Ayman Ababneh PHD Thesis

Download as pdf or txt
Download as pdf or txt
You are on page 1of 226

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

net/publication/342563852

The Coupled Effect of Moisture Diffusion, Chloride Penetration, and Freezing-


Thawing on Concrete Durability

Thesis · December 2002

CITATIONS READS

15 457

1 author:

Ayman N. Ababneh
Jordan University of Science and Technology
72 PUBLICATIONS 1,142 CITATIONS

SEE PROFILE

All content following this page was uploaded by Ayman N. Ababneh on 01 July 2020.

The user has requested enhancement of the downloaded file.


THE COUPLED EFFECT OF MOISTURE DIFFUSION CHLORIDE
PENETRATION AND FREEZING-THAWING ON CONCRETE
DURABILITY

By

Ayman Nureddin Ababneh

B.Sc., Jordan University of Science and Technology, 1991

M.Sc., Jordan University of Science and Technology, 1994

Irbid – Jordan

A thesis submitted to the

Faculty of the Graduate School of the

University of Colorado in partial fulfillment of the requirement for the degree of

Doctor of Philosophy

Department of Civil, Environmental and Architectural Engineering

2002
ii

This dissertation entitled:

The Coupled Effect of Moisture Diffusion, Chloride Penetration and Freezing-Thawing


On Concrete Durability

written by Ayman Nureddin Ababneh

has been approved for the Department of

Civil, Environmental and Architectural Engineering

________________________________
Yunping Xi

________________________________
Kaspar Willam

________________________________
Dan Frangopol

________________________________
Benson Shing

________________________________
Kevin Rens

Date_______________________

The final copy of this thesis has been examined by the signatories, and we find that both
the content and the form meet acceptable presentation standards of scholarly work in the
above mentioned discipline.

HRC protocol # ___________________


iii

Ababneh, Ayman Nureddin (Ph.D., Civil, Environmental, and Architectural Engineering)

The Coupled Effect of Moisture Diffusion, Chloride Penetration and Freezing-Thawing

on Concrete Durability.

Thesis directed by Associate Professor Yunping Xi

In this thesis, both theoretical and experimental studies were conducted on some

of the processes related to concrete durability. Those processes include moisture

diffusion, chloride penetration, freezing-thawing and their coupled effects.

The first part of the study includes theoretical modeling of the coupled effect of

moisture diffusion and drying shrinkage and modeling of chloride penetration in partially

saturated concrete. This part encompasses the formulation of the governing equations and

boundary conditions, identification and modeling of the material parameters and the

numerical solution. The governing equations of moisture diffusion and chloride

penetration in concrete were characterized by the nonlinear diffusion equation derived

based on Ficks’ laws. The material parameters involved in the governing equations and

the drying shrinkage of concrete were modeled based on a multiscale method. In the

multiscale method, different theoretical models are used at different scale levels.

Diffusion mechanisms and shrinkage mechanisms are considered at the nanoscale; a

composite mechanics method (generalized self-consistent method, in particular) is used at

the microscale and mesoscale; and continuum approach is applied at the macroscale. The

numerical models were solved by the Alternating-Direction Implicit (ADI) finite

difference method. Several numerical examples were solved by the proposed models and
iv

the results obtained predicted well the basic trends and the effects of concrete material

parameters. The models can be used in the durability studies of concrete structures to

predict concrete deterioration and the onset of steel corrosion and thus the service life and

the time for rehabilitation.

The second part of this study includes experiments on the coupled effect of freezing-

thawing and chloride penetration, effect of concrete material parameters on chloride

penetration, the coupling effect of moisture diffusion and chloride penetration and the

effect of shrinkage-induced cracking on chloride penetration. The experimental study

presented the effect of water-cement ratio and aggregate volume fraction on chloride

penetration, the two-way coupling between moisture diffusion and chloride penetration

and the effect of the damage induced by drying shrinkage and freezing-thawing on

chloride penetration. The experimental results obtained in this study can be used to

validate and demonstrate the accuracy of the numerical models developed in this study

and the future studies.


v

DEDICATION

To my father, mother, wife, bothers and sisters and to my whole family who made

this work possible.


vi

ACKNOWLEDGMENT

I would like to express my deep appreciation and thanks to my advisor Professor

Yunping Xi for his support, guidance and patience during the course of my study. His

fruitful suggestions and valuable ideas have been encouraging and subsidizing this work.

Grateful acknowledgment is made to Professors Kaspar Willam, Dan Frangopol,

Benson Shing and Kevin Rens for reviewing this work and for their valuable suggestions

and comments.

I would like to extend my appreciation to my colleagues Ammar Nakhi, Zhaohui

Xie, A. Suito, Renee Cusson, Andi Asiz and our former colleague Dr. Young-Sook Roh

for their cooperation, help and encouragement. Also my sincere thanks are extended to

my friends Hassan Al-sukhni, Ahmad Ababneh, Mahmoud Shaheen, Fadi Mansour and

my sister in law Rula Abu-Dalo for their support and encouragement during my study.

I shall always be grateful to my parents, my wife and my family for their love and

encouragement without which this work has never been accomplished.

Finally the financial support under NSF grant CMS-9872379 to University of

Colorado at Boulder is gratefully acknowledged.


vii

TABLE OF CONTENTS

CHAPTER Page

1 INTRODUCTION

1.1 Durability of Concrete Structures …….………………………………1


1.2 Objectives of the Study …………………...……………………….….5
1.3 Thesis Organization …………..……………..………...………...……6

2 BACKGROUND INFORMATION AND LITERATURE REVIEW

2.1 Transport Properties of Concrete …..…………………..……………10


2.1.1 Concrete permeability ………...……..………………………12
2.1.2 Concrete diffusivity ……..…..…………..…………………..15
2.1.3 Concrete sorptivity …..…….….………………..…...……….16
2.2 Chloride Ions Penetration into Concrete ……...………….….………18
2.2.1 Mechanism of the reinforcement corrosion …….……….…..20
2.2.2 Onset of the steel corrosion ……………………………….…21
2.3 Heat Transfer in Concrete ….....…………………………………….22

3 COUPLED MOISTURE DIFUSSION AND DRYING SHRINKAGE

3.1 Introduction ……….. ….………………………………......……..… 25


3.2 Modeling Moisture Diffusion in Concrete ……………………....…..27
3.2.1 Basic formulation of the governing equation ….…………….27
3.2.2 Initial and boundary conditions …………………………..…29
3.2.3 Material parameters ………………………………………....30
3.3 Numerical Solution …………...…………………….…….................32
3.3.1 Finite element method …………………...……………….... 32
3.3.2 Finite difference method …...…………….…...…………......35
3.4 Multiscale Modeling of Drying Shrinkage ….……………....……...37
3.5 The Interaction of Drying Shrinkage and Moisture Diffusion.............44
3.5.1 Effect of shrinkage-induced damage on diffusion coefficient.45
3.5.2 Effect of drying shrinkage on moisture capacity……...…......49
3.6 Numerical Applications ..…………………………………................53
viii

3.6.1 Effect of damage ………………………………………….. 61


3.6.2 Effect of water-cement ratio ……………………………… 63
3.6.3 Effect of aggregate volume fraction ………………………. 65
3.7 Conclusion ………………………………………………………… 67

4 COUPLED MOISTURE AND CHLORIDE DIFFUSIONS

4.1 Introduction ………..……………………………………….……… 69


4.2 Basic Formulation ……………………………………………….… 71
4.3 Material Parameters ……………………………………………..… 74
4.3.1 Chloride binding capacity ……….………………………… 75
4.3.2 Chloride diffusion coefficient…………………………........ 80
4.4 Numerical Solution ………………………………………...……… 83
4.5 Numerical Applications and Discussions ………………….……… 85
4.5.1 Numerical example of chloride penetration ….…………… 85
4.5.2 Comparison with test data………… ..………………...…… 88
4.5.3 Effect of the moisture diffusion on chloride penetration…... 88
4.5.4 The Effect of the concrete mix parameters………………….91
4.6 Conclusions ………………………………………...……………… 96

5 THE COUPLED EFFECT OF FREEZING-THAWING AND CHLORIDE


PENETRATION IN CONCRETE

5.1 Introduction ………..……………………………………….……… 98


5.2 Ultrasonic Testing of Materials……………………..…….……….101
5.3 Experimental program …………………………………...………...104
5.3.1 Preparation of test specimens……………………………....104
5.3.2 Cyclic freezing-thawing and chloride penetration………....105
5.3.3 Measurement of Transit time and resonant frequency .…....106
5.3.4 Chemical analysis for total chloride concentration………...107
5.4 Test results and Discussion………………………………………...108
5.4.1 Cracking patterns and visual inspection……………………108
5.4.2 Ultrasonic test results……………………………………….111
5.4.3 Chloride penetration test results…………………………….113
5.5 Conclusion......……………………………………….……………..118
ix

6 EXPERIMENTAL STUDY ON THE EFFECT OF CONCRETE MATERIAL


PARAMETERS ON CHLORIDE PENETRATION

6.1 Introduction ………..……………………………………….……...120


6.2 Specimen Preparation and Experimental Parameters.…….………. 121
6.3 Test Procedure……………………………………………………...123
6.4 Test Results and Discussion……….…………………… ………... 124
6.4.1 Chloride penetration profile ………………………...……...124
6.4.2 Effect of water-cement ratio ………………………...……..124
6.4.3 Effect of aggregate volume fraction………………………..129
6.5 Conclusion………………………………………………………….134

7 EXPERIMENTAL STUDY ON THE COUPLING EFFECT OF MOISTURE


DIFFUSION AND CHLORIDE PENETRATION IN CONCRETE

7.1 Introduction…………………………………………………………135
7.2 Specimens Preparation……………………………………………...136
7.3 Effect of Moisture Diffusion on Chloride Penetration ……...……..137
7.3.1 Experimental procedure…………………………………….137
7.3.2 Experimental results and discussion………………………..139
7.4 The Effect of Chloride Penetration on Moisture Diffusion………...144
7.4.1 Experimental setup and procedure………………........…….144
7.4.2 Results and discussion……………………………………...146
7.5 Conclusions…………………………………………………………149

8 EXPERIMENTAL STUDY ON THE EFFECT OF SHRINKAGE-INDUCED


DAMAGE ON CHLORIDE PENETRATION

8.1 Introduction…………………………………………………………151
8.2 Experimental Program……………………………………………...153
8.2.1 Material……………………………………………………..153
8.2.2 Test specimens preparation………………………………....154
8.2.3 Experimental procedure…………………………………….155
8.3 Experimental results and discussion………………………….…….157
x

8.3.1 Shrinkage strain………………………………………….....157


8.3.2 Restrained shrinkage cracking………………………….…..159
8.3.3 Chloride profiles……………………………………….…...161
8.4 Conclusions…………………………………………………….…..163

9 CONCLUSIONS AND FUTURE RESEARCH RECOMMENDATIONS

9.1 Conclusion…………………………………………………….……165
9.2 Future Research Recommendations………………………….…….167

BIBLIOGRAPHY ………………..………………………………......…..….. 169

APPENDIX A: ADSORPTION ISOTHERMS AND MOISTURE CAPACITIES OF


CEMENT PASTE AND AGGREGATE…......………….178

APPENDIX B: VOLUME PERCENTAGES OF DIFFERENT HYDRATION


PRODUCTS.............................................................…………181

APPENDIX C: MIX DESIGN FOR THE EXPERIMENTAL


PROGRAM………………......................…….…………..…..185

APPENDIX D: COMPRESSION TEST RESULTS ……….......…………..…193

APPENDIX E: EXPERIMENTAL PROCEDURE FOR OBTAINING CHLORIDE


PROFILES..........….…………………………....195
xi

LIST OF FIGURES

Figure Page

1.1 Bridge deck…..………………………………………………………...........4

1.2 Crack on the bridge deck……….. ……………….............………………....4

1.3 Cracks on the shoulder (splash zone) of the bridge deck.………….………..5

2.1 A simple setup for measuring permeability of concrete…………………...13

2.2 Setup for measuring sorptivity of concrete………………………………...18

2.3 Concrete deterioration due to reinforcement corrosion……………………19

2.4 Mechanism of induced-chloride reinforcement corrosion…………………21

3.1 Finite difference mesh of concrete slab……………………………………36

3.2 Different scale levels of concrete…………………………………………..39

3.3 Composite action of concrete………………………………………………40

3.4 Microstructral features of cement paste……………………………............40

3.5 Three-phase effective media model…………………………………..........41

3.6 Four-phase composite model ……………………………….……………..43

3.7 Plane stress problem……………………………………………………….46

3.8 Tensile stress strain diagram ………………………………………………54

3.9 Variation of relative humidity with the depth ……………………………..56


xii

3.10 Variation of shrinkage strain with the depth ………………………….…...56

3.11 Variation of scalar damage parameter with the depth …………….………57

3.12 Variation of hygral stresses with the depth ………………………………..57

3.13 Variation of relative humidity with time of exposure ……….…………….59

3.14 Variation of shrinkage strain with time of exposure ………………………59

3.15 Variation of scalar damage parmater with time of exposure ……………...60

3.16 Variation of hygral stresses with time of exposure ………………………..60

3.17 Effect damage on drying of concrete…………………………….………...62

3.18 Effect of damage on shrinkage of concrete…………………………….......63

3.19 Effect of w/c on drying of concrete……………………………….……....64

3.20 Effect of w/c on shrinkage of concrete……………………………….........64

3.21 Effect of aggregate volume fraction on drying of concrete…………….....66

3.22 Effect of aggregate volume fraction on shrinkage of concrete……………66

4.1 Concrete slab used in the numerical solution……………………………...84

4.2 Depth profile of free chloride concentration………………………….........86

4.3 Time profile of free chloride concentration……………….……………….87

4.4 Comparison with the experimental results…………………………………89

4.5 Effect of the moisture diffusion on free chloride penetration……...………90

4.6 Moisture diffusion in non-saturated concrete……………………………...91


xiii

4.7 Effect of water-cement ratio on chloride penetration ……………………..92

4.8 Effect of water cement ratio on the chloride diffusion coefficient….……..93

4.9 Effect of water-cement ratio on humidity diffusion coefficient…………...93

4.10 Effect of aggregate volume fraction on the chloride penetration ……........94

4.11 Effect of aggregate volume fraction on the chloride diffusion coefficient...95

4.12 Effect of aggregate volume fraction on humidity diffusion coefficient…...95

5.1 Specimen for coupled freezing-thawing and chloride penetration….........104

5.2 Placing of concrete samples in the environmental chamber………….......105

5.3 The freezing and thawing temperature cycle………………….……….....106

5.4 Transit time measurement setup (V-meter)……………….………...........107

5.5 Setup for measurement of fundamental resonant frequency (E-meter)…..107

5.6 The concrete sample (46) after it was exposed to the freezing-thawing….109

5.7 The concrete sample (56) after it was exposed to the freezing-thawing.…109

5.8 The concrete sample (66) after it was exposed to the freezing-thawing.....110

5.9 The concrete sample (55) after it was exposed to the freezing-thawing… 111

5.10 The concrete sample (57) after it was exposed to the freezing-thawing….111

5.11 Chloride profiles for the concrete specimens made of mix 46…………...116

5.12 Chloride profiles for the concrete specimens made of mix 56…………...116

5.13 Chloride profiles for the concrete specimens made of mix 66…………...117
xiv

5.14 Chloride profiles for the concrete specimens made of mix 55…………...117

5.15 Chloride profiles for the concrete specimens made of mix 57…………...118

6.1 Concrete sample…………………………………………………………..122

6.2 Ponding with chloride solution…………………………………………...123

6.3 Chloride profiles for different (w/c) after 50 days exposure time………..126

6.4 Chloride profiles for different (w/c) after 100 days exposure time….…...127

6.5 Chloride profiles for different (w/c) after 200 days exposure time………128

6.6 Segregation of aggregate………………………………………………….128

6.7 Chloride profiles for different (gi) after 50 days of exposure……………131

6.8 Chloride profiles for different (gi) after 100 days of exposure…………..132

6.9 Chloride profiles for different (gi) after 200 days of exposure…………..133

7.1 Sample for coupled moisture diffusion and chloride penetration………...137

7.2 Setup for concrete humidity measurement…………………….…..……..138

7.3 Immersion bath of the concrete samples……………...…………….…….139

7.4 Chloride profiles after 30 days of exposure………...………....……….…142

7.5 Chloride profiles after 90 days of exposure …………………….…....…. 143

7.6 Chloride profiles after 180 days of exposure…………….…..………….. 144

7.7 Experimental setup for the effect of chloride diffusion on moisture


diffusion……………………………………………………………….….146
xv

7.8 The weight of accumulated solution in the center hole for the three different
solutions………………………………………………………………..…148

8.1 Restrained and free shrinkage concrete specimens…………………….…154

8.2 Stresses in the restrained shrinkage concrete sample………………....… 155

8.3 Data acquisition system and personal computer……………………….…156

8.4 Ponding with sodium chloride solution……………………………….….157

8.5 Shrinkage strains development in restrained and free shrinkage concrete.159

8.6 One of the microcracks in the restrained shrinkage ring……….………...160

8.7 Microcrack on the surface of the restrained ring………………………....160

8.8 Microcracking of the surface of the restrainedring…………………….....161

8.9 Chloride profiles after 50 days of exposure…...……………………….…163

E.1 Drilling of concrete sample………………………………………………195

E.2 Chloride concentration measurement kit…………………………..……..196

E.3 Sampling of 1.5 gram of concrete dust and the extraction liquid vial…....197

E.4 Calibration curve for obtaining chloride concentrations………………....199

E.5 Calibration curves for obtaining chloride profiles………………………..203

E.6 Calibration curves for obtaining chloride concentrations………….……..206

E.7 Calibration curve for obtaining chloride concentrations………………....207


xvi

LIST OF TABLES

Table Page

5.1 Ultrasonic test results………………………………..……………………113

5.2 Total chloride concentration as percentage of weight of concrete……….114

5.3 Summary of the experimental results…………………………………….118

6.1 Chloride concentration for different (w/c) after 50 days exposure time….125

6.2 Chloride concentration for different (w/c) after 100 days exposure time...126

6.3 Chloride concentration for different (w/c) after 200 days exposure time...127

6.4 Chloride concentration for different (gi) after 50 days exposure time…....131

6.5 Chloride concentration for different (gi) after 100 days exposure time......132

6.6 Chloride concentration for different (gi) after 200 days exposure time…..133

7.1 Total chloride concentration after 30 days of exposure…………………..141

7.2 Total chloride concentration after 90 days of exposure…………………..142

7.3 Total chloride concentration after 180 days of exposure………………....143

7.4 Weight of the accumulated solution……………………………………...147

8.1 Measured shrinkage strains (in microstrains)……………………………158


xvii

8.2 Total chloride concentration after 50 days of exposure…………………..162

B.1 Parameters related to the hydration of different compounds of cement…183

C.1 Concrete mix design proportions………………………………………....190

C.2 Mix proportions used in the experimental study of Chapters 5, 6 and 7…192

C.3 Concrete mix proportions used in experimental study of the effect of moisture
diffusion on chloride penetration………………..…..………….192

C.4 Proportions of the concrete mix used in the experiment of the effect of induced-
shrinkage cracking on chloride penetration….………………….192

D.1 Compression test results for the concrete mixes used in the experimental study
of Chapters 5, 6 and 7…………………………..……………..…....193

D.2 Compression test results for the concrete used in the experiment of the effect of
moisture diffusion on chloride penetration …………………...……….194

D.3 Compression test results for the concrete used in the experiment of the effect
induced-shrinkage cracking on chloride penetration…………..................194

E.1 The mV readings after 400 freezing-thawing cycles………………......…198

E.2 The mV readings for the standard solutions………………………….…..198

E.3 The mV readings after 50 days of exposure……………………….……..200

E.4 The mV readings after 100 days of exposure……………………….……201

E.5 The mV readings after 200 days of exposure……………………….……202

E.6 The mV readings for the standard solutions………………….…….…….203

E.7 The mV readings after 30 days of exposure……………………….……..204


xviii

E.8 The mV readings after 90 days of exposure…………………….………..204

E.9 The mV readings after 180 days of exposure………………….…………205

E.10 The mV readings for the standard solutions…………………….………..205

E.11 The mV readings after 50 days of exposure…………………….………..206

E.12 The mV readings for the standard solutions…………………….………..207


CHAPTER 1

INTRODUCTION

1.1 Durability of Concrete Structures

Durability of concrete structures can be defined as its ability to maintain a

certain strength and serviceability requirements during its designed service life. Thus,

durability design of concrete shall concern with enhancing its capability to resist

excessively applied mechanical and environmental loads, without significant

damages. Lack of durability in terms of concrete deterioration and reinforcing steel

bars corrosion can cause cracking, spalling, delamination and loss of strength of the

structural members affecting its integrity and ultimately leading to major cracking or

failure. As a result, the durability of concrete has received an increasing attention in

recent years and the developments in cement and concrete technologies have focused

on producing not only stronger but also more durable concrete.

