. One node on each rebar was simultaneously

linked to the concrete using the MPC type beam. This MPC type provides a rigid

beam between two nodes to constrain the displacement and rotation at the first node to

the displacement and rotation at the second node, corresponding to the presence of a

rigid beam between the two nodes (Simulia 2011). It was thought that doing this

would help prevent rotation of the rebar within the concrete. Again, the rebar node

was linked to the concrete node and constrained to move with the concrete. It was

thought that adding this extra constraint would prevent the rebar from rotating within
the beam. It was found, after running models, that the presence of this MPC beam

made no difference to the results, whereas using the MPC slider provided the required

constraint.

When running this first version of the model, warning messages were produced

stating that the plasticity/creep/connector friction algorithm did not converge. These

models also resulted in strength LPFs that were only 11% of the calculated strength.

82
Since no commands referencing plasticity or creep were included in the model, it was

determined that this warning was due to the presence of the “friction” command. For

this reason, in contrast to the previous approach of modeling the friction

exponentially, a constant friction value (µ) was also used in attempt to help with

convergence. The µ values which were used were 0.5, 1.0, and 1.5, as a typical value

of 1.0 for the interface between concrete and rebar was suggested (Amleh and Ghosh

2006). Select results from these models can be seen in Table 4.15; all of the models

which were tested, including both input and results, can be found in Appendix C.

These results show that the strengths of these models was less than 1/3 of the

calculated strength LPF of 0.914.

Table 4.15. Results of varying friction input for 3-D concrete beam models with 3-D
rebar elements.
µ 0.5 1.0 1.5 None
Concrete
CDP SC CDP SC CDP SC CDP SC
Model
Strength
0.139 0.276 0.103 0.350 0.104 0.328 0.084 0.247
(LPF)
Deflection
0.0097 0.0175 0.0071 0.0223 0.0071 0.0207 0.0068 0.0184
(in)

To help determine the cause of the low strengths of the beams, the stress-strain

data for the beam was analyzed. The element located at the center of the bottom of the

beam was chosen, as this location is expected to have the largest tensile stress values.

The mirror element on the top of the beam was also analyzed to compare compression

and tension data. The results of the stress-strain data was then compared to the input

data. The compression concrete comparison and tensile concrete comparison can be

seen in Figure 4.17 and Figure 4.18, respectively, where the stress values reported are

83
for the integration points of the elements. It should be noted that none of the models

reach the nonlinear input values. It was determined that using this approach would

not produce the desired strength results while still inputting material properties which

were realistic.

3500
3000
2500
Stress (psi)

2000
1500
1000
500
0
0.0000 0.0020 0.0040 0.0060 0.0080 0.0100 0.0120 0.0140
Strain

Friction 0.5 Friction 1.5 Friction 1.0 Input


Figure 4.17. Comparison of compression concrete output to input values for 3-D
beam with 3-D rebar.

84
450
400
350
300
Stress (psi)
250
200
150
100
50
0
0.0000 0.0002 0.0004 0.0006 0.0008 0.0010 0.0012
Strain

Friction 0.5 Friction 1.5 Friction 1.0 Input

Figure 4.18. Comparison of tensile concrete output to input values for 3-D beam with
3-D rebar.

4.5.2.2.2 Brittle Cracking Model


In contrast to the ABAQUS/Standard models, errors regarding excessive

distortion were not an issue when applying the brittle cracking technique; therefore, no

“MPC” commands were required. However, a few of the commands defining the

contact surface between the concrete and rebar, which were described in Section

4.5.2.2, have more input values when used in conjunction with ABAQUS/Explicit.

The first of these commands is the “surface behavior” command. For this, in addition

to defining c0 and p0, as described in Section 3.2.2, a kmax value is also defined. The

default value in ABAQUS is infinity, and this value was used (Simulia 2011). In

addition to the “surface behavior” command, the “contact pair” command also used

additional parameters. In this case, a penalty contact algorithm was used by including

the parameter for mechanical constraint and setting it to penalty, as suggested by Val

85
et al. (2009). By employing the penalty contact algorithm, ABAQUS/Explicit uses

pure master-slave weighting, which reduces computational time. There are no extra

input values which are associated with including this parameter.

The first factor which was analyzed for the brittle cracking model was a time

comparison for the loading. It was thought that the slower the load is applied, the

more static the analysis results would be and therefore more consistent with the

physical testing on which the models and targeted performances are based. For this

reason, 1 second, 10 seconds, 60 seconds, and 300 seconds were analyzed. The results

of this analysis are shown in Figure 4.19, shown as deflection versus loading, and

including the theoretical results for the uncracked and cracked sections. It can be seen

in these results that 60 seconds and 300 seconds failed prematurely; the cause for this

is unknown. In addition, all of the results follow the theoretical uncracked section

results relatively well, until becoming nonlinear. For these reasons, both 1 second and

10 seconds were used during subsequent analyses.

86
0.8
0.7
0.6

Load (LPF)
0.5
0.4
0.3
0.2
0.1
0
0 0.05 0.1 0.15
Deflection (in)
1 Second 10 Seconds 60 Seconds
300 Seconds Uncracked Cracked

Figure 4.19. Time comparison for 3-D concrete beam models with 3-D rebar using
brittle cracking.

As with the CDP and SC models, warning messages were produced regarding

the plasticity/creep/connector friction algorithm not converging. Again, friction was

input as a constant value to attempt to reduce computational complexity; however, in

this case it was held constant at 1.0 as this was suggested as a typical value for the

interface between the concrete and rebar (Amleh and Ghosh 2006). It was also

considered that the pressure-overclosure may be the cause of the warning message and

variations of the model were analyzed that did not include this parameter. The results

of these variations can be seen in Table 4.16 where Exp indicates an exponential

pressure-overclosure or friction relationship. These results indicate that the strength

values are again lower than those expected through calculations; the results varied

from being 18-73% below the theoretical strength LPF of 0.914. The model which

produced the highest strength results of an LPF of 0.750 is the same model indicated

in Figure 4.19 as the 1 second model. These changes did not have a positive effect on

87
the results. Specifically, removing the pressure-overclosure relationship significantly

decreased the resulting strength.

Table 4.16. Comparison of variations in friction and pressure-overclosure of 3-D


concrete beam models with 3-D rebar using brittle cracking.
Pressure-
Time Friction Strength
Overclosure
(sec) Used (LPF)
Used
Exp Exp 0.750
N/A Exp 0.250
1
Exp 1.0 0.650
N/A 1.0 0.300
Exp Exp 0.460
N/A Exp 0.290
10
Exp 1.0 0.570
N/A 1.0 0.290

In attempts to more quantitatively determine the failure load, a different

approach was used to determine the step at which the beam failed. It was hoped that

by quantitatively determining the failure load, the results may be higher than initially

determined. A row of elements along the bottom of the beam, halfway between the

center of the beam and the loading, was used, as described in Section 4.2. The

maximum principal stress at each of these elements was determined using the output

file, and this value for each element was added together for each step. The step where

the highest stresses were seen was then considered to be the failure step. Due to the

time required to perform this more detailed analysis, this was only done for 3 models.

The models this was done for and the results can be seen in Table 4.17. It should be

noted that this new technique of determining the load at failure produced significantly

88
lower strength values than the previous technique. Considering that the initial strength

values were lower than the calculated values, these results were not ideal.

Table 4.17. Comparison of strength values utilizing different analysis techniques of 3-


D concrete beams with 3-D rebar using brittle cracking.
Maximum
Pressure- Original
Time Friction Stress
Overclosure Strength
(sec) Used Strength
Used (LPF)
(LPF)
1 Exp Exp 0.750 0.350
10 Exp Exp 0.460 0.330
60 Exp Exp 0.441 0.364

After attempting many different modeling variations, it was determined that

modeling the beams using 3-D concrete elements would need further investigation and

any remaining analyses were outside the scope of this project. More complicated

modeling techniques could be analyzed, including utilizing user subroutines or other

approaches to modeling the bond between the concrete and rebar. Additional

comparison between similar models available in the literature and these models may

also reveal further insights.

4.6 Conclusions
A 2-D beam was calibrated to determine the input parameters required to

model an uncorroded and corroded concrete beam. This was done by first determining

the input parameters required by ABAQUS, then varying these parameters in attempts

to achieve a 50% decrease in strength and 82% increase in deflection. The results of

this calibration process indicated use of a model with a 20% decrease in the

compressive rebar area, 62.5% decrease in the tensile rebar area, 62% decrease in the

89
compressive strength of the concrete, and a 40% decrease in both the tensile strength

of the concrete and the modulus of elasticity of the concrete from the uncorroded base

model. The model utilizing this input produced a 45.2% decrease in strength and

11.6% increase in deflection. These were the input values to be applied to the full-

scale bridge models.

In addition to 2-D beam models, 3-D beam models were created. This was

done in hopes of creating a model which could take more reinforced concrete

parameters into consideration and ultimately allow for more accurate modeling of

corrosion. However, suitable input compatible with the techniques available in

ABAQUS were not able to be identified to reproduce the desired reinforced concrete

beam behavior. Difficulty was found when creating the base model, and calibrating

for corrosion was never completed. When employing the smeared crack and concrete

damaged plasticity approaches, the model terminated prematurely, with the concrete

elements not reaching a nonlinear response. When utilizing the brittle cracking

approach, a base model was created using 2-D rebar elements, but could not be

calibrated to simulate corrosion because, although the strength decrease close to 50%

was achieved, there was a corresponding deflection decrease rather than increase.

When using 3-D rebar, the strength results of the uncorroded model did not reach the
expected strength based on the experimental values or the theoretical values. The

cause of this was unknown and therefore this modeling approach was not utilized.

The 2-D beam models were straightforward to both create and calibrate. This

is because there were fewer variables to affect the results. The 2-D beam input is also

simple to apply to full-scale bridge model as these models were previously created

using 2-D deck elements. However, the material effects of corrosion within the 2-D

90
beams was not modeled as directly as possible; only the gross effects of corrosion in

terms of estimated strength and deflection changes are able to be considered rather

than more refined changes in the mechanics of the interaction between the concrete

and rebar. The final corroded model resulted in a 45.2% decrease in strength and

11.6% increase in deflection at the maximum loading when compared to the final

uncorroded model at its maximum loading. Thus, the strength goals were deemed to

be satisfied to a reasonable extent. Although the deflection increase was less than the

goal value, this difference in deflection was based on the maximum loading for each

model, while in the experimental results the corresponding change in deflection was

evaluated at the ultimate condition, which would theoretically produce a greater

discrepancy. As a simple means to evaluate this hypothesis, deflections were

compared at a common loading, using the maximum loading for the corroded beam;

the percent difference then becomes 123%, as the uncorroded beam had a deflection of

0.0213 in. and the corroded beam a deflection of 0.0477 in. Thus, the deflection

targets were also deemed to be generally satisfied.

The 3-D beam models could include more of the effects of corrosion, including

the bond deterioration at the concrete and rebar interface and the pressure caused by

the creation of corrosion by-products. However, this greater modeling complexity


resulted in the optimal method of modeling reinforced concrete using 3-D elements

not being determined. Nonetheless, it is thought that his approach has the potential to

produce more accurate results.

91
Chapter 5

BRIDGE MODELS

The techniques and input which were calibrated in Section 4.4 were applied to

full-scale bridge models. These bridge models, previously created and calibrated to

field test results in complementary research (Michaud 2011; McConnell et al. in

review; Radovic, personal communication), assume linear-elastic concrete deck

properties throughout the loading range. In the present work, the concrete material

properties were the only variations made in order to more accurately account for

potential degraded concrete deck conditions. Three bridges were utilized: Bridge 7R,

SR 1 over US 13, and SR 299 over SR 1. These are described in Sections 5.1, 5.1.2.2,

and 5.3, respectively.

5.1 7R
The first bridge which was analyzed was Bridge 7R. A 3-D model of this

bridge using ABAQUS was previously created (Ross 2007) and calibrated (Michaud

2011; McConnell et al. in review). This model was modified to include non-linear

behavior of the deck concrete, as described in Section 4.4, and was then used for

analysis. A description of the bridge geometry and location can be found in Section

5.1.1 and the results of the modeling are discussed in Section 5.1.2.

5.1.1 Bridge Information


Bridge 7R is a three span steel girder bridge with a composite concrete deck;

each of the three spans are simply supported. The main span, which was the subject

span, is 105’-3 9/16” in length with 2 approach spans of similar length. The bridge

was designed in 1961 and built soon after and consists of a four girder cross-section.

92
The plate girders have 60” x 3/8” web plates and the exterior girders have a 20” x 1”

top flange and a 20” x 1 1/4” bottom flange. The interior girders have an 18” x 7/8”

top flange, 20” x 1” bottom flange, and cover plates of 1 1/2" which were attached to

the bottom flanges of the interior girders from 32’-9” on either side of centerline; the

exterior girders use 1 7/8” plates from 32’-3” on each side of centerline (Ross 2007).

The reinforced concrete deck has a total thickness of 8”. Shear connectors are

located on the girders in order to create a composite section and transverse and

longitudinal stiffeners are also present. The bridge girders and abutments are on a 63°

skew from tangent to the supports. Cross frames are located at intermediate locations

along the length of the girders. The “K” shaped cross frames are composed of 4” x 3

1/2” x 3/8” double angles. Additional built-up I-shapes serve as end diaphragms to

connect the girders together at each end of the structure. Concrete parapets with steel

handrails and sidewalks are located along both sides of the bridge. Two concrete piers

provide vertical support (Ross 2007).

