reliability of corroding reinforced concrete structures

122
Louisiana State University LSU Digital Commons LSU Master's eses Graduate School July 2019 Reliability of Corroding Reinforced Concrete Structures Travis Honore Louisiana State University and Agricultural and Mechanical College, [email protected] Follow this and additional works at: hps://digitalcommons.lsu.edu/gradschool_theses Part of the Structural Engineering Commons is esis is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Master's eses by an authorized graduate school editor of LSU Digital Commons. For more information, please contact [email protected]. Recommended Citation Honore, Travis, "Reliability of Corroding Reinforced Concrete Structures" (2019). LSU Master's eses. 4979. hps://digitalcommons.lsu.edu/gradschool_theses/4979

Upload: others

Post on 09-Apr-2022

6 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Reliability of Corroding Reinforced Concrete Structures

Louisiana State UniversityLSU Digital Commons

LSU Master's Theses Graduate School

July 2019

Reliability of Corroding Reinforced ConcreteStructuresTravis HonoreLouisiana State University and Agricultural and Mechanical College, [email protected]

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_theses

Part of the Structural Engineering Commons

This Thesis is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSUMaster's Theses by an authorized graduate school editor of LSU Digital Commons. For more information, please contact [email protected].

Recommended CitationHonore, Travis, "Reliability of Corroding Reinforced Concrete Structures" (2019). LSU Master's Theses. 4979.https://digitalcommons.lsu.edu/gradschool_theses/4979

Page 2: Reliability of Corroding Reinforced Concrete Structures

RELIABILITY OF CORRODING REINFORCED CONCRETE STRUCTURES

A Thesis

Submitted to the Graduate Faculty of the Louisiana State University and

Agricultural and Mechanical College in partial fulfillment of the

requirements for the degree of Master of Science

in

The Department of Civil and Environmental Engineering

by Travis Joseph Honore’

B.S., Louisiana State University, 2017

August 2019

Page 3: Reliability of Corroding Reinforced Concrete Structures

ii

ACKNOWLEDGEMENTS

First, I would like to recognize my family who has provided me with unbelievable

support and patience through the duration of this research and completing my educational

studies. I would also like to sincerely thank my advisor, Dr. Ayman Okeil, for the

consistent guidance and supervision through this process as well as the mentorship he has

provided through my undergraduate and graduate studies. Lastly, another special thanks

to my committee members: Dr. Mostafa Elseifi, Dr. Chao Sun, and Dr. Steve Cai. The

research described in this thesis was financially supported and sponsored by the

Transportation Consortium of South Central States (Tran-SET Project # 17STLSU03).

Additional support from the Louisiana Transportation Research Center and the

Department of Civil and Environmental Engineering at Louisiana State University is

greatly acknowledged. Any opinions, findings, and conclusions or recommendations

expressed in this material are those of the authors and do not necessarily reflect the views

of the sponsoring agencies. Additional efforts were given from researchers at Texas

A&M University and University of New Mexico to further enhance the exploration of

this topic. Their work from laboratory support, field results and other technical expertise

is also very much appreciated.

Page 4: Reliability of Corroding Reinforced Concrete Structures

iii

TABLE OF CONTENTS ACKNOWLEDGEMENTS ................................................................................................ ii

LIST OF TABLES ............................................................................................................. iv

LIST OF FIGURES ............................................................................................................ v

ABSTRACT ....................................................................................................................... ix

1. INTRODUCTION .......................................................................................................... 1 1.1. Introduction .................................................................................................................. 1 1.2. Scope of Work ............................................................................................................. 2 1.3. Research Approach ...................................................................................................... 3 1.4. Objective ...................................................................................................................... 3 1.5. Organization ................................................................................................................. 4

2. LITERATURE REVIEW ............................................................................................... 5 2.1. Corrosion ..................................................................................................................... 5 2.2. Detection and protection techniques against corrosion ............................................. 11 2.3. Structural Reliability and Probabilistic Evaluation ................................................... 13 2.4. Reliability Analysis .................................................................................................... 19

3. METHODOLOGY ....................................................................................................... 27 3.1. Design Space .............................................................................................................. 27 3.2. Design Approach ....................................................................................................... 29 3.3. Reliability Analysis .................................................................................................... 42

4. RESULTS ..................................................................................................................... 45 4.1. Time to Corrosion Initiation ...................................................................................... 45 4.2. Moment values for each Dead to Live Ratio ............................................................. 51 4.3. General Corrosion ...................................................................................................... 53 4.4. Pitting Corrosion ........................................................................................................ 78

5. SUMMARY AND CONCLUSION ........................................................................... 103 5.1. Summary .................................................................................................................. 103 5.2. Conclusions .............................................................................................................. 103 5.3. Recommendations for Future Research ................................................................... 103

REFERENCES ............................................................................................................... 105

VITA ............................................................................................................................... 112

Page 5: Reliability of Corroding Reinforced Concrete Structures

iv

LIST OF TABLES

Table 1. Chloride diffusion model parameters and statistics...................................................... 31

Table 2. Water to cement ratio correction factor ............................................................................ 32

Table 3. Curing day correction factor ................................................................................................ 32

Table 4. Interpolated chloride threshold values .............................................................................. 33

Table 5. Load modifiers and factors for LRFD design equation .............................................. 41

Table 6. Statistical characteristics of random variables for each design space .................... 44

Table 7. Moment loads for design space 1 ....................................................................................... 52

Table 8. Moment loads for design space 2 ....................................................................................... 52

Table 9. Moment loads for design space 3 ....................................................................................... 53

Table 10. Pitting factor for each design space ................................................................................. 79

Page 6: Reliability of Corroding Reinforced Concrete Structures

v

LIST OF FIGURES

Figure 1. A schematic illustration of steel corrosion cell in concrete (Markeset and Mydral 2008) ....................................................................................................................................... 6

Figure 2. Definitions for chloride thresholds (Angst et al. 2009) ............................................. 10

Figure 3. PDF’s of resistance, load effects, and the safety margin (Nowak and Collins, 2000) .................................................................................................................................................... 20

Figure 4. Reliability index defined as the shortest distance in the space of reduced variables (Nowak and Collins, 2000) ....................................................................................... 21

Figure 5. Relationship between β and Pf for a normally distributed function (Allen et al., 2005) .................................................................................................................................................... 22

Figure 6. Details of typical bridge used to cover design space (Okeil et al, 2013) ............. 27

Figure 7. Typical girder for design space 1 (Okeil et al, 2013) ................................................. 28

Figure 8. Typical girder for design space 2 (Okeil et al, 2013) ................................................. 28

Figure 9. Typical girder for design space 3 (Okeil et. al, 2013) ................................................ 29

Figure 10. Pit configuration (Val, 2007) ........................................................................................... 38

Figure 11. Pit configuration (Darmawan, 2010) ............................................................................. 38

Figure 12. Time to Corrosion Initiation design space 1 with 12 centimeter cover ............. 46

Figure 13. Time to Corrosion Initiation design space 1 with 15 centimeter cover ............. 47

Figure 14. Time to Corrosion Initiation design space 1 with 18 centimeter cover ............. 48

Figure 15. Time to Corrosion Initiation design space 2 with 11 centimeter cover ............. 48

Figure 16. Time to Corrosion Initiation design space 2 with 13 centimeter cover ............. 49

Figure 17. Time to Corrosion Initiation design space 2 with 15 centimeter cover ............. 49

Figure 18. Time to Corrosion Initiation design space 3 with 10 centimeter cover ............. 50

Figure 19. Time to Corrosion Initiation design space 3 with 12 centimeter cover ............. 50

Figure 20. Time to Corrosion Initiation design space 3 14 with centimeter cover ............. 51

Figure 21. Reliability Index vs. Time design space 1 WC46C12Cl35 ................................... 54

Figure 22. Reliability Index vs. Time design space 1 WC50C12Cl35 ................................... 55

Figure 23. Reliability Index vs. Time design space 1 WC53C12Cl35 ................................... 55

Figure 24. Reliability Index vs. Time design space 1 WC46C15Cl35 ................................... 56

Figure 25. Reliability Index vs. Time design space 1 WC50C15Cl35 ................................... 56

Figure 26. Reliability Index vs. Time design space 1 WC53C15Cl35 ................................... 57

Figure 27. Reliability Index vs. Time design space 1 WC46C18Cl35 ................................... 57

Page 7: Reliability of Corroding Reinforced Concrete Structures

vi

Figure 28. Reliability Index vs. Time design space 1 WC50C18Cl35 ................................... 58

Figure 29. Reliability Index vs. Time design space 1 WC53C18Cl35 ................................... 58

Figure 30. Reliability Index vs. Time design space 2 cover WC46C11Cl30 ....................... 59

Figure 31. Reliability Index vs. Time design space 2 WC50C11Cl30 ................................... 59

Figure 32. Reliability Index vs. Time design space 2 WC53C11Cl30 ................................... 60

Figure 33. Reliability Index vs. Time design space 2 WC46C13Cl30 ................................... 60

Figure 34. Reliability Index vs. Time design space 2 WC50C13Cl30 ................................... 61

Figure 35. Reliability Index vs. Time design space 2 WC53C13Cl30 ................................... 61

Figure 36. Reliability Index vs. Time design space 2 WC46C15Cl30 ................................... 62

Figure 37. Reliability Index vs. Time design space 2 WC50C15Cl30 ................................... 62

Figure 38. Reliability Index vs. Time design space 2 WC53C15Cl30 ................................... 63

Figure 39. Reliability Index vs. Time design space 3 WC46C10Cl30 ................................... 63

Figure 40. Reliability Index vs. Time design space 3 WC50C10Cl30 ................................... 64

Figure 41. Reliability Index vs. Time design space 3 WC53C10Cl30 ................................... 64

Figure 42. Reliability Index vs. Time design space 3 WC46C12Cl30 ................................... 65

Figure 43. Reliability Index vs. Time design space 3 WC50C12Cl30 ................................... 65

Figure 44. Reliability Index vs. Time design space 3 WC53C12Cl30 ................................... 66

Figure 45. Reliability Index vs. Time design space 3 WC46C14Cl30 ................................... 66

Figure 46. Reliability Index vs. Time design space 3 WC50C14Cl30 ................................... 67

Figure 47. Reliability Index vs. Time design space 3 WC53C14Cl30 ................................... 67

Figure 48. Reliability Index vs. Concrete Cover design space 1 at T=0 years ..................... 68

Figure 49. Reliability Index vs. Concrete Cover design space 1 at T=10 years .................. 68

Figure 50. Reliability Index vs. Concrete Cover design space 1 at T=25 years .................. 69

Figure 51. Reliability Index vs. Concrete Cover design space 1 at T=50 years .................. 69

Figure 52. Reliability Index vs. Concrete Cover design space 1 at T=75 years .................. 70

Figure 53. Reliability Index vs. Concrete Cover design space 2 at T=0 years ..................... 70

Figure 54. Reliability Index vs. Concrete Cover design space 2 at T=10 years .................. 71

Figure 55. Reliability Index vs. Concrete Cover design space 2 at T=25 years .................. 71

Figure 56. Reliability Index vs. Concrete Cover design space 2 at T=50 years .................. 72

Figure 57. Reliability Index vs. Concrete Cover design space 2 at T=75 years .................. 72

Figure 58. Reliability Index vs. Concrete Cover design space 3 at T=0 years ..................... 73

Page 8: Reliability of Corroding Reinforced Concrete Structures

vii

Figure 59. Reliability Index vs. Concrete Cover design space 3 at T=10 years .................. 73

Figure 60. Reliability Index vs. Concrete Cover design space 3 at T=25 years .................. 74

Figure 61. Reliability Index vs. Concrete Cover design space 3 at T=50 years .................. 74

Figure 62. Reliability Index vs. Concrete Cover design space 3 at T=75 years .................. 75

Figure 63. Time to Reliability Index Threshold design space 1 ................................................ 76

Figure 64. Time to Reliability Index Threshold design space 2 ................................................ 77

Figure 65. Time to Reliability Index Threshold design space 3 ................................................ 77

Figure 66. PDF vs. Location Parameter of Pitting Factor ........................................................... 78

Figure 67. Reliability Index vs. Time design space 1 WC46C12Cl35 ................................... 80

Figure 68. Reliability Index vs. Time design space 1 WC50C12Cl35 ................................... 80

Figure 69. Reliability Index vs. Time design space 1 WC53C12Cl35 ................................... 81

Figure 70. Reliability Index vs. Time design space 1 WC46C15Cl35 ................................... 81

Figure 71. Reliability Index vs. Time design space 1 WC50C15Cl35 ................................... 82

Figure 72. Reliability Index vs. Time design space WC53C15Cl35 ....................................... 82

Figure 73. Reliability Index vs. Time design space 1 cover WC46C18Cl35 ....................... 83

Figure 74. Reliability Index vs. Time design space 1 WC50C18Cl35 ................................... 83

Figure 75. Reliability Index vs. Time design space 1 WC53C18Cl35 ................................... 84

Figure 76. Reliability Index vs. Time design space 2 WC46C11Cl30 ................................... 84

Figure 77. Reliability Index vs. Time design space 2 WC50C11Cl30 ................................... 85

Figure 78. Reliability Index vs. Time design space 2 WC53C11Cl30 ................................... 85

Figure 79. Reliability Index vs. Time design space 2 WC46C13Cl30 ................................... 86

Figure 80. Reliability Index vs. Time design space 2 WC50C13Cl30 ................................... 86

Figure 81. Reliability Index vs. Time design space 2 WC53C13Cl30 ................................... 87

Figure 82. Reliability Index vs. Time design space 2 WC46C15Cl30 ................................... 87