Concrete deteriorations can be classified based on the factors attributing to the

deteriorations, such as physical, chemical, biological and reinforcement corrosion.

Some of the major deterioration mechanisms of concrete will be introduced briefly in

this chapter. Also, some common features of the deterioration mechanisms will be

identified.

Physical deteriorations can be attributed to physical actions such as, wetting


2

and drying cycles, freezing and thawing, temperature changes and other physical

actions such as waves, movement of solid particles in water or air, wear and tear

actions. Wetting and drying cycles may result in cracking due to excessive shrinkage.

Freezing and thawing can produce internal stresses inside concrete causing spalling

and cracking. Physical actions such as waves, movement of solid particles in water or

air and/or wear and tear can cause abrasion of concrete. Temperature changes can

produce thermal stresses causing damage and loss of strength of concrete members.

Chemical deteriorations can be attributed to chemical attacks from the

environment such as acid attack and sulfate attack or chemical reactions between the

constituents of the concrete such as alkali silica reaction and alkali carbonate reaction.

Acids can attack concrete by reacting with calcium silicate hydrate (C-S-H) gel

producing water-soluble calcium compounds, which may leach out causing surface

efflorescence. Acids also can attack reinforcing steel bars causing corrosion, which

may ultimately lead to cracking, spalling and delamination of concrete. Sulfates can

attack concrete reacting with calcium hydroxide (CH) to form gypsum. Gypsum

reacts with hydrated calcium aluminates to form ettringite. Both reactions cause

volume expansion and produce stresses causing damage and cracking of concrete.

Alkali silica reaction between the active silica in aggregates and alkalis in cement

results in the formation of alkali-silica complex, which absorbs water causing volume

expansion and producing internal stresses which may lead to map or pattern cracks.

Alkali-carbonate reaction occurs between some dolomitic limestone aggregates and

alkalis in cement. The products of the reaction have less volume than the original

materials and susceptible to swelling and volume changes. Those volume changes
3

result in internal stresses and deterioration of concrete.

Biological deterioration (Kawai, 1996) can be attributed to aerobic and

anaerobic sulfur bacteria and fungi, which metabolize organic and carbonic acids.

Those acids react with hydration products of cement causing leaching and volume

changes of concrete.

Reinforcement corrosion can be attributed to chloride attack or carbonation. In

chloride-induced corrosion, chloride ions breakdown the passive oxide film on the

steel bars and allow the corrosion process to start. In carbonation-induced corrosion,

carbonation process lowers the value of the pH by forming carbonic acid and induces

the steel corrosion.

One example on the lack of durability of concrete structures is the

deterioration of bridge deck shown in Fig. (1.1). The repeated overloading of the

bridge deck in terms of cyclic mechanical loading, wetting-drying, freezing-thawing

and chloride ions ingress from deicing salts causes fatigue cracks as shown in Figs.

(1.2) and (1.3), which result in two types of damage. The first one, which occurs in

most cases, is the reduction of the durability of concrete in a way that permeability of

the concrete is drastically increased. There has not been much research done in this

area. The other one is the immediate failure of the structure, after certain cycles of

loading, due to unstable propagation of the major fatigue crack. In the latter case,

fatigue fracture has been studied extensively by experiments (Nordby 1958, Hilsdorf

et al. 1966, Baluch et al. 1989).


4

Fig. 1.1: bridge deck (part of highway US-285 over North Turkey
Creek Road) near Evergreen, Colorado.

Fig. 1.2: Crack on the bridge deck.


5

Fig. 1.3: Cracks on the shoulder (splash zone) of the bridge deck.

As one can see, diffusion processes such as, temperature, moisture and

chemicals control all physical and chemical deteriorations of concrete. The common

feature of these concrete deteriorations is that they are all diffusion-controlled

processes. Therefore, accurate characterization of the diffusion processes becomes

very important.

1.2 Objectives of the Study

The main objective of this study is to quantify some of the different processes

that affect the durability of concrete structures. These processes include moisture

diffusion, chloride penetration, freezing-thawing and their coupling effects. The study

includes theoretical modeling and experimental study. The theoretical modeling


6

encompasses the formulation of the governing equations and the boundary conditions,

identification of the material parameters involved, solution of the governing equations

using one of the numerical techniques such as finite element method or finite

difference method, application of the numerical solution in a computer program and

performing a parametric study to analyze the effects of different influential

parameters. The experimental study include the effect of the concrete material

parameters on chloride penetration, the coupling effect of moisture diffusion and

drying shrinkage, the coupling effect of moisture diffusion and chloride penetration

and the coupling effect of freezing-thawing and chloride penetration in concrete.

1.3 Thesis Organization

This thesis has nine chapters. This chapter includes introductory information

about the durability of concrete structures, motivation and the objectives of the study.

Chapter 2 presents detailed background information about some issues

related to the durability of concrete structures such as its diffusion properties, chloride

ions penetration and heat conduction.

Chapter 3 discusses the interaction effect between the drying shrinkage and

the moisture diffusion process in concrete. This chapter contains seven sections. The

first section presents introductory information about the two processes, drying

shrinkage and moisture diffusion in concrete and their interaction. The second section

discusses various issues of the moisture diffusion in concrete such as the basic

formulation of the governing partial differential equation, the material parameters


7

involved in the governing equation. The third section discusses the finite element and

the finite difference numerical solutions. The fourth section includes the modeling of

drying shrinkage of concrete. The fifth section discusses the interaction of drying and

shrinkage in concrete. The sixth section presents some numerical applications on the

different influential parameters and the conclusions are in the last section.

Chapter 4 discusses the coupling effect of moisture and chloride diffusion

processes in concrete. This chapter contains six sections. The first section is an

introductory about the interaction effect of the two processes. The second section

presents the basic formulation of the governing partial differential equation. The third

section presents the material parameters involved in the governing equation. The

fourth section presents the finite difference numerical solutions. The fifth section

discusses some numerical applications on the different influential parameters and

section six presents the conclusions.

Chapter 5 discusses the coupling effect of freezing-thawing and chloride

penetration in concrete. This chapter contains five sections. The first section

introduces the coupling effect of the two processes. The second section presents a

brief review about the ultrasonic testing of materials. The third section presents the

experimental program, which includes the preparation of test specimens and the

description of the testing methods and techniques. The fourth section present the

experimental results, which include the visual inspection of the cracking of the test

specimens, the damage factors resulted from the ultrasonic testing and the chloride

profiles resulted from the electrochemical analysis. The fifth section presents
8

summary of the experimental results and conclusions.

Chapter 6 presents an experimental study on the effect of concrete material

parameters on chloride penetration. This chapter contains five sections. The first

section presents introductory information about the effect of different material

parameters on chloride penetration in concrete. The second section describes the

specimen and parameters of the experimental study. The third section describes the

test procedure. The fourth section presents the experimental results for the effect of

water-cement ratio and aggregate volume fraction and the fifth section contains the

conclusions.

Chapter 7 describes an experimental study on the coupled effect of moisture

diffusion and chloride penetration in concrete. This chapter contains five sections.

The first section introduces some information about the coupling effect of the two

processes. The second section describes the concrete specimen used in this study. The

experimental program, experimental results and discussions are presented in section

three for the effect of moisture diffusion on chloride penetration. Section four is for

the effect of chloride penetration on moisture diffusion, while section five contains

the conclusions.

Chapter 8 describes an experimental study of the effect of shrinkage-induced

damage on chloride penetration in concrete. This chapter contains four sections. The

first section contains an introductory information and literature review on the effect of

cracking on chloride penetration in concrete. The second section describes the

materials, test specimens and experimental procedures for measurement of shrinkage


9

strains and the chloride profiles. The third section contains the test results. The

conclusions are presented in section four.

Chapter 9 contains conclusions and recommendations for future research.


CHAPTER 2

BACKGROUND INFORMATION AND LITERATURE REVIEW

2.1 Transport Properties of Concrete

Durability of concrete structures is mainly affected by the transport of fluids

and some deleterious chemicals into or through concrete. There are many fluids (in a

broad sense) that can flow in concrete and affect its durability. Those fluids include

gases such as carbon dioxide (CO2) and oxygen (O2) and liquids such as pure water or

water carrying some deleterious chemicals. The deleterious chemicals include acids

and aggressive ions such as sulfates and chlorides. Freezing and thawing is mainly

affected by the diffusion of moisture in concrete. Corrosion of embedded steel bars is

affected collectively by gas diffusion (CO2 and O2), moisture diffusion and chloride

ions penetration in concrete. Chemical attacks such as acids; sulfate and chloride ions

are mainly affected by the diffusion of these chemicals in concrete, as well as the

subsequent chemical reactions between the chemicals and some components in

concrete.

Transport properties of concrete depend primarily on its pore structure, which

can be characterized by its porosity, pore size distribution, and morphology of the

pore structure (such as pore shape, pore connectivity and pore tortuosity). High

porosity does not always correspond to high permeability unless the pores are
11

continuous and effective with respect to flow. Discontinuous pores, long and tortuous

pore systems, pores contain adsorbed water and pores have narrow entrances are

ineffective with respect to the flow of fluids. Concrete is a particular composite

material composed of aggregates and cement paste. Aggregates are porous materials,

but their contribution to the diffusion of fluids is minor because of the discontinuity in

their pore structure and also, they are enveloped by the cement paste. Hydrated

cement is composed of a very poorly crystalline material called calcium silicate

hydrate (C-S-H) gel, unhydrated cement particles, calcium hydroxide (CH) crystals,

ettringite crystals and pores. Those pores include air voids and their size ranges

100µm−10mm, water-filled spaces called capillary pores and their size ranges

0.05−100µm, and gel pores and their size is smaller than 50 nm. Transport of fluids in

capillary pores is much easier than in gel pores. As a result, the permeability of

hardened cement paste is mainly controlled by its capillary porosity.

Transport of mass in concrete, differing from fracture failure and strength

failure of a concrete structure, involves not only conventional mechanics methods but

also material theories at different scale levels. For instance, the strength of concrete

can be correlated to the size of defects as small as 50 micrometers. But, the fine

pores with smaller sizes have almost no effect on strength of the materials. Mass

transport, on the other hand, is controlled by the internal pores with a very broad

range of size distribution from millimeter down to nanometer levels. Therefore, the

prediction models for mass transport must be developed based on the influential

parameters and mechanisms on multiple scale levels.


12

Transport of fluids and deleterious chemicals into or through concrete is

governed by several mechanisms and each mechanism is related to specific transport

property. Fluids and deleterious chemicals can flow into or through concrete under

pressure gradient, concentration gradient, capillary suction and potential gradient and

the relevant diffusion properties are the permeability, diffusivity, sorptivity and ionic

diffusivity.

2.1.1 Concrete permeability

Permeability of concrete refers to the ease of flow of fluids into or through

concrete under pressure gradient. In this work we concern about the moisture

diffusion in concrete and so water permeability of concrete will be discussed.

The flow of water through saturated concrete and under drop of a hydraulic

head (∆h) is governed by Darcy’s law as follows:

dq 1 ∆h
=K (2.1)
dt A L

where: dq/dt = flow rate of water in m3/sec,

A = cross-sectional area of flow in m2,

K = coefficient of permeability of concrete in m/sec,

∆h = drop in hydraulic head in m,

L = depth of flow in m.
13

The coefficient of permeability of concrete can be measured using the simple

setup shown in Fig. (2.1), in which water is allowed to flow through saturated

concrete under pressure gradient.

Inflow

Sealant
L Concrete
Sample

Outflow

Fig. 2.1: A simple setup for measuring permeability of concrete.

The intrinsic permeability of concrete, which is independent of the type of the

liquid and temperature, can be calculated using the following formula:

Q ηL
K′ = (2.2)
A (P1 − P2 )

where, K ′ = intrinsic permeability of concrete in m2,

Q = measured flow rate of water in m3/sec,

A = cross-sectional area of the concrete sample in m2,

η = dynamic viscosity of water in N . sec/m2,


14

L = depth of concrete sample in m,

P1 = upstream pressure in N/m2,

P2 = downstream pressure in N/m2.

Then the coefficient of permeability of concrete can be calculated using the

following formula:

ρg
K = K′ (2.3)
η

where, K = coefficient of permeability in m/sec,

ρ = water density in kg/m3,

g = gravitational acceleration in m/sec2,

η = dynamic viscosity of water in N . sec/m2,

The permeability of concrete is mainly affected by water-cement ratio, age,

temperature, curing condition, curing period, cement properties, cement composition

and the volume fraction of aggregates. The water-cement ratio affects the

permeability of concrete by introducing more capillary pores. As the water-cement

ratio increases, the capillary porosity increases resulting in an increase in the

permeability of concrete. The age of concrete affects its permeability due to the

progress of the hydration process. As the hydration process continues more calcium

silicate hydrate (C-S-H) gel is produced which fill the available capillary pores
15

resulting in a rapid decrease of the permeability of concrete (Neville 1996). The

temperature affects the permeability of concrete by changing the dynamic viscosity of

the flowing fluid. Also temperature changes produce volume changes in the concrete

body, when restrained by either internal or external restraints they produce stresses

and micro-crack the concrete increasing its permeability. Moist curing of concrete

will reduce its permeability by increasing the rate of cement hydration process and

decreasing the capillary porosity. Curing of concrete in dry condition will result in

moisture losses. Those moisture losses will produce internal strains in the concrete

body and if restrained, stresses will develop, causing cracking and thus increasing the

permeability of concrete. Cement properties such as degree of fineness can affect the

permeability of concrete by decreasing its porosity and thus decreasing its

permeability. Cement composition affects the permeability of concrete by increasing

the rate of cement hydration. And as a result it will not affect the ultimate

permeability of concrete. The volume fraction of aggregates in concrete affects its

permeability by changing the volume of the cement paste, which is the effective

media for the flow of fluids. Also aggregates block the flow paths in concrete making

them longer and more tortuous. And as a result an increase in the volume fraction of

aggregates will result in a decrease in the permeability of concrete.

2.1.2 Concrete diffusivity

Diffusivity of concrete is related to the diffusion of mass into or through

concrete under concentration gradient. In this work we concern about the moisture

and chloride ion diffusion processes in concrete and so moisture and chloride ion
16

diffusivities of concrete will be discussed. The diffusion of moisture and chloride ions

into or through concrete under concentration gradient is governed by Fick’s first law

as follows:

J = − D ∇C (2.4)

where: J = mass diffusion rate in kg/m2,

D = diffusivity of concrete in m2/sec,

∇C = concentration gradient in kg/m4.

The moisture diffusivity of concrete is affected by many factors such as pore

structures, humidity level and temperature. The chloride diffusivity of concrete

affected by many factors such as features of pore structures, level of chloride

concentration, humidity level and temperature. The pore structure of concrete is

affected by many factors such as water-cement ratio, age, temperature, curing

condition and period, cement properties and cement composition and the volume

fraction of aggregates.

2.1.3 Concrete sorptivity

Sorptivity of concrete is related to the diffusion of fluids into or through

concrete due to capillary forces in the pores. Sorption of water into concrete can be

expressed by the following formula:

M (t ) = M o + ρAS t (2.5)
17

where: M(t) = mass of fluid gained by the concrete at time t,

M0 = initial mass gain in kg,

ρ = water density in kg/m3,

A = cross-sectional area of sorption in m2,

S = sorptivity of concrete in m/sec0.5,

t = time in seconds.

Sorptivity of concrete can be estimated using the setup shown in Fig. (2.2). In

which a concrete specimen is exposed to water at one of the faces while the other

faces are sealed. The gain in the mass of concrete sample and the time of exposure are

reported. Also, the initial gain in the mass of concrete sample should be reported. The

sorptivity of concrete can be estimated using Eq. (2.5) and knowing the cross-

sectional area and the density of the water. It is important to make sure that the

concrete sample has uniform moisture content before exposing to water. This can be

done either by keeping the sample in oven for 7 days at 50 oC followed by another 7

days in sealed container which will produce a uniform moisture content for 100 mm

sample or by oven drying the sample at 105 oC until there is less than 0.1% change in

the mass after 24 hours of drying.


18

Sealed
Surfaces

Exposed Concrete
Surface Sample
Water

Fig. 2.2: Setup for measuring sorptivity of concrete.

2.2 Chloride Ions Penetration into Concrete

Chloride ions penetration is one of the major problems that affect the

durability of reinforced concrete structures. Chloride solution causes corrosion of

embedded steel bars, which in turn causes cracking, spalling and delimination of

concrete as shown in Fig. (2.3). It may also reduce the load carrying capacity of the

structural members by reducing the cross-sectional area of the steel bars.

There are three stages in the deterioration process of reinforced concrete

structures due to steel bars corrosion. The first stage is dominated by the penetration

of chloride ions through concrete towards the reinforcing bars. Once the chloride

concentration reaches a critical value, corrosion of steel starts. The second stage is

controlled by rust expansion, during which the rust forms in the interface transition

zone between aggregate and cement paste. The formation of rust is associated with

large volume expansion.


19

Steel
Bars

Steel
Bar
(a) (b)
Steel Bars

(c)
Fig. 2.3: Concrete deterioration due to reinforcement corrosion (a) cracking
(b) spalling and (c) delimination.

The third stage is concrete cracking. The volume expansion of the rust causes

the volumetric mismatch, which leads to cracking, spalling and delamination of

concrete. Among the three stages, the first one may be the longest one, which could

last for 15 to 20 years. The second stage is shorter than the first one, depending on the

porosity in the interface transition zone. The third stage is the shortest one among the

three stages, since the typical thickness of concrete cover is 5 cm, and crack

propagate through the concrete cover very fast once the crack is formed around the

steel bars.
20

2.2.1 Mechanism of the reinforcement corrosion

During the hydration process of the cement a thin layer of oxide forms on the

surface of the embedded steel bars called passivity layer. This layer protects the steel

from reacting with oxygen and moisture and thus prevents the rust formation and

corrosion of steel bars. Once the chloride ions concentration reaches a certain level, it

destroys the thin passive film around the reinforcing steel bar and the electrochemical

cell will setup as shown in Fig. (2.4). At the anode the oxidation of iron will occur,

and the ferrous ions (Fe++) will produce:

Fe ⇒ Fe + + + 2e − (2.6)

The ferrous ions will pass into the pore solution and the electrons will flow

through the steel bar to the cathode. At the cathode the electrons will react with

moisture and oxygen to form the hydroxyl ions (OH-).

2 H 2 O + O2 + 4e − ⇒ 4OH − (2.7)

Then the hydroxyl ions will travel through the pore solution and combine with

the ferrous ions to form ferrous hydroxide (Fe(OH)2 ):

Fe + + + 2OH − ⇒ Fe(OH ) 2 (2.8)

Further oxidation of ferrous hydroxide will result into the ferric hydroxide,

which is the rust:

4 Fe (OH ) 2 + 2 H 2O + O2 ⇒ 4 Fe (OH ) 3 (2.9)


21

Chloride Moisture Oxygen

Concrete
Cover
Cl- Ions -
Fe++ + 2(OH) Fe(OH)
2

Fe(OH)3 (Rust)
- Steel Bar
Fe++ OH

Anode Electrons ( e- ) Cathode

- -
Fe Fe++ + 2e 2H2O + O2 + 4e- 4(OH)
Passivity
Layer

Fig. 2.4: Mechanism of induced-chloride reinforcement corrosion.