Bridge 7R served as an exit ramp for Interstate 295 North through Delaware,

just south of the Delaware Memorial Bridge. Figure 5.1 shows the location of the

bridge. When the bridge was in service, the traffic exiting the interstate continued

onto Route 13 south. The bridge carried one lane of traffic with no existing shoulders;
however, 18” concrete parapets and 30” sidewalks were found on either side of the

travel lane. Baylor Boulevard passed under the bridge and received a minimal amount

of traffic daily. The vertical clearance below the bridge during its service life was 14’-

6”.

93
Figure 5.1. Map of location of Bridge 7R (Ross 2007).

5.1.2 Results
The results of the finite element models were analyzed. Initially, this was done

by reviewing the results files to determine the maximum loading and the general

qualitative location of any stress concentrations within the deck. Then, in order to

express the results in terms meaningful for bridge design and rating applications,

distribution factors (DF) were calculated from the girder stress results. The strength

and stress results of the model are discussed in Section 5.1.2.1 and the DF values are

discussed in Section 5.1.2.2.

5.1.2.1 Finite Element Modeling


In the previously created model, the reinforced concrete deck already utilized

2-D elements and the “rebar” command. This model can be seen in Figure 5.2. The

94
uncorroded (Tables 4.3 and 4.6) and corroded (Table 4.9) non-linear input which was

calibrated, as described in Sections 4.4.2 and 4.4.3, was then applied directly to the

bridge deck. As a means of comparison, the bridge model was analyzed in three

conditions: without any non-linear concrete or plastic rebar commands, with

uncorroded non-linear concrete and plastic rebar commands, and with corroded non-

linear concrete and plastic rebar commands.

(a) (b) (c)

Figure 5.2. Finite element model of bridge 7R viewed from (a) the top, (b) the bottom,
and (c) isoparametrically.

The results of this initial analysis can be seen in Table 5.1. The general trend is

as expected; the elastic model with no non-linear concrete and rebar commands has

the highest strength, while the uncorroded non-linear model has the next highest

strength and the corroded non-linear model has the lowest strength. However, it

should be noted that the elastic model is the only model which achieved a peak load

before decreasing, whereas both the uncorroded and corroded models terminated at the

peak load due to inability to converge.

95
Table 5.1. Original capacity results for Bridge 7R.
Maximum Load Equivalent Number of
Model
(kips) Trucks
Elastic 2178.4 30.3
1
Uncorroded 1402.4 19.5
1
Corroded 1084.5 15.1
1
Model terminated at peak load, indicating inability to converge.

In hopes of determining the cause of the convergence problems, the results

files were analyzed visually; the 3 model versions were then compared. These results

can be seen in Figure 5.3, showing (a) the elastic model with no non-linear concrete

and rebar commands, (b) the uncorroded non-linear model, and (c) the corroded non-

linear model. In this figure, the lighter colors represent tension and the darker colors

represent compression. It can be seen that the deck changes from tension to

compression over one of the girders in all 3 models. The exact cause of this is

unknown, but it is thought that this may contribute to the inability to converge which

was experienced.

96
(a)

(b)

(c)

Figure 5.3. Results of bridge 7R at maximum loading for the (a) elastic version with
no non-linear concrete or rebar commands, (b) the uncorroded version,
and (c) the corroded version.

97
In addition, the stress-strain response of a deck element was analyzed for each

of the bridge models and compared to determine if the bridge was reaching the non-

linear stress range. The deck element which was chosen was located in a stress

concentration which was common between all the different models. The location of

this element in the elastic, uncorroded, and corroded deck is shown in Figure 5.4 (a),

(b), and (c), respectively. In this figure, the red square indicates the element which

was analyzed. The results of this stress-strain response analysis, along with those of

subsequent models, are described in detail later in the section.

98
(a)

(b)

(c)
Figure 5.4. Representative example of location of deck element used for stress
distribution analysis in (a) the elastic, (b) uncorroded and (c) corroded
deck.

99
In hopes of achieving convergence, the input for the CDP approach being

utilized was further researched. It was found that the model is sensitive to the tension

stiffening value; a small value for tension stiffening gives early numerical problems

and therefore could be causing the inability to converge (Baskar et al. 2002). It is also

suggested that the value for tension stiffening be calibrated separately for each

individual problem (Baskar et al. 2002). For this reason, different values of tension

stiffening were used. The tension stiffening value models the load transfer across

cracks through the rebar by defining the stress-strain response after cracking.

Originally, the tension stiffening input was defined as a curved line. However,

Baskar et al. (2002) suggested defining only 2 points; the maximum strength of the

concrete and an associated direct cracking strain of 0, followed by 0 stress and the

tension stiffening value. Baskar et al. (2002) also suggests using a value of 0.1. The

values of tension stiffening which were analyzed for the present bridge models were

0.1, 0.01, and 0.035, as seen in Figure 5.7, along with the original tension stiffening

input. It was thought that a value larger than 0.1 would be too unrealistic; the tension

stiffening value is typically around 10 times the failure tensile strain of the plain

concrete, which equates to 0.027 for the input assumed herein (Basker et al. 2002).

100
800

Direct Stress After Cracking (psi)


700
600
500
400
300
200
100
0
0 0.02 0.04 0.06 0.08 0.1
Direct Cracking Strain

Original 0.01 0.035 0.1


Figure 5.5. Tension stiffening values for uncorroded model.

In addition to modifying the tension stiffening value, the tension damage

coefficient commands were removed. As stated previously, the strain values defined

in the “concrete compression hardening” and “concrete tension stiffening” commands

must correspond to the strain input in the “concrete compression damage” and

“concrete tension damage” commands, respectively. The damage coefficients

previously used were taken from literature where they were calibrated to the original

strain values in the tension stiffening input. The relationship is not well enough

understood to appropriately modify these commands for the new tension stiffening

input. It was also previously shown in Section 4.4.3.1 that including or removing

these commands does not affect the strength results.

101
These tension stiffening values were applied to the bridge models and the results

of these models can be seen in Table 5.2, along with the previous results utilizing the

original tension stiffening input for comparison. Although the results of these models

produced higher strengths than with the previous tension stiffening values, all of the

corroded models reached a higher strength than their uncorroded equivalents. It

should be also noted that, for each of these models with modified tension stiffening

values, the model reached a peak load before decreasing. This indicates that

convergence is no longer an issue. Lastly, the results between the different tension

stiffening values are relatively close to one another, also indicating convergence.

Table 5.2. Results of bridge 7R when varying tension stiffening values.


Tension Maximum Equivalent
Model
Stiffening Loads (kips) Number of Trucks
N/A Elastic 2178.4 30.3
1
Uncorroded 1402.4 19.5
Original 1
Corroded 1084.5 15.1
Uncorroded 1795.0 24.9
0.1
Corroded 1813.7 25.2
Uncorroded 1795.0 24.9
0.035
Corroded 2047.5 28.4
Uncorroded 1888.5 26.2
0.01
Corroded 2047.5 28.4
1
Model terminated at peak load, indicating inability to converge.

As stated previously, the stress-strain response of a deck element located in a

stress concentration was analyzed. This was done for the elastic model, as well as the

original tension stiffening values and a tension stiffening value of 0.035. The stress-

strain response using the original tension stiffening and a value of 0.035 can be seen in

Figures 5.6 and 5.7, respectively, where the elastic input stress-strain response is also

102
shown for reference. In these graphs, the principal stress and strain values are

presented. It should be noted that all of these values are tensile stresses. These results

indicate that the stresses within the element are consistent between the uncorroded and

elastic models, and the effects of corrosion can be seen in the offset of the stress

response of the corroded model. However, contrary to the understanding of the

nonlinear input commands, the elements exhibit tensile stresses up to 7 times higher

than the input maximum tensile stress. It was expected that the element stress would

reach the maximum specified value (756.7 psi for the uncorroded and 302.7 psi for the

corroded), then the nonlinear response would enforce a decrease in stress proportional

to the calculated strain. The cause of this discrepancy between the input and the

results is unknown. However, it was determined that the stresses in the deck were not

the cause of the model to reach its maximum load and begin decreasing.

7000
6000
5000
Stress (psi)

4000
3000
2000
1000
0
0 20 40 60 80 100 120
Strain (x105)

Elastic Uncorroded Corroded


Figure 5.6. Stress-strain response of representative deck element of Bridge 7R at a
stress concentration location using the original tension stiffening input.

103
7000

6000

Stress (psi) 5000

4000

3000

2000

1000

0
0 50 100 150 200 250
Strain (x105)

Elastic Uncorroded Corroded


Figure 5.7. Stress-strain response of element in the deck of Bridge 7R at a stress
concentration location using a tension stiffening value of 0.035.

After this stress-strain response analysis was performed, the cause of the

bridge to begin unloading was investigated. The cause of this unloading appears to be

yielding of the web elements to which the cross-frame members are connected. These

cross-frame elements are attached to a node on the web of the girders, which connect

to four surrounding web elements. The cross-frames have no plastic commands

associated with the elements; however, as the cross-frames transfer loading to the

surrounding web elements, they begin to yield and lose the ability to redistribute load,

simulating a realistic response. For this reason, the forces within the cross-frames at

the maximum loading of each model was investigated. These forces can be seen in

Table 5.3. In each of these cases, with the exception of the models with the original

tension stiffening values which did not converge, the force in the cross-frames well

exceeds the yield stress.

104
Table 5.3. Peak stress values in cross-frames as maximum loading in different 7R
bridge models.
Tension Stiffening Model Stress (psi)
N/A Elastic 62889
1
Original Uncorroded 36000
1
Corroded 34551
Uncorroded 43127
0.1
Corroded 73095
Uncorroded 53599
0.035
Corroded 88669
Uncorroded 59906
0.01
Corroded 88778
1
Model terminated at peak load, indicating inability to converge.

5.1.2.2 Distribution Factors


Another factor which was analyzed was the live load distribution factors.

These factors take into consideration how the live load is distributed, or the system

effect of bridges. Many different equations exist which estimate the DF using bridge-

specific geometry and material properties, including, but not limited to: span length,

modulus of elasticity, and girder spacing. Previously, the DF for Bridge 7R was

calculated using the AASHTO equations for one lane traffic then applying the skew

reduction factors (McConnell et al. in review). These calculations were also


performed using the field test data recorded when the bridge was loaded until failure,

as described in Section 2.1.1. The girders are numbered G1 through G4; this

numbering can be associated with Figure 5.2 (b), where the girders are numbered 1

through 4 from left to right. The results of the previously calculated DF values can be

seen in Table 5.4.

105
Table 5.4. Distribution factors of bridge 7R previously determined in research
(McConnell et al. in review).
Distribution Factor
Analysis Method G1 G2 G3 G4
AASHTO (Elastic loading) 0.480 0.371 0.371 0.480
In-Service Field Test (Elastic loading) 0.334 0.368 0.330 0.289
Theoretical Inelastic Distribution Factor 0.250 0.250 0.250 0.250

For comparison, the DFs of the finite element models were calculated. This was

done by averaging the Mises stress in the elements along the bottom flange at the

centerline of each girder at the time of maximum loading. It was previously shown

that this technique and location are valid approaches for determining DF (Sparacino,

draft internal report, 2013). The results of these calculations can be seen in the

following Table 5.5. These results indicate that, as the tension stiffening value is

increased, the DF values approach those of the theoretical inelastic values. The

models with lower tension stiffening values have DF values closer to those which

were experienced during the field test. It can also be seen that, in general, the

corrosion in the deck caused a more uniform distribution between the girders than with

an uncorroded deck, which is the opposite of the predicted results. This indicates that

a more thorough analysis should be performed regarding the best technique for

calculating the DF values from the model.

106
Table 5.5. DF values for finite element models created of Bridge 7R.
Model Distribution Factor
Tension Stiffening Type G1 G2 G3 G4
N/A Elastic 0.257 0.267 0.252 0.224
Uncorroded1 0.265 0.321 0.224 0.190
Original 1
Corroded 0.260 0.314 0.223 0.203
Uncorroded 0.268 0.270 0.256 0.205
0.01
Corroded 0.260 0.262 0.252 0.226
Uncorroded 0.274 0.277 0.251 0.198
0.035
Corroded 0.260 0.262 0.252 0.226
Uncorroded 0.287 0.215 0.278 0.221
0.1
Corroded 0.272 0.274 0.253 0.201
1
Model terminated at peak load, indicating inability to converge.

It should be noted that the DF results shown in Table 5.5 do not exactly follow

the expected trend. It was thought that, if corrosion decreases the load that can be

distributed through the deck, the elastic model should have the lowest DF values for

G1 and G2, followed by the uncorroded and then the corroded. However, it can be

seen that this is not the case. The results of the corroded model resulted in DF values

between those of the elastic and those of the uncorroded models. The cause of this

was analyzed by visually reviewing the .odb model result files. Here, it was seen that

the deck of the corroded model has more widespread inelastic response at the same

loading as the corresponding uncorroded model; the uncorroded and corroded

responses are shown in Figure 5.8 (a) and (b), respectively. In this figure, the dark

grey indicates any elements with a stress higher than the input f’t value. This

nonlinear response provided greater load sharing between the girders than expected in

the corroded condition, and caused the DF values to be lower than the uncorroded

values. This greater load sharing is also evidenced when comparing the amount of

cross-frame yielding between the models; the uncorroded model shows more yielding

107
than the elastic, and the corroded model displays more yielding than the uncorroded.

In addition, when analyzing the yielding in the girder visually, the uncorroded girder

exhibited slightly more yielding than the uncorroded, which was almost exactly the

same as the elastic model at a comparable loading.

(a)

(b)
Figure 5.8. Comparison of nonlinear response of concrete for bridge 7R for the (a)
uncorroded and (b) corroded deck.

Thus, from this combined evaluation, it is concluded that it is logical for the

corroded deck to result in a lower distribution factor than the uncorroded model.

108
However, additional evaluation of the exact cause of termination and corresponding

stress distribution in each model is needed to provide a complete understanding of the

relative differences in DF values observed between the elastic, uncorroded, and

corroded models. Specifically, a plot of DF for each girder versus load throughout the

loading range may be a beneficial first step to provide insight on this topic.