Figure 83. Reliability Index vs. Time design space 2 WC50C15Cl30 ................................... 88

Figure 84. Reliability Index vs. Time design space 2 WC53C15Cl30 ................................... 88

Figure 85. Reliability Index vs. Time design space WC46C10Cl30 ....................................... 89

Figure 86. Reliability Index vs. Time design space 3 WC50C10Cl30 ................................... 89

Figure 87. Reliability Index vs. Time design space 3 WC53C10Cl30 ................................... 90

Figure 88. Reliability Index vs. Time design space 3 WC46C12Cl30 ................................... 90

Figure 89. Reliability Index vs. Time design space 3 WC50C12Cl30 ................................... 91

Page 9: Reliability of Corroding Reinforced Concrete Structures

viii

Figure 90. Reliability Index vs. Time design space 3 WC53C12Cl30 ................................... 91

Figure 91. Reliability Index vs. Time design space 3 WC46C14Cl30 ................................... 92

Figure 92. Reliability Index vs. Time design space 3 WC50C14Cl30 ................................... 92

Figure 93. Reliability Index vs. Time design space 3 WC53C14Cl30 ................................... 93

Figure 94. Reliability Index vs. Concrete Cover design space 1 at T=0 years ..................... 93

Figure 95. Reliability Index vs. Concrete Cover design space 1 at T=10 years .................. 94

Figure 96. Reliability Index vs. Concrete Cover design space 1 at T=25 years .................. 94

Figure 97. Reliability Index vs. Concrete Cover design space 1 at T=50 years .................. 95

Figure 98. Reliability Index vs. Concrete Cover design space 1 at T=75 years .................. 95

Figure 99. Reliability Index vs. Concrete Cover design space 2 at T=0 years ..................... 96

Figure 100. Reliability Index vs. Concrete Cover design space 2 at T=10 years ................ 96

Figure 101. Reliability Index vs. Concrete Cover design space 2 at T=25 years ................ 97

Figure 102. Reliability Index vs. Concrete Cover design space 2 at T=50 years ................ 97

Figure 103. Reliability Index vs. Concrete Cover design space 2 at T=75 years ................ 98

Figure 104. Reliability Index vs. Concrete Cover design space 3 at T=0 years .................. 98

Figure 105. Reliability Index vs. Concrete Cover design space 3 at T=10 years ................ 99

Figure 106. Reliability Index vs. Concrete Cover design space 3 at T=25 years ................ 99

Figure 107. Reliability Index vs. Concrete Cover design space 3 at T=50 years ............. 100

Figure 108. Reliability Index vs. Concrete Cover design space 3 at T=75 years ............. 100

Figure 109. Time to Reliability Index Threshold design space 1 .......................................... 101

Figure 110. Time to Reliability Index Threshold design space 2 .......................................... 102

Figure 111. Time to Reliability Index Threshold design space 3 .......................................... 102

Page 10: Reliability of Corroding Reinforced Concrete Structures

ix

ABSTRACT

The enhanced performance due to the addition of steel within concrete elements

has thrust reinforced concrete structures to be one of the most commonly used

construction materials in the world. Reinforced concrete structures are exposed to a

variety of deterioration mechanisms during their operational existence, each of which

contributes to reducing their service life, durability, and load bearing capacity. The

longevity of these reinforced concrete structures is strongly connected to degradation

processes, whose origin is environmental and/or functional. Among these processes is the

initiation and evolution of corrosion on the steel reinforcement. Steel reinforcement has

direct effects on the durability, safety, and useful service life of a reinforced concrete

structure. The reliability of an engineering design is measured by the probability that a

design meets certain demands under certain conditions, or its resistance will be able to

sustain given loads. The main focus of this study is to develop a reliability-based service

life prediction model using the uncertainties characteristically recognized in the

parameters included in established corrosion detection processes. Reliability algorithms,

such as Advanced First Order Reliability Method (AFORM), were used to analyze the

effects of key factors within both uniform and localized corrosion conditions on time-

dependent reliability. The results clearly illustrate the decrease in reinforced concrete

element reliability after exposure to corrosion due to the reduction in flexural capacity of

the steel reinforcement.

Page 11: Reliability of Corroding Reinforced Concrete Structures

1

1. INTRODUCTION

1.1. Introduction

Concrete advantages, such as versatility and competitive cost, have given concrete

a practical advantage when compared to other structural materials. With the enhanced

addition of steel for structural ductility and tensile resistance, reinforced concrete has

been the most widely used construction materials in the world since the 1950’s (Liberati

2014). Construction methods, mathematical/analysis models and design codes for the

design of reinforced concrete structures seem to be well developed in the modern era of

engineering. Safety, economic and functionality requirements are heavily considered

throughout the design process. Another important factor to plan for when designing a

structure is the structural durability. The durability of reinforced concrete structures is

strongly connected to degradation processes whose origin is environmental and/or

functional. Among these processes is the development of corrosion on the steel

reinforcement. Corrosion is a complex, mechanical degradation process that occurs in

metallic materials as a result of chemical or electrochemical actions (Farsani 2015).

Reinforcement corrosion has a direct effect on the durability of reinforced concrete

structures. Corrosion of reinforcing bars is one of the major causes of deterioration of

reinforced concrete structures; affecting the useful service life of a given structure. Two

types of corrosion, general and pitting, are possible (Kioumarsi 2017). General corrosion

affects the cross section of reinforcement with nearly uniform metal loss over the

perimeter of reinforcing bars. It also causes cracking and eventual spalling of the concrete

cover and produces rust staining on the concrete surface, making it easily detectable

during inspection of the structure. Pitting, also referred to as localized corrosion,

Page 12: Reliability of Corroding Reinforced Concrete Structures

2

concentrates over small areas of reinforcement. Pitting corrosion often does not cause

disruption to the concrete cover and produces little rust staining on the concrete surface,

making it slightly more difficult to be discovered during inspections (Ghanooni-Bagha

2017).

Structural reliability and risk analysis aim at quantifying the probability of failure

taking into account relevant uncertainties. These uncertainties in a system vary from

design to manufacturing to environmental conditions. The modeling of structural systems

considering these uncertainties has become useful regarding the durability of a structure.

Probability models can embrace knowledge from multiple fields to allow for a more

consistent, comprehensive and dependable understanding concerning the reliability of a

structure as opposed to using a purely deterministic approach. This method acknowledges

the addition of uncertainties in numerous analyses in a consistent theoretical manner

through statistical associations (Liberati 2014).

1.2. Scope of Work

This research aims to develop a framework of analysis focusing on corrosion-

affected reinforced concrete members and systems, specifically evaluating the

reliability levels stemming from deterioration of steel reinforcement undergoing

general and localized corrosion degradation processes within reinforced concrete

girders of multiple highway bridges. The developed framework can be applied to

any design limit states that can be affect by corrosion. However, only flexural

strength limit states are considered in this study.

Page 13: Reliability of Corroding Reinforced Concrete Structures

3

1.3. Research Approach

Computational tools, such as MATLAB, were employed using the First Order

Reliability Method to gauge the serviceability of the selected design space of

reinforced concrete girders enduring uniform and pitting corrosion conditions. The

different corrosion conditions and varying design spaces, further discussed in

Chapter 3 of this paper, used probabilistic corrosion models from literature review

as a basis of material to be input into MATLAB to evaluate the integrity or

serviceability of a reinforced concrete girder. No concrete deterioration or loss of

bond strength at the concrete to steel interface was taken into account. The

developed framework can be used by infrastructure owners to assist in their

decision making process to project the existing life of reinforced concrete materials

experiencing corrosion degradation. Based on these projections, owners can also

plan intervention strategies to mitigate corrosion or strengthen the affected

members.

1.4. Objective

As of 1998, there were 583,000 bridges in the United States. Of these, 200,000

were steel; 235,000 were standard reinforced concrete; and 108,000 were prestressed

concrete. At this time, it was estimated that nearly 15 percent of bridges were structurally

deficient, predominantly due to the corrosion of steel and steel reinforcement. The annual

direct cost of corrosion for highway bridges in 1998 was estimated at $8.3 billion (Koch

et al., 2002). Indirect costs to users of this infrastructure can be account with time and

lost productivity spent in traffic delays.

Page 14: Reliability of Corroding Reinforced Concrete Structures

4

With an understanding of the corrosion process as well basic chemical and

reinforced concrete concepts, reliability algorithms can be established and analyzed using

established methodologies such as the First Order Reliability Method (FORM). These

methods have the ability to determine the probability of structural failure. The

approximation of structural failure probability can be used for estimating the performance

and lifetime of structures. The primary goal of this research is to analyze the corrosion of

steel reinforcement within a reinforced concrete structure, specifically within an

individual girder of a bridge under flexure loading conditions, and the concerns it will

yield regarding the structures’ reliability.

1.5. Organization

This thesis is organized into five chapters. In Chapter 1, the problem of corrosion

is introduced and the objective of the study is laid out. Chapter 2 presents a review of

relevant published literature regarding the topic. Procedures and techniques are

established in order to simulate the structural integrity of given design spaces that witness

corrosion in Chapter 3. The results of the simulations are calculated and shared in

Chapter 4. Summaries, basic conclusions and proposals of further research are made in

Chapter 5.

Page 15: Reliability of Corroding Reinforced Concrete Structures

5

2. LITERATURE REVIEW

In this chapter, previous studies from the literature regarding corrosion and

the subject of structural reliability are reviewed. Described in this chapter will be an

overview of what corrosion is, the origins of its inception, different types of

corrosion, etc. Also outlined is the background and framework of structural

reliability and how the integrity of infrastructure can be modeled using simulation

techniques.

2.1. Corrosion

Corrosion of the reinforcement embedded in concrete causes most of the failures

of reinforced concrete structures. Up until about half a century ago, carbonation of the

concrete was considered to be the main cause of corrosion. Since then, chloride-induced

corrosion has become much more renowned, especially for structures exposed to

chloride-containing environments, i.e. deicing salt, marine climate, or salt-contaminated

aggregates (Böhni 2005). As previously mentioned, corrosion of the reinforcement can

present itself through different forms, ranging from widespread, general corrosion to a

more localized attack, referred to as pitting corrosion. General corrosion is associated

mostly in cases of carbonated concrete, whereas pitting corrosion is commonly instigated

by chloride ions resulting in pits randomly distributed along steel bars. The depth of

these pits in steel bars may become large before any signs of deterioration appear on

concrete surfaces, making it an extremely serious issue. Deterioration caused by

reinforcement corrosion is normally divided into two main timeframes: initiation and

propagation.

Page 16: Reliability of Corroding Reinforced Concrete Structures

6

Along the surface of an embedded steel bar, when there exists a difference in

electric potential, the surrounding concrete acts as an electrochemical cell consisting of

anodic and cathodic regions of the steel with the pore water in the hardened cement paste

acting as an electrolyte. Goyal et al. (2018) describes that this process generates a flow of

current through the system, causing an attack on the metal with more negative electrode

potential, i.e. the anode, while the cathode remains undamaged. Thus, the corrosion of the

reinforcement is initiated as depicted in Figure 1 Böhni (2005) notes the occurrence of

the corresponding chemical reactions in the following equations.

Anodic Reaction: 2𝐹𝑒 → 2𝐹𝑒 4𝑒 (1)

Cathodic Reaction: 𝑂 2𝐻 𝑂 4𝑒 → 4𝑂𝐻 (2)

Reaction Sum: 2𝐹𝑒 2𝐻 𝑂 𝑂 → 2𝐹𝐸 𝑂𝐻 (3)

Figure 1. A schematic illustration of steel corrosion cell in concrete (Markeset and Mydral 2008)

The anodic reaction from Equation 1 represents the dissolution of the metal. The flux of

ions and electrons respectively can be taken as a measure of the corrosion rate. The

corrosion rate is customarily given as mass lost per unit of time and area, reduction of the

thickness per unit of time or as current density (current per unit of area). Figure 1 and

Equations 1-3 provide a chemical approach of defining corrosion initiation. A more basic

Page 17: Reliability of Corroding Reinforced Concrete Structures

7

explanation for the period of corrosion initiation can be labeled as the time from the end

of construction until the activation of the corrosion process, i.e. when the protective layer,

which ensures the adhesion between concrete and steel, is destroyed in the presence of

moisture and oxygen. This period can be modeled using a pure diffusion process using

Fick’s second law under the assumption that concrete is a homogenous and isotropic

material (Bamforth 2000; Tutti 1982).

Following the corrosion initiation period is the propagation phase. This stage is

recognized as the time required for corrosion to cultivate to the extent at which the load

capacity of the structural member becomes insufficient (Shayanfar et al. 2015). During

this phase, the corrosion of the steel reinforcements leads to the reduction of

reinforcement cross-sections prompting a drop in strength of the reinforced concrete

member (Markeset and Mydral 2008). The formation of rust on the reinforcement then

leads to concrete cracking in the proximity of other steel bars. The development of cracks

stimulates a loss of bond strength at the steel to concrete interface and remains until

spalling of the concrete cover transpires (El Hassan et al. 2010).

2.1.1. Carbonation Induced Corrosion

Porosity, a measure of the void spaces in a given material, of concrete ranges on a

minuscule scale, typically from the micrometer to nanometer level (Nóvoa 2016). Within

the pores of concrete, there is obviously some liquid water, but there is also adsorbed and

structural water. Goyal et al. (2018) explains that the presence of the adsorbed and

structural water can affect the structural and mechanical properties of the concrete. The

natural reactivity and porous structure of concrete make it prone to natural degradation.