2.2.2 Onset of the steel corrosion

The chloride concentration at which, corrosion of steel in concrete starts, is

called critical chloride concentration. Research results on the critical value of chloride

content showed some variations and also, they are not in the same unit. We first

convert different critical values from literature into a unified unit, i.e. gram of

chloride per gram of concrete, (g/g), then select a specific critical value to be used in

this study. This unit is considered to be convenient because the test results of

chloride content by AASHTO T259 and AASHTO T260 are in this unit.

Berke (1986) obtained the value of 0.9 - 1.0 kg of chloride per cubic meter of

concrete. Using 2300 kg/m3 as the density of concrete, the critical content in gram of

chloride per gram of concrete will be 0.039% to 0.043% (g/g). Browne (1982)

showed the value of 0.4% by weight of cement used in the concrete. Assuming 300
22

Kg of cement for a cubic meter of concrete, 0.4% of 300 Kg cement is 1.2 Kg, and

then using 2400 Kg/m3 as the density of concrete, the critical value will be 0.05%

(g/g). Cady and Weyers (1992) gave the value of 0.025 - 0.05% by weight of

concrete. Therefore, the range of the critical content is from 0.025 – 0.05% in gram

of chloride per gram of concrete, which is a quite broad range. Coggins and French

(1990) determined the chloride profiles in three girders (two facial and one interior)

and the deck of a twenty-year-old prestressed concrete bridge. The bridge deck and

the girders were exposed to chloride penetration during its service life. Dust samples

were collected at ½ in. (13 mm) depth increments and tested for total chloride

concentration. It was found that total chloride concentration at reinforcement depth

does not exceed 120 ppm (by concrete weight) in girders and in the range of 400 to

1500 ppm in bridge deck. Although, the chloride concentration in bridge deck

exceeds the corrosion threshold value (200-450 ppm), but no evidence of corrosion

was reported.

2.3 Heat Transfer in Concrete

In practice, concrete is exposed to temperature variations during mixing,

placing, hardening and its service life. Temperature variations can be attributed

mainly to the heat transferred into or from concrete by surface convection and

radiation and in concrete by conduction, or to the heat produced during cement

hydration. Those temperature variations affect the properties of both fresh and

hardened concrete and thus affect its durability performance.

Reasonable rise in the curing temperature would speeds up the rate of cement
23

hydration and subsequently increases the early strength of concrete without harmful

effect on its later strength. On the other hand, higher temperature variations during

placing and hardening of concrete would result in poorly structured cement paste with

higher porosity. The poor structure of the hydrated cement paste is due to the non-

uniform distribution of the hydration products within the cement paste. This non-

uniform distribution of hydration products would reduce locally the strength of

concrete. The higher porosity and the microcracking due to the lower strength would

accelerate the transport of moisture and some deleterious chemicals, which affect

adversely the durability and reduce the service life of concrete structures.

Heat developed during hydration of cement may have harmful effects on the

durability of concrete structures. The heat evolved during hydration of cement and the

loss of heat at the surface may results in steep temperature gradients between the

interior and the exterior of concrete resulting in cracking. On the other hand, the heat

produced during hydration of cement may prevent the freezing of pore water resulting

in continuous hydration of cement and thus more gain in the strength of concrete and

reduction in the porosity will be achieved.

High temperature rise in the surface layer of the hardened concrete due to fire

or some industrial applications such as take-off of airplane could result in steep

temperature gradients between the outside and the inside layer of concrete. This

could results in cracking and spalling of concrete at joints or poorly compacted areas

or delemination of concrete at the concrete-steel interface.

Very low ambient temperatures (−200 oC - 0 oC) could result in the formation
24

of ice in the hydrated cement paste. The formation of ice during setting and hardening

of concrete could seize the hydration process of cement. Upon thawing the hydration

of cement continues and concrete need to be vibrated again. Lack of vibration might

result in setting and hardening of concrete in such a way that high porosity will be

created due to expansion of mixing water upon freezing. The high porosity will

decrease the strength and accelerate the diffusion of moisture and some deleterious

chemical into concrete. If ice formed after setting and before an appreciable strength

gained, concrete might undergoes larger deteriorations. Those deteriorations can be

attributed to the low strength of concrete and the availability of more water, which is

not consumed yet by the hydration of the cement. The formation of ice in the

capillary pores of the hardened cement paste during the service life of concrete results

in volume expansion of those pores. Upon thawing and refreezing more volume

expansions take place in the capillary pores. Subsequent freezing and thawing cycles

would result in progressive damage and thus more cracking of concrete would occur.

The cracking will accelerate the diffusion of the moisture and chloride ions into

concrete and those affect adversely its durability.


CHAPTER 3

COUPLED MOISTURE DIFUSSION AND DRYING SHRINKAGE

3.1 Introduction

Drying shrinkage of concrete occurs due to the loss of moisture and thus, it is

controlled by moisture diffusion process. On the other hand, shrinkage causes

cracking of concrete and affects its moisture diffusion properties. And thus moisture

diffusion and drying shrinkage are two coupled processes and their interactive effect

is very important for durability of concrete structures, especially structures with large

surface area to volume ratio, such as highway pavements, bridge decks, parking

garages and industrial garages.

In recent years, many researchers have investigated the interactive effect of

the mass diffusion and the cracking resulted from moisture gradient, temperature

gradient and mechanical loading. All the research work indicated that the

microcraking produced during hardening, drying and heating of concrete and the

cracking resulted from the mechanical loading could increase the diffusion coefficient

and thus, accelerate the mass diffusion process. The diffusion properties of cracked

concrete were related to several damage indicators. One indicator is the critical crack

width (Aldea et al., 1999; Reinhardt et al., 1997; Wang et al., 1997; Bazant et al.,

1987). Another indicator is the critical monotonic or cyclic loading levels (Gontar et
26

al., 2000; Meziani and Skoczylas, 1999; Konin et al., 1997; Saito and Ishimori, 1995;

Kermani, 1991). Gerard and Marchand (2000) conducted a theoretical study on the

effect of both isotropic and anisotropic traversing crack networks on the steady state

diffusion properties of severely damaged concrete. The diffusion equation was

simplified by assuming that the cracks are uniform in size and distributed evenly on

one or two-dimensional grid. The study concluded that the cracking could accelerate

the diffusion process in concrete. Also, they proposed a simple model for the effect of

cracking on the diffusion coefficient of concrete based on the crack density and mean

crack aperture. Bazant et al. (1987) investigated the effect of crack width and crack

spacing on the permeability and diffusion coefficient of concrete. It was observed that

the permeability and diffusion coefficient of cracked concrete is proportional to the

crack width cubed and inversely proportional to the crack spacing. Bentz et al. (1995)

developed structural models for the properties of the calcium silicate hydrated (C-S-

H) gel at the nanometer level. In this model the calcium silicate hydrate (C-S-H) gel

particles are represented by hard core-soft shell concentric spheres. Each one of those

spheres can overlap with other sphere on the soft core as well as the hard core. This

nanostructural models can be used in the multi-scale modeling of the drying

shrinkage of concrete. Also, the developed models are validated by the images

obtained from the high-resolution transmission electron microscopy (TEM) and the

experimental data measured by sorption test.

In the following sections, the diffusion model and the shrinkage model are

briefly described first, and then the coupling of the model parameters is discussed in

detail. The moisture distribution in concrete was modeled using the nonlinear
27

moisture diffusion equation derived based on Ficks’ law (Bazant and Najjar 1972,

Bazant and Thonguthai, 1978, Xi et al. 1994). The nonlinear moisture diffusion

equation involves two diffusion parameters, the moisture capacity and humidity

diffusion coefficient. Those two parameters were coupled with the material

parameters related to drying shrinkage. Specifically, the moisture capacity is related

to the shrinkage coefficients of the constituents in concrete, and the moisture

diffusion coefficient is related to the damage parameter induced by the drying

shrinkage. Both of the diffusion model and the shrinkage model used in the study are

multiphase and multiscale models, in which the influential parameters at macro-,

meso-, and micro-scales are taken into account; and concrete design parameters are

included as model inputs. Finally, the coupled problem was solved using finite

difference method. The influential parameters such as water-cement ratio and

aggregate volume fraction were studied through several numerical applications.

3.2 Modeling Moisture Diffusion in Concrete (without considering shrinkage-

induced damage)

3.2.1 Basic formulation of the governing equation

The isothermal moisture flux in concrete can be formulated using Fick’s first

law by two different methods. In the first method, the moisture flux (J) is

proportional to the gradient of the water content and can be written as following:

J = − k grad (we ) (3.1)

In the second method, the moisture flux (J) is proportional to the gradient of the pore
28

relative humidity, and can be written as following:

J = − k grad (H ) (3.2)

where, we and H are the evaporable water content and the pore relative humidity,

respectively. Also k has two different physical meanings in above two formulations.

In Eq. (3.1), k is the moisture diffusion coefficient and in Eq. (3.2), k is the humidity

diffusion coefficient.

The isothermal rate of change of total water content, w, in concrete can be

expressed using the mass balance equation or Fick’s second law as follows:

∂w
= − div ( j ) (3.3)
∂t

in which, w=we+wne, where wne is the nonevaporable or chemically combined water

content, The substitution of the moisture flux (J), described in Eqs. (3.1) and (3.2)

into the mass balance equation, Eq. (3.3), will resulted in two formulations of the

nonlinear moisture diffusion equation as follows:

∂ (we + wne )
= div ( Dw grad we ) (3.4)
∂t

∂w ∂H
= div ( DH grad H ) (3.5)
∂H ∂t

where, H is the relative humidity, Dw and DH are the moisture diffusion coefficient

and the humidity diffusion coefficient, ∂w/∂H is the moisture capacity and t is the
29

time. The nonlinearity of the moisture diffusion equation, Eq. (3.4), resulted from the

dependence of the moisture diffusion coefficient, Dw, on the moisture content, w. And

the nonlinearity of the moisture diffusion equation, Eq. (3.5), resulted from the

dependence of the moisture capacity, ∂w/∂H and the humidity diffusion coefficient,

DH on the relative humidity, H.

In the present work the second formulation in terms of relative humidity H,

which described by Eqs. (3.2) and (3.5) will be used to describe the moisture

diffusion in concrete. This is due mainly to the fact that amounts of we and wne are

functions of time even if there is no moisture exchange with the surrounding

environment (i.e. total water, w, is a constant). Otherwise another model must be

included to characterize the effect of hydration of cement on we and wne.

3.2.2 Initial and boundary conditions

The initial condition can be written as follows:

H ( x, y, z, t = 0) = H o ( x, y, z ) (3.6)

where, Ho(x,y,z) is the initial distribution of relative humidity in the concrete body.

In this work only convective type boundary conditions will be considered.

This type of boundary conditions is prescribed at the surfaces of the body exposed to

the surrounding environment. In the case there is a difference in the relative humidity

between the inside of the exposed body and the surrounding environment and the

surfaces of the body are not sealed, an exchange of moisture will occur. If the
30

surrounding relative humidity is larger, the moisture will flux through the exposed

surfaces from the environment to the body and this case called wetting. On the other

hand, if the relative humidity inside the body is larger, the moisture will flux out of

the body to the surrounding environment and this case called drying. In both cases the

boundary conditions are imposed on the surface as follows:

q.n = 0 (3.7)

q.n = β (H − H env ) (3.8)

Where, q is the humidity flux, n is a unit vector normal to the exposed surface and

directed towards the exterior environment, β is the surface moisture diffusion

coefficient, H is the unknown surface humidity and Henv is the environmental relative

humidity. Eq. (3.7) represents the case of no moisture exchange between the body and

the environment on the surface of that body.

3.2.3 Material parameters

In order to find a numerical solution for the governing partial differential

equation of moisture diffusion in concrete, Eq. (3.5), the material parameters involved

should be determined.

Moisture capacity, ∂w/∂H

The moisture capacity of concrete can be calculated based on the multiphase

and multiscale model. Concrete is a two-phase material, in which the aggregates are

considered as the inclusions and represent one phase. The cement paste is the matrix
31

and represents the other phase. Then, without considering the effect of shrinkage of

concrete, the moisture capacity of concrete can be calculated as a certain average of

the moisture capacities of the two constituent phases:

⎛ ∂w ⎞ ⎛ ∂w ⎞ ⎛ ∂w ⎞
⎜ ⎟ = f agg ⎜ ⎟ + f cp ⎜ ⎟ (3.9)
⎝ ∂H ⎠ avg ⎝ ∂H ⎠ agg ⎝ ∂H ⎠ cp

where fagg and fcp are the weight percentages of the aggregates and cement paste. The

moisture capacities of the aggregate and cement paste in Eq. (3.9) can be evaluated by

using the model developed by Xi et al. (1994a, 1994b) and Xi (1995a, 1995b) as

shown in Appendix A.

Humidity diffusion coefficient, DH

Humidity diffusion coefficient depends strongly on the diffusion mechanisms,

which depend on the pore structure of the concrete body. There are three different

diffusion mechanisms of moisture into concrete; molecular or ordinary diffusion,

Knudsen diffusion and surface diffusion. Thus, the total diffusivity is a complex

property that often includes contribution from multiple mechanisms. The moisture

diffusion coefficient of concrete can be predicted by the composite model developed

by Christensen (1979):

⎛ gi ⎞
DH = DHcp ⎜1 + ⎟ (3.10)
⎜ [1 − g ] / 3 + 1/[(D / D ) − 1] ⎟
⎝ i Hagg Hcp ⎠

in which, gi is the aggregate volume fraction, DHcp is the humidity diffusion

coefficient of the cement paste and DHagg is the humidity diffusion coefficient of the
32

aggregates.

The humidity diffusion coefficient of aggregates in concrete is very small

comparing with the diffusion coefficient of concrete and can be neglected in Eq.

(3.10). The humidity diffusion coefficient of cement paste can be predicted by the

empirical model developed by Xi et al. (1995b).

γ h ( H −1)
D Hcp = α h + β h [1 − 2 −10 ] (3.11)

where, αh, βh and γh are coefficients to be calibrated from the test data and given as:

α h = 1.05 − 3.8 ( w c) + 3.56 ( w c) 2 (3.12)

β h = −14.4 + 50.4 ( w c) − 41.8 ( w c) 2 (3.13)

γ h = 31.3 − 136 ( w c ) + 162 ( w c ) 2 (3.14)

3.3 Numerical Solution

The nonlinearity of the partial differential equation, which governs the moisture

diffusion in concrete, makes the numerical solution favorable. The numerical solution

can be done either by finite element method or finite difference method.

3.3.1 Finite element method

The moisture diffusion equation, Eq. (3.5), can be solved using finite element

method. In the finite element solution the standard residual procedure may be applied
33

as follows:

⎡ ∂w ∂H ⎤
∫Ω
wi ⋅ ⎢
⎣ ∂H ∂t
− div ( D H grad H ) ⎥ d Ω = 0

(3.15)

By choosing standard Galerkian method in which, wi = Ni .

⎡ ∂w ∂H ∂H ⎤
∫Ω
NT ⋅ ⎢
⎣ ∂H ∂t
− div ( DH ) dΩ = 0
∂X ⎥⎦
(3.16)

Rearranging Eq. (3.16) results in:

∂w ∂H ∂H ∂N T ∂H
∫Ω
NT ⋅
∂H ∂t
dΩ − ∫ div ( N T ⋅ DH
Ω ∂X
) dΩ + ∫
Ω ∂X
DH
∂X
dΩ = 0 (3.17)

By using divergence theorem, Eq. (3.17) becomes:

∂w ∂H ∂H ∂N T ∂H
∫Ω
NT ⋅
∂H ∂t
dΩ − ∫ N T ⋅ DH
Γ ∂X
⋅ n dΓ + ∫
Ω ∂X
DH
∂X
dΩ = 0 (3.18)

The flux of moisture, q can be written as:

∂H
q = − DH (3.19)
∂X

Substituting Eq. (3.19) into Eq. (3.18) resulting in:

∂w ∂H ∂N T ∂H
∫Ω
NT ⋅
∂H ∂t
dΩ + ∫ N T ⋅ q ⋅ n dΓ + ∫
Γ Ω ∂X
DH
∂X
dΩ = 0 (3.20)

Substituting the boundary condition described in Eq. (3.8) into Eq. (3.20) resulting in:
34

∂w ∂H ∂N T ∂H
∫ NT ⋅ dΩ + ∫ N T ⋅ β (H − H env ) dΓ + ∫ DH dΩ = 0 (3.21)
Ω ∂H ∂t Γ Ω ∂X ∂X

The relative humidity at any point in an element can be determined using shape

functions, N as follows:

H ( X , t ) = ∑ N i ( X )H i (t ) (3.22)

By applying the finite element method with the shape functions Ni of the ith order we

have:

⎛ T ∂w ⎞ ∂H ⎛ ∂N T ∂N ⎞
⎜∫N N dΩ ⎟ + ⎜∫ D dΩ + ∫ N T
β N dΓ ⎟ H = ∫ N T β H env dΓ (3.23)
⎜ ∂H ⎟ ∂t ⎜ ∂X H
∂X ⎟
⎝Ω ⎠ ⎝Ω Γ ⎠ Γ

The moisture diffusion in concrete is described by the following general expression:

∂H
C + KH = F H (3.24)
∂t

Where, C is the moisture capacity matrix, K is the moisture diffusion coefficient

matrix and FH is the hygral load vector and can be written as follows:

∂w
C = ∫ NT N dΩ (3.25)

∂H

T
∂N ∂N
K =∫ DH dΩ + ∫ N T β N dΓ (3.26)

∂X ∂X Γ

FH = ∫ N T βH env dΓ (3.27)
Γ
35

Eq. ( 3.24) can be rewritten as follows:

∂H
+ AH = P (3.28)
∂t

where,

A = C −1 K (3.29)

P = C −1 FH (3.30)

Eq. (3.28) can be solved using one the step-by-step integration schemes as

follows (Argyris et al. 1977):

[I + α∆tA]H i +1 = [I − (1 − α )∆tA]H i + ∆t (1 − α ) Pi + ∆tαPi +1 (3.31)

The value of α should lies in the interval [0,1]; for α = 0, we have Forward Euler

Method (Explicit scheme), for α = 1, 0.5, and 2/3, we have Backward Euler Method,

Trapezoidal and Galerkian Methods, respectively (Implicit schemes).

3.3.2 Finite difference method

The Alternating-Direction Implicit (ADI) finite difference method developed

by Peaceman and Rachford (Carnahan et al. 1969) can be employed to solve the

moisture diffusion problem described in Eq. (3.5). A two-dimensional finite

difference mesh as shown in Fig. (3.1) can be used to descretize this problem. The

pore relative humidity, H(i,j,n), at different points of the mesh at different time

intervals need to be estimated.


36

∆X

Slab
∆Y Depth

Slab Width

Fig. 3.1: Finite difference mesh of concrete slab

For the two dimensional problem, see Fig. (3.1), Eq. (3.5) can be rewritten as

follows:

∂w dH ∂ ⎛ ∂H ⎞ ∂ ⎛ ∂H ⎞
= ⎜D ⎟+ ⎜D ⎟⎟ (3.32)
∂H dt ∂x ⎝ ∂x ⎠ ∂y ⎜⎝ ∂y ⎠

where, the moisture capacity, ∂w/∂H and the humidity diffusion coefficient, D

(replace DH) are functions of H, x, y and t. And, the relative humidity, H is function of

x, y and t.

Further simplification of Eq. (3.32) results in:

∂w dH ∂ 2 H ∂D ∂H ∂ 2 H ∂D ∂H
=D 2 + +D 2 + (3.33)
∂H dt ∂x ∂x ∂x ∂y ∂y ∂y

The principle of Alternating-Direction Implicit (ADI) method can be applied

to solve Eq. (3.33) by using two difference equations over two successive time-steps

each of duration ∆t/2:


37
n
H i*, j − H in, j
∆t / 2
⎛ ∂H ⎞
=⎜
∂ w
n
[
⎟ * Di , jδ x H i , j + δ x Di , jδ x H i , j + Di , jδ y H i , j + δ y Di , jδ y H i , j
2 * n * n 2 n n n
] (3.34)
⎝ ⎠i , j

H in, +j 1 − H i*, j n

∆t / 2
⎛ ∂H ⎞
=⎜
∂w
n
[
⎟ * Di , jδ x H i , j + δ x Di , jδ x H i , j + Di , jδ y H i , j + δ y Di , jδ y H i , j
2 * n * n 2 n +1 n n +1
] (3.35)
⎝ ⎠i , j

Where, δx, δ2x and δy, δ2y are the first and second order central-difference operators in

x and y directions, respectively. Eq. (3.34) is implicit only in x-direction and used to

solve for Hi,j* (intermediate value at the end of the first time-step) by knowing Hi,jn

starting from the initial condition. Eq. (3.35) is implicit only in y-direction and used

to solve for Hi,jn+1 (final value at the end of the whole time step) after calculating Hi,j*

in the first step.