5.2 SR 1 over US 13
The next bridge which was analyzed was SR 1 over US 13, subsequently

referred to as US13. A 3-D model of this bridge using ABAQUS was created and

calibrated by Ambrose (2012) and Radovic (personal communication 2013). This

model was modified to include non-linear behavior of the deck concrete and was used

for analysis in the same manner as Bridge 7R. A description of the bridge geometry

and location is located in Section 5.2.1 and the results are described in Section 5.2.2.

5.2.1 Bridge Information


SR 1 over US 13 is a 65 degree skew steel I-girder bridge on Delaware State

Route 1. Twin spans carry the north- and south-bound lanes. The field-tested bridge

used for model validation by Radovic (personal communication 2013) carries the

southbound lanes of State Route 1 over US 13 approximately 5 miles south of the

Chesapeake and Delaware Canal in Delaware, immediately south of Road 423 and just

north of Boyd’s Corner, Delaware. Figure 5.9 indicates the location of this bridge. It

consists of two continuous spans of equal (165’) lengths. There are five girders spaced

9’-6” on center with exterior girders spaced 2’-10” and 3’-10” away from the outer

edge of the bridge parapets on the west and east sides, respectively; therefore, the total

width of the bridge is 44’-8”, carrying two 12’ lanes, a 12’ shoulder on the west side,

109
and a 6’ shoulder on the east side, while also having parapets 1’-4” in width on each

side of the bridge (Ambrose 2012).

Figure 5.9. Satellite View of SR 1 over US 13 (Ambrose 2012).

X-type cross-frames are used to laterally brace girders of the bridge and are

spaced 20’ on center with the exception of the first cross-frame from the end and the

first cross-frame from the support which are spaced at 22’-6” on center. The cross-

frames consist of two 3 1/2” x 3 1/2" x 3/8” steel angles that comprise the inclined
members of the cross-frame and a 4” x 4” x 1/2” steel angle serves as the bottom

chord. The two inclined members are bolted at their intersection by a 1/2” x 6” x 1’-1”

fill plate. All of the angles are bolted to the girders with Type 1, 7/8” diameter A325

high strength mechanically galvanized friction bolts via a 1/2” x 10” connection plate

fillet welded along the full height of the web. All structural steel is AASHTO M270

Grade 50 (specified minimum yield strength of 50,000 psi) painted with a urethane

paint. The steel girder is composite with the bridge deck (Ambrose 2012).

110
5.2.2 Modeling Results
As with bridge 7R, the deck was already created using 2-D elements and

utilizing the “rebar” command. This model can be seen in Figure 5.10. Again, the

uncorroded (Tables 4.3 and 4.6) and corroded (Table 4.9) non-linear input which was

calibrated, as described in Sections 4.4.2 and 4.4.3, was then applied directly to the

bridge deck. As a means of comparison, the bridge model was analyzed in three

conditions: without any non-linear concrete or plastic rebar commands, with

uncorroded non-linear concrete and plastic rebar commands, and with corroded non-

linear concrete and plastic rebar commands. The results of this initial analysis can be

seen in Table 5.6. It can be seen in these results that the models which include non-

linear concrete and plastic rebar commands terminated prematurely. The results files

were analyzed visually in hopes of determining the cause of this early termination.

The results files indicated that inability to converge was again the cause of

termination.

(a)

(b)

(c)
Figure 5.10. Finite element model of bridge US13 viewed from (a) the top, (b) the
bottom, and (c) in cross-section.

111
Table 5.6. Initial results from analyzing bridge US13 including non-linear concrete
and rebar commands.
Equivalent Number of
Model Maximum Load (kips)
Trucks
1
Elastic 3461.5 48.1
1
Uncorroded 41.2 0.6
1
Corroded 30.0 0.4
1
Model terminated at peak load, indicating inability to converge.

In hopes of determining the cause of the convergence problems, the results

files were analyzed visually; the 3 model versions were then compared. These results

can be seen in Figure 5.11, showing (a) the elastic model with no non-linear concrete

and rebar commands, (b) the uncorroded non-linear model, and (c) the corroded non-

linear model. In this figure, the lighter colors represent tension and the darker colors

represent compression. Due to the maximum load not reaching a high enough value,

very little information can be obtained from these results. However, when more

closely analyzing these models, two different elements were isolated as causing

convergence problems. The stress-strain results for these elements can be seen in

Figures 5.12 and 5.13 for compression and tension, respectively. It can be noted in the

tensile results that the stress-strain response does not follow the response which was

input. The cause of this is unknown, but it is thought that this difference may

contribute to the cause of termination.

112
(a)

(b)

(c)

Figure 5.11. Results of bridge US13 for (a) the elastic version at s similar loading as
the maximum for the uncorroded/corroded model, (b) the uncorroded
version at the maximum loading, and (c) the corroded version at the
maximum loading.

113
6000
5000
Stress (psi) 4000
3000
2000
1000
0
0 200 400 600 800 1000 1200 1400
Strain (x105)

Element 1 Element 2 Input


Figure 5.12. Compressive stress-strain response of elements in deck of US13 causing
inability to converge.

700
600
500
Stress (psi)

400
300
200
100
0
0 20 40 60 80 100 120
Strain (x105)

Element 1 Element 2 Input


Figure 5.13. Tensile stress-strain response of elements in deck of US13 causing
inability to converge.

114
As with bridge 7R, and as described in Section 5.1.2.1, the values of tension

stiffening were modified. The values which were analyzed were 0.1, 0.01, and 0.035.

Again, the tension damage coefficient commands were removed. The results of these

models can be seen in Table 5.7. Although the results of these models produced

higher strengths than with the previous tension stiffening values, as with original

tension stiffening input, these models terminated at the peak load; this again indicated

convergence being the cause of termination. The values in Table 5.7 are the

maximum loading applied which was at the terminating step of the model. For this

reason, these results were not considered to be an accurate representation of the

response of a corroded concrete deck, nor can relative strengths of the differing

models be reliably predicted from these results.

Table 5.7. Results of Bridge US13 when varying tension stiffening values.
Tension Maximum Equivalent
Stiffening Model Loads (kips) Number of Trucks
1
Uncorroded 2049.2 28.5
0.1 1
Corroded 924.9 12.8
1
Uncorroded 1340.3 18.6
0.035 1
Corroded 947.1 13.2
1
Uncorroded 1301.5 18.1
0.01 1
Corroded 786.5 10.9
1
Model terminated at peak load, indicating inability to converge.

5.3 SR 299 over SR 1


The final bridge which was analyzed was SR 299 over SR 1, subsequently

referred to as SR299. A 3-D model of this bridge using ABAQUS was created and

validated by Ambrose (2012) and Radovic (personal communication 2013). This

model was modified to include non-linear behavior of the concrete deck in the same

115
manner as performed for the models discussed above. A description of the bridge

geometry and location is located in Section 5.3.1 and the results are described in

Section 5.3.2.

5.3.1 Bridge Information


SR 299 over SR 1 is a 32º skew (measured from tangent to the supports) steel

I-girder bridge on Delaware State Route 299 over Delaware State Route 1. It is located

in the Middletown-Odessa area of Delaware, approximately 9 miles south of the

Chesapeake and Delaware Canal in Delaware. The location of Bridge SR299 is shown

in Figure 5.14. It consists of two continuous spans, of 128’ and 134’. There are eleven

girders in the cross-section, spaced 9’-1” with exterior girders spaced 2’-11” away

from the outer edge of the bridge parapets; therefore, the total width of the bridge is

95’-11”, carrying four 12’ lanes of traffic, two 12’ outside shoulders, a 22’ median and

turning lane which varies position along the length of the bridge, and two 1’-4”

parapets (Ambrose 2012).

116
Figure 5.14. Satellite view of SR 299 over SR 1 (Ambrose 2012).

K-type cross-frames are used to laterally brace the girders of the bridge and are

spaced 18’-3” on center on the west span and 19’-6” on the east span, with the

exception of the first cross-frame from each support, where the spacing varies. The

typical cross-frames consist of two 3 1/2” x 3 1/2” x 3/8” steel angles that comprise

the inclined members of the cross-frame and one 4” x 4” x 1/2” steel angle that serves

as the bottom chord. The two steel angles of the inclined members are welded with a

5/16” fillet weld on both sides to a 1/2” gusset plate, which is also connected by a

5/16” fillet weld on both sides to the midspan of the bottom chord. All fillet welds are

at least 4” in length. All of the angles are connected with 5/16” fillet welds to 1/2”

gusset plates that are connected to the 1/2” x 7” connection plate fillet welded to the

girders along the full height of the web. All structural steel is AASHTO M270 Grade

50 (minimum specified yield strength of 50,000 psi) and is painted (Ambrose 2012).

117
5.3.2 Modeling Results
As with bridges 7R and US13, the deck was already created using 2-D

elements and utilizing the “rebar” command. This bridge is so large and the elements

used to create it so small, a picture of the model is not included as no details can be

discerned at the size required to fit within these margins. As with 7R and US13, the

uncorroded (Tables 4.3 and 4.6) and corroded (Table 4.9) non-linear input which was

calibrated, as described in Sections 4.4.2 and 4.4.3, was then applied directly to the

bridge deck. As a means of comparison, the bridge model was analyzed in three

conditions: without any non-linear concrete or plastic rebar commands, with

uncorroded non-linear concrete and plastic rebar commands, and with corroded non-

linear concrete and plastic rebar commands. The results of this initial analysis can be

seen in Table 5.8. As with Bridge US13, these models all terminate prematurely.

Again, the results files indicated that a convergence problem was the cause of

termination. In hopes of determining the cause of the convergence problems, the

results files were analyzed visually and the 3 model versions were compared.

However, as with US13, due to the maximum load not reaching a high enough value,

very little information can be obtained from these results.

Table 5.8. Initial results from analyzing Bridge US299.


Equivalent Number of
Model Maximum Load (kips) Trucks
1
Elastic 1157.5 16.1
1
Uncorroded 40.9 0.6
1
Corroded 26.4 0.4
1
Model terminated at peak load, indicating inability to converge.

As with bridges 7R and US13, the values of tension stiffening were modified.

The values which were analyzed were 0.1, 0.01, and 0.035. Again, the tension

118
damage coefficient commands were removed. The results of these models can be seen

in Table 5.9. Although the results of these models produced higher strengths than

with the previous tension stiffening values, all of the models again terminated at the

maximum load rather than decreasing and failed to converge. The only exception was

the corroded model with a tension stiffening value of 0.01, which reached the

maximum number of increments. Due to time constraints, this model was not able to

be run with a larger number of increments to determine the maximum load or if

convergence was again a problem. As a result of this inability to converge, the

relative accuracy of the results could not be analyzed.

Table 5.9. Results of Bridge US299 when varying tension stiffening values.
Tension Maximum Equivalent
Model
Stiffening Loads (kips) Number of Trucks
1
Uncorroded 620.3 8.6
0.1 1
Corroded 692.3 9.6
1
Uncorroded 620.3 8.6
0.035 1
Corroded 664.6 9.2
1
Uncorroded 969.2 13.5
0.01 2
Corroded 603.7 8.4
1
Model terminated at peak load, indicating inability to converge.
2
Model terminated at maximum number of analysis increments specified.

5.4 Conclusions
It can be seen in Sections 5.1 through 5.3 that the modeling of an uncorroded

and corroded deck on a full-scale finite element bridge model still needs to be refined.

Although the input using the smaller 2-D beam was calibrated, directly applying this

to the full-scale bridges did not work as desired. Many of the bridges experienced

problems with convergence, terminating prematurely. This caused the models to be

analyzed using different tension stiffening values in order to aid in convergence. In

119
Bridge 7R, these new tension stiffening values enabled convergence; however, the

SR299 and US13 models did not converge using the updated tension stiffening values.

Due to this lack of convergence, and because more information about the bridge was

known, more emphasis was placed on the analysis of 7R. In addition, both SR299 and

US13 were multiple spans and statically indeterminate; it was thought that this likely

contributed to the convergence problems which were experienced.

After modifying the tension stiffening input, the models of 7R converged,

reaching a maximum loading before decreasing. The maximum loading of these

models was determined. In addition to the strength results, the distribution factors of

the bridge models were analyzed. It was seen in these results that the models with

lower tension stiffening values have distribution factors closer to those which were

expected based on theoretical inelastic bridge behavior. It can also be seen that, in

general, the corrosion in the deck caused a more uniform distribution between the

girders than with an uncorroded deck. Consequently, the results also indicated that the

corroded models reached higher strengths than their uncorroded counterparts. It was

thought that the cause of both the strength and DF value discrepancies was due to

greater load sharing in the corroded model when compared to the corresponding

uncorroded model.
Due to time constraints, the proper modeling technique for bridges US13 and

SR299 was not able to be determined. It is hypothesized that one of the reasons for

this is that both bridges are continuous and therefore are statically indeterminate; due

to this, they contain a large region of concrete in tension over the center piers. More

effort should be placed on determining the appropriate modeling technique to allow

for convergence in these situations. The lack of results for these models prevented the

120
validation of the relationship between deck corrosion and the system effects to the

bridge. Although the results of 7R were able to be analyzed thoroughly, more models

are needed for comparison to validate whether the trends observed in these results are

indicative of a general phenomenon. In addition, it is suggested that more work be

performed to determine a relationship between the current rating systems used for

inspecting bridge decks and how these ratings may correlate to the corrosion variables

input into these analyses. Determining this information on a general basis would

allow the relationship between deck corrosion and system capacity to be employed

when evaluating structurally deficient bridges, which could ultimately serve as a

means for prioritizing bridges for repairs and rehabilitation.