Of these degradation processes exists carbonation, which refers to the penetration of

Page 18: Reliability of Corroding Reinforced Concrete Structures

8

carbon dioxide into the concrete layer and the subsequent neutralization of alkalis, such

as calcium carbonate, in the pore fluid (Zhou et al. 2015). Consequently, carbonation

causes a reduction in the pH level of concrete. A necessary pH of concrete for sufficient

corrosion protection is usually around 11-11.5; however, after the carbonation process,

the pH could fall to around 9.0 (Böhni 2005). At a pH level of this magnitude, the passive

layer will not be stable and thus, corrosion is likely to occur if sufficient oxygen and

water are present in the vicinity of the reinforcement (Heiyantuduwa et al. 2006).

Corrosion induced by carbonation is expected to be much more uniform than corrosion

that is chloride-induced (Cairns et al. 2005).

2.1.2 Chloride induced corrosion

The steel reinforcement will remain in a passive sate so long that it is embedded

in a sound concrete layer; however, the passive state no longer exists when the concrete

around it begins to deteriorate. With a lack of surrounding concrete, chloride ions may

penetrate from the environment and be able reach the reinforcement. As chloride ions

penetrate into concrete, the alkalinity near the steel reinforcement will increase as

represented in aforementioned Equation 2. Goyal et al. (2018) expresses that to maintain

electro-neutrality, chloride and hydroxide ions diffuse to the interface. With these

diffusion movements, the chloride ion concentration will build up close to the surface,

yielding saturation at the interface of Fe , iron, and Cl , chloride. This will shift the

electric potential in a more cathodic direction (Popov 2015). This process is known as

chloride-induced corrosion.

The chloride content required for steel depassivation and corrosion initiation is

referred to as the critical chloride content or the chloride threshold value (Goyal et al.

Page 19: Reliability of Corroding Reinforced Concrete Structures

9

2018). Once the chloride ion concentration goes beyond the threshold value, the passive

layer will locally be destroyed, which results in localized corrosion (Bertolini et al. 2018).

The critical threshold depends on various factors including but not limited to: concrete

quality, temperature, moisture content, and concrete cover (El Hassan et al. 2010).

According to Angst et al. (2009), there are actually multiple ways to define the critical

chloride threshold. Scientifically speaking, the critical chloride content can be defined as

the chloride content required for the depassivation of steel. From an engineering

standpoint, the chloride threshold value is the chloride content associated with visible or

tolerable deterioration of the reinforced concrete structure. Yet, both of these critical

chloride definitions relates to different circumstances. The scientific designation of a

chloride threshold only includes a depassivation criterion, which means only the

corrosion initiation stage is considered, while the engineering description of the threshold

considers part of the propagation stage in addition to the initiation stage. Therefore, the

two definitions of chloride threshold lead to different critical chloride content values as

shown in Figure 2. Figure 2 clearly shows that different definitions of the critical chloride

lead to separate critical chloride values, where the engineering outlook leads to a higher

chloride threshold value. The scientific view of the threshold is more detailed and

approximate since it relates only to the chloride content related to depassivation.

However, the chloride content related to the engineering view remains unclear since the

amount of tolerable deterioration is ambiguous.

Page 20: Reliability of Corroding Reinforced Concrete Structures

10

Figure 2. Definitions for chloride thresholds (Angst et al. 2009)

For these reason, the critical chloride threshold value is an unclear component of this

corrosion process. Since the pH of concrete varies according to the concrete composition

of the concrete, e.g. type of cement or water to cement ratio, results reported in the

literature scatter over a large range, and therefore, no fixed or single value for this

threshold is comprehensively accepted (Böhni 2005). Though, a popular approach to

quantify the critical chloride threshold level is based on Fick’s second law, which

requires knowledge of the diffusion coefficient of the cement-based material (Poupard et

al. 2004).

The effects of chloride diffusion are more thoroughly examined due to its’

threatening nature. Therefore, accurately modeling the chloride diffusion will be essential

to better evaluate the corrosion on steel reinforcement. Due to the inherent randomness

regarding chloride diffusion and corrosion, these processes should be properly modeled

Page 21: Reliability of Corroding Reinforced Concrete Structures

11

using probabilistic approaches (Liberati et al. 2014). Ghanooni-Bagha et al. (2017)

introduce a probabilistic model for predicting corrosion initiation and propagation. The

time to corrosion initiation, using Fick’s second diffusion law due to its’ accuracy, can be

calculated using Equations 4 and 5 by simulating the chloride ingress in concrete

(Almusallam 2001; Bastidas-Arteaga et al. 2009).

𝐷 (4)

𝑇 𝑒𝑟𝑓 1 (5)

where

C = Concrete cover (cm)

erf = Error function

𝐶 = Chloride threshold

𝐶 = Constant amount of chloride on surface

𝐷 = Chloride diffusion coefficient

The chloride diffusion coefficient used by Ghanooni-Bagha et al. (2017), instituted from

Bastidas-Arteaga et al. (2008) and Bhargava et al. (2011), is directly correlated with the

water-cement ratio as seen in Equation 6.

𝐷 1.249 5.051𝑤/𝑐 8.941𝑤/𝑐 ∗ 10 (6)

2.2. Detection and protection techniques against corrosion

It is important to be proactive in identifying and rehabilitating corrosion-affected

reinforced concrete members. Repairs to corrosion-damaged concrete structures are

typically categorized into two methods: conventional repair and electrochemical (Goyal

Page 22: Reliability of Corroding Reinforced Concrete Structures

12

et al. 2018). Conventional methods for concrete repair aim at removing all chloride-

contaminated or carbonated concrete, cleansing the steel and reorganizing the surface

with fresh, chloride-free and alkaline, cementitious concrete or mortar. Geiker and Polder

(2016) concluded that though these are generally temporary techniques for corrosion

prevention, there is downside to these methods as they can lead to acceleration of

corrosion in nearby repaired areas. Unless all chloride is removed, especially from steel

surfaces where corrosion pits may lie, corrosion will likely reinitiate after a short duration

(Böhni 2005). For these reasons, Elsener and Angst (2007) declares that these methods

are commonly costly and not as effective as electrochemical methods. In electrochemical

techniques, the chemical reactions and current flows due to corrosion are restrained by

the application of an external direct current supply with the assistance of an anode. The

direct current passes from the artificial anode as a flow of ions through the pore water of

the concrete to the reinforcing steel (Drewett et al., 2011). Some effective

electrochemical techniques for corrosion prevention and mitigation include: cathodic

protection, cathodic prevention, electrochemical realkalization and electrochemical

chloride removal.

Furthermore, some steps or practices can be taken earlier than expected, including

during the construction of reinforced concrete structures. To prevent corrosion of

reinforcement in concrete, an alternative reinforcement that is made of corrosion-resistant

material could be utilized, e.g. stainless steel or fiber-reinforced plastic amongst other

types (Goyal et al. 2018). The corrosion resistance of stainless steel bars is significantly

higher than carbon or mild steel because of the higher stability of their passive film,

which is high in chromium and is self-healing in an oxygen rich environment (The

Page 23: Reliability of Corroding Reinforced Concrete Structures

13

Concrete Society, 1998). Alternatively, fiber-reinforced plastic (FRP reinforcement) are

composite materials; which have high corrosion resistance, are light weight, and

generally have a high tensile strength (ACI, 2007) and can be used as internal

reinforcement for concrete structures. Additionally, adding coatings to steel or concrete

can serve as a physical barrier to corrosion. Cicek (2018) states that coatings, sometimes

referred to as sealers, have higher adhesion and are non-reactive in a corrosive

environment and protect steel from any mechanical damages.

2.3. Structural Reliability and Probabilistic Evaluation

Structures are complex systems that are susceptible to a number of uncertainties at

all phases of design execution as well as practical usage. Some uncertainties can never be

completely eliminated and must, therefore, be taken into account when designing or

validating the integrity of a structure. Some common sources of uncertainties include:

lack of reliable experimental data (i.e., statistical uncertainty), environmental influences,

geometrical data, randomness of actions, and gross errors in design. Numerous design

methods and operative techniques have been proposed and utilized worldwide in an effort

to predict and control the unfavorable effects of underlying uncertainties within the

duration of a structure’s service life. As a by-product, the theory of structural reliability

has been developed to evaluate uncertainties within a structural system in a rational way

and to consider when designing and monitoring structural performance.

Reliability is the ability of a structure to satisfactorily meet given requirements

under specified conditions during the intended design life (Holický, 2009). Essentially, a

structure should maintain its’ integrity for the usage of which it was required as well as

Page 24: Reliability of Corroding Reinforced Concrete Structures

14

sustaining all actions and influences likely to occur during construction and usage during

its’ intended working life.

2.3.1. Background and Development

Design methods that recognize uncertainties have been continuously developed

using available empirical, experimental, and theoretical knowledge of mechanics and the

theory of probability. The development of various empirical methods for structural design

evolved throughout the twentieth century (Holický, 2009) seeing three generally used

methods gaining popularity. These methods, with various modifications since initial

inception, are still employed to this day.

The first design method, the method of permissible stresses, is based on the

condition that the local maximum load effect is less than the factored resistance of an

element or structural system (Equations 7 and 8).

𝜎 𝜎 (7)

𝜎 (8)

where the coefficient k is gaged with regard to uncertainties in the determination of the

local load effect, 𝜎 , resistance, 𝜎 , and critical stress that initiates motion toward

failure, 𝜎 ,. The main deficiency of this method is analyzing the location verification of

reliability as well as the difficulty of considering separately the uncertainties of multiple

basic quantities and parameters, especially with respect to the demands or loads. The

probability of failure is controlled only by the k coefficient.

The second design method is the method of the global safety factor.

𝑠 𝑠 (9)

Page 25: Reliability of Corroding Reinforced Concrete Structures

15

The calculated safety factor, S, must be greater than its’ specified value, 𝑆 . The global

safety factor design method tries to indicate the behavior of elements as well as their

material and geometric properties through the evaluation of the structural resistance, 𝑋 ,

and action effect, 𝑋 . Similarly to the method of permissible stresses, the probability of

failure is controlled by only one quantity, i.e. the global safety factor.

Lastly, methods of probabilistic modeling were introduced, implemented, and

developed to better investigate the reliability of a structure. This third approach is

covered within the next section featuring utilized data.

2.3.2. Probabilistic Methods

Probabilistic design methods are based on the condition that the probability of

failure, 𝑃 , does not exceed a specified target value during the service life of a structure

(Holický, 2009). The probability of failure can be assessed using stronger and quicker

computational models. These models can be defined with basic quantities called random

variables, 𝑋 𝑋 , 𝑋 , … , 𝑋 , for uncertainties including actions, mechanical properties,

and geometrical data. The condition of a structure is defined by the limit state function

(performance function), g(X), for which the ‘safe’ state of a respective element or

structural system is held positive, represented in Equation 10.

𝑔 𝑋 0 𝑔 𝑋 , 𝑋 , … , 𝑋 0 (10)

Similarly, the ‘unsafe’ state, corresponding to failure and indicated in Equation 11,

occurs when the limit state function is negative.

𝑔 𝑋 0 𝑔 𝑋 , 𝑋 , … , 𝑋 0 (11)

A random variable, X, is noted to be a variable that attains one singular previously

unknown value, x, when a certain random event occurs. Within the concept of structural

Page 26: Reliability of Corroding Reinforced Concrete Structures

16

reliability, there are both continuous random variables and discrete random variables.

Continuous random variables shall attain any value within a given interval or domain.

Discrete random variables attain strictly distinct values only. Most computational

probabilistic models utilize continuous random variables.

The cumulative distribution function, CDF, noted symbolically as Φ 𝑥 for the

standard form of a normal distribution, gives the probability for each value at x that the

random variable, X, will be less than or equal to x (Equation 12).

Φ 𝑥 𝑃 𝑋 𝑥 (12)

The probability density function, PDF, and recognized symbolically as 𝜙 𝑥 , of a

continuous random variable x is the derivative of the distribution function.

The measure of distribution for a random variable relative to the mean, μ, is given

by the central moment of the second order, and is referred to as the variance (Equation

13).

𝜎 𝑥 𝜇 𝜙 𝑥 𝑑𝑥 (13)

The square root of the variance, σ, denotes the standard deviation of the random variable.

The ratio of the standard deviation to the mean yields a dimensionless parameter of a

population recognized as the coefficient of variation, computed by Equation 14, which is

an important measure of relative dispersion.

𝐶𝑂𝑉 (14)

One of the most significant types of distribution of a continuous random variable

is the normal (Laplace-Gauss) distribution. Normal distribution of a random variable, X,

is symmetric on an unlimited interval from -∞<x<∞. The probability density function of

Page 27: Reliability of Corroding Reinforced Concrete Structures

17

normal random variable X can be found by Equation 15 using parameters of the

distribution with mean, 𝜇 , and standard deviation, 𝜎 .

𝜙 𝑥√

𝑒𝑥𝑝 (15)

Similarly, using parameters of the distribution as well as knowing the distribution type,

the cumulative probability function is documented using Equation 16.

Φ 𝑥12

1 erf𝑥 𝜇

𝜎√2

(16)

Where ‘erf’ is considered to be the error function, which in the case of this distribution,

gives the probability of a random variable of the mean with value 0 and the variance of

value 0.5 falling in the range of [-x, +x]. With knowledge of distribution parameters with

respect to the normal distribution, CDF and PDF values can commonly be found on

readily available numerical tables.

Another common distribution type is the lognormal distribution, an asymmetric

distribution that lies on a finite definite domain. A random variable X is considered to be

a lognormal random variable if Y=ln(X) is normally distributed, where a lognormal

random variable is defined for positive values only (Nowak and Collins, 2000). Since the

distribution is not normally distributed, the relationship of standard normal random

variables is used to give calculations for the CDF in Equations 17 - 21.