The stability of the Alternating-Direction Implicit (ADI) method was

discussed by von Neumann method (Carnahan et al. 1969). It was shown that this

method is unconditionally stable and the solution converges within a discretization

error of O[(∆t)2+(∆x)2+(∆y)2].

3.4 Multiscale Modeling of Drying Shrinkage

Shrinkage of concrete has been evaluated in two different ways. It can be

estimated for structural members, and thus represents the average strain over a large

scale. In this case, the size and shape of the structure, properties of the concrete, and

the effect of the environment have to be taken into account. As a result, the prediction

has been based on some empirical equations.


38

On the other hand, the shrinkage can be evaluated for a small representative

volume of concrete. In this case, the predicting model is a constitutive relationship

among the shrinkage, the humidity, and related material parameters. The structural

configuration is not included. The first type of models (for structural members) is

used in structural analysis, and the second type (constitutive laws) is mainly used for

numerical analysis such as finite element method and finite difference method.

One of the constitutive models to predict free shrinkage of concrete is

suggested by Bazant and Chern (1985). In this model the rate of change of the local

shrinkage strain is related to the rate of change of the relative humidity as follows:

∆ε sh = β sh ∆H (3.36)

in which, βsh is shrinkage coefficient of concrete, which can be characterized by a

multiscale model developed by Xi and Jennings (1997).

Concrete is a multiscale material and can be partitioned into four different scale

levels ranging from nanometer to meter level as shown in Fig. (3.2). The first scale

level is the meter level where most of the civil engineering structures are controlled

by theories and laws applicable on this level. The second is the millimeter level where

most of the physical and mechanical properties of concrete are obtained from samples

and testing procedures applicable on this level. The third is the micrometer level

where the pore structure of cement paste can be measured at this level. Also, the

microstructural features of the cement paste such unhydrated cement particles,

calcium hydroxide (CH) crystals, calcium silicate hydrate (C-S-H) gel and other
39

hydration products are observed at this scale level. The fourth is the nanometer level

where many models for the structure of calcium silicate hydrate (C-S-H) are

developed at this level (Bentz et al. 1995).

Meter Level
(Real Structure)

Millimeter Level
(Concrete)

Micrometer Level
(Cement Paste)

X1000

Nanometer Level
(Calcium Silicate Hydrate)

X1000

X1000

Fig. 3.2: Different scale levels of concrete.

Multiscale mathematical modeling can be used to predict the drying shrinkage

of cement paste and concrete. Concrete as shown in Fig. (3.3) is a particulate

composite material in which the aggregates are the inclusions, and the cement paste is

the matrix. Also, cement paste as shown in Fig. (3.4) is a particulate composite

material composed of unhydrated cement particles; calcium hydroxide (CH) and

ettringite crystals, and calcium silicate hydrate (C-S-H) gel.


40

Fig. 3.3: Composite action of concrete.

Fig. 3.4: Microstructral features of cement paste.

The three-phase model developed by Christensen (1979) to determine

effective elastic properties of a two-phase composite can be used in the multiscale

modeling for concrete and cement paste. This model was originally developed for

elastic properties only, but many researchers (Herve and Zaoui 1990) proved that this
41

model could be extended to nonlinear materials.

The shrinkage of concrete at the millimeter level, see Figs. (3.2) and (3.3), is

equivalent to the effective shrinkage of the three- phase model shown in Fig. (3.5),

where the inclusions (aggregates) represent phase 1 and the matrix (cement paste)

represents phase 2.

Phase 1 (Aggregate)
Phase 2 (Cement Paste)
Effective homogeneous medium (Concrete)

Fig. 3.5: Three-phase effective media model.

The shrinkage coefficient and bulk modulus of concrete (at mesoscale level),

βsh and Kc, can be evaluated (Xi and Jennings 1997) for the composite system:

K1`ε 1sh c12 (3K 2 + 4G2 ) + K 2ε 2sh (1 − c12 )(3K1 + 4G2 )


β sh = (3.37)
K 2 (3K1 + 4G2 ) − 4c12 G2 (K 2 − K1 )

K 2 (3K1 + 4G2 ) − 4c12 G2 (K 2 − K1 )


Kc = (3.38)
(3K1 + 4G2 ) + 3c12 (K 2 − K1 )

where, ε 1sh and ε 2sh , K1 and K2, and G1 and G2 are the shrinkage, bulk modulus and
42

shear modulus of phase 1 (aggregates) and phase 2 (cement paste), respectively; c12 is

the volume fraction of phase 1 (aggregates). The shrinkage of concrete is mainly due

to the shrinkage of cement paste, where the shrinkage of the aggregates can be

neglected ( ε 1sh = 0) and so Eq. (3.37) becomes:

⎧ ⎫
⎪⎪ c12 (K1 / K 2 ) ⎪⎪
β sh = ε 2sh ⎨1 − ⎬ (3.39)
⎪ 1 + (K1 / K 2 − 1) 3 + 4c12 G2 / K 2 ⎪
⎩⎪ 3 + 4G2 / K 2 ⎭⎪

The shrinkage strain of concrete can be calculated using Eq. (3.39) and knowing the

bulk modulus K1, shear modulus G1 and volume fraction of aggregates, shrinkage of

cement paste ( ε 2sh ), bulk modulus K2, shear modulus G2 of cement paste.

Applying the same concept at the micrometer level, see Figs. (3.2) and (3.4),

the elastic properties (K and G) and the shrinkage strain of cement paste can be

calculated using the four-phase composite model shown in Fig. (3.6). In this model

phase 1 is the solid core, which includes the unhydrated cement particles, calcium

hydroxide (CH) and other crystalline products, phase 2 is inner hydration products,

phase 3 is the outer hydration products. Both the inner product and the outer product

are C-S-H (Jennings and Tennis 1994) but the density of the inner product is higher

than that of the outer product (Diamond and Bonen 1993), and the shrinkage of C-S-

H is mainly from the outer product. The bulk modulus and shrinkage strain of cement

paste is equivalent to the bulk modulus and shrinkage strain of the effective

homogeneous four-phase model as follows:


43

K eff
12
(ε effsh )12 ( f 1 + f 2 )(3K 3 + 4G3 ) + K 3ε 3sh (1 − f 1 − f 2 )(3K eff
12
+ 4G3 )
ε sh
eff = (3.40)
K 3 (3K eff
12
+ 4G3 ) − 4( f 1 + f 2 )( K 3 − K eff
12
)G3

( f 1 + f 2 )( K eff
12
− K3 )
K eff = K 3 + (3.41)
1 + (1 − f 1 − f 2 )(( K eff
12
− K 3 ) /( K 3 + 4G3 / 3))

where,

K 1ε 1sh ( f 1 /( f 1 + f 2 ))(3K 2 + 4G 2 ) + K 2 ε 2sh (1 − f 1 /( f 1 + f 2 ))(3K 1 + 4G 2 )


(ε effsh ) 12 = (3.42)
K 2 (3K 1 + 4G 2 ) − 4( f 1 /( f 1 + f 2 ))( K 2 − K 1 )G 2

( f 1 /( f 1 + f 2 ))( K 1 − K 2 )
K eff
12
= K2 + (3.43)
1 + (1 − f 1 /( f 1 + f 2 ))((K 1 − K 2 ) /( K 2 + 4G 2 / 3))

in which, f1, f2 and f3 are the volume fraction of phases 1, 2 and 3, K1, K2 and K3 are

the bulk modulus of phases 1, 2 and 3, G1, G2 and K3 are the shear modulus of phases

1, 2 and 3 and ε 1sh , ε 2sh and ε 3sh are the shrinkage of phases 1, 2 and 3.

Phase 1 (Unhydrated cement and crystalline phases)


Phase 2 (Inner products of cement hydration)
Phase 3 (Outer products of cement hydration)
Effective homogeneous medium (Cement paste)

Fig. 3.6: Four-phase composite model.


44

From test results using the environmental scanning electron microscopy

(Neubauer et al. 1997), it was found that the shrinkage of the phase 1 (unhydrated

cement particles, calcium hydroxide crystals and other crystalline products) is very

small and can be neglected ( ε 1sh =0). Also, It was found that the shrinkage of phase 2

(inner hydration products) in the range of relative humidity between 30% and 100%

is very small and can be neglected ( ε 2sh =0). As a results Eq. (3.42) vanishes and Eq.

(3.40) becomes:

K 3 ε 3sh (1 − f 1 − f 2 )(3K eff


12
+ 4G3 )
ε sh
eff = (3.44)
K 3 (3K eff
12
+ 4G3 ) − 4( f 1 + f 2 )( K 3 − K eff
12
)G3

The volume fraction of the three phases f1 (solid core), f2 (inner hydration

products) and f3 (outer hydration products) can be calculated using Jenning and

Tennis model (1994) as described in Appendix B.

Some of the elastic properties (bulk modulus, K and shear modulus, G) for the

constituent phases of the multiscale shrinkage model are difficult to evaluate,

especially at the microscale level. An inverse method was developed by Xi and

Jennings (1997), in which the elastic properties were evaluated based on the elastic

properties of concrete and cement paste at the millimeter level.

3.5 The Interaction of Drying Shrinkage and Moisture Diffusion

The interactive effect of drying shrinkage and moisture diffusion can be

studied using one of the two alternatives. The first alternative is to consider that the

stresses or strains induced by drying shrinkage are one of the driving forces, and thus,
45

there will be an additional term in the diffusion equation, which corresponds to the

effect of stresses and strains (Majorana and Mazars 1997). The second alternative is

to consider the effect of the drying shrinkage on the two material parameters of the

diffusion equation, Eq. (3.5), i.e. moisture capacity and moisture diffusion coefficient.

In this case, the stress or strain is not appeared explicitly in the diffusion equation, but

as an internal parameter. We applied the second alternative in the present study.

The effects of drying shrinkage on the two material parameters are

incorporated into the diffusion equation by two different methods. For the moisture

diffusion coefficient, we consider that the effect of the damage can be treated by a

way similar to the degradation in the secant modulus of elasticity, where the

degradation of the secant modulus can be characterize by the scalar damage

parameter. For the moisture capacity, we consider that the effect of the damage can

be treated by a way similar to the degradation in the heat capacity of thermoelasticity,

where one of the influential factors to the heat capacity is the coefficient of thermal

expansion.

3.5.1 The effect of shrinkage-induced damage on moisture diffusion coefficient

Let us first consider the drying shrinkage of a concrete layer, which can be

simplified as a plane stress problem, as shown in Fig. (3.7), in which x direction is

perpendicular to the plane, and the stresses in y and z direction are equal and can be

written as:

σ x = τ xy = τ xz = 0 (3.45)
46

γ xy = γ xz = 0 (3.46)

z z

y x

x
Fig. 3.7: Plane stress problem.

Since shrinkage strain is related to volumetric change only,

τ yz = 0 and γ yz = 0 (3.47)

The total strain is the summation of the mechanical strain and the shrinkage strain:

ε xt = ε xm + ε sh
ε ty = ε ym + ε sh (3.48)
ε = ε + ε sh
t
z
m
z

The shrinkage strain can be written in a similar form to the thermal strain as follows:

∆ε sh = β sh ∆H (3.49)
47

where, βsh is shrinkage coefficient, similar to thermal expansion, coefficient and ∆Η

is the change of relative humidity.

Since the total strains vanishes in y and z directions:

ε ym = −ε sh
(3.50)
ε zm = −ε sh

The stresses in y and z direction are equal and can be written as:

Es
σ y =σz = ε sh (3.51)
1 −υ

where, E s is the secant elastic modulus and υ is the Poissons’ ratio of concrete. The

secant elastic modulus can be written in terms of the scalar damage parameter, d as

follows:

E s = (1 − d )E o (3.52)

where, E o is the initial tangent elastic modulus of concrete? Combining Eq. (3.51)

and (3.52)

(1 − d )E o
σy =σz = ε sh (3.53)
1−υ

Since drying shrinkage produce tensile volumetric strains (εx=εy=εz=εsh), the scalar

damage parameter, d can be written in terms of the shrinkage strain (εsh) using a
tensile stress-strain relationship of concrete, and so Eq. (3.53) can be rewritten as:
48

σy =σz =
(1 − d (ε sh ) )Eo ε
1−υ
sh
(3.54)

The incremental stress can be written as:

Eo ⎛ ∂d ⎞
dσ y = dσ z = ⎜⎜ [1 − d (ε sh ) ] − ε sh ⎟⎟ dε sh (3.55)
1−υ ⎝ ∂ε sh ⎠

Starting from a certain time step, n where the stress and the strain are known, the

stress and the strain at the time n+1 can be written as:

ε shn +1 = ε shn + dε sh

σ yn +1 = σ yn + dσ y (3.56)

σ yn +1 = σ yn + dσ y

The scalar damage model, was first introduced by Kachanov (1958), based on

the assumption that the area (or the volume) with damage cannot hold any load, so,

there is a reduction in the effective cross section area as loading level increase, which

can be generally expressed by a reduction in the modulus of elasticity. Following the

same concept, we can assume that, for moisture diffusion in concrete, the damaged

area (or volume) cannot resist the penetration of moisture. So, there is also a

reduction in the effective cross section area, which can be generally expressed as an

increase in the volume of pathway for the moisture to diffuse in concrete.

Therefore, with the increase of damage, the concrete diffusion coefficient

increases. Similar to the effect of damage on modulus of elasticity, the effect of


49

damage on the moisture diffusion coefficient, DH, can be expressed in terms of the

scalar damage parameter, d:

D H (H )
D H (H , d ) = (3.57)
1− d

in which, DH(H) is the diffusion coefficient of intact of intact concrete, and DH(H,d)

is the diffusion coefficient with damage.

3.5.2 The effect of drying shrinkage on moisture capacity

First, we compare the similarities between the effect of thermal stress on heat

capacity of composite and the effect of drying shrinkage stress on moisture capacity.

Eq. (3.5) for moisture diffusion has the same form as the equation for heat

conduction, Eq. (3.58)

∂T
Ct = div[ K T grad (T )] (3.58)
∂t

where, T = temperature; KT = thermal conductivity, Ct = heat capacity. Moreover, the

constitutive equations for thermal stress, Eq. (3.59), and the constitutive equations for

shrinkage stress, Eq. (3.60), also have the same form

(
σ ij = Cijkl ε ij − α ijθ ) (3.59)

(
σ ij = Cijkl εij − βij η ) (3.60)

in which σij = stress tensor; εij = strain tensor; Cijkl = elastic tensor; αij = coefficients
50

of thermal expansion; β = coefficients of drying shrinkage; θ is the change of


ij

0
temperature in Kelvin from the reference state T , T = T + θ ; η is the change of
0

pressure from the reference state p0, p=p0+η, where p0 = reference pressure in Pa.

The similarities of the corresponding coefficients for heat conduction and

moisture transfer are obvious: αij verses βij; KT verses DH; and Ct verses ∂w/∂H. The

idea of the comparative study is that the composite models developed for heat

conduction may possibly be borrowed for evaluating the effective parameters of

moisture transfer. Based on Eqs. (3.58) and (3.59), the effect of thermal strain on

heat capacity has been determined by Rosen and Hashin (1970) using a generalized

method based on extreme energy principle. Rosen and Hashin model turned out to be

the only model for the effect of thermal strain on effective heat capacity. Similarly,

using Eqs. (3.5) and (3.60) together with the extreme energy principle, Xi (1995a and

1995b) derived an analytical model for the effect of shrinkage strain on moisture

capacity for two-phase composite materials

2
⎛ ⎞ ⎡ _ ⎤
⎜ ⎟ ⎢⎛⎜ 1 ⎞⎟ 1 ⎥

p ⎜ βi − βm ⎟
0
n, p = n, p + 9
p p
⎢⎜ ⎟− ⎥
RT 0 ⎜ 1 1 ⎟ ⎢⎜ K ⎟ K ⎥
(3.61)
⎜ − ⎟
⎝ Km Ki ⎠ ⎢⎣⎝ ⎠ ⎥⎦

p0
n −n =
v
,p
p
,p C ijkl β ij β kl (3.62)
RT 0

with,
51

⎛ _ ⎞
⎜ 1 ⎟ gm gi
_
n = gmn
p
,p
p
, pm + gin p
, pi and ⎜⎜ K ⎟⎟ = K + K (3.63)
m i
⎝ ⎠

p v
in which n, p is the moisture capacity at constant stress in moles/m3, n, p is the

moisture capacity at constant deformation in moles/m3; the subscript p means that the

moisture capacity is in terms of pressure p; p = p0+ η; n= moles of water adsorbed

per volume of the composite material; subscript m is for matrix and i for inclusion;

n,ppm and n,ppi are the moisture capacities at constant stress for matrix and inclusion,

respectively; K = bulk modulus in Pa; R = gas constant in Joules/mole/K; β =

shrinkage coefficient; where β i and β m are in terms of η; and g = volume fraction.

Eqs. (3.61) through (3.63) are general expressions for moisture capacities of two

phase composites. In order to apply these equations to concrete, some modifications

are needed.

First of all, the distinction between the moisture capacities at constant stress

and constant deformation has not been directly identified by experiments, especially

for cement paste and aggregate. In current engineering practice, experiments of

equilibrium isotherm are carried out in a condition of neither constant deformation

nor constant stress. After a detailed analysis, Xi (1995a) concluded that the

v p
difference between n, p and n, p are negligible, which means n,vp ≈ n,pp = n, p .

Superscripts v and p may be dropped. Furthermore, all adsorption test data and

shrinkage coefficients for concrete have been obtained in terms of relative humidity H

rather than pore pressure p, hence, Eq. (3.61) must be expressed in terms of H, too. H
52

= p/ps, where p s is saturation pressure and depends on temperature. At isotherm

condition, ps = constant, so, n, p may be replaced by ∂w/∂H, and p by H . Then, the


0 0

final form of the moisture capacity suitable for concrete is

⎧ ⎛ ⎞
2
⎡ _ ⎤⎫
⎪ ⎜ ⎟ ⎢⎜ 1 ⎟ 1 ⎥ ⎪⎪
⎛ ⎞
∂w ⎛ ∂w ⎞
=⎜ ⎟
⎪ H0
+ ψ ⎨9 ⎜ βi − βm ⎟ ⎢⎜ ⎟− ⎥⎬
∂H ⎝ ∂H ⎠ avg ⎜ 1 1 ⎟ ⎢⎜ K ⎟ K ⎥ ⎪ (3.64)
⎪ RT
0
⎜ − ⎟
⎪ ⎝ Km Ki ⎠ ⎢⎣⎝ ⎠ ⎥⎦ ⎪
⎩ ⎭

in which (∂w/∂H)avg is the moisture capacity of concrete without considering

the effect of shrinkage as shown in Eq. (3.9); β i and β m are in terms of η, which

now stands for the change of relative humidity H, and ψ is a factor used to convert

the units of the second term in the right hand side of Eq. (3.61) from mole/m3 to

gram/gram. In Eq. (3.64), all material parameters are for an individual phase (either

cement paste or aggregate) except K, which is the effective bulk modulus of concrete.

K may be related to the properties of cement paste and aggregate by a proper

composite model. In the present study, we used the so-called three-phase model

(Christensen 1979)

g i (K i − K m )
K = Km +
Ki − Km (3.65)
1+ gm
4
K m + Gm
3

where Gm = shear modulus of the matrix (cement paste) in Pa. As one can see

from Eq. (3.64), the moisture capacity of concrete is not just the weighted average of
53

the moisture capacities of the matrix and the inclusion, it depends strongly on the

shrinkage coefficients of the two constituents (see the second term).