121
Chapter 6

CONCLUSIONS

6.1 Summary
The objectives of this thesis were to create and calibrate a reinforced concrete

model that considers the effects of corrosion due to deicing agents using finite element

analysis, then apply this modeling method to full-scale bridge models to determine

how deck corrosion effects the system capacity of bridges. The system capacity of

bridges is a topic not well quantified; however, the system capacity can significantly

affect the strength of a bridge and cause it to be higher than that of the design strength.

Previously, a literature review was performed to determine how corrosion

effects the behavior of reinforced concrete members (McCarthy 2012), which was

described in Section 2.1.2. The results of this review suggested that reasonable

expectations for corroded reinforced concrete, after 25 years of deterioration, were a

50% decrease in maximum strength and an 82% increase in ultimate deflection

(McConnell et al. 2012). These material properties were used as baseline goals for

calibrating a reinforced concrete beam finite element model.

Different modeling techniques for modeling reinforced concrete were also

researched. Specifically, this search was narrowed down to include modeling

techniques using the commercial finite element software ABAQUS. These modeling

techniques included the brittle cracking, smeared crack (SC), and concrete damaged

plasticity (CDP) approaches (Simulia 2011). The input commands and values for

these modeling techniques were investigated to determine the values which needed to

be utilized. These commands and associated input for the concrete and rebar are

described in Sections 3.1 and 3.2, respectively.

122
Once the modeling approaches were determined, finite element models of

reinforced concrete beams were created based on actual beams that were corroded and

tested in literature (Oyado et al. 2010). Two-dimensional and 3-D beam models were

created. Initially, 2-D beams were utilized, as this modeling approach is simpler and

directly applicable to previously created bridge models. This modeling technique is

described in Section 4.4. However, it was thought that using 3-D beams would be a

more accurate representation of the physical changes caused by corrosion in reinforced

concrete members; the 3-D modeling allowed for encompassing more properties,

including the bond between the concrete and rebar and the pressure caused by the

creation of corrosion by-products. This modeling technique is described in Section

4.5. For the 3-D beam models, all 3 of the approaches were analyzed. However, in

the 2-D beam models, the material points never reach the failure point for the brittle

cracking technique due to the elements carrying both compressive and tensile forces

and therefore the brittle cracking technique was not analyzed.

In addition to these modeling techniques and material properties, hand

calculations were performed to determine the theoretical strength and deflection

values associated with the experimental specimen and the corresponding values as the

material and geometric properties were varied in the calibration models. A description
of these calculations is presented in Section 4.3. The results of the models were

compared to these theoretical values to assess the accuracy of the modeling

approaches and input values.

The 2-D beam models were calibrated to simulate the strength and deflection

effects of corrosion to the extent possible. This calibration process is described in

Section 4.4.3. This was done by varying the different input parameters. These

123
parameters included the values used for the modulus of elasticity of concrete, the

tensile strength of concrete, the compressive strength of concrete, the area of

reinforcing bars, the yield strength of the reinforcing bars, and the dilation angle,

along with including or disregarding the use of the optional damage coefficients in the

CDP technique. A mesh sensitivity analysis was performed to determine the

appropriate element sizes and establish a strength and deflection baseline. The results

of the literature review (i.e., 50% decrease in maximum strength with 82% increase in

ultimate deflection) were then targets for the strength and deflection results of this

corroded model. Once the input parameters were calibrated, these input values were

then applied as the corroded concrete input for future bridge models.

The 3-D beam models which were created used two different types of rebar: 2-

D rebar and 3-D rebar elements. Using 3-D rebar elements allowed for the inclusion

of the effects of the varying bond between the concrete and rebar as the rebar

corrodes. As with the 2-D beams, the results of the literature review were the criteria

applied to the uncorroded base model to attempt to determine the new material

properties for a corroded model. This process for the 2-D and 3-D rebar is described

in Sections 4.5.1 and 4.5.2, respectively.

Once the input values for the uncorroded and corroded reinforced concrete
models were determined, the input was applied to full-scale bridge models. These

models were previously created and calibrated based on existing bridges within

Delaware and prior field testing performed on the bridges (Ross 2007; Michaud 2011;

McConnell et al. in review; Ambrose 2012; Radovic, personal communication 2013).

The concrete decks of these models were already created using 2-D elements allowing

for direct application of the concrete material property input which was calibrated for

124
2-D uncorroded and corroded reinforced concrete. The application of the uncorroded

and corroded deck input are described in Chapter 5.

6.2 Results
A 2-D reinforced concrete beam model was created. A mesh sensitivity

analysis determined that the optimal mesh size to be 0.7874 in (20 mm) squares. The

mesh sensitivity analysis also indicated that the SC model results were erratic and not

analogous to the expected value along with an inability to converge. The input values

for the uncorroded model were determined; this model used a dilation angle of 38,

flow potential eccentricity (ϵ) of 1, ratio of initial equibiaxial compressive yield stress

to initial uniaxial compressive stress (σb0/σc0) of 1.12, ratio of the second stress

invariant on the tensile meridian to that of the compressive meridian at initial yield

(Kc) of 2/3, area of compressive steel (A’s) of 0.0438 in2, area of tensile steel (As) of

0.20563 in2, modulus of elasticity of compressive steel (E’s) and modulus of elasticity

of tensile steel (Es) of 29,000 ksi, yield strength of compressive steel (f’y) of 47,000

psi, yield strength of tensile steel (fy) of 55,000 psi, compressive strength of concrete

(f’c) of 3147.3 psi, tensile strength of concrete (f’t) of 420.8 psi, and modulus of

elasticity of concrete (Ec) of 3,197,746 psi. This model resulted in a load


proportionality factor (LPF) of 1.10 (6,491 lbs) and a corresponding deflection at mid-

span of 0.0435 inches at the same level of load.

Through the literature review, it was determined that the corroded model

should experience a 50% decrease in strength and an 82% increase in deflection. This

is synonymous with a targeted LPF of 0.55 and a targeted deflection of 0.0792 in.

Through a parametric calibration analysis, the corroded model input best achieving

these targets was determined; the corroded model used an Ec of 1,918,648 psi, f’t of

125
252.4 psi, f’c of 1,133 psi, As of 0.07711 in2, and an A’s of 0.03504 in2. These values

correspond to a: 40% decrease in Ec, 40% decrease in f’t, 64% decrease in f’c, 20%

decrease in A’s, and 61.5% decrease in As. All the remaining values were unchanged

from the uncorroded model. These Ec, f’t, and f’c were determined by decreasing the

modulus of elasticity and calculating the tensile and compressive strengths based on

Equations 3.1 and 3.2, respectively. This model resulted in a 45.2% decrease in

strength and 11.6% increase in deflection at the maximum loading when compared to

the uncorroded model at its maximum loading. Thus, the strength goals were deemed

to be satisfied to a reasonable extent. Although the deflection increase was less than

the goal value, this difference in deflection was based on the maximum loading for

each model, while in the experimental results the corresponding change in deflection

was evaluated at the ultimate condition, which would theoretically produce a greater

discrepancy. As a simple means to evaluate this hypothesis, deflections were

compared at a common loading, using the maximum loading for the corroded beam;

the percent difference then becomes 123%, as the uncorroded beam had a deflection of

0.0213 in and the corroded beam a deflection of 0.0477 in. Thus, the deflection

targets were also deemed to be generally satisfied.

The 3-D model with 2-D rebar elements was analyzed next. A mesh
sensitivity analysis was performed using the brittle cracking, CPD, and SC concrete

models; it was determined to use 0.7874 in (20 mm) cubic elements with 3.937 in (100

mm) long rod elements for the rebar. Base input values for the uncorroded model

were determined using the brittle cracking technique; these values were an f’c of

3,147.3 psi, f’t of 420.8 psi, Ec of 3,197,746 psi, direct cracking strain of 0.0027, direct

cracking failure strain of 0.0027, crack opening strain of 0.0027, A’s of 0.0438 in2, and

126
an As of 0.20563 in2. These resulted in a LPF of 0.850 (5,016 lbs) and a deflection at

mid-span of 0.616 in. This leads to a targeted LPF of 0.425 LPF (2,508 lbs) and a

deflection goal of 1.12 in for the corroded model.

A parametric study was performed. The results of this study indicated that

utilizing 3-D beam elements with 2-D rebar elements was not conducive to accurately

modeling corrosion. Although the models did display a decrease in strength, the

deflection also decreased proportionately to the strength; this is the opposite of the

desired change in deflection. For this reason, 3-D concrete with 3-D rebar models

were analyzed in attempts to more accurately model corrosion.

The final reinforced concrete model which was analyzed was the 3-D concrete

beam elements using 3-D rebar elements. A mesh sensitivity analysis was not

performed, as the cross-sectional area of the rebar limited the beam element sizes.

These elements measured 0.3937 in x 0.5249 in x 0.3937 in (10 mm x 13.33 mm x 10

mm); all of these dimensions of the elements are smaller than the previously

determined optimal mesh sizes. Initially, these models were analyzed using the SC

and CDP techniques. However, when using these techniques, the model failed to

converge and terminated early; the results of these models were less than 1/3 of the

calculated expected strength using the theoretical equations as well as compared to the
previously created models. A technique which allowed these models converge was

not determined; for this reason, this modeling technique was not used.

A time analysis was performed using the brittle cracking technique. As

opposed to the SC and CDP modeling techniques, the brittle cracking approach

utilizes ABAQUS/Explicit rather than ABAQUS/Standard. For this reason, the time

over which the loading is applied is included in the modeling and an analysis was

127
performed to determine the appropriate length of time to apply the loading. These

results concluded with both 1 second and 10 being utilized for subsequent analyses.

As with the models using SC and CDP, the brittle cracking models also produced

strengths significantly lower than the calculated theoretical strength and those

produced using the 2-D beam models, ranging from 27% to 82% of the theoretical

strength. For this reason, the corroded input was not calibrated and this modeling

approach was not applied to the full-scale bridge models.

Once the input for the uncorroded and corroded beams was determined, it was

applied to 3 different full-scale bridge models previously created based on actual

bridges located in Delaware. Due to time constraints and convergence difficulties,

only one of these models was thoroughly analyzed; this bridge is referred to as 7R and

served as an exit ramp for Interstate 295 North, just south of the Delaware Memorial

Bridge before being decommissioned in 2010. Initially, the input which was

calibrated using the 2-D beam models was applied. However, this input prevented the

models from converging on a final result. For this reason, the tension stiffening input

associated with the “concrete tension stiffening” command was modified and strain

values at which the stress decreased to zero of 0.01, 0.035, and 0.1 were analyzed.

Modifying these values allowed all of the 7R models to converge, while the remaining
bridges still experienced an inability to converge.

The maximum loading of the 7R bridge models was determined. In addition to

the strength results, the distribution factors (DF) of the bridge models were analyzed.

It was seen in these results that the models with lower tension stiffening values have

distribution factors closer to those which were experienced during the field test. It can

also be seen that, in general, the corrosion in the deck caused a more uniform

128
distribution between the girders than with an uncorroded deck. The results of these

models also indicated that the corroded models reached higher strengths than their

uncorroded counterparts. It was thought that the cause of the differences in both the

strength and DF values was due to greater load sharing in the uncorroded model when

compared to the corresponding corroded model.

The DF values between different models were compared; it was thought that, if

corrosion decreases the load that can be distributed through the deck, the elastic model

should have the lowest DF values for the girders closest to the location of the applied

loading, followed by the uncorroded and then the corroded. However, it was seen that

this is not the case. The results of the corroded model produced DF values between

those of the elastic and those of the uncorroded models due to the corroded deck

distributing the load more uniformly throughout the deck as the peak stress for any

individual element was limited to a smaller value. However, for all three models

(elastic, uncorroded, and corroded deck models), the distribution factor results for the

two-most heavily loaded girders (which are the two girders whose loading is

synonymous with valid load positions for calculating distribution factors) are within

11% of the theoretical inelastic values and are a 25 to 46% reduction relative to the

current AASHTO (2013) elastic distribution factors; the percent values reported here
are for the 0.035 tension stiffening model but similar results are obtained for models

with other tension stiffening values.

6.3 Future Work


A modeling technique utilizing 3-D elements that was able to converge was

not able to be concluded; the 3-D modeling technique offers a more direct approach to

modeling the physical effects of corrosion within concrete, as opposed to 2-D models

129
which require the general strength and deflection effects of corrosion to be modeled

indirectly. Therefore, future work should be done to determine the appropriate means

of modeling reinforced concrete using 3-D elements. The results of the uncorroded

beam models utilizing the brittle cracking technique did not meet the strength

expectations and the CDP and SC techniques were unable to converge. More effort

should be put into researching modeling techniques and input parameters for 3-D

beam modeling to potentially determine a more accurate corroded reinforced concrete

beam model.

When applying the 2-D reinforced concrete input to the full-scale bridge

models, the continuous bridge models did not converge. A more thorough analysis

into the cause which prevented convergence should be performed. The results from

additional bridge models would help to determine a general relationship between deck

corrosion and the system capacity of bridges. It is suggested that, when analyzing

additional bridges, simple span bridges (to initially minimize potential convergence

problems) be used as a starting point to determine if the results of 7R are consistent

amongst simple span steel I-girder bridges. In addition, future work is required to

determine the correlation of the deck condition based on current rating systems to the

material input parameters used in the model. This would help create a system to better
prioritize the repair of bridges categorized as structurally deficient or functionally

obsolete based on their actual system capacity.

In addition, the results of Bridge 7R should be further investigated.

Specifically, the results indicated that the input was not performing in an anticipated

way and more effort should be placed in determining why the concrete elements were

not unloading after reaching the defined maximum stress. It is thought that this may

130
be due to discrepancies between the direct input values and the evolution of the failure

surface as the multi-axial stresses develop in the model and further evaluation of this

hypothesis should be considered. Alternatively, additional analysis of the stress

distribution throughout the bridge at different loadings would help to better understand

the system behavior with different deck conditions. Evaluating the regions of the deck

in tension and compression over a girder line is also suggested to determine if the

changes from tension to compression observed in the present results are realistic.