𝐹 𝑥 𝑃 𝑋 𝑥 𝑃 𝑙𝑛 𝑋 𝑙𝑛 𝑥 𝑃 𝑌 𝑦 𝐹 𝑦 (17)

Random variable Y is considered to be normally distributed; therefore, standard normal

functions are used to relate lognormal and normal distribution types.

𝐹 𝑥 𝐹 𝑦 Φ (18)

where

Page 28: Reliability of Corroding Reinforced Concrete Structures

18

𝑦 ln 𝑥 (19)

𝜎 𝜎 𝑙𝑛 1 (20)

𝜇 𝜇 ln 𝜇 𝜎 (21)

Nowak and Collins (2000) note that though not commonly utilized, the derivative of the

CDF can be used to find values for the PDF using Equation 22.

𝑓 𝑥 𝐹 𝑥 Φ 𝜑 (22)

Extreme value distribution is useful when characterizing the probabilistic nature

of extreme values, such as exceptionally large or small values of some random

occurrence over a span of time. Three types of distribution of extreme values are

denoted, and each type has two separate versions – one for the distribution of minimal

values, and another for the distribution of maximal values. Each type has simple

exponential shapes. Type I, Gumbel distribution, is useful when evaluating a maximum

value of a recognized phenomenon. Though it can be useful to model minimum values as

well, any reference to Gumbel distribution from here on will be used with respect to

modeling the distribution of maximum values. The scale parameter, 𝜇, and the location

parameter, 𝛼, are related to the mean and standard deviation of the distribution as seen in

Equations 23 and 24 (Nowak and Collins, 2000).

𝜇 𝜇 . (23)

𝜎 . (24)

Therefore, if the mean and standard deviation are known, the equations above can be

used to solve for the location and scale parameters.

Page 29: Reliability of Corroding Reinforced Concrete Structures

19

Other useful distribution types include: Gamma, Triangular, Poisson, Weibull,

and Chi-Square. Each type has independent distribution parameters and describes specific

events as well as differing techniques to find the CDF and PDF curves.

2.4. Reliability Analysis

The fundamental task of the theory of structural reliability is the investigation of

the constraint that the load effects on a structure, Q, will be less than the structural

resistance, R, of which a structure has the ability to provide (Holický, 2009).

𝑔 𝑅, 𝑄 𝑅 𝑄 (25)

This constraint describes the states of a considered structural component, aforementioned

as limit states. It is assumed that structural failure occurs when the resistance of the

structure, R, is less than the load effects, Q (Holický, 2009). Essentially, thought it might

not relate to physical failure, or collapse, of a structure, structural failure here means that

a structure cannot perform its’ intended function (Nowak and Collins, 2000). Within

structural and civil engineering design, there are distinctions between types of limit state

functions with the two most recognizable are ultimate limit states and serviceability limit

states. The ultimate limit states are associated with various forms of structural failure.

These limit states represent the loss of equilibrium or excessive deformation or

settlement, rupture, or loss of stability. Serviceability limit states are associated with the

criteria related to the usage of the structure under normal operation conditions. These

limit states require consideration of deformation or deflection, vibration limiting

structural use, and detrimental cracking (Holický, 2009).

Using methods within the theory of structural reliability, the probability of failure

can be computed into an explicit value. The probability of failure, 𝑃 , is equivalent to the

Page 30: Reliability of Corroding Reinforced Concrete Structures

20

probability that the limit state function will fall in the unsafe state, i.e. when g(R,Q) is

less than 0 (Nowak and Collins, 2000) as modeled in Equation 26.

𝑃 𝑃 𝑅 𝑄 0 𝑃 𝑔 𝑋 0 (26)

If both the resistance, R, and load effects, Q, are continuous random variables,

then each has a probability density function. Therefore, the quantity of the limit state

function can itself be considered to be a random variable with its’ own PDF and CDF as

depicted in Figure 3.

Figure 3. PDF’s of resistance, load effects, and the safety margin (Nowak and Collins, 2000)

The probability of failure of this arbitrary limit state function is depicted as the shaded

area.

In further structural reliability analyses, it will be advantageous to convert all

random variables to their standard form, which is recognized as a non-dimensional form

of each variable. Equation 27 displays that for a basic variable, X, the standard form

utilizes the random variable’s mean and standard deviation.

𝑍 (27)

Page 31: Reliability of Corroding Reinforced Concrete Structures

21

The variable 𝑍 can now be referred to as a reduced variable. Reduced variables are

utilized, with respect to the resistance and load effects, within the limit state function to

further assess the probability of failure, depicted in Equation 28.

𝑔 𝑍 , 𝑍 𝜇 𝑍 𝜎 𝜇 𝑍 𝜎 𝜇 𝜇 𝑍 𝜎 𝑍 𝜎 (28)

Figure 4 portrays how this limit state function now represents a straight line in the space

of all reduced variables within the calculation.

Figure 4. Reliability index defined as the shortest distance in the space of reduced variables (Nowak and Collins, 2000)

The reliability index, β, has a geometrical interpretation from Figure 4 paralleled to

Equation 29.

𝛽 (29)

where 𝛽 is the inverse of the coefficient of variation of the limit state function. With

random variables that are normally distributed, the reliability index is related to the

probability of failure as seen in Equation 30 and in Figure 5.

𝛽 Φ 𝑃 or 𝑃 Φ 𝛽 (30)

Page 32: Reliability of Corroding Reinforced Concrete Structures

22

Figure 5. Relationship between β and 𝑷𝒇 for a normally distributed function (Allen et al., 2005)

If the resistance, R, is not normally distributed then the reliability index, β, is formally

defined as a negative value of a standardized random variable corresponding the failure

probability (Holický, 2009).

2.4.1. Historically used Methods

There are various statistical methods and simulation techniques utilizing the

reliability theories to quantify the structural reliability of a structure within its’ service

life. Notable procedures include Mean Value First Order Second Moment (MVFOSM),

First Order Second Moment (FOSM), First Order Reliability Method (FORM), Monte

Carlo Simulation, and Latin Hypercube Sampling.

Mean Value First Order Second Moment (MVFOSM)

Page 33: Reliability of Corroding Reinforced Concrete Structures

23

Within situations where the only information about random variables is their

second moments, i.e. their means, standard deviations, and correlation coefficients,

MVFOSM provides a useful method in order to obtain closed form solutions to basic

limit state functions. Haukaas (2018) states that MVFOSM utilizes the mean and standard

deviation of the limit state function in order to evaluate the measure of reliability. For the

limit state function from Equation 25, the MVFOSM reliability index is concordant with

Equation 29. MVFOSM is not one of the most approximate methods as results can be

inaccurate if the tails of the distribution functions cannot be approximated by normal

distribution or also if there is an invariance problem; and therefore, it is used as more of a

preliminary approach (Nowak and Collins, 2000).

First Order Second Moment (FOSM)

The first order second moment has become one of the most often and willingly

used approaches to evaluate structural safety characteristics. This method originates from

the partial derivatives or Taylor series method (Dolinski, 1982). It is designed mainly for

analysis by hand, the second moment method is based on point estimates of the system

response (Wong, 1984). In a procedure outlined by Nowak and Collins (2000), the

reliability index depends only on the means and standard deviation of the random

variables; hence why it is referred to as a second-moment measure of structural safety

since only the first two moments (mean and variance) are required to calculate the

reliability index, 𝛽. Considering a limit state function from Equation 31, an expression

for the reliability index can be analyzed Equation 32.

𝑔 𝑋 , 𝑋 , … , 𝑋 𝑎 𝑎 𝑋 𝑎 𝑋 ⋯ 𝑋 𝑎 ∑ 𝑎 𝑋 (31)

Page 34: Reliability of Corroding Reinforced Concrete Structures

24

𝛽∑

(32)

First Order Reliability Method (FORM)

Perhaps one of the most efficient reliability methods is FORM. Developed by

Hasofer and Lind (1974), FORM is used as a fundamental procedure by a number of

software products and researchers for reliability analyses. In this method, the basic

random variables must be transformed into a space of standardized normal variables. The

transformation into equivalent normal variables at a given point, 𝑥∗, is based on the equal

distribution functions from Equations 33 and 34 and equal probability increments yielded

from Equations 35 an 36.

Φ 𝑥∗ Φ∗

(33)

𝜙 𝑥∗ 𝜙∗

(34)

𝜇 𝑥∗ 𝜎 Φ Φ 𝑥∗ (35)

𝜎 ∗ 𝜙 𝜙 Φ 𝑥∗ (36)

Instead of using mean values of variables, the design point, i.e. the closest point to the

origin, is found throughout an iterative procedure. Then, the reliability index, 𝛽, is

denoted as the distance of the design point from the origin. Holický (2009) and Nowak

and Collins (2000) describe the iterative procedure as:

1. The initial assessment of the design point, 𝑥∗, is made by the mean values of n-1

basic variables and the last one is determined from the limit state function of

𝑔 𝑥∗ 0

2. Equivalent normal distributions are found for all basic variables using Equations

35 - 36

Page 35: Reliability of Corroding Reinforced Concrete Structures

25

3. The transformed design point, 𝑧∗, or standardized random variable, Z,

corresponding to the design point is determined using Equation 37

𝑧∗∗

(37)

4. Partial derivatives denoted as a vector, 𝐺 , of the limit state function with respect

to the standardized values, Z, are evaluated at the design point as seen in Equation

38

𝐺 𝜎 (38)

Where 𝐺

⎩⎪⎨

⎪⎧

𝐺𝐺..

𝐺 ⎭⎪⎬

⎪⎫

5. The reliability index is estimated with Equation 39.

𝛽∗

(39)

Where 𝑧∗

⎩⎪⎨

⎪⎧

𝑧∗

𝑧∗

.

.𝑧∗ ⎭

⎪⎬

⎪⎫

6. Sensitivity factors, α, are determined in Equation 40

𝛼 (40)

7. A new design point is determined for n-1 standardized and original basic variables

from Equations 41 and 42.

𝑧∗ 𝛼 ∗ 𝛽 (41)

𝑥∗ 𝜇 𝑢∗ ∗ 𝜎 (42)

Page 36: Reliability of Corroding Reinforced Concrete Structures

26

8. Similarly to Step 1, the design value of the remaining basic variable is determined

from the limit state function 𝑔 𝑥∗ 0

9. Steps 2 to 8 are to be repeated until the reliability index and the design point have

the required accuracy or convergence of β.

Monte Carlo Simulation Technique

The Monte Carlo Simulation technique uses statistical parameters to simulate a

respective limit state function (Nowak and Collins, 2000). This method is employed to

solve complex problems that can be approximately solved in closed form by using many

simplifying assumptions. By simulating the limit state function a number of times, the

number of failures throughout those simulations can be obtained. For the most accurate

results, there would have to be a very large number of simulations. With the aid of a

random number generator as well knowledge of the mean, standard deviation, and

distribution type, random values can be generated for each random variable within a limit

state function. With the collected values of each random variable, the limit state function

is to be calculated and stored. After a sufficient amount of simulations, the amount of

limit state function values that fall in the unsafe zone are counted. The probability of

failure, Equation 43, is assessed using the ratio of number of simulations relating to

failure, 𝑁 , divided by the total number of simulations, N.

𝑃 (43)

Other simulation techniques are available such as Latin Hypercube Sampling,

Rosenbluth 2k+1, etc. However, they were not used in this study, and thus are beyond the

scope of this research.

Page 37: Reliability of Corroding Reinforced Concrete Structures

27

3. METHODOLOGY

Now knowing fundamental information regarding corrosion and structural

reliability, this chapter displays the design spaces and methods that are set up in order to

evaluate the ramifications that the corrosion process has on steel reinforcement.

3.1. Design Space

With the goal of evaluating cases of both uniform and localized corrosion

surrounding varying design specifications, approaches and methods of past and current

researchers are herein utilized. A design space used by Okeil et al. (2013) was selected to

ensure that the design equations and methods convey acceptable reliability levels over a

wide range of distinctive design scenarios. Three different bridges were chosen for these

scenarios. Each bridge varies in span length, but each contains five, typical reinforced

concrete girders spaced transversally seven feet laterally from each other, as can be seen

in Figure 6.

Figure 6. Details of typical bridge used to cover design space (Okeil et al, 2013)

The span length for each considered bridge is 45 feet, 60 feet, and 75 feet, or 13.72

meters, 18.29 meters, and 22.86 meters, respectively. A singular interior reinforced

concrete girder from each particular bridge, as shown in Figures 7-9, will be analyzed

Page 38: Reliability of Corroding Reinforced Concrete Structures

28

using the following approaches. The nominal flexural demands on the selected beams are

found in Tables 7-9. More details on these design spaces can be found elsewhere in Okeil

et al. 2013.

Figure 7. Typical girder for design space 1 (Okeil et al, 2013)

Figure 8. Typical girder for design space 2 (Okeil et al, 2013)

Page 39: Reliability of Corroding Reinforced Concrete Structures

29

Figure 9. Typical girder for design space 3 (Okeil et. al, 2013)

3.2. Design Approach

The diffusion coefficient is dependent on conditions such as concrete material

properties, water to cement ratio, and exposure conditions. However, Castaneda et al.

(2018) shifted focus on developing an effective, probabilistic model using the least

number of parameters based on model selection criteria for developing the diffusion

coefficient in a manner where it would not be a deterministic, constantly held value such

as portrayed by Equation 6. Bayesian methodology was utilized within the evidence-

based models using experimental data. The impacts of each parameter including material

properties, environmental conditions, and the structural geometries were invetigated

considering the uncertianties within the system. A statistical approach was added utilizing

Bayesian updating technique to decide the model parameters. The model parameters, 𝜃 ,

𝜃 , and 𝜃 , were estimated by the maximization function that maximizes the

concentration profile likelihood when previous reference case data set was given. Still

Page 40: Reliability of Corroding Reinforced Concrete Structures

30

using Fick’s law (Liberati, 2014, Nogueira, 2012), Equations 44 and 45 were established

by Castaneda et al. (2018).