In the case that the shrinkages of the inclusion and matrix are the

same, the second term in Eq. (3.64) vanishes, which means that there is no coupling

effect. Otherwise, the coupling effect must be taken into account. For concrete,

β m ≥ β i and K i ≥ K m (the matrix shrinks more than the inclusion and the inclusion is

stronger than the matrix), the additional term depends strongly on shrinkage values of

the constituent phases. With increasing β m β i the value of the additional term

increases drastically. Numerical analyses by Xi (1995a and 1995b) showed that the

influence of the shrinkage ratio is more significant than that of the ratio of bulk

modulus, and furthermore, the second term in the right hand side of Eq. (3.64) is

always positive when β m ≥ β i and K i ≥ K m , which means that the differential

shrinkage between the two constituent phases increases the moisture capacity of the

composite.

3.6 Numerical Applications

The coupled problem of moisture diffusion and drying shrinkage discussed in

the previous sections was implemented and solved in a computer program. The

moisture diffusion was solved using the Alternating-Direction Implicit (ADI) finite

difference method discussed in section 3.3.2. The drying shrinkage was modeled

according to the model shown in Eq. (3.36). As a numerical example, a 10 cm depth

concrete slab similar to the slab shown in Fig. (3.1) was analyzed. The concrete has a
54

water-cement ratio (w/c) of 0.55, aggregate volume fraction (gi) of 0.65, average

compressive strength (28 days) of 34 MPa and moist cured for 28 days before drying.

The slab, which is initially saturated (Hini=100%), is exposed to drying on the top

surface (Henv= 50 %) after curing. The effect of micro-cracking due to drying

shrinkage is taken into account through the scalar damage parameter. The hygral

stresses are calculated based on the assumption of plane stress for concrete. The

tensile stress strain curve used in the example is the one suggested by Zhen-hai and

Xiu-qin (1987) and it is shown in Fig. (3.8). The graph shows that the strain-strain

relationship is composed of two parts. The ascending (hardening) part is described by

high order polynomial. The descending (softening) part is described by an equation

with α = 0.312 f t ' 2 and β = 1.7, where f t ' is the ultimate tensile stress in (MPa).

f ' = 0.5 f'


t c
(MPa)

f ' [1.2 (εt /εtu) - 0.2 (εt /εtu)6 ] ε [ ε


t t tu

f =
t
Tensile stress, f

t f ' (εt /εtu)


t
β
(εt /εtu) + α[(εt /εtu) −1] εt > εtu

Eo = 4700 f'
c

ε tu = 44x10 -6 ft'
Tensile strain, ε t

Fig. 3.8: Tensile stress strain diagram.


55

The variation of the pore relative humidity (H), shrinkage strain (εsh), scalar

damage parameter (d) and hygral stresses (σy or σz) with the depth from the exposure

surface (top surface) are shown in Figs. (3.9) through (3.12). From Fig. (3.9), we can

see that the drying process becomes slower for concrete at deeper locations from the

exposed surface. The shrinkage is related to the moisture loss, and as a result the

exterior part of concrete slab has larger strain as shown in Fig. (3.10). The build up of

strains near the exposed surface results in more degradation of the elastic modulus

and thus more damage is expected in the exterior skin of concrete and the damage

become less as going deeper from the exposed surface as shown in Fig. (3.11). The

hygral stresses are tensile stresses produced by the restraining of the drying shrinkage

by interior concrete, which shrinks less. Fig. (3.12), shows that stresses are larger in

the outer skin of the concrete and they vanish as going deeper. Therefore, the drying

process, drying shrinkage, the local damage and hygral stresses in the concrete

structural members are related to their surface area-volume ratio. Shallow structural

members such as slabs, decks and walls have larger surface area with respect to their

volume, and thus are subjected to more moisture loss and more deterioration due to

drying shrinkage. On the other hand, other structural members of less surface area

with respect to their volume have less drying shrinkage and the damage is limited

only to surface layer.


56

0.95 t = 1 month
t = 2 months
0.9 t = 4 months

Pore relative humidity, H


0.85
0.8

0.75

0.7
0.65

0.6
0.55

0.5
0 1 2 3 4 5 6 7 8 9 10
Depth (cm)

Fig. 3.9: Variation of relative humidity with the depth.

450

400 t = 1 month
Shrinkage strain, ε sh (microstrain)

t = 2 months
350 t = 4 months

300

250

200

150

100

50

0
0 1 2 3 4 5 6 7 8 9 10
Depth (cm)

Fig. 3.10: Variation of shrinkage strain with the depth.


57

0.8

0.7 t = 1 month
t = 2 months

Scalar damage parameter, d


0.6 t = 4 months

0.5

0.4

0.3

0.2

0.1

0
0 1 2 3 4 5 6 7 8 9 10
Depth (cm)

Fig. 3.11: Variation of scalar damage parameter with the depth.

3.5
Hygral stress, σy or σz (MPa)

2.5

1.5

1
t = 1 month
0.5 t = 2 months
t = 4 months
0
0 1 2 3 4 5 6 7 8 9 10
Depth (cm)

Fig. 3.12: Variation of hygral stresses with the depth.


58

The variation of pore relative humidity, the shrinkage strain, the scalar

damage parameter and the hygral stresses with time of exposure are shown in Figs.

(3.13) through (3.16). The time of exposure is started after the curing period of

concrete. Fig. (3.13) shows the drying processes at different depths in the concrete.

One can see that the rate of drying is higher near the exposed surface. A 20% drop in

the relative humidity occurred approximately within the first 5 days of exposure at 1

cm depth, 25 days of exposure at 3 cm depth and 50 days of exposure at 5 cm depth.

Fig. (3.14) shows that the variations of shrinkage in the concrete specimen at different

depths with increasing time of exposure. It shows that, the rate of the strain

accumulation increases with an increase of exposure time. The strain reaches 0.0001

approximately within 10 days at 1 cm depth, 50 days at 3 cm depth, and 85 days at 5

cm depth. The difference in the shrinkage strain at different depths (i.e. local strain

gradient) results in shrinkage stress and the local damage. Fig. (3.15) shows the

increase of the scalar damage parameter due to the shrinkage stress. From the figure

we can see the damage starts approximately after 5 days at 1 cm depth, 40 days at 3

cm depth and 70 days at 5 cm depth. Also the figure shows that a 40% loss in the

initial elastic modulus occurs approximately after 20 days of exposure at 1 cm depth,

95 days of exposure at 3 cm depth and 125 days of exposure at 5 cm depth. Fig.

(3.16), shows the stress history at different depths in the concrete. The stress in

concrete reaches the maximum tensile strength after approximately 25 days at 1 cm

depth, 105 days at 3 cm depth and 135 days at 5 cm depth. Based on these numerical

results, we can see that long period of drying should be avoided. This is especially

important for shallow structures and early age concrete, where the effect of drying
59

becomes very severe and the damage due to drying reduces the service life of the

concrete structures.

Depth = 1.0 cm
0.9 Depth = 3.0 cm
Pore relative humidity, H

Depth = 5.0 cm

0.8

0.7

0.6

0.5
0 30 60 90 120 150 180
Time (days)

Fig. 3.13: Variation of relative humidity with time of exposure.

350
Depth = 1.0 cm
Depth = 3.0 cm
300
Shrinkage strain, ε sh (microstrain)

Depth = 5.0 cm

250

200

150

100

50

0
0 30 60 90 120 150 180
Time (days)

Fig. 3.14: Variation of shrinkage strain with time of exposure.


60

0.7

0.6

Scalar damage parameter, d


0.5

0.4

0.3

0.2

0.1 Depth = 1.0 cm


Depth = 3.0 cm
Depth = 5.0 cm
0
0 30 60 90 120 150 180
Time (days)

Fig. 3.15: Variation of scalar damage parameter with time of exposure.

3.5
Hygral stress, σy or σz (MPa)

2.5

1.5

1 Depth = 1.0 cm
Depth = 3.0 cm
0.5 Depth = 5.0 cm

0
0 30 60 90 120 150 180
Time (days)

Fig. 3.16: Variation of hygral stresses with time of exposure.


61

Another important observation is that strain softening induced by shrinkage

reaches 5 cm or more in the concrete, which is the thickness of concrete cover in most

of concrete structures. Of course, the depth of softening front depends on the concrete

quality. This numerical example gives us an important estimation on the damage in

concrete due to drying shrinkage.

3.6.1 Effect of damage

The drying shrinkage creates local damage in concrete, which is taken into

account by the scalar damage parameter. The effect of the damage on the diffusion of

moisture is taken into account by the two models developed in the present study for

the moisture diffusion coefficient and the moisture capacity. Figs. (3.17) and (3.18)

show the coupling effect: profiles of the relative humidity and shrinkage strain in the

concrete at different exposure times with and without considering the effect of the

damage. The starting point of the profiles is from the exposed surface. In Fig. (3.17),

all curves eventually reach the equilibrium distribution of 50%, which is the

environmental relative humidity. One can see that the humidity profiles including the

shrinkage-induced damage approach the equilibrium distribution in a faster rate,

which indicates that the shrinkage-induced damage accelerates the diffusion process

and increases the rate of drying. In turn, as shown in Fig. (3.18), the accelerated

drying process increases the shrinkage strain. For example, comparing the two curves

exposed after three months with and without the damage, the curve with the damage

is much higher. One can also see that, at a fixed depth (say 4 cm), the incremental

shrinkage strain with the damage (e.g. the difference between one month and three

months) is much higher than the corresponding increment without the damage, which
62

indicates that the rate of drying shrinkage is increased due to the damage.

Figs. (3.17) and (3.18) show very clear that there is an active interplay between

the drying process and the resulting shrinkage. Basically the moisture loss causes

drying shrinkage, which generates the damage. In turn, the damage increases

diffusion coefficient of concrete, and thus accelerates the moisture transfer.

0.9
Pore relative humidity, H

0.8

0.7

0.6

0.5
1 month (Damage Considerd)
0.4 3 months
6 months
1 month (Damage Not Considerd)
0.3 3 months
6 months
0.2
0 2 4 6 8 10
Depth (cm)

Fig. 3.17: Effect damage on drying of concrete.


63

600

1 month (Damage Considerd)


500 3 months
6 months
1 month (Damage Not Considerd)
3 months
ε sh (microstrains)
400
6 months

300

200

100

0
0 2 4 6 8 10
Depth (cm)

Fig. 3.18: Effect of damage on shrinkage of concrete.

3.6.2 Effect of water-cement ratio

The water-cement ratio affects both moisture diffusion and drying shrinkage of

concrete. The effect of water-cement ratio on drying and shrinkage of concrete is

shown by Figs. (3.19) and (3.20). The figures show the time variation of the internal

relative humidity and shrinkage strains at two different depths from the exposed

surface (1 cm and 3 cm) for two different water-cement ratios (0.45 and 0.65). Both

figures show that as the water-cement ratio increases the rate and the magnitude of

drying increases, thus resulting in larger shrinkage strains and higher rate of

shrinkage. By increasing water-cement ratio, the porosity of concrete increases

which, accelerates the moisture diffusion and thus increases the rate of drying.
64

w/c = 0.45
0.9
w/c = 0.65

Pore relative humidity 0.8

at 3 cm depth
0.7

0.6
at 1 cm depth

0.5
0 10 20 30 40 50 60 70 80 90
Time (days)

Fig. 3.19: Effect of w/c on drying of concrete.

450
w/c = 0.45
400
w/c = 0.65
350

300
ε sh (microstrains)

250
at 1 cm depth
200

150
at 3 cm depth
100

50

0
0 10 20 30 40 50 60 70 80 90
Time (days)

Fig. 3.20: Effect of w/c on shrinkage of concrete.


65

3.6.3 Effect of aggregate volume fraction

The aggregate volume fraction in concrete affects both moisture diffusion and

drying shrinkage processes of concrete. The effect of aggregate volume fraction on

drying of concrete is shown by Fig. (3.21). The figure shows the variation of relative

humidity with time at two different depths from the exposed surface (1 cm and 3 cm)

for two different aggregate volume fractions (0.55 and 0.75). From the figure, one

can see that adding more aggregate to concrete decreases the rate and the magnitude

of the loss of relative humidity. In general, aggregate has lower diffusion coefficient

than cement paste, which is mainly due to the discontinuity in pore system of

aggregate. Therefore, the addition of more aggregates to concrete reduces its

moisture diffusion coefficient. At the same time, more aggregate in concrete makes

the flow paths longer and more tortuous, which also help to reduce the diffusion

coefficient of concrete.

The effect of aggregate volume fraction on shrinkage of concrete is shown in

Fig. (3.22). The figure shows the time variation of shrinkage strain at two different

depths from the exposed surface (1 and 3 cm) for two different aggregate volume

fractions (0.55 and 0.75). The graph shows that addition of more aggregate to

concrete decreases the rate and the magnitude of the shrinkage of concrete. This can

be attributed to two reasons. One is the reduction of diffusion coefficient of the

concrete, and the other is that aggregates do not shrink.


66

gi = 0.55
0.9 gi = 0.75

Pore relative humidity 0.8


at 3 cm depth

0.7

0.6 at 1 cm depth

0.5
0 10 20 30 40 50 60 70 80 90
Time (days)

Fig. 3.21: Effect of aggregate volume fraction on drying of concrete.

800
gi = 0.55
700 gi = 0.75

600
ε sh (microstains)

500

400

300
at 1 cm depth
200
at 3 cm depth

100

0
0 10 20 30 40 50 60 70 80 90
Time (days)

Fig. 3.22: Effect of aggregate volume fraction on shrinkage of concrete.


67

3.7 Conclusions

1. The interactive effect of moisture diffusion and drying shrinkage in concrete was

studied systemically in this chapter. Multiscale modeling (at micro-, meso- and

macro-scale) was used for modeling the moisture diffusion and drying shrinkage

processes of concrete. The diffusion equation and the drying shrinkage of

concrete are established for a representative volume of element at the macro-

scale. At the meso-scale, both models consider concrete as a two-phase

composite with the aggregate as inclusion and the cement paste as matrix. At the

micro-scale, the influential parameters on microstructure of cement paste, such as

water-cement ratio and curing time, are included.

2. The effect of shrinkage-induced damage is characterized by modifying the two

material parameters in the moisture diffusion equation, i.e. the moisture capacity

and the humidity diffusion coefficient. In this manner, the shrinkage-induced

stress or strain is not shown in the diffusion equation as an explicit driving force,

but as an implicit parameter.

3. Since the microcracks induced by drying shrinkage are randomly distributed

without a preferential orientation, the shrinkage-induced damage may be

considered as isotropic and simply evaluated by the isotropic scalar damage

parameter, d. The effect of d on humidity diffusion coefficient is characterized by

a similar manner as the effect of d on the secant modulus proposed by Kachanov.

4. The effect of drying shrinkage on the moisture capacity is included by an

analytical model based on non-equilibrium thermodynamics and minimum

potential energy principle for a two-phase composite. The influence of the drying
68

shrinkage depends mainly on the difference between the shrinkage coefficients of

the two constituent phases: the aggregate and the cement paste. When the two

shrinkage coefficients are the same, the moisture capacity of the concrete is

simply the volumetric average of the moisture capacities of the two components;

when the two shrinkage coefficients are different, the coupling effect must be

taken into account. With increasing differential shrinkage, the moisture capacity

of concrete increases.

5. A numerical solution for the moisture diffusion equation, taken into account the

coupling effect of the drying shrinkage is developed by using Alternating-

Direction Implicit (ADI) finite difference method. The solution is stable and

convergent.

6. The present model can predict the basic trend for the coupling effect between the

drying shrinkage of concrete and the moisture diffusion in concrete. The

shrinkage-induced damage increases the humidity diffusion coefficient; the

differential shrinkage between the aggregate and the cement paste increases the

moisture capacity of concrete. As a result, the moisture diffusion process in

concrete is accelerated by the drying shrinkage. In turn, the accelerated drying

process increases the shrinkage of concrete.

7. The present model can predict the effects of water-cement ratio and aggregate

volume fraction on the drying shrinkage of concrete, the moisture diffusion and

the coupling effect between the two. Both water-cement ratio and aggregate

volume fraction have major impact on the internal humidity distribution and

shrinkage strain development of concrete.


CHAPTER 4

COUPLED MOISTURE AND CHLORIDE DIFFUSIONS

4.1 Introduction

There are mainly three driving forces of the ingress of chloride ions in non-

saturated concrete. The first driving force is the non-uniform distribution of chloride

ions, which is important for both saturated and non-saturated concrete. The second

driving force is the diffusion of moisture, which is important in the case of non-

saturated concrete. The third driving force is the ionic migration driven by an electric

potential gradient, which is important only in some special cases. The rapid chloride

permeability test in ASTM C1202 and AASHTO T277 involves the effect of the

second and the third driving forces, but not the first one. AASHTO T259-80 (90-day

ponding test) involves the first and the second driving forces, but not the third one. In

reality, the dominant driving forces for chloride penetration in the concrete of

highway pavements and bridge decks are the first and the second ones, i.e. moisture

diffusion and non-uniform chloride distribution.

Many researchers have investigated the diffusion of chloride ions into

concrete. The review of the available literature on this topic shows that there have

been two ways for formulating the chloride penetration in concrete. One is based on

the conventional diffusion theory at the macroscopic level, that is, Fick's first law and

mass conservation (Saetta et al., 1993; Frey et al., 1994; Wee et al., 1997;
70

Swaddiwudhipong et al., 2000). The other is based on the basic laws of

electrochemistry at the nanometer level, i.e. Nernst-Plank equation and Nernst-

Einstein equation (Andrade 1993 and Chatterji 1994). A new chloride penetration

model has recently been developed for saturated concrete (Xi and Bazant, 1999; Xi et

al., 2000) using the multiscale modeling approach (Xi and Jennings, 1997). In this

model, the conventional diffusion theory was used to formulate the diffusion equation

at the macroscopic level (for the so-called representative volume element); and the

composite theory was used at the mesoscopic level to incorporate the effect of

aggregate on chloride diffusion coefficient; and furthermore, the diffusion

mechanisms of chloride ions were taken into account at the microscopic level. The

comprehensive model can be used to simulate and to predict the chloride diffusion in

saturated concrete satisfactorily.

In reality, concrete is often found in a non-saturated condition rather than the

saturated condition. Swaddiwudhipong et al. (2000) developed a mathematical model

to predict the effect of moisture transport on chloride ingress in partially saturated

concrete. Their model includes the effect of moisture transport due to concentration

gradient (diffusion), capillary action (sorption) and pressure gradient (permeability).

Costa and Appleton (1999) presented simplified models for predicting diffusion

coefficient for chloride ions, surface chloride concentration and their dependence on

the exposure time, exposure conditions and concrete properties. Saetta et al. (1993)

analyzed the diffusion of chloride ions in non-saturated concrete. The non-linear and

coupled partial differential equations were solved using the finite element method.

The solution took into account several factors involved in the partial differential
71

equations and their dependence on the concrete mix parameters and the

environmental conditions.

In this chapter, a comprehensive model for chloride penetration in non-

saturated concrete was developed. This work is a continuation and further extension

of the research work of Xi and Bazant (1999) for the chloride penetration in saturated

concrete. The framework of the model is along the line of the multiscale modeling

(Xi et al., 2000). The coupled moisture and chloride diffusion is a complex problem;

it is not uncommon to find mathematical models with too many empirical parameters

to be determined by curve fitting, whereas the real physical meanings of the

parameters are lost. Our strategy is to establish the material models based on

analytical results first, and if not possible, empirical models are introduced based on

dominant physical or chemical mechanism(s) at different scale levels, and then each

individual empirical model is calibrated by related test data. A new numerical

method is used to solve for the coupled partial differential equations of chloride

penetration and moisture diffusion. The prediction of the theoretical model is

compared with available experimental results. Finally, the effect of moisture

diffusion on the chloride penetration is analyzed by numerical examples.

4.2 Basic Formulation

First of all, the governing equations representing the three driving forces for the

chloride diffusion must be established. When chloride ions enter concrete, some of

the chlorides react chemically with the cement paste components (like the reaction

with C3A and C4AF to form Friedels salt) and become chemically bound and some
72

become physically adsorbed to the cement gel or physically trapped in gel pores or

capillary pores. These chlorides are called bound chloride, while the others, called

free chloride, diffuse around in concrete. The sum of the two types of chloride is the

total chloride content in the concrete. Apparently, the steel corrosion is related only

to the free chloride content not to the total chloride content, because the bound

chloride is not free to move and never reaches the surface of embedded steel bars.