Lastly, the results presented here were calibrated based on relatively small,

simple beam and cylinder tests with limited physical specimens. Additional research

should be performed to more accurately understand how corrosion effects the material

properties of and load transfer mechanisms within reinforced concrete decks and

members along with the quantification of the variables affecting these responses in

terms consistent with FEA input, as well as the global response generally expected for

these members. Specifically, more large-scale tests based on realistic exposure

conditions would provide valuable insight. Furthermore, calibrating FEA models to

existing corroded cylinder data could provide additional insights.

131
REFERENCES

Adelaide, L., Richard, B., Ragueneua, F., and Cremona, C. (2012). “A simplified
numerical approach of global behavior of RC beams degraded by corrosion.”
European J. of Envr. And Civ. Eng., 16(3-4), 414-439.

Ambrose, K. (2012). “Field Measurements and Corresponding FEA of Cross-Frame


Forces in Skewed Steel I-Girder Bridges.” M.C.E. thesis, Univ. of Delaware,
Newark, DE.

American Association of State Highway and Transportation Officials (AASHTO).


(2013). “AASHTO LRFD Bridge Design Specifications.” Washington DC, 6th
Edition.

American Society of Civil Engineers (ASCE). (2009). “Bridges.” Report Card for
America’s Infrastructure, <http://www.infrastructurereportcard.org/fact-
sheet/bridges> (Nov. 22, 2012).

Amleh, L., and Ghosh, A. (2006). “Modeling the effect of corrosion on bong strength
at the steel-concrete interface with finite-element analysis.” Can. J. Civ. Eng.,
33, 673-682.

Barth, K. E., White, D. W., Righman, J. E., and Yang, L. (2005). “Evaluation of web
compactness limits for singly and doubly symmetric steel I-girders.” J. Constr.
Steel Res., 61(10), 1411-1434.

Baskar, K., Shanmugam, N. E., and Thevendran, V. (2002). “Finite-Element Analysis


of Steel-Concrete Composite Plate Girder.” J. of Struct. Eng., 128(9), 1158-
1168.

Bechtel, A., McConnell, J., and Chajes, M. (2011). “Ultimate Capacity Destructive
Testing and Finite-Element Analysis of Steel I-Girder Bridges.” J. Bridge
Eng., 16, 197-206.

Biondini, F. and Vergani, M. (2012). “Damage modeling and nonlinear analysis of


concrete bridges under corrosion.” Proc., Int. Conf. on Bridge Maintenance,
Safety, and Management, Taylor and Francis, Florence, KY.

Chajes, M., McConnell, J., Shenton, H., Michaud, K., Ross, J., and Russo, C. (2010).
“Full-Scale Destructive Bridge Test Allows Prediction of Ultimate Capacity”,
J. of the Transportation Research Board, 2200, 117-124.

132
Chen, D., and Mahadevan, S. (2007). “Chloride-induced reinforcement corrosion and
concrete cracking simulation.” Cement and Conc. Composites, 30, 227-238.

Coronelli, D. and Gambarova, P. (2004). “Structural Assessment of Corroded


Reinforced Concrete Beams: Modeling Guidelines.” J. of Struct. Eng., 130(8),
1214-1224.

Dekoster, M., Buyle-Bodin, F., Maurel, O., and Delmas, Y. (2003). “Modelling of the
flexural behavior of RC beams subjected to localized and uniform corrosion.”
Eng. Struct., 25, 1333-1341.

Fang, C., Lundgren, K., Plos, M., and Gylltoft, K. (2006). “Bond behavior of corroded
reinforcing steel bars in concrete.” Cement and Conc. Research, 36, 1931-
1938.
Gu, X. L., Zhang, W. P., Shang, D. F., and Wang, X. G. (2010). “Flexural Behavior of
Corroded Reinforced Concrete Beams.” Earth and Space 2010: Engineering,
Science, Construction, and Operations in Challenging Environments, ASCE,
Reston, VA, 3545-3552.

Jankowiak, T. and Lodygowski, T. (2005). “Identification of Parameters of Concrete


Damage Plasticity Constitutive Model.” Foundations of Civ. and Env. Eng., 6,
53-69.

Kallias, A. N. and Rafiq, M. I. (2010). “Finite element investigation of the structural


response of corroded RC beams.” Eng. Struct., 32, 2984-2994.

Li, C.Q. (2000). “Corrosion Initiation of Reinforcing Steel in Concrete under Natural
Salt Spray and Service Loading – Results and Analysis”. ACI Materials
Journal, 97(6), 690-697.

Li, C. Q. and Zheng, J. J. (2005). “Propagation of Reinforcement Corrosion in


Concrete and its Effects on Structural Steel.” Mag. of Conc. Research, 57(5),
261-271.

McCarthy, G. (2012). “Improved Bridge Rating Procedures Integrating Load Path


Redundancy: A Cost-Effective Simulation.” M.C.E. thesis, Univ. of Delaware,
Newark, DE.

McConnell, J., Chajes, M. and Michaud, K. (in review). “Destructive Testing of a


Decommissioned Skewed Steel I-Girder Bridge: Analysis of System Effects.”
J. of Struct. Eng.

133
McConnell, J, McCarthy, G, and Wurst, D. (2012). “An Approach to Evaluating the
Influences of Aging on the System Capacity of Steel I-Girder Bridges.” Proc.,
Int. Conf. on Bridge Maintenance, Safety, and Management, Taylor and
Francis, Florence, KY.

Melchers, R. E., Li, C. Q., Lawanwisut, W. (2006). "Modelling deterioration of


structural behavior of reinforced concrete beams under saline environment
corrosion." Mag. of Concrete Research, 58(9), 575-587.

Michaud, K. M. (2011). “Evaluating Reserve Bridge Capacity Through Destructive


Testing of a Decommissioned Bridge.” M.C.E. thesis, Univ. of Delaware,
Newark, DE.

Morales, E. M. (n.d.). “Significance of the Ratio of Tensile Strength to Yield Stress


(TS/YS) of Reinforcing Bars.” <
http://www.pgatech.com.ph/documents/Significance%20of%20the%20Ratio%
20of%20Tensile%20Strength%20to%20Yield%20Stress.PDF> (November 15,
2013).

Oyado, M., Kanakubo, T., Sato, T., and Yamamoto, Y. (2010). “Bending performance
of reinforced concrete member deteriorated by corrosion.” Struct. and
Infrastructure Eng., 7(1-2), 121-130.

Richard, B., Ragueneau, F., Adelaïde, L., and Cremona, C. (2011). “A multi-fiber
approach for modeling corroded reinforced concretes structures.” European J.
of Mech., 30, 950-961.
Ross, J. H. (2007). “Evaluating Ultimate Bridge Capacity Through Destructive
Testing of Decommissioned Bridges.” M.C.E. thesis, Univ. of Delaware,
Newark, DE.

Shirai, N., Watanabe, K., Oh-oka, T., Hakuto, S., and Fujita, T. (2002). “Finite
element analysis of shear wall specimens made of ductile fiber reinforced
cementitious composites subjected to lateral loading.” Proc., Finite Elements
in Civ. Eng. Applications, Taylor & Francis, Tokyo, Japan, 219-227.

Simulia. (2011). ABAQUS/CAE User Manual v. 6.11. ABAQUS, Providence, RI.

Takahashi, Y. (2008). “Flexural Strengthening of RC Beams with CFRP Sheets and


U-Jackets.” Proc., Int. Conference on FRP Composites in Civ. Eng., Zurich,
Switzerland.

134
Val, D. V., Chernin, L, and Stewart, M. G. (2009). “Experimental and Numerical
Investigation of Corrosion-Induced Cover Cracking in Reinforced Concrete
Structures.” J. of Struct. Eng., 135, 376-385.

Wight, J. K., and MacGregor, J. G. (2009). "Reinforced Concrete Mechanics And


Design." Pearson Education, Inc., Upper Saddle River, NJ, 158-163.

135
Appendix A

SAMPLE STRENGTH CALCULATIONS

Variable Definitions:
T = tensile force
C = total compressive force
Cc = compressive force in concrete
Cs = compressive force in steel
As = area of tensile steel
A’s = area of compressive steel
f’c = compressive strength of concrete
f’t = tensile strength of concrete
fy = ultimate strength of tensile rebar
f’y = ultimate strength of compressive rebar
b = length of base of concrete cross-section
β1 = 0.85
c = distance from top of beam to neutral axis
f’s = force in compressive rebar
Ec = modulus of elasticity of concrete
Es = modulus of elasticity of tensile steel
E’s = modulus of elasticity of compressive steel
εs = strain in tensile steel
ε’s = strain in compressive steel
εcu = strain in concrete at cracking
d = distance from top of beam to tensile rebar
d’ = distance from top of beam to compressive rebar
h = total height of beam
I = moment of inertia of reinforced concrete beam
Ic = cracked moment of inertia of reinforced concrete beam
Ast = transformed area of tensile steel
A’st = transformed area of compressive steel
n = modular ratio of tensile steel
n’ = modular ratio of compressive steel
P = force applied at one point on beam during loading
L = total length of beam
x = distance from support to P

136
Generic Equations

Neutral Axis:
𝑇 = 𝐶 = 𝐶𝑐 + 𝐶𝑠

𝐴𝑠 𝑓𝑦 = (0.85)𝑓 ′ 𝑐 ∗ 𝑏 ∗ 𝛽1 ∗ 𝑐 + 𝐴′ 𝑠 (𝑓 ′ 𝑠 − 0.85 ∗ 𝑓 ′ 𝑐)

𝑓 ′ 𝑠 = 𝐸′𝑠 𝜀′𝑠

𝑐 − 𝑑′
𝜀′𝑠 = ( ) 𝜀𝑐𝑢
𝑐

𝐴𝑠 𝑓𝑦 = 0.85 ∗ 𝑓 ′ 𝑐 ∗ 3.937 ∗ 0.85 ∗ 𝑐 + 𝐴′𝑠


𝑐 − 0.7874
∗ {[( ) ∗ 0.003] ∗ 𝐸′𝑠 − 0.85 ∗ 𝑓′𝑐}
𝑐
Solve for c:

𝐴2𝑠 𝑓𝑦2 − 0.006𝐴𝑠 𝐴′ 𝑠 (𝐸 ′ 𝑠 − 283.333𝑓 ′ 𝑐 )𝑓𝑦


√+0.000009𝐴′ 𝑠 (𝐴′ 𝑠 (𝐸 ′ 2𝑠 − 566.667𝐸 ′ 𝑠 𝑓 ′ 𝑐 + 80277.8𝑓 ′ 2𝑐 )
+2986.33𝐸′𝑠 𝑓′𝑐 )
𝐴𝑠 𝑓𝑦 − 0.003𝐴′ 𝑠 (𝐸 ′ 𝑠 − 283.333𝑓 ′ 𝑐 )
𝑐 = 0.175779 ∗
𝑓 ′𝑐

Moment of Inertia:
𝐸𝑠
𝑛=
𝐸𝑐

𝐸′𝑠
𝑛′ =
𝐸𝑐

𝐴𝑠𝑡 = 𝑛 ∗ 𝐴𝑠

𝐴′𝑠𝑡 = 𝑛′ ∗ 𝐴′𝑠
2
𝑏𝑐 3 𝑐 2 𝑏(ℎ − 𝑐)3 ℎ − 𝑐 − 𝑑′
𝐼= + 𝑏𝑐 ( ) + + 𝑏(ℎ − 𝑐) ( ) + 𝐴𝑠𝑡 (ℎ − 𝑐 − 𝑑′)2
12 2 12 2
+ 𝐴′𝑠𝑡 (𝑐 − 𝑑′)2

137
𝑏𝑐 3 𝑐 2
𝐼𝑐 = + 𝑏𝑐 ( ) + 𝐴𝑠𝑡 (ℎ − 𝑐 − 𝑑′)2 + 𝐴′𝑠𝑡 (𝑐 − 𝑑′)2
12 2

Moment Capacity:

𝐶𝑐 = 0.85 ∗ 𝑓 ′ 𝑐 ∗ 𝑏 ∗ 𝛽1 ∗ 𝑐

𝐶𝑠 = 𝐴′𝑠 (𝑓 ′ 𝑠 − 0.85𝑓′𝑐)

𝛽1 𝑐
𝑀𝑛 = 𝐶𝑐 (𝑑 − ) + 𝐶𝑠 (𝑑 − 𝑑′)
2

𝑀𝑛 = 𝑃𝑥

Deflection:

𝑃𝑥
∆𝑚𝑎𝑥 = (3𝐿2 − 4𝑥 2 )
24𝐸𝑐 𝐼

Check Rebar Yielded:

𝑓𝑦
𝜀𝑦 =
𝐸𝑠

𝑑−𝑐
𝜀𝑠 = ( ) 𝜀𝑐𝑢
𝑐

𝜀𝑠 ≥ 𝜀𝑦

138
Sample Calculation of Base Model

Given Values:
As = 0.20563in2 β1 = 0.85 f’c = 3147.2psi
A’s = 0.0438in2 Es = E’s = 29000000psi Ec = 3197746psi
fy = 55,000psi d’ = 0.7874in h = 7.874in
f’y = 47,000psi d = 7.0866in x = 27.5586in
b = 3.937in εcu = 0.003

𝐴𝑠 𝑓𝑦 = (0.85)𝑓 ′ 𝑐 ∗ 𝑏 ∗ 𝛽1 ∗ 𝑐 + 𝐴′ 𝑠 (𝑓 ′ 𝑠 − 0.85 ∗ 𝑓 ′ 𝑐)