𝐶 𝑥, 𝑡 𝐶 1 𝑒𝑟𝑓√

(44)

D 𝑡 𝜃 𝜃 𝑡⁄ (45)

where,

C 𝑥, 𝑡 = Chloride concentration (% mass),

𝐶 = Chloride concentration on concrete surface (% mass),

𝑥 = Distance from surface (mm),

𝐷 = Chloride diffusivity (mm2/month),

𝑡 = Time (month), and

𝜃 , 𝜃 , 𝜃 = Parameters in the diffusion model

Table 1 displays the consequent statistics of the parameters for the chloride

diffusion model using the reference case of WC = 0.46 and 7 days of curing, based on

the diffusion equation without the correction factor. Sigma, 𝜎, represents the model

error. The correlation coefficient shows the parameters’ correlation with one

another. In this case, 𝜃 and 𝜃 shows relatively higher correlation compared to

other parameters’ correlation, which means that those two parameters are possibly

merged to increase the equation efficiency, but there is also other possibility that

such a merging process increases the model uncertainty. It was observed that the

model has low statistical uncertainty as is evident by the generally low standard

deviation values.

Page 41: Reliability of Corroding Reinforced Concrete Structures

31

Table 1. Chloride diffusion model parameters and statistics

Correlation Coefficient

Mean Std. Dev. 𝜃 𝜃 𝜃

𝜃 500 1 1 0.013 -0.213

𝜃 30 1.13 0.013 1 -0.26

𝜃 0.56 0.027 -0.213 -0.26 1

𝜎 0.342 0.019 0.019 -0.336 0.044

In order to express the different concrete conditions compared to the

reference case, the correction factors, α and α , from Tables 2 and 3 were added on

the diffusivity equation. The correction factor helped improve the model fit to the

given data set, and reflects the trend that may not be observed in the posterior

statistics with reference case. The factors, α and α , representa water to cement

ratio correction factor andcuring day correction factor, respectively.

Page 42: Reliability of Corroding Reinforced Concrete Structures

32

Table 2. Water to cement ratio correction factor

WC Correction Factor, 𝛼  

WC Model  Distribution  𝛼 Mean  𝛼 Std. Dev. 

0.46  Normal  1  0.4 

0.5  Normal  1.719  0.576 

0.53  Normal  2.283  0.673 

0.7  Normal  2.678  ‐ 

0.76  Normal  2.783  0.984 

Table 3. Curing day correction factor

Curing Days Correction Factor, α  

C  Model  Distribution  α Mean  α Std. Dev. 

1  Normal  1.265  0.221 

3  Normal  0.716  0.073 

7  Normal  0.875  0.09 

Using these correction factors and model parameters, Castaneda et al. (2018)

probabilistically generated the following time to corrosion initiation Equation 46,

similar to the commonly used Equation 5.

𝑇 𝑒𝑟𝑓 1 (46)

As previously stated, there is no known single, precise value for the chloride

threshold value. Therefore, the experimental work of Castro-Borges et al. (2013)

involving polarization resistance techniques provided interpolated chloride

threshold values at five respective water to cement ratios for Castaneda et al.

(2018), which will be the assumed critical chloride values herein. These values are

provided in Table 4.

Page 43: Reliability of Corroding Reinforced Concrete Structures

33

Table 4. Interpolated chloride threshold values

Water to Cement Ratio, WC Interpolated Chloride Threshold Value, 𝐶 (% per weight of

concrete) 0.46 0.112

0.50 0.1447

0.53 0.1265

0.70 0.06797

0.76 0.0586

3.2.1. General corrosion

Numerical definition of corrosion propagation is provided primarily in the form of

corrosion rate, which is the amount of steel loss per unit time and unit area. Most current

non-destructive techniques detect corrosion utilizing an electromechanical measurement

of current rate, 𝑖 . Analytical models, developed by Vu and Stewart (2000), for

estimating corrosion rate depend on parameters such as quality of concrete and rebar

cover as given in Equation 47.

𝑖 𝑇 0.85𝑖 1 𝑇 . (47)

The time, T, indicates the time span since initiation, i.e. the initiation time of corrosion

deducted from structure life. Equation 47 is viable to be used in cities located within

Asian, European, American, and Australian regions due to the average relative humidity

of many locations in these areas being over 70%; therefore, it is assumed that the

corrosion rate is limited by the availability of oxygen at the steel surface (Vu et al. 2005).

One key disadvantage of this equation is related to corrosion within makeups having

short periods of structural life. If the time, T, is too small, the corrosion rate will

Page 44: Reliability of Corroding Reinforced Concrete Structures

34

theoretically tend to reach infinity; therefore, modified Equation 48 was put into place

(Vu and Stewart 2000).

𝑖 𝑇𝑖 1 0 𝑇 1 𝑦𝑒𝑎𝑟

0.85 ∗ 𝑖 1 ∗ 𝑇 . 𝑇 1𝑦𝑒𝑎𝑟

(48)

where

𝑖 1 = One-year corrosion rate at the start of corrosion propagation

Developed and suggested by Dai and Wang (2009), the one-year corrosion rate is

defined at the start of corrosion initiation and dependent upon variables such as the water

to cement ratio and the cover of reinforcement (Equation 49).

𝑖 1 . . (49)

According to classifications recommended by BRITE/EURAM (1995), corrosion rates

between 0.1 and 0.5 microamperes per square centimeter (μA/ cm2) correspond to low

corrosion rates, a range of 0.5 to 1.0 μA/ cm2 correspond to moderate corrosion, and

above 1.0 μA/ cm2 correspond to a high rate of corrosion. It is important to note that

most reliability analyses have assumed a constant corrosion rate during the propagation

period. However, Vu and Stewart (2000) rationalize that it is expected that the formation

of products on the steel surface will reduce the diffusion of the iron ions away from the

steel surface. Furthermore, the area ratio between the anode and the cathode is reduced.

These assessments suggest that the corrosion rate will diminish with time; explicitly,

more rapidly during the first few years after corrosion initiation but then at a decelerating

rate as it approaches a more uniform level (Yalcyn and Ergun 1996). Data from Liu and

Weyers (1998) was used in order to develop the relationship between time since

corrosion initiation and the corrosion rate as seen in the aforementioned Equation 48.

Page 45: Reliability of Corroding Reinforced Concrete Structures

35

Bushman (2000) acknowledges that Faraday’s law of electrochemical equivalence

is employed into the loss of steel reinforcement when analyzing general corrosion. The

reduction of diameter through the general corrosion process is indicated with Equation

50.

Δ𝐷 𝑇 0.0232 ∗ 𝑖 𝑇 (50)

where

Δ𝐷 𝑇 = Reduction of steel reinforcement diameter (mm)

As previously mentioned, many reliability analyses rely on a constant corrosion

rate within investigations. If using a constant corrosion rate, the reduction of

reinforcement diameter would simply be a function of consistent corrosion multiplied by

the time. Nevertheless, only a corrosion rate with respect to time will be used herein. If

the corrosion current density is assumed to be identical for “n” reinforcement bars, the

residual cross sectional area of rebars could be computed using Equation 51.

𝐴 𝑇 𝑛 ∗ 0 (51)

However, if the corrosion is localized as a product of pitting corrosion, the

residual cross section of the rebars is found using Equation 52.

𝐴 𝑇 𝑛 𝐴 0 (52)

3.2.2. Pitting corrosion

Tutti (1982) and Gonzalez et al. (1995) have been significant developers of

modeling pitting corrosion within chloride-contaminated structures. Stewart (2004)

mentions that the maximum pit depth, 𝑃 , will normally exceed the penetration

calculated based on general corrosion, 𝑃 . Researchers indicate, however, that there is a

significant uncertainty associated with this with this ratio. Gonzalez et al. (1995) found in

Page 46: Reliability of Corroding Reinforced Concrete Structures

36

a study where concrete specimens were exposed to natural environments that the

maximum ratio of maximum pit depth to penetration of general corrosion, denoted as R,

tended to vary from 4 to 8. The results of Gonzalez et al. (1995) were in broad agreement

with Tutti (1982) who suggested that the ratio usually falls within a range of 4 to 10.

Darmawan and Stewart (2003) have found the distribution of maximum pit depths for

prestressing wires has found to be best represented by the Gumbel (Extreme Value-Type

I) distribution. Extreme value statistics have been widely used to characterize the pitting

of steel plates and pipes as well as prestressing strands. Therefore, it is reasonable that the

Gumbel distribution be appropriate for modeling maximum pit depths of reinforcing bars

(Stewart 2004).

A popular approach in modeling pit depth using extreme value theory originates

from Turnbull (1993). The ratio of maximum pit depth to average penetration from

general corrosion, R, is treated as a random variable modeled by the Gumbel distribution

using Equation 53. This random variable is also referred to as the pitting factor.

𝐹 𝑅 𝑒 (53)

where

𝜇 = Location parameter

𝛼 = Scale parameter

The location and scale parameters characterize the spread of the distribution.

According to Stewart (2004), these parameters can be determined, grounded on the

results of Gonzalez et al. (1995), that for an 8 millimeter diameter bar of 125 millimeter

length, R=4 and R=8 represent the 5th and 95th percentiles of the distribution, respectively.

The mean and coefficient of variation are then considered to correspondingly be 5.65 and

Page 47: Reliability of Corroding Reinforced Concrete Structures

37

0.22, which relate to the parameters of the Gumbel distribution 𝜇 =5.08 and 𝛼 =1.02.

Turnbull (1993) suggests that for a reinforcing bar with different dimensions, the Gumbel

distribution parameters can be determined using Equations 54 and 55. The mean and

standard deviation are related to these parameters by Equations 23 and 24.

𝜇 𝜇 𝑙𝑛 (54)

𝛼 𝛼 (55)

where

𝐴 = surface area of the respective bar

𝐴 = surface area of an 8mm diameter bar of 125 mm length

With the distribution factors of random variable of R, the maximum pit depth along a

reinforcing bar can be evaluated using Equation 56.

𝑃 𝑇 0.0116 ∗ 𝑖 𝑇 ∗ 𝑅 ∗ 𝑇 (56)

This equation considers Faraday’s law where a unit current density induces a

uniform corrosion penetration of 11.6 micrometers per year (El Hassan et al. 2010). A

model to predict the loss of the cross-sectional area of a reinforcing bar due to pitting has

been proposed by Val and Melchers (1997). The pit configurations, as seen in Figures 10

and 11, the width of the pit, and cross sectional area of the pit can be expressed using

Equations 57-62.

Page 48: Reliability of Corroding Reinforced Concrete Structures

38

Figure 10. Pit configuration (Val, 2007)

Figure 11. Pit configuration (Darmawan, 2010)

𝑏 2 ∗ 𝑃 𝑇 ∗ 1 (57)

𝐴 0.5 𝜃 𝑏 (58)

𝐴 0.5 𝜃 ∗ 𝑃 𝑇 𝑏𝑃 𝑇

𝐷

(59)

𝜃 2arcsin (60)

𝜃 2𝑎𝑟𝑐𝑠𝑖𝑛 (61)

Page 49: Reliability of Corroding Reinforced Concrete Structures

39

𝐴 𝑇

⎩⎨

⎧ 𝐴 𝐴 𝑃 𝑇√

𝐴 𝐴 𝐴 √

𝑃 𝑇 𝐷

𝐴 𝑃 𝑇 𝐷

(62)

Using the residual area definitions from either general or pitting corrosion from

Equations 51 and 52, respectively, the flexural limit state function formed for the

reliability analysis of each design space is molded from basic bending theory. Nominal

moment flexural strength, 𝑀 , recognized from reinforced concrete concepts, exhibits a

correlation with the area of steel rebars and the flexural resistance can be estimated using

Equations 63 and 64.

𝑀 𝐴 ∗ 𝑓 ∗ 𝑑 (63)

𝑎∗

. ∗ ∗ (64)

where

𝑎 = neutral axis of cross section

The limit state function for flexural capacity is generally conveyed as the ultimate

flexural moment due to external loads, 𝑀 , subtracted from the nominal moment flexural

resistance as seen in Equation 65.

𝑔 𝑀 𝑀 𝑀 (65)

With this relationship, a probabilistic model, Equation 66, developed by

Ghanooni-Bagha et al. (2017) was used as the limit state function for evaluation of each

design space.

𝑔 𝑀 𝑀𝐸 ∗ 𝐴 𝑇 ∗ 𝑓 𝑇 ∗ 𝑑 𝐾 ∗∗

∗𝑀 (66)

Where

𝑀𝐸 = flexural model uncertainty coefficient

Page 50: Reliability of Corroding Reinforced Concrete Structures

40

𝐴 𝑇 = residual area of steel reinforcement (m2)

𝑓 𝑇 = residual yielding steel strength of steel reinforcement (N/ m2)

𝑑 = effective depth of cross section (cm)

𝐴 = initial area of steel reinforcement (m2)

𝑏 = width of cross section (m)

𝑓 = compressive strength of concrete (N/ m2)

The residual yielding steel strength is computed using Equation 67.