Therefore, the proposed governing equation for chloride diffusion is formulated in

terms of the free chloride concentration. Similarly, for the second driving force of the

diffusion process (i.e. the moisture diffusion), the moisture flux can only carry the

free chloride ions not the bound chloride.

The flux of chloride ions (J) through a unit area of porous media in a unit time

can be expressed using Fick’s first law:

J = − Dcl grad (C f ) − Dele grad (φ ) (4.1)

where, Dcl = chloride diffusion coefficient (cm2/day), Cf = free chloride concentration

(in gram of free chloride per gram of concrete, g/g), Dele = ionic diffusion coefficient

and φ = electric potential.

The mass balance of chloride ions can be expressed using Fick’s second law

as follows:

dCt dw
= − div ( J ) + µC f (4.2)
dt dt
73

in which Ct = total chloride concentration (in gram of total chloride per gram of

concrete, g/g); w = moisture content of concrete (in gram of moisture per gram of

concrete, g/g) and µ is a unit conversion factor for converting the unit of the second

term in the right hand side of Eq. (4.2) to gram of chloride per gram of concrete.

µ = 1/(ρsol βsol), where ρsol is the density of pore solution in gram of pore solution per

liter, and βsol is the ratio of the volume of pore solution to the weight of concrete. βsol

will be explained in detail in later sections. By substituting Eq. (4.1) into Eq. (4.2)

and decoupling the free chloride concentration (Cf) from the total chloride

concentration (Ct), the Partial Differential Equation (PDE), that governs the diffusion

of chloride ions in partially saturated concrete, can be written as:

∂Ct ∂Ct ∂C f ∂w
= = div [ DCl grad (C f )] + µ C f + div [ Dele grad (φ )] (4.3)
∂t ∂C f ∂t ∂t

in which, one can see clearly that the first term in the right hand side represents the

diffusion of chloride ions from high concentration to low concentration

(concentration gradient); the second term is for the diffusion of moisture, which

serves as a carrier of the chloride ions; and the third term describes the migration of

chloride ions driven by an electric potential gradient. In non-saturated concrete,

∂w/∂t in the second term of the governing PDE involves the diffusion of moisture,

which is described as following:

∂w ∂w ∂H
= = div [ DH grad ( H )] (4.4)
∂t ∂H ∂t
74

in which, H=Pore relative humidity; ∂w/∂H=Moisture capacity; DH=humidity

diffusion coefficient.

It is important to note that the third term in Eq. (4.3) becomes important only

when a strong external electric potential is imposed on the concrete member, as in the

case of cathodic protection. This term will not be investigated in the present work

and will be dropped from the governing equation, Eq. (4.3). Thus the penetration of

chloride ions in a non-saturated concrete is governed by the coupled system of PDEs:

∂w ∂w ∂H
= = div [ DH grad ( H )] (4.5)
∂t ∂H ∂t

∂C t ∂C t ∂C f ∂w
= = div [ DCl grad (C f )] + µ Cf (4.6)
∂t ∂C f ∂t ∂t

4.3 Material Parameters

In order to develop numerical solutions for the governing equations, Eqs. (4.5)

and (4.6), the material parameters involved in the equations must be determined first.

The material parameters are the moisture capacity (∂w/∂H) and the humidity diffusion

coefficient (DH) in Eq. (4.5), chloride binding capacity (∂Cf/∂Ct) and the chloride

diffusion coefficient (Dcl) in Eq. (4.5). The moisture capacity and the humidity

diffusion coefficient are discussed in the previous chapter (chapter 3). The chloride

binding capacity and the chloride diffusion coefficient will be discussed in this

chapter.
75

4.3.1 Chloride binding capacity

The total chloride concentration, Ct, is the summation of the free chloride
concentration, Cf, and the bound chloride concentration, Cb

Ct = C f + Cb (4.7)

The chloride binding capacity is the ratio between the change of free chloride

concentration and the total chloride concentration. The chloride binding capacity can

be expressed as follows:

dC f 1
= (4.8)
dCt dCb
1+
dC f

The ratio dCb/dCf can be obtained experimentally. Based on available test data, a

prediction model for the binding capacity has been developed based on Freundlich

isotherm (Xi and Bazant, 1999):

dC 1
=
f
A −1
(4.9)
dC t A 10 β C − S − H ⎛
B
Cf ⎞
1+ ⎜ ⎟
35450 β sol ⎜⎝ 35 .45 β sol ⎟

where, A and B are two material constants related to chloride adsorption and equal to

0.3788 and 1.14, respectively (Tang and Nilson 1993). The binding capacity depends

on the two parameters, βsol and βC-S-H (Xi and Bazant, 1999).

The parameter βsol, as explained for Eq. (4.2), is the ratio of pore solution to
76

concrete (L/g). This parameter reflects the effect of the structure of hydration

products and can be written as:

Vsol wsol n( H , T )
β sol = = = (4.10)
wconc ρ sol wconc ρ sol

in which Vsol and wsol are the volume and the weight of pore solution, wconc is

the weight of concrete, ρsol = density of the pore solution in (g/L) and it depends on

the chloride concentration. For simplicity, ρsol can be used as the density of the pore

water. The weight ratio of pore solution to concrete (wsol/wcoc) is the so-called

chloride adsorption isotherm, which depends on relative humidity H, temperature T

and the pore structure of concrete. Because of the lack of experimental results on the

chloride adsorption isotherms for concrete, the water adsorption isotherm is used

instead. Thus, n(H,T) can be expressed as follows:

n( H , T ) = f cp n cp ( H , T ) + f agg n agg ( H , T ) (4.11)

in which, fcp and fagg are weight percentages of cement paste and aggregates and

ncp(H,T) and nagg(H,T) are the water adsorption isotherms of cement paste and

aggregate. The adsorption isotherm of aggregate and cement paste can be calculated

as shown in Appendix A.

The parameter βC-S-H is the weight ratio of C-S-H gel to concrete (g/g). This

parameter determines the effect of the cement composition and age on the volume

fraction of C-S-H gel and can be written as:


77

wC − S − H
β C −S −H = (4.12)
wtotal

in which wC-S-H and wtotal are the weight of C-S-H gel and the total weight of concrete.

Since there is no difference in the densities of the C-S-H gel and the cement paste, the

parameter βC-S-H is approximately equal to fC-S-H and can be calculated using Eq. (B.2)

of Appendix B.

In Eq. (4.9), one can see that, since the parameter A < 1, ∂Cf/∂Ct = 0 when the

free chloride concentration Cf is zero. For saturated concrete, the second term

vanishes in Eq. (4.6), it is apparent that ∂Cf/∂Ct = 0 leads to ∂Cf/∂ t = 0. This means

that Cf is a constant at all time steps, and it equals to the initial free chloride

concentration, which is zero. This means that the chloride diffusion will never start.

This is the problem when Freundlich isotherm is used when the free chloride

concentration is low. According to Tang and Nilsson (1993), Freundlich isotherm is

valid only when Cf is large (> 0.01 mol/L), and Langmuir isotherm should be used

when Cf is very small (< 0.05 mol/L). Therefore, in the present study, the binding

capacity based on Langmuir isotherm is used for the initial condition with Cf = 0, and

Eq. (4.9) based on Freundlich isotherm is used for Cf > 0. The following is the

derivation of the binding capacity based on Langmuir isotherm.

First of all, Langmuir isotherm is of the form

1 1 1 1
= + (4.13)
Cb′ k ′Cbm C ′f Cbm
78

in which k’ is an adsorption constant, and Cbm is the bound chloride content at

saturated monolayer adsorption (Tang and Nilsson 1993). k’ and Cbm can be

determined by curve fitting based on adsorption test data; C ′b and C ′f are the bound

and free chloride concentrations, respectively, but in a unit system differing from Cb

and Cf. C ′b is in milligrams of bound chloride per gram of calcium silicate hydrate

(C-S-H) gel (mg/g); C ′f is in moles of free chloride per liter of pore solution (mol/L).

The advantage of using such a unit system in Eq. (4.13) is that the resulting Langmuir

isotherm is independent of the water-cement ratio and the aggregate content of

concrete, because it is based on the amounts of C-S-H gel and pore solution.

However, for practical engineering problems, it is necessary to use the unit system

defined for Cb and Cf, because the amounts of C-S-H gel and the pore solution vary

with the hydration process and are not easy to measure in conventional laboratories.

Using the two factors, βsol and βC-S-H, as defined in Eqs. (4.10) and (4.12), the two

different unit systems can be correlated

Cf
C ′f = (4.14)
35.45β sol

1000Cb
C′b = (4.15)
β C −S −H

Substituting Eqs. (4.14) and (4.15) into Eq. (4.13), we obtain

1 1000 ⎡ 35.45β sol 1 1 ⎤


= ⎢ + ⎥ (4.16)
Cb β C − S − H ⎣⎢ k ′Cbm C f Cbm ⎥⎦
79

Eq. (4.16) can be re-written in a simple form as

1
Cb = (4.17)
1
β+
αC f

in which,

k ′Cbm β C − S − H
α= (4.18)
35450 β sol

1000
β= (4.19)
β C − S − H Cbm

Derivative of Eq. (4.17) with respect to Cf gives

dCb 1
= (4.20)
dC f α (C ) 2 ( β + 1 ) 2
αC f
f

then, by substituting Eq. (4.20) into Eq. (4.8), we obtain the binding capacity based

on Langmuir isotherm:

dC f 1 1
= = (4.21)
dCt dCb 1
1+ 1+
dC f 1
α ( βC f + ) 2
α

when the free chloride concentration Cf approaches zero, the binding capacity
approaches to 1/(1+α). α can be calculated from Eq. (4.18). Although α is not a
constant but depends on many parameters, it is definitely a non-zero number.
80

According to the test data of Tang and Nilsson (1993), 1/ Cbm = 0.1849, and 1/(k’Cbm)
= 0.002438, and thus, k’ = 75.841 and Cbm = 5.4083 are used in the numerical model.

4.3.2 Chloride diffusion coefficient

The diffusion coefficient of chloride ions in concrete can be estimated using the

multifactor method as follows:

Dcl = f cl f 2 ( g i ) f 3 ( H ) f 4 (T ) f 5 (C f ) (4.22)

in which f1 is a factor that accounts for the effect of water-cement ratio and curing

time of concrete (t0) on chloride diffusion coefficient. An expression for f1 was

suggested (Xi and Bazant, 1999):

28 − t0 ⎛ 1 ( 28 − t0 ) ⎞⎛ w ⎞
6.55

f cl = +⎜ + ⎟⎜ ⎟ (4.23)
62500 ⎝ 4 300 ⎠⎝ c ⎠

The second factor, f2(gi) is to account for the effect of composite action of the

aggregates and the cement paste on the diffusion coefficient of concrete. This factor

can be formulated using the three phase composite model developed by Christensen

(1979):

⎛ gi ⎞
f1 ( g i ) = Dcp ⎜1 + ⎟ (4.24)
⎜ [1 − g ] / 3 + 1 /[( D / D ) − 1] ⎟
⎝ i agg cp ⎠

in which, gi is the volume fraction of aggregates in the concrete, Dagg and Dcp are the

diffusivities of aggregates and cement paste, respectively. Both Dagg and Dcp can be

calculated using the general model proposed by Martys (1994) for porous media:
81

D=
(
2 1 − (V p − V pc ) ) (V − V pc )
4.2
(4.25)
2 p
S

where, Vp is the porosity, S is the surface area and V pc is the critical porosity (the

porosity at which the pore space is first percolated). When Eq. (4.25) is used for the

diffusion coefficient of cement paste Dcp, Vp, S, and V pc in the equation should be the

parameters for cement paste. The critical porosity V pc may be taken as 3% for cement

paste (Martys 1994). The porosity of cement paste Vp can be approximately

estimated by using the water content at saturation, that is, the value of the adsorption

isotherm at H = 100%, ncp(H=1,T), which can be calculated from Eq. (A.1) and

Eqs.(A3) through (A.5) in Appendix A. One must keep in mind that the water

content such calculated is a weight fraction, which must then be converted into the

volume fraction, Vp = BSGcp*ncp(H=1,T), where BSGcp is the bulk specific gravity of

the cement paste (the estimated value for BSGcp ≈ 2.5). The surface area S of cement

paste is proportional to the monolayer capacity Vm and it can be estimated from Eq

(A.3) of Appendix A. The diffusion coefficient of aggregates Dagg can also be

calculated by Eq. (4.28) if the values for the porosity Vp, surface area S and critical

porosity V pc are known; or Dagg can be simply taken as a constant, and a suggested

value is 1x10-12 cm2/sec. The diffusion coefficient of aggregates Dagg can be

calculated using equation (4.25) and knowing the values for the porosity Vp, surface

area S and critical porosity V pc or simply may be taken as constant, a suggested value

is 1x10-12 cm/sec.
82

The third factor, f3(H) is to account for the effect of relative humidity on the

chloride diffusion coefficient. A model proposed by Bazant and Najjar (1972) can be

used, which was developed initially for moisture diffusion. In this work assuming a

full analogy between the moisture and chloride diffusion, the model is used:

−1
⎛ (1 − H ) 4 ⎞
f3 ( H ) = ⎜⎜1 + ⎟
4 ⎟
(4.26)
⎝ (1 − H C ) ⎠

in which Hc is the critical humidity level at which the diffusion coefficient drops
halfway between its maximum and minimum values (Hc = 0.75). Eq. (4.26) is very
important in terms of characterizing the coupling between the moisture diffusion and
the chloride penetration. Numerical examples will be shown later for detailed
analyses.

The fourth factor, f4(T) is to account for effect of temperature on the diffusion

coefficient of concrete. The temperature effect can be incorporated using Arrhenius’

law

⎡U ⎛ 1 1 ⎞ ⎤
f 4 (T ) = exp ⎢ ⎜⎜ − ⎟⎟ ⎥ (4.27)
⎣⎢ R ⎝ T0 T ⎠ ⎦⎥

in which U = the activation energy of the diffusion process, R = gas constant (8.314 J

mol-1K-1), T and T0 are the current and reference temperatures, respectively, in Kelvin

(T0 = 296 K). It was found that the activation energy of the diffusion process

depends on the porosity and cement type (Page et al., 1981). The values of the

activation energy, U, for a cement paste made of ordinary Portland cement are

41.8±4.0 (Kj/mol) for w/c of 0.4, 44.6±4.3 (Kj/mol) for w/c of 0.5 and
83

32.0±2.4 (Kj/mol) for w/c of 0.6.

The fifth factor, f5(Cf) is to account for the dependence of the chloride

diffusion coefficient on the free chloride concentration. For the diffusion of ions,

especially diffusion of chloride ions in the present study, the movement of ions is

restricted by the electrostatic field induced by the other ions present in the solution

f 5 (C f ) = 1 − k ion (C f ) m (4.28)

in which kion and m are two constants. kion and m were calibrated by Xi and Bazant

(1999) kion = 8.333 and m = 0.5.

4.4 Numerical Solution

The non-linearity of the material parameters involved and the coupling between

the chloride diffusion and the moisture diffusion in the governing equation, Eq. (4.6),

make the numerical solution favorable. In the present study the Alternating-Direction

Implicit (ADI) finite difference method developed by Peaceman and Rachford was

employed for solving the moisture and chloride diffusion problem (Carnahan et al.

1969). A two-dimensional finite difference mesh as shown in Fig. (4.1) was used in

this study. The free chloride concentration, Cf(i,j,n), at different points of the mesh at

different time intervals are estimated.

In general, the following partial differential equation was solved in two

consecutive steps.
84

df ( x, y, t )
dt
[( ) ]
= a( f , x, y, t ) ∇ D( f , x, y, t ) ∇ f ( x, y, t ) + µb( x, y, t ) f ( x, y, t ) (4.29)

Exposed Surface Chloride Solution


Boundary Condition

Initial Condition
Hini and Cfini 10cm

Width=30cm

Boundary Conditions
Henv and Cfenv

Fig. 4.1: Concrete slab used in the numerical solution.

As the first step, Eq. (4.29) is solved for the moisture diffusion problem, Eq.

(4.5), with a(f,x,y,t)=dH/dw, D(f,x,y,t) = DH, and b(x,y,t) =0. In the second step, Eq.

(4.29) is solved for the chloride penetration problem, Eq. (4.6), with f(x,y,t)=Cf(x,y,t),

a(x,y,t)=dCf/dCt,, D(f,x,y,t)=Dcl and b(x,y,t)=dw/dt.

For a two-dimensional finite difference mesh, see Fig. (4.1), Eq. (4.29), can be
rewritten as follows:

df ⎡ ∂ 2 f ∂D ∂f ∂ 2 f ∂D ∂f ⎤
= a ∗ ⎢D ∗ 2 + ∗ + D∗ 2 + ∗ + µb ∗ f ⎥ (4.30)
dt ⎣ ∂x ∂x ∂x ∂y ∂y ∂y ⎦

The principle of Alternating-Direction Implicit (ADI) Method can be applied to solve

Eq. (4.30) by using two difference equations over two successive time-steps each of

duration ∆t/2:
85

fi*,j − fi ,nj
∆t / 2
[
= ain, j * Din, jδ x2 fi*,j + δ x Din, jδ x fi*,j + Din, jδ y2 fi ,nj + δ y Din, jδ y fi ,nj + bin, j fi ,nj ] (4.31)

fi ,nj+1 − fi*,j
∆t / 2
[
= ain, j * Din, jδ x2 fi *,j + δ x Din, jδ x fi *,j + Din, jδ y2 fi ,nj+1 + δ y Din, jδ y fi ,nj+1 + bin, j fi *,j ] (4.32)

in which, δx, δ2x and δy, δ2y are the first and second order central-difference operators

in x and y directions, respectively. Eq. (4.31) is implicit only in x-direction and used

to solve for f i *,j (intermediate value at the end of the first time-step) by knowing

f i ,nj starting from the initial condition. Eq. (4.32) is implicit only in y-direction and

used to solve for f i ,nj+1 (final value at the end of the whole time step) after calculating

f i *,j in the first step.

4.5 Numerical Applications and Discussions

4.5.1 Numerical example of chloride penetration in non-saturated concrete

The numerical solution algorithm described in the previous section solving for

the governing partial differential equation is implemented in a computer program,

which is used to solve for the diffusion of chloride ions in partially saturated concrete

slabs. As a numerical example, a 10 cm depth concrete slab similar to the slab shown

in Fig. (4.1) was exposed to 5% chloride solution (8.24% sodium chloride solution)

and 100% relative humidity on the upper. The water-cement ratio of the concrete is

0.6 and the curing period is 28 days. The initial free chloride concentration in the

concrete slab Cfini = 0 and the initial relative humidity in the concrete Hini = 50%.
86

The variation of free chloride concentration with the depth at different times

of exposure is shown in Fig. (4.2). The depth of penetration is measured from the

exposed surface of the concrete slab. Fig. (4.2), shows that the free chloride

concentration decreases with the increases of the depth from the exposed surface.

The variation of free chloride concentration with the time of exposure at different

depths is shown in Fig. (4.3). The time of exposure starts right after the curing period.

Fig. (4.3) shows that at a fixed depth inside of the concrete slab, the free chloride

concentration increases with the increase of the time of exposure.

-3
x 10
7

Time = 1 month
6
= 3 months
Free chloride concentration, C f (g/g)

= 6 months
= 1 year
5
= 2 years
= 3 years
4

0
0 1 2 3 4 5 6 7 8 9 10
Depth from exposed surface (cm)

Fig. 4.2: Depth profile of free chloride concentration.


87

-3
x 10
7
Depth = 1 cm
= 2 cm
6
= 3 cm

Free chloride concentration, C f , (g/g)


= 4 cm
5 = 5 cm

0
0 200 400 600 800 1000 1200
Exposure time (days)

Fig. 4.3: Time profile of free chloride concentration.