𝑓 ′ 𝑠 = 𝐸′𝑠 𝜀′𝑠

𝑐 − 𝑑′
𝜀′𝑠 = ( ) 𝜀𝑐𝑢
𝑐

𝑐 − 0.7874𝑖𝑛
𝜀 ′𝑠 = ( ) ∗ 0.003
𝑐

𝑐 − 0.7874𝑖𝑛
𝑓′𝑠 = 29,000,000𝑝𝑠𝑖 ∗ ( ) ∗ 0.003
𝑐

0.20563𝑖𝑛2 ∗ 55,000𝑝𝑠𝑖
= 0.85 ∗ 3147.3𝑝𝑠𝑖 ∗ 3.937𝑖𝑛 ∗ 0.85 ∗ 𝑐 + 0.0438𝑖𝑛2
𝑐 − 0.7874𝑖𝑛
∗ [29,000,000𝑝𝑠𝑖 ( ) ∗ 0.003 − 0.85 ∗ 3147.3𝑝𝑠𝑖]
𝑐

3693.43(𝑐 − 0.81238)
11,309.7 = 8942.44𝑐 +
𝑐
𝑐 = 1.14377𝑖𝑛
𝐸𝑠 29,000,000𝑝𝑠𝑖
𝑛 = 𝑛′ = =
𝐸𝑐 3197746𝑝𝑠𝑖
𝑛 = 𝑛′ = 9.06889

𝐴𝑠𝑡 = 𝐴𝑠 ∗ 𝑛 = 0.20563𝑖𝑛2 ∗ 9.06889

𝐴𝑠𝑡 = 1.86484𝑖𝑛2

𝐴′𝑠𝑡 = 𝐴′𝑠 ∗ 𝑛′ = 0.0438𝑖𝑛2 ∗ 9.06889

𝐴′𝑠𝑡 = 0.397217𝑖𝑛2

139
2
𝑏𝑐 3 𝑐 2 𝑏(ℎ − 𝑐)3 ℎ − 𝑐 − 𝑑′
𝐼= + 𝑏𝑐 ( ) + + 𝑏(ℎ − 𝑐) ( ) + 𝐴𝑠𝑡 (ℎ − 𝑐 − 𝑑′)2
12 2 12 2
+ 𝐴′𝑠𝑡 (𝑐 − 𝑑′)2

(3.937𝑖𝑛) ∗ (1.14377𝑖𝑛)3 1.14377𝑖𝑛 2


𝐼= + (3.937𝑖𝑛) ∗ (1.14377𝑖𝑛) ∗ ( )
12 2
(3.937𝑖𝑛) ∗ (7.874𝑖𝑛 − 1.14377𝑖𝑛)3
+ + (3.937𝑖𝑛)
12
7.874𝑖𝑛 − 1.14377𝑖𝑛 − 0.7874𝑖𝑛 2
∗ (7.874𝑖𝑛 − 1.14377𝑖𝑛) ∗ ( )
2
+ 1.86484𝑖𝑛2 ∗ (7.874𝑖𝑛 − 1.14377𝑖𝑛 − 0.7874𝑖𝑛)2 + 0.397217𝑖𝑛2
∗ (1.14377𝑖𝑛 − 0.7874𝑖𝑛)2
𝐼 = 401.851𝑖𝑛4

𝑏𝑐 3 𝑐 2
𝐼𝑐 = + 𝑏𝑐 ( ) + 𝐴𝑠𝑡 (ℎ − 𝑐 − 𝑑′)2 + 𝐴′𝑠𝑡 (𝑐 − 𝑑′)2
12 2
(3.937𝑖𝑛) ∗ (1.14377𝑖𝑛)3 1.14377𝐼𝑁 2
𝐼𝑐 = + (3.937𝑖𝑛) ∗ (1.14377𝑖𝑛) ∗ ( )
12 2
+ 1.86484𝑖𝑛2 ∗ (7.874𝑖𝑛 − 1.14377𝑖𝑛 − 0.7874)2 + 0.397217𝑖𝑛2
∗ (1.14377𝑖𝑛 − 0.7874𝑖𝑛)2

𝐼𝑐 = 67.875𝑖𝑛4

𝐶𝑐 = 0.85 ∗ 𝑓 ′ 𝑐 ∗ 𝑏 ∗ 𝛽1 ∗ 𝑐 = 0.85 ∗ 3147.3𝑝𝑠𝑖 ∗ 3.937𝑖𝑛 ∗ 0.85 ∗ 1.14377𝑖𝑛

𝐶𝑐 = 10239.5𝑙𝑏
1.14377𝑖𝑛 − 0.7874𝑖𝑛
𝜀′𝑠 = ( ) ∗ 0.003
1.14377𝑖𝑛
𝜀′𝑠 = 0.000935
𝑓 ′ 𝑠 = 29,000,000𝑝𝑠𝑖 ∗ 0.000935

𝑓′𝑠 = 27,107𝑝𝑠𝑖
𝐶𝑠 = 𝐴′ 𝑠 ∗ (𝑓 ′ 𝑠 − 0.85 ∗ 𝑓 ′ 𝑐 ) = 0.0438𝑖𝑛2 ∗ (27,107𝑝𝑠𝑖 − 0.85 ∗ 3147.3𝑝𝑠𝑖)

𝐶𝑠 = 1070.1𝑙𝑏
𝛽1 ∗ 𝑐
𝑀𝑛 = 𝐶𝑐 ∗ (𝑑 − ) + 𝐶𝑠 ∗ (𝑑 − 𝑑 ′ )
2

141
0.85 ∗ 1.14377𝑖𝑛
= 10239.5𝑙𝑏 ∗ (7.0866𝑖𝑛 − ) + 1070.1𝑙𝑏
2
∗ (7.0866𝑖𝑛 − 0.7874𝑖𝑛)

𝑀𝑛 = 74,327𝑙𝑏 − 𝑖𝑛
𝑀𝑛 74,327𝑙𝑏 − 𝑖𝑛
𝑃= =
𝑥 27.5586𝑖𝑛
𝑃 = 2,697.1𝑙𝑏

𝑃𝑥
∆𝑢𝑛𝑐𝑟𝑎𝑘𝑐𝑒𝑑 = ∗ (3 ∗ 𝐿2 − 4 ∗ 𝑥 2 )
24 ∗ 𝐸𝑐 ∗ 𝐼
2697.1𝑙𝑏 ∗ 27.5586𝑖𝑛
= ∗ (3 ∗ (82.677𝑖𝑛)2 − 4 ∗ (27.5586𝑖𝑛)2 )
24 ∗ 3197749𝑝𝑠𝑖 ∗ 401.41𝑖𝑛4
∆𝑢𝑛𝑐𝑟𝑎𝑐𝑘𝑒𝑑 = 0.0421𝑖𝑛
𝑃𝑥
∆𝑐𝑟𝑎𝑐𝑘𝑒𝑑 = ∗ (3 ∗ 𝐿2 − 4 ∗ 𝑥 2 )
24 ∗ 𝐸𝑐 ∗ 𝐼𝑐
2697.1𝑙𝑏 ∗ 27.5586𝑖𝑛
= 4
∗ (3 ∗ (82.677𝑖𝑛)2 − 4 ∗ (27.5586𝑖𝑛)2 )
24 ∗ 3197749𝑝𝑠𝑖 ∗ 67.871𝑖𝑛
∆𝑐𝑟𝑎𝑐𝑘𝑒𝑑 = 0.2493𝑖𝑛

𝑓𝑦 55,000𝑝𝑠𝑖
𝜀𝑦 = =
𝐸𝑠 29,000,000𝑝𝑠𝑖
𝜀𝑦 = 0.001897
𝑑−𝑐 7.0866𝑖𝑛 − 1.14377𝑖𝑛
𝜀𝑠 = ( ) 𝜀𝑐𝑢 = ( ) ∗ 0.003
𝑐 1.14377𝑖𝑛
𝜀𝑠 = 0.015587
𝜀𝑠 ≥ 𝜀𝑦

0.015587 ≥ 0.001897 √

142
Appendix B

COMPLETE RESULTS FOR 2-D BEAM CALIBRATION VARIATIONS

Table B.1. Results of mesh sensitivity analysis for 2-D beam models.
Tension Coefficient Y N N/A Y N N/A
Compression
Input

Coefficient Y N N/A Y N N/A


Concrete Model CDP CDP SC CDP CDP SC
Mesh Size (mm) 100 100 100 50 50 50
Highest Load (LPF) 1.75 1.74 4.72 2.55 2.47 3.02
Results

Highest Deflection (in) 0.053 0.053 0.383 0.282 0.283 1.413


Converged N N N Y Y N
Table B.1. cont’d
Tension Coefficient Y N N/A Y N N/A
Compression
Input

Coefficient Y N N/A Y N N/A


Concrete Model CDP CDP SC CDP CDP SC
Mesh Size (mm) 25 25 25 20 20 20
Highest Load (LPF) 0.890 0.931 4.10 1.10 1.12 2.98
Results

Highest Deflection (in) 0.023 0.022 0.211 0.043 0.041 0.136


Converged N Y Y Y Y N
Table B.1. cont’d
Tension Coefficient Y N N/A Y N N/A
Compression
Input

Coefficient Y N N/A Y N N/A


Concrete Model CDP CDP SC CDP CDP SC
Mesh Size (mm) 10 10 10 5 5 5
Highest Load (LPF) 1.08 1.08 2.39 3.69 3.65 10.20
Results

Highest Deflection (in) 0.042 0.042 0.099 0.002 0.002 0.007


Converged Y N N Y N N

143
Table B.2. Complete input and results of all concrete damaged plasticity models
tested during 2-D beam calibration.
Dilation Angle 38 38 38 38 36 40
Flow Potential
1 1 0.1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.16 1.12 1.12
Tension
Y N Y Y Y Y
Coefficient
Compression
Y N Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.2056 0.2056 0.2056 0.2056 0.2056 0.2056


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 29000 29000 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Ec (psi) 3197746 3197746 3197746 3197746 3197746 3197746
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 1.10 1.10 1.11 1.10 1.10 1.10


Results

Deflection (in) 0.043 0.041 0.044 0.043 0.043 0.045


Converged N Y Y N Y Y

144
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.1542 0.1028 0.2056 0.2056 0.2056 0.2056


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 29000 29000 29000 29000
f'y (psi) 47000 47000 47000 47000 42300 37600
fy (psi) 55000 55000 49500 44000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Ec (psi) 3197746 3197746 3197746 3197746 3197746 3197746
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 1.06 0.767 1.10 1.10 1.10 1.10
Results

Deflection (in) 0.045 0.032 0.043 0.043 0.043 0.043


Converged N Y Y N N N

145
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.03285 0.0438 0.0438 0.0438
Input

As (in2) 0.2056 0.2056 0.1542 0.1542 0.1028 0.1542


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 29000 29000 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 2832.57 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 378.72 420.8 378.72 378.72 315.6
Ec (psi) 3197746 3197746 3197746 3197746 3197746 3197746
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 1.10 1.02 1.01 0.976 0.705 0.856


Results

Deflection (in) 0.043 0.041 0.043 0.041 0.029 0.034


Converged Y N Y Y Y Y

146
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.2056 0.2056 0.2056 0.2056 0.2056 0.2056


E's (ksi) 29000 29000 29000 26100 29000 26100
Es (ksi) 29000 29000 29000 29000 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Ec (psi) 3037859 2877971 2398310 3197746 3197746 3197746
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 1.11 1.11 1.06 1.08 1.09 1.07


Results

Deflection (in) 0.045 0.047 0.050 0.042 0.044 0.043


Converged Y N Y Y N N

147
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.2056 0.2056 0.2056 0.2056 0.2056 0.2056


E's (ksi) 29000 29000 29000 31900 29000 31900
Es (ksi) 29000 29000 29000 29000 31900 31900
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Ec (psi) 3357633 3517521 3997183 3197746 3197746 3197746
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 1.1 1.09 1.08 1.12 1.12 1.13


Results

Deflection (in) 0.042 0.040 0.036 0.044 0.043 0.043


Converged N Y Y Y N N

148
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.1542 0.1028 0.2056 0.2056 0.1542 0.1542


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 29000 29000 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 378.72 315.6 378.72 315.6
Ec (psi) 2398310 2398310 2398310 2398310 2398310 2398310
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 0.929 0.803 0.985 0.976 1.00 0.908


Results

Deflection (in) 0.046 0.042 0.046 0.046 0.053 0.046


Converged Y Y N N Y N

149
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.1028 0.1028 0.1542 0.1542 0.1542 0.1542


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 26100 26100 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 378.72 315.6 378.72 315.6 378.72 378.72
Ec (psi) 2398310 2398310 2398310 2398310 2238422 1918648
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 0.734 0.523 0.932 0.843 0.994 1.11


Results

Deflection (in) 0.039 0.026 0.050 0.042 0.055 0.069


Converged Y N Y Y Y N

150
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.1542 0.1542 0.1542 0.1028 0.2056 0.2056


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 29000 29000 29000 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 2840.4 2549.3
f't (psi) 315.6 315.6 315.6 315.6 399.718 378.68
Ec (psi) 2398310 2238422 1918648 1918648 3037859 2877971
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 0.843 0.853 0.567 0.509 1.06 1.02


Results

Deflection (in) 0.042 0.044 0.031 0.030 0.043 0.043


Converged Y Y Y N Y N

151
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.2056 0.1542 0.1542 0.1028 0.1028 0.1542


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 29000 29000 29000 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1770.4 1542.2 1133.0 1542.2 1133.0 1133.0
f't (psi) 315.6 294.5 252.5 294.5 252.5 252.5
Ec (psi) 2398310 2238422 1918648 2238422 1918648 1918648
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 0.852 0.825 0.719 0.772 0.694 0.726