𝑓 𝑇 1 𝛼 ∗ ∗ 𝑓 (67)

Where

𝛼 = yielding strength uncertainty coefficient

𝑓 = initial yielding strength of steel reinforcement (N/ m2)

The ultimate flexural moment was estimated using the conceptual nature of

reinforced concrete and stated by AASHTO LRFD (2017) in Equations 63 and 68. For a

T-beam, as those in each design space of note within this research, if the neutral axis is

assumed to fall within the flange of the beam, the value of the ‘a’ can be computed

similarly using rectangular beam cross section equations, e.g. Equation 64. Also,

Equation 68 describes that a resistance factor is applied to the nominal moment in order

to find the ultimate moment capacity (AASHTO LRFD, 2017).

𝜙 𝑀 𝑀 (68)

where

𝜙 = flexure resistance factor

From AASHTO LRFD (2017), the ultimate moment could be further broken

down into a general design equation portraying the effects of both dead and live loads.

Page 51: Reliability of Corroding Reinforced Concrete Structures

41

The general AASHTO LRFD design equation, Equation 69 displays how loads,

distribution factors, and ultimate resistance correspond to the strength limit state.

∑ 𝜂 ∗ 𝛾 ∗ 𝑄 ∑ 𝜂 ∗ 𝛾 ∗ 𝑄 𝜑 ∗ 𝑅 𝑅 (69)

Where

𝑛 = load modifier, relating to ductility, redundancy, and operational importance

𝛾 = dead load factor

𝛾 = live load factor

𝑅 = nominal resistance

𝑅 = factored resistance

The modifiers and distribution factors taken from AASHTO LRFD (2017) for a bridge

evaluated for flexure are displayed in Table 5.

Table 5. Load modifiers and factors for LRFD design equation

𝑛 1.00

𝛾 1.25

𝛾 1.75

𝜑 0.9

Furthermore, eleven dead to live load ratios, ranging from 0.5 to 3.0, are put into

place to evaluate the consequences that a more demanding dead load could produce on

the design spaces. Derived from Equation 69, Equations 70 and 71 will herein be used as

the conclusive limit state function.

∑ 𝜂 ∗ 𝛾 ∗ 𝑀 ∑ 𝜂 ∗ 𝛾 ∗ 𝑀 𝜑 ∗ 𝑀 𝑀 (70)

Page 52: Reliability of Corroding Reinforced Concrete Structures

42

𝑔 𝑀 𝑀𝐸 ∗ 𝐴 𝑇 ∗ 𝑓 ∗ 𝑑 𝐾 ∗∗

∗𝑀 𝑀 (71)

3.3. Reliability Analysis

MATLAB, software allowing for matrix multiplication, plotting of functions and

data, implementation of algorithms and symbolic computing amongst many other

capabilities, was primarily used to compute results and outcomes for each respective

design space. Random variables established by Castaneda et al. (2018) were implemented

into the MATLAB software in order to identify the time to corrosion initiation at chloride

surface concentrations for each design space using different parameters, such as concrete

cover and water to cement ratio, using Equation 46.

Similarly, more complex scripts were input into MATLAB software to evaluate

the corrosion propagation period of each design space. Using the FORM method, it is

necessary to have the Symbolic Math Toolbox within the MATLAB software as this

toolbox provides the abilities to perform differentiation, integration, simplification,

transforms, and equation solving. From Equations 66 and 67 within the Design

Approach section above, it is noteworthy that the resistance of the limit state function

decreases with time with respect only to the residual area and yielding strength, also a

function of residual strength, of the steel reinforcement of each respective section. The

aforementioned definitions of residual area will show that this parameter is affected by

the diameter loss of steel reinforcement for general corrosion and the pitting area of steel

reinforcement for pitting corrosion. Both the diameter loss and pitting area are found to

result from the value of corrosion current density with respect to time, icorr(T). From

Equations 48 and 49, the corrosion current density is affected by two constraints within a

cross section, the water to cement ratio and the cover of the steel reinforcement. Provided

Page 53: Reliability of Corroding Reinforced Concrete Structures

43

this information, the effects of three varying concrete covers and three water to cement

ratios were investigated with respect to each design space at varying dead to live load

ratios for both general and pitting corrosion. In total, 54 cases were analyzed for this

study evaluating three different concrete covers at three water to cement ratios for all

three design spaces under both uniform and pitting corrosion cases. In the next section,

results for each case are depicted. In order to easily represent each case, designations for

each scenario are given in order to relay which parameters the case describes. These

parameters include: water to cement ratio (WC), cover (C), and surface chloride content

(Cl). For example, when analyzing design space 2 at a water to cement ratio of 0.46 with

a concrete cover of 13 centimeters and chloride surface concentration of 0.3%, the

designation will read “Design space 2 WC46C13Cl30”.

Table 6 displays the statistical characteristics of random variables within the limit

state function, both uncertainty variables and variables influencing mechanical features,

for each design space.

Page 54: Reliability of Corroding Reinforced Concrete Structures

44

Table 6. Statistical characteristics of random variables for each design space

Parameter Mean Coefficient of

Variation

Distribution Reference

𝑀𝐸 1.1 0.12 Normal Nowak et al.

(2005)

K 0.59 0.05 Normal Stewart (2009)

𝛼 0.5 0.12 Log-Normal Stewart (2004)

𝑓 (MPa) 1.155666615*27.6 0.10 Log-Normal Nowak (1999)

𝑓 (MPa) 1.145*414 0.05 Log-Normal Nowak (1999)

Dead Load

Moment

(N.m.)

1.05*𝑀 0.10 Normal Ellingwood et

al. (1980)

Live Load

Moment

(N.m.)

1.35*𝑀 0.18 Normal Nowak (1999)

d (cm) Vary with each

design space

0.02 Normal Ostlund

(1991)

Concrete

cover (cm)

Vary with each

design space

0.12 Normal Darmawan et

al. (2010)

Page 55: Reliability of Corroding Reinforced Concrete Structures

45

4. RESULTS

Using the methods and design models from Chapter 3, this chapter exhibits the

reaction of steel reinforcement sustaining corrosion and the corresponding changes

to the structural integrity to each design space at varying times. First, the time to

corrosion initiation is evaluated at three different water to cement ratios for varying

chloride surface concentrations. Following this, the results from both general and

pitting corrosion can be seen for multiple water to cement ratios as well as differing

concrete covers.

4.1. Time to Corrosion Initiation

The time to corrosion initiation was estimated using Equation 46 and Tables 1-6.

However, in Equation 46, there are multiple random variables with a normal distribution

type, which allow for irregularity within a design space. Though realistic and

representative to the premise that no design space with the same materials or mechanical

properties will have the same precise time to corrosion initiation, the average value from

10,000 simulations of Equation 46 was utilized for each design space with three varying

concrete covers and three separate water to cement ratios. For the purpose of consistency,

these average values were modeled as the beginning of the life of a structure before

corrosion propagation within instances of general and pitting corrosion, which can be

seen in the following sections. Using MATLAB scripts, the approximate values of time

to corrosion initiation for each considered occurrence can be seen in Figures 12-20

below. Whether the design space undergoes pitting or general corrosion is redundant;

therefore, the only parameters necessary to analyze the time to corrosion initiation is the

water to cement ratio and concrete cover. These figures clearly portray that the time to

Page 56: Reliability of Corroding Reinforced Concrete Structures

46

corrosion initiation decreases with a lower concrete cover and a higher water to cement

ratio.

Figure 12. Time to Corrosion Initiation design space 1 with 12 centimeter cover

Page 57: Reliability of Corroding Reinforced Concrete Structures

47

Figure 13. Time to Corrosion Initiation design space 1 with 15 centimeter cover

Page 58: Reliability of Corroding Reinforced Concrete Structures

48

Figure 14. Time to Corrosion Initiation design space 1 with 18 centimeter cover

Figure 15. Time to Corrosion Initiation design space 2 with 11 centimeter cover

Page 59: Reliability of Corroding Reinforced Concrete Structures

49

Figure 16. Time to Corrosion Initiation design space 2 with 13 centimeter cover

Figure 17. Time to Corrosion Initiation design space 2 with 15 centimeter cover

Page 60: Reliability of Corroding Reinforced Concrete Structures

50

Figure 18. Time to Corrosion Initiation design space 3 with 10 centimeter cover

Figure 19. Time to Corrosion Initiation design space 3 with 12 centimeter cover

Page 61: Reliability of Corroding Reinforced Concrete Structures

51

Figure 20. Time to Corrosion Initiation design space 3 14 with centimeter cover

4.2. Moment values for each Dead to Live Ratio

The dead and live load moments, considered as the load effects of the limit state

function, are estimated for each considered design using Equations 63, 68, and 70. Both

the dead and live loads are represented as random variables within the limit state

function. Different dead to live load ratios were evaluated for each design space, and

respective dead and live load values that satisfy the design equation were computed at

each ratio. Thus, producing new values to each variable already containing randomness,

it is expected that the FORM procedure will produce distinctions in the reliability index

for each ratio because of their difference statistical characteristics. The values for each

dead and live load moment, as well as a consistent ultimate moment value, can be

reviewed for each design space from Tables 7-9.

Page 62: Reliability of Corroding Reinforced Concrete Structures

52

Table 7. Moment loads for design space 1

𝑀 /𝑀   𝑀  (kN*m)  𝑀  (kN*m) 𝑀  (kN*m) 𝜑 ∗ 𝑀

𝑀 (kN*m) 

0.5  1245  2490 

6572  5914 

0.75  1651  2201 

1  1971  1971 

1.25  2232  1785 

1.5  2447  1632 

1.75  2629  1502 

2  2783  1392 

2.25  2917  1296 

2.5  3033  1213 

2.75  3135  1140 

3  3226  1075 

Table 8. Moment loads for design space 2

𝑀 /𝑀   𝑀  (N*m)  𝑀  (N*m)  𝑀  (N*m) 𝜑 ∗ 𝑀 𝑀0.5  736.0  1472 

3885  3496 

0.75  975.7  1301 

1  1165  1165 

1.25  1319  1055 

1.5  1447  964.5 

1.75  1554  887.9 

2  1645  822.6 

2.25  1724  766.3 

2.5  1793  717.2 

2.75  1853  674.0 

3  1907  635.7 

Page 63: Reliability of Corroding Reinforced Concrete Structures

53

Table 9. Moment loads for design space 3

𝑀 /𝑀   𝑀  (N*m)  𝑀  (N*m)  𝑀  (N*m) 𝜑 ∗ 𝑀 𝑀0.5  431.7  863.4 

2278  2051 

0.75  572.3  763.0 

1  683.5  683.5 

1.25  773.8  619.0 

1.5  848.5  565.7 

1.75  911.4  520.8 

2  965.0  482.5 

2.25  1011  449.4 

2.5  1052  420.6 

2.75  1087  395.3 

3  1118  372.8 

4.3. General Corrosion

The first corrosion scenario identified and analyzed was general corrosion. Within

each evaluation, the reliability index of almost each respective setting ranged from

around 4 to 4.5 before the onset of corrosion. These results of general corrosion cases

exhibit high rates of corrosion initially after the conclusion of the corrosion initiation

period; however, the corrosion rate drops to moderate levels throughout the corrosion

propagation phase, and therefore diminishes the reliability index in decelerating fashion.

This incidence can be explained by a previously stated assertion by Vu and Stewart

(2000) that the development of rust-like products on the steel surface can essentially

reduce the corrosion rate with progression of time. Therefore, as the corrosion rate

decreased at a decelerating rate, the residual area, correlating residual yield strength, and

the correlated flexural moment resistance also decreased at a slowing rate. As expected,

the results seen in Figures 21-47 display that a smaller concrete cover and a higher water

Page 64: Reliability of Corroding Reinforced Concrete Structures

54

to cement ratio will have a significant impact on the reliability index and yield a higher

probability of failure. Comparing the evaluations of three different concrete covers and

water to cement ratios for each respective design space at the most resilient dead to live

ratio of 0.5 in Figures 48-62, only ten of the twenty-seven assessments featured a

reliability index fall below 0, correlating to a 50% probability of failure, after T=75 years

of corrosion propagation.