In practice, penetration front of free chloride has often been used, which is

defined as the depth in the concrete where the free chloride concentration equals to a

predetermined value. The location of the penetration front can be plotted against the

time of exposure. In this way, the movement of the penetration front can be used as

an indicator (similar to a tracer), that follows the movement of the chloride from the

exposed surface to inner part of the concrete. Either Fig. (4.2) or Fig. (4.3) can be

used to predict the location of penetration front for a fixed free chloride

concentration. As an example, a penetration front of 0.06% (g/g) free chloride

concentration is reached approximately to the depth of 2 cm after one month of

exposure; 2.9 cm after three months, 3.7 cm after six months, 4.8 cm after 1 year, 6.4
88

cm after 2 years, and 8 cm after 3 years of exposure. If a reinforcing steel bar is

embedded at the depth of 4 cm, and 0.06% free chloride concentration is the critical

chloride concentration that triggers the onset of steel corrosion, then the relationship

(among the penetration front, time, and depth) can be used to predict the starting point

of the steel corrosion as well as the development of the corrosion process. Therefore,

it provides very useful information about the service life of a concrete structure and

the time needed for rehabilitation.

4.5.2 Comparison with test data

The experimental results obtained from the 90-day ponding test (Andrade and

Whiting 1996) for a concrete slab similar to the concrete slab shown in Fig. (4.1), is

compared with the numerical results obtained in the present study, see Fig. (4.4).

Two types of concrete with different water-cement ratios of 0.4 and 0.6 are compared.

One can see from Fig. (4.4) that the predicted results by the present model agree very

well with the test data for the chloride penetration into partially saturated concrete.

4.5.3 Effect of the moisture diffusion on chloride penetration

The effect of moisture diffusion becomes very important in partially saturated

concrete. To study the effect of the moisture diffusion on the diffusion of chloride

ions, a concrete slab of 10 cm thick, as shown in Fig. (4.1), was exposed to 5%

chloride solution (8.24% sodium chloride solution) on the top surface. The chloride

diffusion in the concrete slab is analyzed for two different moisture conditions. In

one of the cases, the moisture boundary condition H = 100% on the top surface, the
89

0.007

0.006 w/c = 0.4 (Experimental)

t (g/g)
w/c = 0.6
Total chloride concentration, C 0.005 w/c = 0.4 (Present)
w/c = 0.6

0.004

0.003

0.002

0.001

0
0 1 2 3 4 5
Depth from exposed surface (cm)

Fig. 4.4: Comparison with the experimental results.

initial pore relative humidity inside the slab is 50%, and therefore, the moisture

diffuses from the top of the slab into the concrete, the same direction as the chloride

diffusion. In the other case, the concrete is fully saturated, and thus, there is only

chloride diffusion due to the concentration gradient, and there is no moisture

diffusion occurs. The plots of depth versus free chloride concentration after 1 month,

6 months and 6 years of exposure for the two cases are shown in Fig. (4.5). The plots

of depth versus pore relative humidity after 1 month, 6 months and 6 years of

exposure for the two cases are shown in Fig. (4.6). One can see from Fig. (4.5) that, at

any fixed depth, the free chloride concentration is greater in the partially saturated

concrete (with 50% initial humidity condition). This is due to the coupled moisture

diffusion and chloride penetration, which is shown in the second term of the right
90

hand side of Eq. (4.6). So, the coupling effect contributes significantly to the chloride

penetration in the partially saturated concrete. The basic trend of the coupling effect

predicted by the present numerical model agrees very well with the test data of

chapter 7. Fig. (4.6) also shows that the contribution of the moisture diffusion to the

chloride penetration becomes less significant as the time of exposure increases. This

is due to that the fact that, the concrete becomes fully saturated after long-term

exposure and so there is no moisture exchange inside the concrete, see Fig. (4.6).

One can see from Fig. (4.6) that, after the long-term exposure of six year, the partially

saturated concrete becomes saturated, and the free chloride concentrations are the

same for both cases, see Fig. (4.5).

-3
x 10
7

P artially saturated, H in i = 50%


6
Fully saturated, H = 100%
Free chloride concentration, C f (g/g)

4
6 years

3
1 month

2
6 months

0
0 1 2 3 4 5 6 7 8 9 10
Depth from exposed surface (cm)

Fig. 4.5: Effect of the moisture diffusion on free chloride penetration.


91

0.95

0.9

0.85
P ore relative humidity, H

0.8

0.75

0.7

0.65

0.6 Time = 1 month


= 6 months
0.55 = 6 years

0.5
0 1 2 3 4 5 6 7 8 9 10
Depth from exposed surface (cm)

Fig. 4.6: Moisture diffusion in non-saturated concrete.

4.5.4 The Effect of concrete mix parameters

The material models developed for chloride and humidity diffusion

coefficients can handle the effect of concrete mix parameters, such as water-cement

ratio and volume fraction of aggregate. Fig. (4.7) shows the profiles of the free

chloride concentration after 1 year of exposure for a concrete slab made with 70%

aggregate volume fraction and three different water-cement ratios, 0.5, 0.55 and 0.6.

The concrete slab is 10 cm thick and exposed to 5% chloride solution from the top

surface as shown in Fig. (4.1). Fig. (4.7) shows that the higher the water-cement ratio

the higher the free chloride concentration (at a fixed depth and fixed time). This is

due to the fact that both chloride and humidity diffusion coefficients are increased as

water-cement ratio increases.


92

-3
x 10
7

6
w/c = 0.50
Free chloride concentration, C f (g/g)

= 0.55
= 0.60
5

0
0 1 2 3 4 5 6 7
Depth from exposed surface (cm)

Fig. 4.7: Effect of water-cement ratio on chloride penetration.

Figs. (4.8) and (4.9) show the chloride and the humidity diffusion coefficients

as functions of water-cement ratio, the higher the water-cement ratio, the higher the

diffusion coefficients. This is because the porosity of cement paste is higher with

higher water-cement ratios, and the connectivity of the pore system is also increased

which results in a reduction in the resistance to the diffusions of both moisture and

chloride ions.
93

-1 0
x 10
1

0.9 w/c = 0.50

Chloride diffusion coefficent, D cl (m /sec)


= 0.55
0.8 = 0.60

2
0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
P ore relative humidity, H

Fig. 4.8: Effect of water cement ratio on the chloride diffusion coefficient.

-1 0
x 10
1.2

w/c = 0.4
Humidity diffusion coefficent, D H (m /sec)

1 = 0.5
= 0.6
2

0.8

0.6

0.4

0.2

0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
P ore relative humidity, H

Fig. 4.9: Effect of water-cement ratio on humidity diffusion coefficient.


94

The effect of the volume fraction of aggregates on the free chloride

concentration is shown in Fig. (4.10). The same concrete slab in the previous example

was used, but the water-cement ratio is fixed as 0.6 and the volume fraction of

aggregate varies, 0.5, 0.6 and 0.7. The effects of the volume fraction of aggregate on

the chloride and humidity diffusion coefficients are shown in Figs. (4.11) and (4.12),

in which both chloride and humidity diffusion coefficients decrease with the increase

of aggregate volume fraction. This is resulted from the low diffusivity of aggregate.

0.01

0.009 gi = 0.50
= 0.60
0.008
Free chloride concentration, C f (g/g)

= 0.70

0.007

0.006

0.005

0.004

0.003

0.002

0.001

0
0 1 2 3 4 5 6 7 8 9 10
Depth from exposed surface (cm)

Fig. 4.10: Effect of aggregate volume fraction on the chloride penetration.


95

-1 1
x 10
5

4.5 gi = 0.50
= 0.60

Chloride diffusion coefficent, D cl (m /sec)


4 = 0.70

2
3.5

2.5

1.5

0.5

0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
P ore relative humidity, H

Fig. 4.11: Effect of aggregate volume fraction on the chloride diffusion coefficient.

-1 0
x 10
2.4

2.1
Humidity diffusion coefficent, D H (m /sec)

gi = 0.5
2

1.8 = 0.6
= 0.7
1.5

1.2

0.9

0.6

0.3

0
0.5 0.55 0.6 0.65 0.7 0.75 0.8 0.85 0.9 0.95 1
P ore relative humidity, H

Fig. 4.12: Effect of aggregate volume fraction on humidity diffusion coefficient.


96

4.6 Conclusions

1. The multiscale modeling of material parameters involved in the coupled moisture

and chloride diffusion equations results in a robust mathematical model for

characterizing chloride ion ingress into non-saturated concrete. There are four

materials parameters in the model. Moisture diffusion coefficient and moisture

capacity are involved in the moisture diffusion equation, and chloride diffusion

coefficient and binding capacity are involved in the chloride penetration equation.

Four material models are developed for the four material parameters based on

different diffusion mechanisms and their dependence on the environmental

conditions and concrete properties. Each material model is calibrated based on

related test data, which ensure that each model parameter has distinct physical

meaning, not just a results from curve fitting.

2. The model simulates satisfactorily the chloride profiles obtained by the 90-day

ponding test (AASHTO T259–90). The model predictions agree with available

test data very well.

3. The model predicts very well the effect of moisture diffusion on the chloride ion

penetration in partially saturated concrete. The numerical results show that the

moisture diffusion accelerates the chloride penetration.

4. The material models incorporate the effects of various concrete mix parameters

(such as volume fraction of aggregate and w/c ratio) on the diffusion of chloride

ions into partially saturated concrete.

5. The model can be used to predict the onset of steel corrosion in reinforced

concrete structure and thus the service life and the time for rehabilitation.
97

6. The Alternating-Direction Implicit (ADI) finite difference method is used as an

effective tool for solving the coupled partial differential equations. A two-

dimensional finite difference program is developed.


CHAPTER 5

THE COUPLED EFFECT OF FREEZING-THAWING AND

CHLORIDE PENETRATION IN CONCRETE

5.1 Introduction

One of the major effects of freeze-thaw cycles on the deterioration of concrete

is the volume expansion associated with ice formation in the pores of concrete

subjected to freezing and thawing cycles (Philleo 1986). The ice formed in the pores

reduces the effective diffusion coefficient of the concrete, however, excessive volume

expansion due to ice formation results in progressive damage and cracking in

concrete, and the cracks accelerate the diffusion of moisture and chloride ions in

concrete.

The effect of freezing and thawing on concrete durability has been studied

theoretically and experimentally. Mohamed et al. (2000) conducted an experimental

study on the effect of cement type and air-entrainment on freezing-thawing durability

of concrete. Also, they studied the effect of damage caused by freezing and thawing

cycles through the degradation in the compressive strength of concrete. In their study,

they used the benchmark test setup prepared by the Portland Cement Association in

1940. In this test 28 different types of cement were used in a parapet wall panels as

part of Green Mountain Dam (Colorado, USA).


99

Zuber et al. (2000) conducted a theoretical and experimental study on the ice

formation mechanisms in normal and high performance concrete. Theoretically, they

modeled the ice formation mechanisms in the pores of concrete based on the

thermodynamic equilibrium (Gibbs-Duhem equation) between the solid and liquid

phases. Experimentally, the low-temperature calorimetry technique used to

investigate the ice formation in concrete mixes with different water-to-cement ratios.

The test results showed that reducing the water-to-cement ratio resulted in the

refinement of the pore structure and thus reduce the ice formation. Both numerical

modeling and experiments showed that ice formation mechanisms were controlled by

ice front propagation. Also, they concluded that it is difficult to totally impede the ice

formation by enhancing the performance of concrete.

Scherer (1999) studied the mechanism of crystallization of ice and the stresses

development on the walls of the pore network of concrete. Those stresses are

controlled by the pore size distribution, the interfacial energy between the pore wall

and the growing crystal and the yield stress or buckling strength of the crystal. The

crystallization in the pores of concrete was studied based on the principles of

thermodynamics. One of the main conclusions was that cracking is not due to

crystallization in a single pore rather it is due to crystallization in a region large

enough compared to a size of strength-controlling flaws.

Bazant et al. (1988) developed a complete mathematical model for freezing-

thawing durability of concrete. The model included the formulation of sorption

isotherm for concrete below 0 oC based on the well known sorption isotherm for
100

concrete above 0 oC; the formulation of the water diffusion equation in concrete in the

presence of air-entrained bubbles; the calculation of temperature variations in

concrete taking into consideration the latent heat of freezing, and prediction of the

stress produced in concrete due to freezing of pore water.

Saetta et al. (1998) studied the coupled effect of moisture, heat and pollutant

(e.g., Cl-, (SO4)-2 and CO2) flows through reinforced concrete structures. They

introduced two damage parameters. One is the mechanical damage parameter and it is

related to the degradation of the material due to mechanical actions such as

deformations, applied loading, fatigue, drying shrinkage and thermal expansion. And

the other is the chemical damage parameter and it is related to the degradation of the

mechanical strength of the material due to the chemical reaction between the

attacking chemical and the cementitious material, and the reinforcement corrosion.

In this chapter, an experimental study on the coupled effect of freezing–thawing

and chloride penetration on concrete was conducted. Two sets of concrete mixes were

used in this study. In one of the sets, the effect of water-cement ratio was studied,

while the effect of aggregate volume fraction was studied in the other set. The

experimental study resulted in the damage values and chloride profiles for especially

designed concrete specimen exposed to simultaneous action of cyclic freezing-

thawing and continuous chloride penetration. The obtained experimental results can

be used for validation and calibration of numerical results obtained by the

mathematical models.
101

5.2 Ultrasonic Testing of Materials

The experimental program in this chapter involved the assessment of damage in

concrete under the coupled action of freezing-thawing and chloride penetration. The

damage was evaluated using two non-destructive testing techniques, the pulse

velocity and the forced resonance methods. In this section, a brief review about the

ultrasonic testing of materials will be provided.

Ultrasonic testing methods are widely used in the quality assessment of

various materials used in civil engineering structures, including metallic, plastics,

rubber, wood, concrete and cement-based materials. The ultrasonic methods have

many advantages over the traditional testing methods. In addition to their non-

destructive nature, ultrasonic methods can be applied to in-service inspections and for

early detection of internal flaws and defects inside the material, which can be used as

an indicator for routine maintenance, retrofitting or rehabilitation of the structures.

Some characteristics of the ultrasonic pulse waves depend on the density,

elastic properties and the existence of air voids and cracks inside the material. Those

characteristics include the pulse velocity, resonant frequency, pulse amplitude and

frequency content. One of the main drawbacks on the pulse velocity method is the

small variation in the pulse velocity with the variation of concrete material

parameters, such as water-cement ratio and aggregate volume fraction, which may

lead to incorrect prediction of the concrete strength.

In this study, two indicators were estimated for concrete damage resulted from
102

the degradation of the initial dynamic modulus of elasticity due to cyclic freezing and

thawing. One was calculated based on the velocity of the compressive pulse wave

traveling through the concrete sample. The other was calculated based on the

fundamental resonant frequency of longitudinal wave. The damage factor based on

the degradation of the initial dynamic modulus of elasticity can be written as follows:

E ddam
d = 1− (5.1)
E do

where, E do = initial or undamaged dynamic elastic modulus and E ddam = dynamic elastic

modulus of the damaged concrete.

The dynamic modulus of elasticity of concrete can be calculated in terms of the

transmitted ultrasonic pulse velocity, V of compression wave as follows:

(1 + µ )(1 − 2 µ )
E d = ρV 2 (5.2)
(1 − µ )

where, ρ is the material density and µ is the dynamic Poisson’s ratio.

On the other hand, the fundamental resonant frequency (fL in Hz) of longitudinal wave

traveling through a concrete sample of prismatic cross-sectional area can be used to

estimate the dynamic modulus of elasticity (in Pa) as follows:

E d = 4 ρLf L
2
(5.3)

where, ρ is the material density in (Kg/m3) and L is the length of specimen in (m).
103

By substituting Eq. (5.2) in Eq. (5.1) and neglecting the effect of damage on the

Poisson’s ratio (µ) and material density (ρ), the damage factor based on the ultrasonic

pulse velocity can be written as:

2
⎛ V dam ⎞
d = 1− ⎜ o ⎟ (5.4)
⎜ V ⎟
⎝ ⎠

Since, the V-meter (Pulse Velocity Tester) measures the transit time (t in

microseconds), and the pulse velocity, V can be written in terms of transit time (t) and

the length of the sample (L) as follows:

L
V = (5.5)
t

The damage factor can be written in terms of the transit time by substituting Eq. (5.5)

into Eq. (5.4) as follows:

2
⎛ to ⎞
d = 1 − ⎜ dam ⎟ (5.6)
⎜t ⎟
⎝ ⎠

where, to is the initial transit time and tdam is the transit time of the pulse wave

measured after damage.

The damage factor based on the fundamental resonant frequency of longitudinal pulse

wave (fL in Hz) by substituting Eq. (5.3) in Eq. (5.1) as follows:


104

2
⎛ f dam ⎞
d = 1 − ⎜⎜ L 0 ⎟⎟ (5.7)
⎝ fL ⎠

5.3 Experimental Program


5.3.1 Preparation of test specimens

In order to study the effect of freezing and thawing on the chloride penetration,

a specially designed concrete specimen was used for this purpose. As shown in Fig.

(5.1), the concrete specimen is 40x13x13 cm in dimension. The dimensions of the

specimen were modified from the dimensions of the specimen used in the previous

experiment (30x15x15 cm). The reason for the modification was to maintain the

length to maximum width ratio between 3 and 5. Specimens having the length to

maximum width ratio outside this range are difficult to excite for the fundamental

mode of vibration.

Fig. 5.1: Specimen for coupled freezing-thawing and chloride penetration.

Two sets of specimens were used in this experiment. One set has an aggregate
105

volume fraction of 0.65 and three different water-cement ratios of 0.45, 0.55 and

0.65. The other set has a water-cement ratio of 0.55, and three different aggregate

volume fractions of 0.55, 0.65 and 0.75. The same concrete mixes used in the

experimental study of chapter 6 (see section 6.2) were used in the current study. The

mix proportions and the compression test results are shown in Table (C.2) of

Appendix C and Table (D.1) of Appendix D. The specimens were casted, remolded

after 24 hours and stored in the moisture room for 28 days. No air-entraining agent

was used in the concrete mixes.

5.3.2 Cyclic freezing-thawing and chloride penetration

After curing, all specimens were exposed to 8% sodium chloride (NaCl)

solution on the top surface and placed in an environmental chamber with

programmable freezing and thawing cycles as shown in Fig. (5.2).

Fig. 5.2: Placing of concrete samples in the environmental chamber.


106

25

15

Temperature ( C )
o

-5 0 1 2 3 4 5 6

-15

-25
Time (hour)

Fig. 5.3: The freezing and thawing temperature cycle.

5.3.3 Measurement of Transit time and fundamental resonant frequency

The setup for the V-meter (Pulse Velocity Tester) and the concrete sample is

shown in Fig. (5.4). This setup was used to measure the time required for the earliest

part of longitudinal (compression wave) pulse traveling through the concrete sample

from transmitting transducer to receiving transducer. This time is called the transit

time and it is measured in microseconds.

The other method was by measuring the fundamental resonant frequency (in

Hz) of a longitudinal wave traveling through concrete sent by a vibrator on one side

of the specimen and received by the pick-up on the other side. The E-meter (Resonant

Frequency Tester) and the concrete sample setup are shown in Fig. (5.5).
107

Fig. 5.4: Transit time of measurement setup (V-meter).

Fig. 5.5: Setup for measurement of resonant frequency (E-meter)

5.3.4 Chemical analysis for total chloride concentration

After exposed to freezing-thawing and chloride penetration, the total chloride

profiles for concrete specimens were obtained following the experimental procedure

described in Appendix E. The m-Volt readings resulted from the electrochemical

analysis are shown in Table (E.1) of Appendix E. The mV readings of the known

chloride concentration solutions (0.005%, 0.02%, 0.05% and 0.5%) were taken and
108

shown in Table (E.2). The calibration curve for standard concentration solutions is

shown in Fig. (E.4).

5.4 Test Results and Discussion

5.4.1 Cracking patterns and visual inspection

The cracking patterns for the concrete samples after exposed to 400 temperature

cycles of freezing and thawing are shown in Figs. (5.6) through (5.10). Fig. (5.6)

shows the cracking pattern for the top, bottom and the two sides of the concrete

specimen made of mix 46 (0.45 water-cement ratio and 0.65 aggregate volume

fraction). From visual inspection, large number of cracks was noticed but the width

was very small (0.4 mm). There were no cracks on the top surface of the sample

(underneath the chloride solution).

The cracking patterns of the sample made of the concrete mix 56 (have a 0.55

water-cement ratio and 0.65 aggregate volume fraction) are shown in Fig. (5.7). Few

cracks were noticed on the surfaces of the specimen, but their width was larger when

compared to the previous sample (made of mix 46). The average width of the crack

was 1.2 mm and there were no cracks on the top surface of the sample (surface

exposed to chloride solution).