Results

Deflection (in) 0.041 0.046 0.045 0.049 0.049 0.047


Converged N Y N Y Y Y

152
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.1028 0.1542 0.1542 0.1028 0.1028 0.1542


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 29000 29000 29000 29000 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 44000 44000 44000 44000 44000
f'c (psi) 1133.0 1542.2 1133.0 1542.2 1133.0 1133.0
f't (psi) 252.5 294.5 252.5 294.5 252.5 252.5
Ec (psi) 1918648 2238422 1918648 2238422 1918648 1918648
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 0.680 0.825 0.719 0.772 0.681 0.726


Results

Deflection (in) 0.050 0.046 0.045 0.049 0.049 0.047


Converged Y Y N Y Y Y

153
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.03942 0.03504
Input

As (in2) 0.1028 0.0823 0.0617 0.0514 0.1028 0.1028


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 44000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 0.680 0.654 0.439 0.420 0.662 0.644


Results

Deflection (in) 0.050 0.051 0.031 0.030 0.049 0.047


Converged Y Y N Y Y N

154
Table B.2. cont’d
Dilation Angle 38 38 38 38 38 38
Flow Potential
1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03504 0.03942 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0823 0.0823 0.0720 0.0720 0.0720 0.0771


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete Model CDP CDP CDP CDP CDP CDP

Max Load (LPF) 0.619 0.637 0.457 0.457 0.453 0.467


Results

Deflection (in) 0.048 0.050 0.031 0.032 0.032 0.032


Converged N N N Y Y Y

155
Table B.2. cont’d
Dilation
38 38 38 38 38 38
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.03504 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0771 0.0771 0.0781 0.0781 0.0781 0.0802


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.465 0.603 0.649 0.631 0.466 1.37
(LPF)
Results

Deflection
0.032 0.049 0.051 0.050 0.032 0.128
(in)
Converged Y Y Y Y Y N

156
Table B.2. cont’d
Dilation
38 38 38 38 38 38
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.03504 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0802 0.0802 0.0761 0.0761 0.0761 0.0766


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.634 0.617 0.463 0.462 0.611 0.562
(LPF)
Results

Deflection
0.050 0.048 0.031 0.032 0.049 0.052
(in)
Converged N N N Y N N

157
Table B.2. cont’d
Dilation
38 38 38 38 38 38
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.03504 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0766 0.0766 0.0776 0.0776 0.0776 0.0779


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.464 0.462 0.572 0.465 0.613 0.581
(LPF)
Results

Deflection
0.032 0.032 0.043 0.032 0.049 0.044
(in)
Converged Y Y Y Y Y Y

158
Table B.2. cont’d
Dilation
38 38 38 38 38 38
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.03504 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0779 0.0779 0.0777 0.0777 0.0777 0.0775


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.659 0.707 0.567 0.466 0.613 0.647
(LPF)
Results

Deflection
0.053 0.057 0.043 0.032 0.049 0.051
(in)
Converged Y Y Y Y Y Y

159
Table B.2. cont’d
Dilation
38 38 38 38 38 38
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.03504 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0775 0.0775 0.0773 0.0773 0.0773 0.0769


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.464 0.576 0.468 0.683 0.464 0.466
(LPF)
Results

Deflection
0.032 0.045 0.032 0.054 0.032 0.032
(in)
Converged Y Y Y Y Y Y

160
Table B.2. cont’d
Dilation
38 38 38 38 38 38
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.03504 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0769 0.0769 0.0767 0.0767 0.0767 0.0765


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 26100
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 252.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.463 0.463 0.567 0.465 0.463 0.562
(LPF)
Results

Deflection
0.032 0.032 0.043 0.032 0.032 0.043
(in)
Converged Y Y Y N Y Y

161
Table B.2. cont’d
Dilation
38 38 38 38 38 40
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.03504 0.0438 0.03942 0.03504 0.0438
Input

As (in2) 0.0765 0.0765 0.0763 0.0763 0.0763 0.1542


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 26100 26100 26100 26100 26100 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1133.0 3147.3
f't (psi) 252.5 252.5 252.5 252.5 252.5 378.7
Ec (psi) 1918648 1918648 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.462 0.461 0.466 0.463 0.461 0.919
(LPF)
Results

Deflection
0.032 0.032 0.032 0.032 0.032 0.054
(in)
Converged Y Y N Y Y Y

162
Table B.2. cont’d
Dilation
40 45 40 40 40 40
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438 0.0438
Input

As (in2) 0.1542 0.1542 0.1542 0.1542 0.1542 0.2056


E's (ksi) 29000 29000 29000 30000 29000 29000
Es (ksi) 29000 29000 30000 30000 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 3147.3 3147.3 3147.3 3147.3 1133.0 1133.0
f't (psi) 378.7 378.7 378.7 378.7 252.5 252.5
Ec (psi) 2238422 2238422 2238422 2238422 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
1.00 1.01 1.03 1.03 0.717 0.694
(LPF)
Results

Deflection
0.056 0.056 0.057 0.057 0.045 0.040
(in)
Converged Y Y Y Y N Y

163
Table B.2. cont’d
Dilation
40 40 40 40 45 45
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03942 0.0438 0.03504 0.0438 0.0438 0.03504
Input

As (in2) 0.0761 0.1542 0.0771 0.0766 0.0766 0.0771


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 29000 29000 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1542.9 1133.0 1133.0 1133.0 1133.0
f't (psi) 252.5 294.5 252.5 252.5 252.5 252.5
Ec (psi) 1918648 2238422 1918648 1918648 1918648 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.640 0.825 0.624 0.658 0.485 0.484
(LPF)
Results

Deflection
0.049 0.046 0.048 0.051 0.033 0.033
(in)
Converged Y Y Y Y Y N

164
Table B.2. cont’d
Dilation
41 42 43 44 38 38
Angle
Flow
Potential 1 1 1 1 1 1
Eccentricity
σb0/σc0 1.12 1.12 1.12 1.12 1.12 1.12
Tension
Y Y Y Y Y Y
Coefficient
Compression
Y Y Y Y Y Y
Coefficient
A's (in2) 0.03504 0.03504 0.03504 0.03504 0.0438 0.0438
Input

As (in2) 0.0771 0.0771 0.0771 0.0771 0.2056 0.2056


E's (ksi) 29000 29000 29000 29000 29000 29000
Es (ksi) 29000 29000 29000 29000 29000 29000
f'y (psi) 47000 47000 47000 47000 47000 47000
fy (psi) 55000 55000 55000 55000 55000 55000
f'c (psi) 1133.0 1133.0 1133.0 1133.0 1542.2 1133.0
f't (psi) 252.5 252.5 252.5 252.5 294.5 252.5
Ec (psi) 1918648 1918648 1918648 1918648 2238422 1918648
Concrete
CDP CDP CDP CDP CDP CDP
Model
Max Load
0.624 0.624 0.625 0.482 0.798 0.689
(LPF)
Results

Deflection
0.048 0.048 0.048 0.033 0.041 0.039
(in)
Converged N Y Y Y Y Y

165
Appendix C

COMPLETE RESULTS FOR 3-D BEAM CALIBRATION VARIATIONS

Table C.1. Results of mesh sensitivity analysis for 3-D beam with 2-D rebar models
utilizing the brittle cracking technique.
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Direct
Cracking 0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Direct
Cracking 0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Failure Strain
Input

Crack Opening
0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Load Rate
5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
(lb/s)
Rebar Size
10 100 5 10 20 50
(mm)
Mesh Size
5 5 10 10 10 10
(mm)
Last Step
Results

0.40 0.25 0.45 0.45 0.45 0.40


Before Failure
Deflection (in) 0.0384 0.0181 0.0282 0.0279 0.0289 0.0266

166
Table C.1. cont’d
f't (psi) 420.8 420.8 420.8 420.8 420.8
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027
Failure Strain
Input

Crack Opening
0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Load Rate (lb/s) 5901.39 5901.39 5901.39 5901.39 5901.39
Rebar Size
100 100 100 100 100
(mm)
Mesh Size (mm) 10 20 25 50 100
Last Step Before
0.35 0.60 0.45 1.50 2.25
Results

Failure
Deflection (in) 0.0209 0.0520 0.2028 0.1876 0.1230

Table C.2. Results of mesh sensitivity analysis for 3-D beam with 2-D rebar models
utilizing the CDP and SC techniques.
Tension
N/A 0.0027 N/A 0.0027 N/A 0.0027
Stiffening Strain
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
Input

f't (psi) 420.8 N/A 420.8 N/A 420.8 N/A


Model Type CDP SC CDP SC CDP SC
Rebar Size (mm) 5 5 10 10 100 100
Mesh Size (mm) 5 5 5 5 5 5
Results

Max Load (LPF) 0.321 0.171 0.321 0.321 0.29 0.165

Converged N N N N N N

167
Table C.2. cont’d
Tension Stiffening
N/A 0.0027 N/A 0.0027 N/A 0.0027
Strain
f'c (psi) 3147.3 3147 3147.3 3147 3147.3 3147
Input

f't (psi) 420.8 N/A 420.8 N/A 420.8 N/A


Model Type CDP SC CDP SC CDP SC
Rebar Size (mm) 5 5 10 10 20 20
Mesh Size (mm) 10 10 10 10 10 10
Results

Max Load (LPF) 0.334 0.287 0.341 0.287 0.337 0.23


Converged N N N N N Y

Table C.2. cont’d


Tension Stiffening
N/A 0.0027 N/A 0.0027 N/A 0.0027
Strain
f'c (psi) 3147.3 3147 3147.3 3147 3147.3 3147
Input

f't (psi) 420.8 N/A 420.8 N/A 420.8 N/A


Model Type CDP SC CDP SC CDP SC
Rebar Size (mm) 50 50 100 100 100 100
Mesh Size (mm) 10 10 10 10 20 20
Results

Max Load (LPF) 0.339 0.234 0.146 0.234 0.354 0.395


Converged N N Y Y Y N

Table C.2. cont’d


Tension Stiffening
N/A 0.0027 N/A 0.0027 N/A 0.0027
Strain
f'c (psi) 3147.3 3147 3147.3 3147 3147.3 3147
Input

f't (psi) 420.8 N/A 420.8 N/A 420.8 N/A


Model Type CDP SC CDP SC CDP SC
Rebar Size (mm) 100 100 100 100 100 100
Mesh Size (mm) 25 25 50 50 100 100
Results

Max Load (LPF) 0.323 0.434 0.65 1.56 1.25 1.67


Converged N N N N N N

168
Table C.3. Results for 3-D beam with 2-D rebar models utilizing the brittle cracking
techniques.
Load Rate (lb/s) 11240 11240 11240 11240 11240
As (in2) 0.20563 0.20563 0.20563 0.20563 0.20563
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438
f't (psi) 210.4 210.4 210.4 105.2 105.2
Input

Direct Cracking
0.0027 0.002025 0.00135 0.0027 0.0054
Strain
Direct Cracking
0.0027 0.002025 0.00135 0.0027 0.0054
Failure Strain
Crack Opening
0.0027 0.002025 0.00135 0.0027 0.0054
Strain
Mesh Size (mm) 25 25 25 25 25
1st Removed Element 0.150 0.125 0.125 0.100 0.150
Results

Max Load (LPF) 0.400 0.400 0.375 0.400 0.425


Deflection (in) 0.342 0.404 0.347 0.387 0.563

Table C.3. cont’d


Load Rate (lb/s) 11240 11240 11240 11240 11240
As (in2) 0.20563 0.20563 0.20563 0.20563 0.20563
A's (in2) 0.0438 0.0438 0.0438 0.0438 0.0438
f't (psi) 52.6 105.2 105.2 105.2 210.4
Input

Direct Cracking
0.0054 0.0054 0.0054 0.0054 0.00135
Strain
Direct Cracking
0.0054 0.0054 0.108 0.0108 0.0054
Failure Strain
Crack Opening Strain 0.0054 0.0108 0.0054 0.0108 0.00135
Mesh Size (mm) 25 25 25 25 25
1st Removed Element 0.125 0.225 0.150 0.225 0.125
Results

Max Load (LPF) 0.400 0.425 0.425 0.450 0.400


Deflection (in) 0.345 0.476 0.507 0.546 0.418

169
Table C.3. cont’d
Load Rate (lb/s) 11240 11802.78 11240 11240
As (in2) 0.10287 0.20563 0.20563 0.20563
A's (in2) 0.02117 0.0438 0.0438 0.0438
f't (psi) 420.8 420.8 420.8 420.8
Input

Direct Cracking
0.0027 0.0027 0.0027 0.00119
Strain
Direct Cracking
0.0027 0.0027 0.0027 0.00119
Failure Strain
Crack Opening Strain 0.0027 0.0027 0.0027 0.00119
Mesh Size (mm) 25 10 25 25
1st Removed Element 0.225 N/A 0.25 0.2
Results

Max Load (LPF) 0.225 0.45 0.4 0.2


Deflection (in) 0.111 0.143 0.366 0.035

Table C.4. Results for 3-D beam with 3-D rebar models utilizing the CDP techniques.
Tension Damage Y Y Y N N N
Compression Damage Y Y Y N N N
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Hard Hard Hard Hard Hard Hard
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 0.5 1.5 N/A 1.0 0.5 1.5
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
1 1 1 1 1 1
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.135 0.145 0.238 0.319 0.142 0.145
Results

Deflection (in) 0.010 0.010 0.021 0.025 0.011 0.010


Converged N Y N N N N

170
Table C.4. cont’d
Tension Damage N Y Y Y Y N
Compression Damage N Y Y Y Y N
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Hard Hard Hard Hard Hard Hard
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient N/A 1.0 0.5 1.5 N/A 1.0
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
1 N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.258 0.142 0.135 0.145 0.122 0.319
Results