Figure 21. Reliability Index vs. Time design space 1 WC46C12Cl35

Page 65: Reliability of Corroding Reinforced Concrete Structures

55

Figure 22. Reliability Index vs. Time design space 1 WC50C12Cl35

Figure 23. Reliability Index vs. Time design space 1 WC53C12Cl35

Page 66: Reliability of Corroding Reinforced Concrete Structures

56

Figure 24. Reliability Index vs. Time design space 1 WC46C15Cl35

Figure 25. Reliability Index vs. Time design space 1 WC50C15Cl35

Page 67: Reliability of Corroding Reinforced Concrete Structures

57

Figure 26. Reliability Index vs. Time design space 1 WC53C15Cl35

Figure 27. Reliability Index vs. Time design space 1 WC46C18Cl35

Page 68: Reliability of Corroding Reinforced Concrete Structures

58

Figure 28. Reliability Index vs. Time design space 1 WC50C18Cl35

Figure 29. Reliability Index vs. Time design space 1 WC53C18Cl35

Page 69: Reliability of Corroding Reinforced Concrete Structures

59

Figure 30. Reliability Index vs. Time design space 2 cover WC46C11Cl30

Figure 31. Reliability Index vs. Time design space 2 WC50C11Cl30

Page 70: Reliability of Corroding Reinforced Concrete Structures

60

Figure 32. Reliability Index vs. Time design space 2 WC53C11Cl30

Figure 33. Reliability Index vs. Time design space 2 WC46C13Cl30

Page 71: Reliability of Corroding Reinforced Concrete Structures

61

Figure 34. Reliability Index vs. Time design space 2 WC50C13Cl30

Figure 35. Reliability Index vs. Time design space 2 WC53C13Cl30

Page 72: Reliability of Corroding Reinforced Concrete Structures

62

Figure 36. Reliability Index vs. Time design space 2 WC46C15Cl30

Figure 37. Reliability Index vs. Time design space 2 WC50C15Cl30

Page 73: Reliability of Corroding Reinforced Concrete Structures

63

Figure 38. Reliability Index vs. Time design space 2 WC53C15Cl30

Figure 39. Reliability Index vs. Time design space 3 WC46C10Cl30

Page 74: Reliability of Corroding Reinforced Concrete Structures

64

Figure 40. Reliability Index vs. Time design space 3 WC50C10Cl30

Figure 41. Reliability Index vs. Time design space 3 WC53C10Cl30

Page 75: Reliability of Corroding Reinforced Concrete Structures

65

Figure 42. Reliability Index vs. Time design space 3 WC46C12Cl30

Figure 43. Reliability Index vs. Time design space 3 WC50C12Cl30

Page 76: Reliability of Corroding Reinforced Concrete Structures

66

Figure 44. Reliability Index vs. Time design space 3 WC53C12Cl30

Figure 45. Reliability Index vs. Time design space 3 WC46C14Cl30

Page 77: Reliability of Corroding Reinforced Concrete Structures

67

Figure 46. Reliability Index vs. Time design space 3 WC50C14Cl30

Figure 47. Reliability Index vs. Time design space 3 WC53C14Cl30

Page 78: Reliability of Corroding Reinforced Concrete Structures

68

Figure 48. Reliability Index vs. Concrete Cover design space 1 at T=0 years

Figure 49. Reliability Index vs. Concrete Cover design space 1 at T=10 years

Page 79: Reliability of Corroding Reinforced Concrete Structures

69

Figure 50. Reliability Index vs. Concrete Cover design space 1 at T=25 years

Figure 51. Reliability Index vs. Concrete Cover design space 1 at T=50 years

Page 80: Reliability of Corroding Reinforced Concrete Structures

70

Figure 52. Reliability Index vs. Concrete Cover design space 1 at T=75 years

Figure 53. Reliability Index vs. Concrete Cover design space 2 at T=0 years

Page 81: Reliability of Corroding Reinforced Concrete Structures

71

Figure 54. Reliability Index vs. Concrete Cover design space 2 at T=10 years

Figure 55. Reliability Index vs. Concrete Cover design space 2 at T=25 years

Page 82: Reliability of Corroding Reinforced Concrete Structures

72

Figure 56. Reliability Index vs. Concrete Cover design space 2 at T=50 years

Figure 57. Reliability Index vs. Concrete Cover design space 2 at T=75 years

Page 83: Reliability of Corroding Reinforced Concrete Structures

73

Figure 58. Reliability Index vs. Concrete Cover design space 3 at T=0 years

Figure 59. Reliability Index vs. Concrete Cover design space 3 at T=10 years

Page 84: Reliability of Corroding Reinforced Concrete Structures

74

Figure 60. Reliability Index vs. Concrete Cover design space 3 at T=25 years

Figure 61. Reliability Index vs. Concrete Cover design space 3 at T=50 years

Page 85: Reliability of Corroding Reinforced Concrete Structures

75

Figure 62. Reliability Index vs. Concrete Cover design space 3 at T=75 years

There are optimum reliabilities that should be decided on before construction,

both a target reliability index for new structures as well as target lower reliability indices

for existing structures due to economic considerations. Today, these target reliability

indices calculated for beams and columns for buildings and bridges can be found in ACI

318 and AASHTO LRFD design codes (Kaszynska and Nowak, 2005). If the reliability

index is too small, there can be problems including structural failures. If the relibality

index is too large, the structures could be too expensive. The selection criteria for a target

reliability stem from: consequences of failure, economic analysis, previous practices,

human consideration, and social/political decisions (Kaszynska and Nowak, 2005).

Frangopol (1997) suggests that an optimal lifetime repair strategy for existing structures

should be developed by minimizing the expected lifetime repair costs and preserving a

prescribed degree of system safety, found from the reliability index, throughout the

Page 86: Reliability of Corroding Reinforced Concrete Structures

76

remaining service life of the structure. As used by Frangopol (1997), a minimum

allowable, or target, reliability index considered for the design spaces in this research of

2.0 was established. Therefore, anytime the reliability of the evaluated design spaces fell

below this prescribed base reliability, some type of rehabilitation or replacement should

be made. Figures 63-65 give the time after corrosion initiation to this reliability threshold

is reach for each 27 design scenarios with respect to general corrosion. This time appears

to be heavily reliant on the concrete cover for each evaluation, as the varying water to

cement ratios seem to influence the time to reach these minimum reliability values by

only a span of a couple of years for each evaluation. However, the concrete cover seems

to be the more consequential factor as a lower cover shows much quicker times to reach

this minimum reliability threshold, as seen in the below figures.

Figure 63. Time to Reliability Index Threshold design space 1

Page 87: Reliability of Corroding Reinforced Concrete Structures

77

Figure 64. Time to Reliability Index Threshold design space 2

Figure 65. Time to Reliability Index Threshold design space 3

Page 88: Reliability of Corroding Reinforced Concrete Structures

78

4.4. Pitting Corrosion

The pitting factor is derived using the cumulative density function and its

respective probability density function from Equation 53. The developed script calculates

the location parameters of these functions across a region of values, which are then used

to provide information into Equation 56 for the maximum pit depth regarding instances

of pitting corrosion. Similar to time to corrosion initiation, the pitting factor is modeled

by Gumbel distribution. For conservative purposes, the location parameter of the pitting

factor for each design space was assumed at a 50% value of the PDF. With an assumed

location parameter, the mean value of the pitting factor was found using Equation 23. The

PDF of the location scale parameter and its correlated value of pitting factors from each

design space are seen in Figure 66 and Table 10.

Figure 66. PDF vs. Location Parameter of Pitting Factor

Page 89: Reliability of Corroding Reinforced Concrete Structures

79

Table 10. Pitting factor for each design space

Scenarios of pitting corrosion were studied to analyze more extreme corrosion

situations resulting in higher pitting depths and less residual area over time, which can be

seen in Figures 67-93. Similar to the situations of general corrosion during the corrosion

initiation period, the reliability index falls around 4 to 4.5 for each evaluated setting at

each design space initially before the reinforcement starts corroding and still has all of its

steel area. The concrete cover and water to cement ratio are all the more important in

these occasions of corrosion since, unlike general corrosion, the corrosion rate accelerates

over time, catalyzing expedited losses of steel reinforcement with each successive year.

The rapid decrease of residual area lowers the likelihood that there will be enough

flexural strength resisting the flexural stresses. This is evident in Figures 94-108 as

eighteen of the twenty-seven evaluations display a reliability index below -1, correlating

to a probability of failure of about 85% probability of failure, after only T=50 years after

the conclusion of corrosion initiation. Furthermore, twenty-four of the twenty-seven

evaluations portray a probability of failure of at least 95% after T=75 years of

propagation.

Design Space Pitting Factor, R Standard Deviation, 𝜎

1 12.1196

1.2569 2 11.9010

3 11.5028

Page 90: Reliability of Corroding Reinforced Concrete Structures

80

Figure 67. Reliability Index vs. Time design space 1 WC46C12Cl35

Figure 68. Reliability Index vs. Time design space 1 WC50C12Cl35

Page 91: Reliability of Corroding Reinforced Concrete Structures

81

Figure 69. Reliability Index vs. Time design space 1 WC53C12Cl35

Figure 70. Reliability Index vs. Time design space 1 WC46C15Cl35

Page 92: Reliability of Corroding Reinforced Concrete Structures

82

Figure 71. Reliability Index vs. Time design space 1 WC50C15Cl35

Figure 72. Reliability Index vs. Time design space WC53C15Cl35

Page 93: Reliability of Corroding Reinforced Concrete Structures

83

Figure 73. Reliability Index vs. Time design space 1 cover WC46C18Cl35

Figure 74. Reliability Index vs. Time design space 1 WC50C18Cl35

Page 94: Reliability of Corroding Reinforced Concrete Structures

84

Figure 75. Reliability Index vs. Time design space 1 WC53C18Cl35

Figure 76. Reliability Index vs. Time design space 2 WC46C11Cl30

Page 95: Reliability of Corroding Reinforced Concrete Structures

85

Figure 77. Reliability Index vs. Time design space 2 WC50C11Cl30

Figure 78. Reliability Index vs. Time design space 2 WC53C11Cl30

Page 96: Reliability of Corroding Reinforced Concrete Structures

86

Figure 79. Reliability Index vs. Time design space 2 WC46C13Cl30

Figure 80. Reliability Index vs. Time design space 2 WC50C13Cl30

Page 97: Reliability of Corroding Reinforced Concrete Structures

87

Figure 81. Reliability Index vs. Time design space 2 WC53C13Cl30

Figure 82. Reliability Index vs. Time design space 2 WC46C15Cl30

Page 98: Reliability of Corroding Reinforced Concrete Structures

88

Figure 83. Reliability Index vs. Time design space 2 WC50C15Cl30

Figure 84. Reliability Index vs. Time design space 2 WC53C15Cl30

Page 99: Reliability of Corroding Reinforced Concrete Structures

89

Figure 85. Reliability Index vs. Time design space WC46C10Cl30

Figure 86. Reliability Index vs. Time design space 3 WC50C10Cl30

Page 100: Reliability of Corroding Reinforced Concrete Structures

90

Figure 87. Reliability Index vs. Time design space 3 WC53C10Cl30

Figure 88. Reliability Index vs. Time design space 3 WC46C12Cl30

Page 101: Reliability of Corroding Reinforced Concrete Structures

91

Figure 89. Reliability Index vs. Time design space 3 WC50C12Cl30

Figure 90. Reliability Index vs. Time design space 3 WC53C12Cl30

Page 102: Reliability of Corroding Reinforced Concrete Structures

92

Figure 91. Reliability Index vs. Time design space 3 WC46C14Cl30

Figure 92. Reliability Index vs. Time design space 3 WC50C14Cl30

Page 103: Reliability of Corroding Reinforced Concrete Structures

93

Figure 93. Reliability Index vs. Time design space 3 WC53C14Cl30

Figure 94. Reliability Index vs. Concrete Cover design space 1 at T=0 years

Page 104: Reliability of Corroding Reinforced Concrete Structures

94

Figure 95. Reliability Index vs. Concrete Cover design space 1 at T=10 years

Figure 96. Reliability Index vs. Concrete Cover design space 1 at T=25 years

Page 105: Reliability of Corroding Reinforced Concrete Structures

95

Figure 97. Reliability Index vs. Concrete Cover design space 1 at T=50 years

Figure 98. Reliability Index vs. Concrete Cover design space 1 at T=75 years

Page 106: Reliability of Corroding Reinforced Concrete Structures

96

Figure 99. Reliability Index vs. Concrete Cover design space 2 at T=0 years

Figure 100. Reliability Index vs. Concrete Cover design space 2 at T=10 years

Page 107: Reliability of Corroding Reinforced Concrete Structures

97

Figure 101. Reliability Index vs. Concrete Cover design space 2 at T=25 years

Figure 102. Reliability Index vs. Concrete Cover design space 2 at T=50 years

Page 108: Reliability of Corroding Reinforced Concrete Structures

98

Figure 103. Reliability Index vs. Concrete Cover design space 2 at T=75 years

Figure 104. Reliability Index vs. Concrete Cover design space 3 at T=0 years

Page 109: Reliability of Corroding Reinforced Concrete Structures

99

Figure 105. Reliability Index vs. Concrete Cover design space 3 at T=10 years

Figure 106. Reliability Index vs. Concrete Cover design space 3 at T=25 years

Page 110: Reliability of Corroding Reinforced Concrete Structures

100

Figure 107. Reliability Index vs. Concrete Cover design space 3 at T=50 years

Figure 108. Reliability Index vs. Concrete Cover design space 3 at T=75 years

Page 111: Reliability of Corroding Reinforced Concrete Structures

101

As previously mentioned in Section 4.3., a minimum allowable reliability index

was discussed for instances considering general corrosion. The same target reliability

index of 2.0 was considered for pitting corrosion scenarios as well. The time, from the

conclusion of corrosion initiation to the time of exceeding the allowable reliability

threshold, is shown for each pitting corrosion design scenario below in Figures 109-111.

Unlike general corrosion, the water to cement ratio and concrete cover appear to affect

the time to the minimum reliability in a more balanced fashion due to the detrimental

nature of rapid reinforcement loss in pitting corrosion.

Figure 109. Time to Reliability Index Threshold design space 1

Page 112: Reliability of Corroding Reinforced Concrete Structures

102

Figure 110. Time to Reliability Index Threshold design space 2

Figure 111. Time to Reliability Index Threshold design space 3

Page 113: Reliability of Corroding Reinforced Concrete Structures

103

5. SUMMARY AND CONCLUSION

5.1. Summary

Structural reliability methods can provide a probabilistic risk assessment of the

structural integrity of infrastructure and offer insight on the vulnerability of structures

corresponding to particular levels of demand they must resist. There are numerous

statistical methods and simulation techniques that employ the reliability theories to

approximate the structural reliability of a structure during its service life. In this study,

typical reinforced concrete girders were analyzed using the First Order Reliability

Method from three different bridge configurations undergoing particular corrosion

conditions to assess how the mechanical and material properties would affect the

reliability of each respective element over its’ intended service life.

5.2. Conclusions

The results above clearly show the impact of the concrete cover and water to

cement ratio and how these parameters can affect other mechanical properties and

performance of a structure over time within corrosion processes. Increasing the concrete

cover is the most inexpensive and practically feasible way to protect the reinforcing steel

from corrosion and sustain its’ structural integrity over a prolonged period. It is crucial to

identify the nature of corrosion that reaches steel reinforcement as soon as possible in

order to best evaluate methods for repair, especially in instances of pitting corrosion due

to the rapid decline of residual area that can happen in areas of importance.

5.3. Recommendations for Future Research

The proposed methodology could be used as a decision-making tool that allow

infrastructure owners and managers to prioritize repair needs and budget allocations

Page 114: Reliability of Corroding Reinforced Concrete Structures

104

based on time at which a certain reliability threshold will be reached. Additional research

should be done on this topic to evaluate the reliability of a structure with increased

accuracy. First and foremost, this research only analyzed the loss steel reinforcement and

its effects on the flexural strength of structures undergoing corrosion mechanisms. This

framework can be taken a step further by identifying manners on how corrosion impairs

the surrounding concrete strength as well. With regards to this exclusive study, the

definition of the chloride threshold should be specified more clearly in order to be able to

accurately quantify this variable. Also, the model used in this research suggested that the

pitting factor was deterministic for each bar of steel reinforcement across the entirety of

the span length. This is an area within pitting corrosion that can see development, as it is

not likely that each individual bar will have similar pitting depths. Another future

calibration could also include contrasting current densities within the cross section of a

featured area, where reinforcement closer to edges will likely receive a higher corrosion

density and thus, a higher probability of failure for those particular elements. Lastly,

effects of environmental conditions such as temperature, humidity, and location could be

studied to reflect how the climate of a structures’ location will influences serviceability.

Page 115: Reliability of Corroding Reinforced Concrete Structures

105

REFERENCES

AASHTO LRFD Bridge Design Specifications. Washington, D.C. :American Association

of State Highway and Transportation Officials, 2017. Print.

Allen T.M., Nowak A.S., Bathurst R.J. 2005. Calibration to Determine Load and

Resistance Factors for Geotechnical and Structural Design. Transportation

Research E-Circular(E-C079).

Almusallam A A. Effect of degree of corrosion on the properties of reinforcing steel bars.

Construction and Building Materials 2001; 15(8): 361-368,

https://doi.org/10.1016/S0950-0618(01)00009-5.

Bastidas-Arteaga E, Bressolette P, Chateauneuf A, Sánchez-Silva M. Probabilistic

lifetime assessment of RC structures under coupled corrosion–fatigue

deterioration processes. Structural Safety 2009; 31(1): 84-96,

https://doi.org/10.1016/j.strusafe.2008.04.001.

Bastidas-Arteaga E, Sánchez-Silva M, Chateauneuf A, Silva M R. Coupled reliability

model of biodeterioration, chloride ingress and cracking for reinforced concrete

structures. Structural Safety 2008; 30(2): 110-129,

https://doi.org/10.1016/j.strusafe.2006.09.001.

Bertolini, L., Elsener, B., Pedeferri, P., Redaelli, E., and Polder, R. (2018). Corrosion of

Steel in Concrete - Prevention, Diagnosis, Repair (2nd Edition), John Wiley &

Sons.

Bhargava, K., Mori, Y., and Ghosh, A. K. (2011). "Time-dependent reliability of

corrosion-affected RC beams—Part 1: Estimation of time-dependent strengths

and associated variability." Nuclear Engineering and Design, 241, 1371-1384.

Böhni, H. (2005). Corrosion in reinforced concrete structures, Woodhead ;

Page 116: Reliability of Corroding Reinforced Concrete Structures

106

CRC Press, Cambridge, England Boca Raton.

BRITE/EURAM. (1995). “The residual service life of reinforced concrete structures.” Final

Technical Rep. BRUE-CT92-0591, Brite/EURAM, European Union.

Bushman J B, Engineer P P. Calculation of Corrosion Rate from Corrosion Current

(Faraday's Law). Bushman & Associates Inc. 2000.

Cairns, J., Plizzari, G. A., Du, Y. G., Law, D. W., and Franzoni, C. (2005). "Mechanical

properties of corrosion-damaged reinforcement." 256-264.

Castaneda, Homero; Karsilaya, Aydin; Okeil, Ayman; and Taha, Mahmoud Reda,

(2018). Publications. 26. https://digitalcommons.lsu.edu/transet_pubs/26.

Castro-Borges, P., Balancán-Zapata, M., López-González, A. Analysis of Tools to

Evaluate Chloride Threshold for Corrosion Onset of Reinforced Concrete in

Tropical Marine Environment of Yucatán, México, J. Chem. 2013 (2013) 1–8.

doi:10.1155/2013/208619.

Cicek, V.: Corrosion Engineering and Cathodic Protection Handbook. Wliey, Hoboken

(2017).

Dai H, Wang W. Application of low-discrepancy sampling method in structural

reliability analysis. Structural Safety 2009; 31(1): 55-64,

https://doi.org/10.1016/j.strusafe.2008.03.001.

Darmawan MS, Stewart MG. Spatial variability of pitting corrosion and its effect on the

reliability of prestressing wires. In: Der Kiureghain A, Madanat S, Pestana JM,

editors. Nineth International Conference on Applications of Statistics and

Probability in Civil Engineering, 1. Rotterdam: Millpress; 2003. p. 541–8.

Darmawan, M. S. (2010). "Pitting corrosion model for reinforced concrete structures in a

Page 117: Reliability of Corroding Reinforced Concrete Structures

107

chloride environment." 91-101.

Dolinski, K. 1982. First Order Second Moment Approximation in Reliability of

Structural Systems: Critical Review and Alternative Approach.

Drewett, J.; Broomfield, J.: An introduction to electrochemical rehabilitation techniques.

Technical Note 2. CPA Technical Note (2011).

DuraCrete. Probabilistic performance based durability design of concrete structures. The

European Union-Brite EuRam III Project BE95-1347; 2000.

Ellingwood B. Development of a probability based load criterion for American National

Standard A58: Building code requirements for minimum design loads in

buildings and other structures. US Department of Commerce, National Bureau of

Standards; 1980, https://doi. org/10.6028/nbs.sp.577.

Elsener, B., and Angst, U. (2007). "Mechanism of electrochemical chloride removal."

Corrosion Science, 49, 4504-4522.

Frangopol, Dan M., and Allen C. Estes. “Lifetime Bridge Maintenance Strategies Based

on System Reliability.” Structural Engineering International, vol. 7, no. 3, Aug.

1997, p. 193.

Geiker, M. R., and Polder, R. B. (2016). "Experimental support for new electro active

repair method for reinforced concrete." Materials & Corrosion / Werkstoffe und

Korrosion, 67(6), 600-606.

Ghanooni-Bagha, M., Shayanfar, M. A., Reza-Zadeh, O., and Zabihi-Samani, M. (2017).

"The Effect Of Materials On The Reliability Of Reinforced Concrete Beams In

Normal And Intense Corrosions." Wpływ Stosowanych Materiałów Na

Page 118: Reliability of Corroding Reinforced Concrete Structures

108

Niezawodność Belek Żelbetowych W Warunkach Normalnej I Silnej Korozji.,

19(3), 393.

Gonzalez JA, Andrade C, Alonso C, Feliu S. Comparison of rates of general corrosion

and maximum pitting penetration on concrete embedded steel reinforcement.

Cement Concrete Research1995;25(2):257–64.

Goyal, A., Pouya, H. S., Ganjian, E., Claisse, P. J. A. J. f. S., and Engineering (2018). "A

Review of Corrosion and Protection of Steel in Concrete." 43(10), 5035-5055.

Hasaan, J., Bressolette P., Cheateauneuf, A., El Tawil, K. Reliability-based assessment of

the effect of climatic conditions on the corrosion of RC structures subject to

chloride ingress. 2010.

Heiyantuduwa, R., Alexander, M. G., and Mackechnie, J. R. (2006). "Performance of a

Penetrating Corrosion Inhibitor in Concrete Affected by Carbonation-Induced

Corrosion." Journal of Materials in Civil Engineering, 18(6), 842-850.

Holicky, 2009. Reliability Analysis for Structural Design.

Kaszyńska M.,Nowak A.S.,TargetReliabilityforDesignandEvaluationof Bridges,

Proceedings of the Conference on Bridge Management 5, ed. Parke, G.A.R. and

Disney, P. at University of Surrey, U.K., April 2005, pp. 401-408.

Kaukaas. 2018. Mean Value First Order Second Moment Method (MVFOSM).

Koch, G. H. et al. Corrosion Cost and Preventative Strategies in the United States.

Report No. FHWA-RD-01-156, 773 (NACE International, Houston, 2002).

Liberati, E. A. P., Leonel, E. D., and Nogueira, C. G. (2014). "Influence of the

Page 119: Reliability of Corroding Reinforced Concrete Structures

109

reinforcement corrosion on the bending moment capacity of reinforced concrete

beams: a structural reliability approach." Revista IBRACON de Estruturas e

Materiais, Vol 7, Iss 3, Pp 379-413(3), 379.

Liu T, Weyers RE. Modeling the dynamic corrosion process in chloride contaminated

concrete structures. Cement and Concrete Research 1998;28(3):365-79.

Markeset and Mydral. Modeling of Reinforcement Corrosion in Concrete – State of the

Art. 2008.

Marques, P.F.; Costa, A.: Service life of RC structures: carbonation induced corrosion.

Prescriptive vs. performance-based methodologies. Constr. Build. Mater. 24, 258–

265 (2010). https://doi.org/10. 1016/j.conbuildmat.2009.08.039

Nogueria et al. Probabilistic Failure Modeling of Reinforced Concrete Structures

Subjected to Chloride. 2012.

Nóvoa, X. (2016). "Electrochemical aspects of the steel-concrete system. A review."

Journal of Solid State Electrochemistry, 20(8), 2113-2125.

Nowak, A. S. (1999). “Calibration of LRFD bridge design code.” NCHRP Rep. No. 368,

Transportation Research Board, Washington, DC, 37.

Nowak, A. S., and Collins, K. R. (2000). Reliability of structures, McGraw- Hill, New

York.

Nowak, A. S., and Szerszen, M. M. (2003). “Calibration of design code

for buildings (ACI 318): Part 1. Statistical models for resistance.”

ACI Struct. J., 100(3), 377–382.

Okeil, Ayman M., et al. “Reliability Assessment of FRP-Strengthened Concrete Bridge

Page 120: Reliability of Corroding Reinforced Concrete Structures

110

Girders in Shear.” Journal of Composites for Construction, vol. 17, no. 1, Jan.

2013, pp. 91-100. EBSCOhost, doi:10.1061/(ASCE)CC.1943-5614.0000315.

Östlund L. An estimation of γ-values. Bulletin du Comité Euro-international du Béton

1991(202): 38-97.

Popov, B. N. (2015). "Chapter 1 - Evaluation of Corrosion." Corrosion Engineering, B.

N. Popov, ed., Elsevier, Amsterdam, 1-28.

Poupard, O., Aït-Mokhtar, A., and Dumargue, P. (2004). "Corrosion by chlorides in

reinforced concrete: Determination of chloride concentration threshold by

impedance spectroscopy." Cement & Concrete Research, 34(6), 991.

Shayanfar, M. A., Barkhordari, M. A., and Ghanooni-Bagha, M. (2015). "Estimation of

Corrosion Occurrence in RC Structure Using Reliability Based PSO

Optimization." 531-542.

Stewart MG. Spatial variability of pitting corrosion and its influence on structural

fragility and reliability of RC beams in flexure. Structural Safety 2004;26(4):453–

70.

The Concrete Society: Guidance on the use of stainless steel reinforcement. Concrete

Society, TR 51 (1998)

Turnbull, A. (1993). “Review of modeling of pit propagation kinetics.” Br. Corros. J.,

London, 28(4), 297–308.

Tutti, K. (1982). "Corrosion of Steel in Concrete " CBI Swedish Cement and Concrete

Institute 469.

Val DV. Deterioration of strength of RC beams due to corrosion and its influence on

Page 121: Reliability of Corroding Reinforced Concrete Structures

111

beam reliability. J Struct Eng, ASCE 2007;133(9):1297–306

Val DV, Melchers RE. Reliability of deteriorating RC slab bridges. J Struct Eng ASCE

1997;123(12):1638–44.

Vu K, Stewart M G, Mullard J. Corrosion-induced cracking: experimental data and

predictive models. ACI Structural Journal 2005; 102(5): 719-726.

Vu K A, Stewart M G. Structural reliability of concrete bridges including improved

chloride-induced corrosion models. Structural Safety 2000; 22(4): 313-333,

https://doi.org/10.1016/S0167-4730(00)00018-7.

Wong. 1984. First-Order, Second Moment Methods.

Yalsyn H, Ergun M. The prediction of corrosion rates of reinforcing steels in concrete.

Cement and Concrete Research 1996;26(10):1593-9.

Zhou, Y., Gencturk, B., Willam, K., and Attar, A. (2015). "Carbonation-Induced and

Chloride-Induced Corrosion in Reinforced Concrete Structures." Journal of Materials in Civil Engineering, 27(9), 04014245.

Page 122: Reliability of Corroding Reinforced Concrete Structures

112

VITA

Travis Honore’ was born in Baton Rouge, LA, in 1995. Raised in Baton Rouge,

Travis knew he wanted to attend Louisiana State University after spending most of his

life being a fan of LSU sports. After graduating from Catholic High, an all-boys high

school not too far from LSU campus, in 2013, Travis began his undergraduate studies in

the Fall. He obtained his Bachelor’s Degree in Civil Engineering in the Spring of 2017

and was encouraged to further his education by friends, family, and professional peers.

Beginning the Graduate School curriculum for a Master’s in Science in Civil

Engineering, Travis is expected to receive his degree in August of 2019.