Deflection (in) 0.023 0.010 0.010 0.010 0.010 0.025


Converged Y Y N Y N N

Table C.4. cont’d


Tension Damage N N N Y Y Y
Compression Damage N N N Y Y Y
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Hard Hard Hard Penalty Exp Hard
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 0.5 1.5 N/A 1.0 Exp 1.0
MPC Slider
2 2 2 2 2 1
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.142 0.305 0.124 0.134 0.103 0.140
Results

Deflection (in) 0.011 0.023 0.010 0.010 0.007 0.010


Converged N Y N N Y N

171
Table C.4. cont’d
Tension Damage Y Y Y Y Y N
Compression Damage Y Y Y Y Y N
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Penalty Hard Hard Hard Hard Hard
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 1.0 0.5 1.0 1.5 N/A 0.5
MPC Slider
1 4 4 4 4 4
Nodes/Cross-Section
MPC Beam
N/A 1 1 1 1 1
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.136 0.111 0.115 0.116 0.118 0.135
Results

Deflection (in) 0.010 0.008 0.007 0.007 0.009 0.010


Converged Y Y Y Y Y Y

Table C.4. cont’d


Tension Damage N N N Y Y Y
Compression Damage N N N Y Y Y
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Hard Hard Hard Penalty Penalty Penalty
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 1.0 1.5 N/A 0.5 1.0 1.5
MPC Slider
4 4 4 2 2 2
Nodes/Cross-Section
MPC Beam
1 1 1 N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.139 0.141 0.239 0.131 0.134 0.149
Results

Deflection (in) 0.009 0.009 0.020 0.010 0.010 0.011


Converged Y Y Y N N N

172
Table C.4. cont’d
Tension Damage Y N N N N Y
Compression Damage Y N N N N Y
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Penalty Penalty Penalty Penalty Penalty Exp
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient N/A 0.5 1.0 1.5 N/A 0.5
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.125 0.251 0.144 0.111 0.133 0.139
Results

Deflection (in) 0.010 0.021 0.011 0.008 0.011 0.010


Converged N Y Y Y Y Y

Table C.4. cont’d


Tension Damage Y Y Y N N N
Compression Damage Y Y Y N N N
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Exp Exp Exp Exp Exp Exp
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 1.0 1.5 N/A 0.5 1.0 1.5
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.103 0.104 0.084 0.119 0.121 0.117
Results

Deflection (in) 0.007 0.007 0.007 0.008 0.008 0.008


Converged Y Y N N N N

173
Table C.4. cont’d
Tension Damage N Yes Yes Yes Yes Yes
Compression Damage N Yes Yes Yes Yes Yes
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Exp Exp Penalty Hard Hard Hard
Input

No. Surfaces in
2 1 2 2 2 2
Contact
Friction Coefficient N/A Exp 1.0 1.0 0.5 1.5
MPC Slider
2 2 1 1 2 2
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.088 0.103 0.136 0.140 0.135 0.145
Results

Deflection (in) 0.007 0.007 0.010 0.010 0.010 0.010


Converged N N Y N N Y

Table C.4. cont’d


Tension Damage Yes Yes No No No No
Compression Damage Yes Yes No No No No
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Hard Hard Hard Hard Hard Hard
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 0.5 1.5 0.5 1.5 N/A 1.0
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
1 1 1 1 1 1
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.135 0.145 0.142 0.145 0.124 0.319
Results

Deflection (in) 0.010 0.010 0.011 0.010 0.010 0.025


Converged N Y N N Y N

174
Table C.4. cont’d
Tension Damage Yes No No No No Yes
Compression Damage Yes No No No No Yes
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Hard Hard Hard Hard Hard Hard
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient N/A 0.5 1.5 N/A 1.0 N/A
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
1 N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.238 0.142 0.150 0.124 0.319 0.122
Results

Deflection (in) 0.021 0.011 0.011 0.010 0.025 0.010


Converged N N Y N N N

Table C.4. cont’d


Tension Damage Yes Yes Yes Yes Yes Yes
Compression Damage Yes Yes Yes Yes Yes Yes
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Penalty Exp Hard Hard Hard Hard
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 1.0 1.0 1.0 0.5 1.0 1.5
MPC Slider
2 2 2 4 4 4
Nodes/Cross-Section
MPC Beam
N/A N/A N/A 1 1 1
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.134 0.103 0.142 0.110 0.086 0.085
Results

Deflection (in) 0.010 0.007 0.010 0.008 0.007 0.007


Converged N Y Y Y Y Y

175
Table C.4. cont’d
Tension Damage Yes No No No No Yes
Compression Damage Yes No No No No Yes
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Hard Hard Hard Hard Hard N/A
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient N/A 0.5 1.5 N/A 1.0 1.0
MPC Slider
4 4 4 4 4 2
Nodes/Cross-Section
MPC Beam
1 1 1 1 1 N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.119 0.135 0.141 0.093 0.139 0.142
Results

Deflection (in) 0.009 0.010 0.009 0.020 0.009 0.010


Converged Y Y Y Y Y Y

Table C.4. cont’d


Tension Damage Yes Yes Yes Yes Yes Yes
Compression Damage Yes Yes Yes Yes Yes Yes
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Exp Exp Exp Exp Exp Exp
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient Exp 0.5 1.0 1.5 Exp 0.5
MPC Slider
N/A 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.062 0.139 0.125 0.104 0.103 0.139
Results

Deflection (in) 0.005 0.010 0.010 0.007 0.007 0.010


Converged N Y Y Y Y Y

176
Table C.4. cont’d
Tension Damage Yes Yes Yes No No No
Compression Damage Yes Yes Yes No No No
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Exp Exp Exp Exp Exp Exp
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient 1.0 1.5 N/A 0.5 1.0 1.5
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.103 0.104 0.084 0.119 0.121 0.117
Results

Deflection (in) 0.007 0.007 0.007 0.008 0.008 0.008


Converged Y Y N N N N

Table C.4. cont’d


Tension Damage No Yes Yes Yes Yes Yes
Compression Damage No Yes Yes Yes Yes Yes
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Pressure-Overclosure Exp Exp Penalty Penalty Penalty Penalty
Input

No. Surfaces in
2 2 2 2 2 2
Contact
Friction Coefficient N/A Exp 0.5 1.0 1.5 N/A
MPC Slider
2 2 2 2 2 2
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP CDP CDP
Max Load (LPF) 0.088 0.103 0.131 0.134 0.149 0.125
Results

Deflection (in) 0.007 0.007 0.010 0.010 0.011 0.010


Converged N Y N N N N

177
Table C.4. cont’d
Tension Damage No No No No
Compression Damage No No No No
Applied Load (lb) 5901.4 5901.4 5901.4 5901.4
f'c (psi) 3147.3 3147.3 3147.3 3147.3
f't (psi) 420.8 420.8 420.8 420.8
Pressure-Overclosure Penalty Penalty Penalty Penalty
Input

No. Surfaces in Contact 2 2 2 2


Friction Coefficient 0.5 1.0 1.5 N/A
MPC Slider
2 2 2 2
Nodes/Cross-Section
MPC Beam
N/A N/A N/A N/A
Nodes/Rebar
Concrete Model CDP CDP CDP CDP
Max Load (LPF) 0.140 0.144 0.111 0.133
Results

Deflection (in) 0.011 0.011 0.008 0.011


Converged Y Y Y Y

Table C.5. Results for 3-D beam with 3-D rebar models utilizing the SC technique.
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
Friction Coefficient 0.5 1.0 1.5 N/A 0.5 1.0
Pressure-Overclosure Hard Hard Hard Hard Hard Hard
Input

Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4


MPC Slider
4 4 4 4 2 2
Nodes/Cross-Section
MPC Beam
1 1 1 1 1 1
Nodes/Rebar
Max Load (LPF) 0.376 0.391 0.393 0.286 0.340 0.355
Results

Deflection (in) 0.025 0.026 0.025 0.021 0.024 0.024


Converged N Y N N N Y

178
Table C.5. cont’d
f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147.3 3147.3
Friction Coefficient 1.5 N/A 0.5 1.0 1.5 N/A
Pressure-
Hard Hard Penalty Penalty Penalty Penalty
Overclosure
Input

Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.4 5901.4


MPC Slider
Nodes/Cross- 2 2 2 2 2 2
Section
MPC Beam
1 1 1 1 1 1
Nodes/Rebar
Max Load (LPF) 0.325 0.273 0.273 0.277 0.339 0.296
Results

Deflection (in) 0.021 0.020 0.019 0.019 0.024 0.023


Converged N N N N N N

Table C.5. cont’d


f'c (psi) 3147.3 3147.3 3147.3 3147.3 3147 3147
Friction Coefficient 0.5 1.0 1.5 N/A 0.5 1.0
Pressure-
Exp Exp Exp Exp Hard Hard
Overclosure
Input

Applied Load (lb) 5901.4 5901.4 5901.4 5901.4 5901.39 5901.39


MPC Slider
Nodes/Cross- 2 2 2 2 2 2
Section
MPC Beam
1 1 1 1 1 1
Nodes/Rebar
Max Load (LPF) 0.276 0.350 0.328 0.247 0.340 0.355
Results

Deflection (in) 0.018 0.022 0.021 0.018 0.024 0.024


Converged N N N N N Y

179
Table C.5. cont’d
f'c (psi) 3147 3147 3147 3147 3147 3147
Friction
1.5 N/A 0.5 1.0 1.5 N/A
Coefficient
Pressure-
Hard Hard Hard Hard Hard Hard
Overclosure
Input

Applied Load (lb) 5901.39 5901.39 5901.39 5901.39 5901.39 5901.39


MPC Slider
Nodes/Cross- 2 2 4 4 4 4
Section
MPC Beam
1 1 1 1 1 1
Nodes/Rebar
Max Load (LPF) 0.325 0.273 0.376 0.391 0.393 0.286
Results

Deflection (in) 0.021 0.020 0.025 0.026 0.025 0.021


Converged N N N Y N N

Table C.5. cont’d


f'c (psi) 3147 3147 3147 3147 3147
Friction Coefficient 0.5 1.0 1.5 N/A 0.5
Pressure-
Exp Exp Exp Exp Penalty
Overclosure
Input

Applied Load (lb) 5901.39 5901.39 5901.39 5901.39 5901.39


MPC Slider
2 2 2 2 2
Nodes/Cross-Section
MPC Beam
1 1 1 1 1
Nodes/Rebar
Max Load (LPF) 0.276 0.350 0.328 0.247 0.273
Results

Deflection (in) 0.018 0.022 0.021 0.018 0.019


Converged N N N N N

180
Table C.5. cont’d
f'c (psi) 3147 3147 3147
Friction Coefficient 1.0 1.5 N/A
Pressure-Overclosure Penalty Penalty Penalty
Input

Applied Load (lb) 5901.39 5901.39 5901.39


MPC Slider
2 2 2
Nodes/Cross-Section
MPC Beam
1 1 1
Nodes/Rebar
Max Load (LPF) 0.277 0.339 0.296
Results

Deflection (in) 0.019 0.024 0.023


Converged N N N

Table C.6. Results for 3-D beam with 3-D rebar models utilizing the brittle cracking
technique.
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Crack Opening
0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Failure Strain
Input

Load Rate (lb/s) 5901.4 5901.4 5901.4 5901.4 5901.4 4426.0


No. Surfaces in
2 2 2 2 2 2
Contact
Mechanical
Penalty Penalty Penalty Penalty Penalty Penalty
Constraint
Pressure-
Exp Exp N/A N/A Exp Exp
Overclosure
Friction Exp 1 Exp 1 Exp Exp
Results

Max Load (LPF) 0.75 0.65 0.25 0.25 0.75 0.45

Deflection (in) 0.106 0.053 0.019 0.019 0.106 0.035

181
Table C.6. cont’d
f't (psi) 420.8 420.8 420.8 420.8 420.8 420.8
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Crack Opening
0.0027 0.0027 0.0027 0.027 0.0027 0.0027
Strain
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027 0.0027
Failure Strain
Input

Load Rate (lb/s) 11802.8 5901.4 590.1 590.1 98.4 98.4


No. Surfaces in
2 2 2 2 2 2
Contact
Mechanical
Penalty Penalty Penalty Penalty Penalty Penalty
Constraint
Pressure-
Exp Exp Exp Exp Exp Exp
Overclosure
Friction Exp Exp Exp Exp Exp Exp
Results

Max Load (LPF) 0.60 0.45 0.46 0.44 0.27 0.48

Deflection (in) 0.071 0.031 0.037 0.030 0.018 0.032

Table C.6. cont’d


f't (psi) 420.8 420.8 420.8 420.8 420.8
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Crack Opening
0.0027 0.0027 0.0027 0.0027 0.0027
Strain
Direct Cracking
0.0027 0.0027 0.0027 0.0027 0.0027
Failure Strain
Input

Load Rate (lb/s) 19.7 19.7 590.1 590.1 590.1


No. Surfaces in
2 2 2 2 2
Contact
Mechanical
Penalty Penalty Penalty Penalty Penalty
Constraint
Pressure-
Exp Exp Exp N/A N/A
Overclosure
Friction Exp Exp 1 Exp 1
Results

Max Load (LPF) 0.07 0.07 0.57 0.29 0.29

Deflection (in) 0.006 0.005 0.043 0.020 0.020

182
Appendix D

PERMISSION LETTERS

To whom it may concern:

I give Diane Wurst permission to use Figures 2.1 and 2.9 from my thesis, "Field
Measurements and Corresponding FEA of Cross-frame Forces in Skewed Steel I-
girder Bridges".

Sincerely,
Kelly L. Ambrose
(302)-528-2263
1405 Riverside Ave.
Baltimore, MD 21230

183
184
185
186
187
188

Você também pode gostar