109

(a) (b)

(c) (d)
Fig. 5.6: The concrete sample (46) after it was exposed to the freezing-thawing, a) top
of the sample b) bottom of the sample and c and d) the two sides of the sample.

(a) (b)

(c) (d)
Fig. 5.7: The concrete sample (56) after it was exposed to the freezing-thawing, a) top
of the sample b) bottom of the sample and c and d) the two sides of the sample.

Deep cracks were noticed in the concrete sample prepared from mix 66

(water-cement ratio is 0.65 and aggregate volume fraction is 0.65) as shown in Fig.

(5.8). There were a lot of cracks on the top surface of the sample (surface exposed to
110

chloride solution). The average width of the cracks was 2 mm.

The concrete sample shown in Fig. (5.9) was made of mix 55 in which the

water-cement ratio is 0.55 and the aggregate volume fraction is 0.55. From the

observation of the crack patterns, it was noticed that the cracks are deep and had an

average width of 1.4 mm. Also, large number of cracks was observed on the top

surface of the concrete sample.

As shown in Fig. (5.10), few but deeper and wider cracks were noticed on the

surfaces of concrete sample made of mix 57 (aggregate volume fraction=0.75 and

water-cement ratio=0.55). The average width of the cracks was a 2.6 mm. Few

number of cracks were noticed on the top surface underneath the chloride solution.

(a) (b)

(c) (d)
Fig. 5.8: The concrete sample (66) after it was exposed to the freezing-thawing, a) top
of the sample b) bottom of the sample and c and d) the two sides of the sample.
111

(a) (b)

(c) (d)
Fig. 5.9: The concrete sample (55) after it was exposed to the freezing-thawing, a) top
of the sample b) bottom of the sample and c and d) the two sides of the sample.

(a) (b)

(c) (d)
Fig. 5.10: The concrete sample (57) after it was exposed to the freezing-thawing, a)
top of the sample b) bottom of the sample and c and d) the two sides of the sample.

5.4.2 Ultrasonic test results

The measured transit times of the compression pulse wave and fundamental
112

resonant frequencies of longitudinal wave before and after the samples were exposed

to the freezing and thawing cycles are presented in Table (5.1). The damage factor

can be calculated either by Eq. (5.6) using the measured transit times or by Eq. (5.7)

using the measured fundamental resonant frequencies. It is clear that 400 cycles of

freezing and thawing coupled with chloride penetration resulted in a significant

damage in the concrete, and the damage evaluated by the two methods showed

consistent results. From the comparison of damage factors, as shown in Table (5.1),

the concrete samples prepared from mixes 66, 55 and 57 have the most severe

damage among the five samples. Also, during freezing and thawing of the specimens

while monitoring the test and adding chloride solution, it was noticed the cracking

started earlier and propagated faster in those three samples (samples 66, 55 and 57).

In mix 66, the water-cement ratio was 0.65 and it was the highest among the

three mixes 46, 56 and 66 (see Table C.2 in Appendix C). Increasing water-cement

ratio will introduce more capillary pore and coarsen the pore structure of concrete

which increase the amount of ice formed and accelerate the propagation of ice front.

This effect when combined with the decrease in the strength of concrete resulted in

larger damage as shown by the test results.

The effect of aggregate can be evaluated using the present test data. In general,

aggregate has lower permeability than cement paste, therefore, more aggregate in

concrete reduces overall permeability of the concrete. On the other hand, aggregate

has much higher freezing-thawing resistance than cement paste, therefore, more

aggregate in concrete enhances freezing-thawing resistance of the concrete. With


113

these two general criteria in mind, we can see from the present test data that

decreasing the aggregate volume fraction (mix 55 compared to mix 56) resulted in

high permeability; and increase in aggregate volume fraction (mix 57 compared by

mix 56) resulted in a decrease in the permeability.

There is another effect caused by the aggregate that may play a role in the

permeability of concrete. With more aggregate added in the concrete mix, there is a

significant increase in interface transition zone (ITZ) between the cement paste and

aggregate. ITZ has more porosity than the bulk cement paste. So, the increased

porosity (leading to high permeability) may counterbalance the effect of low

permeability of the aggregate.

Table 5.1: Ultrasonic test results.

V-meter E-meter
Mix Transit time (microsec.) Resonant frequency (Hz)
damage damage
before after (%) before after (%)

46 82.4 144.5 67 5116 2537 75


56 86.6 118.3 46 4831 2636 70
66 86.3 214.8 84 4782 983 96
55 84.7 216.8 85 4758 451 99
57 81.7 188.6 81 5026 469 99

5.4.3 Chloride penetration test results

The calibration curve shown in Fig. (E.4) in Appendix E was used to convert

the mVolt readings to total chloride concentration as shown in Table (5.2). The table

presents the values of the total chloride concentration (as a percentage of concrete
114

weight) at different depth ranges (from the exposed surface) for concrete samples

made of different concrete mixes exposed to 400 cycles of freezing and thawing.

Each cycle continues for 6 hours and so the total exposure time is 100 days.

Table 5.2: Total chloride concentration as percentage of weight of concrete.

Depth range (mm)


Mix Sample #
0-6 6-13 13-19 19-25 25-32 32-38 38-45
1 2.013 0.533 0.388 0.519 0.444 0.386 0.275
2 1.732 0.525 0.375 0.499 0.422 0.384 0.275
46 3 1.712 0.522 0.361 0.493 0.420 0.380 0.254
average 1.819 0.527 0.375 0.504 0.428 0.383 0.268
1 2.024 1.165 0.834 0.652 0.542 0.505 0.493
2 1.904 1.127 0.824 0.634 0.531 0.485 0.483
56 3 1.904 1.078 0.820 0.627 0.525 0.480 0.472
average 1.944 1.123 0.826 0.638 0.533 0.490 0.483
1 2.141 0.948 0.705 0.539 - - -
2 2.082 0.896 0.701 0.488 - - -
66 3 1.958 0.862 0.645 0.472 - - -
average 2.060 0.902 0.684 0.500 - - -
1 2.201 1.036 0.857 0.758 0.701 0.536 -
2 2.165 1.013 0.793 0.733 0.656 0.525 -
55 3 2.070 1.002 0.775 0.690 0.610 0.522 -
average 2.145 1.017 0.808 0.727 0.656 0.528 -
1 1.893 0.754 0.390 - - - -
2 1.821 0.610 0.386 - - - -
57 3 1.751 0.593 0.375 - - - -
average 1.822 0.652 0.384 - - - -

The chloride profiles obtained in this study are compared with the profiles

obtained in the experimental study of chapter 6 which was conducted at room

temperature and for the same exposure period (100 days). The comparisons are
115

shown in Figs. (5.10) through (5.20). One can see that the chloride ions in the

concrete experienced freeze-thaw cycles penetrate in a higher rate, which means that

the progressive damage produced by the cyclic freezing and thawing increases the

chloride penetration. The chloride profiles for damaged (due to cyclic freezing-

thawing) and undamaged specimens made of mixes 46, 56 and 55 shown in Figs.

(5.11) through (5.14) present that the concentrations in the damaged concrete is

approximately doubled in the depth range of 0-19 mm and the concentrations are

much higher in the depth range from 19-44 mm. Fig. (5.15), shows that even though

specimen 57 experienced severe damage, the chloride concentration did not increase

significantly. This is due to the high content of aggregate, which decreases diffusivity

and thus impedes chloride penetration.


116

2.00

Chloride conc. (% by concrete weight)


1.75 At room temperature

1.50 After freezing and thawing

1.25

1.00

0.75

0.50

0.25

0.00
0-6 6-13 13-19 19-25 25-32 32-38 38-45
Depth range (mm)
Fig. 5.11: Chloride profiles for the concrete specimens made of mix 46

2.00
Chloride conc. (% by concrete weight)

1.75 At room temperature

1.50 After freezing and thawing

1.25

1.00

0.75

0.50

0.25

0.00
0-6 6-13 13-19 19-25 25-32 32-38 38-45
Depth range (mm)
Fig. 5.12: Chloride profiles for the concrete specimens made of mix 56
117

2.25

Chloride conc. (% by concrete weight)


2.00 At room temperature

1.75 After freezing and thawing


1.50

1.25

1.00

0.75

0.50

0.25

0.00
0-6 6-13 13-19 19-25
Depth range (mm)
Fig. 5.13: Chloride profiles for the concrete specimens made of mix 66

2.25
Chloride conc. (% by concrete weight)

2.00 At room temperature

1.75 After freezing and thawing


1.50

1.25

1.00

0.75

0.50

0.25

0.00
0-6 6-13 13-19 19-25 25-32 32-38
Depth range (mm)
Fig. 5.14: Chloride profiles for the concrete specimens made of mix 55
118

2.00

Chloride conc. (% by concrete weight)


1.75 At room temperature

1.50 After freezing and thawing

1.25

1.00

0.75

0.50

0.25

0.00
0-6 6-13 13-19
Depth range (mm)
Fig. 5.15: Chloride profiles for the concrete specimens made of mix 57

5.5 Conclusions

The experimental results obtained by the ultrasonic testing, chloride profiles

and visual analysis of cracking are summarized in Table (5.3).

Table 5.3: Summary of the experimental results.

*
Damage (%) C dam
t /C undam
t
crack
Mix based on at depth range
pulse resonant 0-19 19-44 width
depth density connectivity
velocity frequency mm mm (mm)
46 67 75 1.5 11 0.4 small high high
56 46 70 2.2 7 1.2 small low low
66 84 96 2.1 - 2 large low low
55 85 99 1.6 19 1.4 large high high
57 81 99 1.3 - 2.6 large low high

* Ratio of total chloride concentrations obtained from damaged and undamaged


concrete specimens.
119

The following can be concluded from the analysis of the experimental results:

1. The exposure to cyclic freezing and thawing resulted in a substantial internal

damage and cracking of concrete samples. The progressive damage resulted in

new cracks, widening and extending of existing micocracks and interconnecting

of the crack network. The chloride penetration is not only related to the level of

damage and crack width, but also to the crack density and connectivity.

2. The damage factors of concrete was measured by the pulse velocity (V-meter) and

the resonant frequency (E-meter). After 400 temperature cycles, the damage

factors obtained by the two methods are consistent.

3. The damage and cracking increase the permeability and accelerate the chloride

penetration into concrete.

4. The damage and cracking resulted in a higher chloride penetration at larger depth

from exposed surface. This effect becomes very important in bridge decks, where

chloride penetration towards the reinforcing bars would increase the concentration

and thus exceeds the threshold value causing corrosion of steel bars.

5. The concrete material parameters such as aggregate volume fraction and water-

cement ratio have significant effect on the resistance to freezing-thawing damage

and chloride penetration and thus durability of concrete structures.


CHAPTER 6

EXPERIMENTAL STUDY ON THE EFFECT OF CONCRETE MATERIAL


PARAMETERS ON CHLORIDE PENETRATION

6.1 Introduction

The resistance of concrete to chloride penetration is determined by its internal

structure, which can characterized by the morphology of the pore structure, including

the total porosity, pore size distribution, spatial distribution of pores, shape,

connectivity and tortuosity of the pores network. These characteristics are controlled

by many factors including concrete material parameters, construction practices,

environmental conditions and the existence of cracking.

In this chapter, an experimental study on the effect of the concrete material

parameters; the water-cement ratio and the aggregate volume fraction, on the chloride

penetration was conducted.

The effect of water-cement ratio on chloride penetration in concrete was studied

by many researchers (Swaddiwudhipong et al., 2000; Suryavanshi et al., 1998;

Mangat and Molloy, 1995; Saetta et all., 1993; Collepardi et. al., 1972; Arya et al.

1990; Midgley and Illston, 1986). All studies show that increasing water-cement ratio

increases the capillary porosity of concrete and thus accelerates moisture diffusion

and chloride penetration. Swaddiwudhipong et al. (2000) found that lowering water-
121

cement ratio resulted in a microstructural densification of the material, which lowers

the moisture and chloride diffusion coefficients and improving the resistance to

chloride ingress in concrete. Suryavanshi et al. (1998) found that concrete mixes with

high water-cement ratio (0.75) have lower resistance to chloride penetration. While

concrete with lower water-cement ratio (0.45 and 0.60) have higher resistance to

chloride penetration. Mangat and Molloy (1995) reported an increase in the amount

of bound chloride with the increase of water-cement ratio.

The volume fraction of aggregate affects the chloride diffusion properties by

adding a low diffusion phase (aggregate) and also makes the flow paths longer and

tortuous (Bentz et al., 1998; Bentz, 2000). Thus, increasing volume fraction of

aggregate in concrete mixes decreases moisture diffusion and chloride penetration.

6.2 Specimen Preparation and Experimental Parameters

In this chapter, a 30x15x15 cm concrete slabs as shown in Fig. (6.1) were

prepared. Two sets of specimens were prepared. One set had the same aggregate

volume fraction (0.65), three different water-cement ratios (0.45, 0.55 and 0.65) and

identified as mixes 46, 56 and 66, respectively. The other set had three different

aggregate volume fractions (0.55, 0.65 and 0.75), the same water-cement ratio (0.55)

and identified as mixes 55, 56 and 57, respectively. Basically, the first digit of the

identification number represents water-cement ratio, and the second aggregate

volume fraction.

Since mix 56 (with 0.55 water-cement ratio and 0.65 aggregate volume
122

fraction) is a common mix in the two sets, five different concrete mixes were

prepared. The proportions of the five mixes are shown in Table (C.1) of Appendix C.

In order to keep a constant volume fraction of aggregate for mixes 46, 56 and 66, the

increase in the weight of mixing water was compensated by a decrease in the weight

of cement. In mixes 55, 56 and 57, the increase in volume fraction of aggregate was

compensated by a decrease in the weights of mixing water and cement, in order to

keep the water-cement ratio constant. As a result, the effect of another material

parameter, i.e. cement content was included in this experimental study. Three

specimens of each mix were casted, demolded after 24 hours and stored in a moist

room for 28 days. The total number of concrete specimens was 15.

Insulation
(four sides) 15c
m
8% NaCl
solution

15cm

Insulation
(two sides)
3 0c m

Fig. 6.1: Concrete sample

The compression test results for the five concrete mixes is shown in Table (D.1)

in Appendix D. The table shows that the compressive strength decreases with an

increase in the water-cement ratio. It also shows a small variation in the strength with

the variation of the aggregate volume fraction.


123

6.3 Testing Procedure

After 28 days of moist curing, the samples were insulated on all surfaces but the

top surface, as shown in Fig. (6.1). Then exposed to 8% sodium chloride (NaCl)

solution on the top surface and stored in a controlled temperature and humidity

conditions for 50, 100 and 200 days as shown in Fig. (6.2). Chloride solution level on

the top surface was monitored and kept constant during the ponding period.

The chloride penetration profiles for the concrete specimens exposed to

scheduled chloride ponding (50, 100 and 200 days) were obtained following the

experimental procedure described in Appendix E. The m-Volt readings resulted from

the electrochemical analysis after 50, 100 and 200 days of exposure are shown in

Tables (E.3, E.4 and E.5) of Appendix E, respectively. The mV readings of the

known chloride concentration solutions (0.005%, 0.02%, 0.05% and 0.5%) were

taken and shown in Table (E.6). The calibration curves for the standard concentration

solutions is shown in Fig. (E.5).

Fig. 6.2: Ponding with chloride solution.


124

6.4 Test Results and Discussion


6.4.1 Chloride penetration profiles

The calibration curves shown in Fig. (E.5) were used to convert the mV

readings to total chloride concentration as shown in Tables (6.1) through (6.6).

6.4.2 Effect of water-cement ratio

The chloride profiles after 50, 100 and 200 days of exposure for different water-

cement ratios 0.45, 0.55 and 0.65 (mixes 46, 56 and 66) are shown in Figs. (6.1)

through (6.3). The comparisons of chloride penetration profiles for mixes 46 and 56

after 50, 100 and 200 of exposure are shown in Figs. (6.3), (6.4) and (6.5),

respectively. One can see that increasing water-cement ratio from 0.45 to 0.55

resulted in a large increase in chloride concentration and this is due to the increase in

the capillary porosity and thus chloride penetration. This is correct except for the

chloride profile obtained from the specimen made of concrete mix 46 after 200 days

of exposure (see Fig. 6.5), where poor compaction of the concrete mix resulted in

high porosity and thus increase the chloride penetration. The poor compaction of mix

46 is a consequence of lowering the amount of mixing water. In the practice,

lowering mixing water should be compensated by adding high range water reducers

for better workability of the mix. But, in the present study, we decided that no

additives are used in all concrete mixes.

Also, the comparison of chloride profiles obtained from specimens made of

mixes 46 and 56 with the specimen made of mix 66 show that for all exposure periods

(50, 100 and 200 days), the chloride penetration in mix 66 is larger. This is true
125

expect in the depth range from 0 to 19 mm in which the specimens were exposed to

chloride solution for 50 and 100 days as shown by Figs. (6.3) and (6.4). In which,

segregation of aggregates away from the exposed surface, as shown in Fig. (6.6),

resulted in high porosity of the concrete in this depth range. The segregation of

aggregates in mix 66 is due to the higher amount of water in the concrete mix.

Table 6.1: Chloride concentration for different (w/c) after 50 days exposure time.

Depth range (mm)


w/c Sample #
0-6 6-13 13-19 19-25 25-32 32-38
1 1.191 0.392 0.159 0.110 0.021 -
2 1.127 0.574 0.222 0.083 0.016 -
0.45 3 0.986 0.521 0.310 0.083 0.065 -
Average 1.101 0.496 0.231 0.092 0.034 -
1 1.153 0.593 0.267 0.074 0.057 -
2 1.169 0.576 0.277 0.086 0.051 -
0.55 3 1.158 0.582 0.305 0.070 0.048 -
Average 1.160 0.584 0.283 0.077 0.052 -
1 1.076 0.471 0.296 0.153 0.134 0.104
2 1.137 0.505 0.202 0.108 0.106 0.116
0.65 3 1.033 0.507 0.166 0.096 0.089 0.103
Average 1.082 0.494 0.222 0.119 0.110 0.107
126

1.4
w/c = 0.45

Chloride conc. (% by concrete weight)


1.2
w/c = 0.55

1.0 w/c = 0.65

0.8

0.6

0.4

0.2

0.0
0-6 6-13 13-19 19-25 25-32
Depth range (mm)

Fig. 6.3: Chloride profiles for different (w/c) after 50 days exposure time.

Table 6.2: Chloride concentration for different (w/c) after 100 days exposure time.

Depth range (mm)


w/c sample #
0-6 6-13 13-19 19-25 25-32 32-38 38-45 45-51
1 1.329 0.473 0.196 0.094 0.037 0.043 0.019 0.028
2 1.298 0.470 0.187 0.086 0.033 0.041 0.017 0.025
0.45 3 1.184 0.468 0.206 0.081 0.028 0.041 0.016 0.025
average 1.270 0.470 0.196 0.087 0.033 0.042 0.017 0.026
1 1.081 0.578 0.299 0.188 0.099 0.072 0.048 0.047
2 1.050 0.616 0.291 0.181 0.099 0.069 0.043 0.040
0.55 3 1.035 0.543 0.282 0.176 0.096 0.063 0.036 0.035
average 1.055 0.579 0.291 0.182 0.098 0.068 0.042 0.040
1 1.273 0.528 0.238 0.170 0.149 0.091 0.131 0.122
2 1.255 0.520 0.238 0.166 0.126 0.089 0.126 0.120
0.65 3 1.225 0.510 0.230 0.156 0.122 0.087 0.125 0.114
average 1.251 0.519 0.235 0.164 0.132 0.089 0.127 0.118
127

1.4
Chloride conc. (% by concrete weight) w/c = 0.45
1.2
w/c = 0.55
w/c = 0.65
1.0

0.8

0.6

0.4

0.2

0.0
0-6 6-13 13-19 19-25 25-32 32-38 38-45 45-51
Depth range (mm)

Fig. 6.4: Chloride profiles for different (w/c) after 100 days exposure time.

Table 6.3: Chloride concentration for different (w/c) after 200 days exposure time.

Depth range (mm)


w/c Sample #
0-6 6-13 13-19 19-25 25-32 32-38 38-45 45-51
1 1.540 0.758 0.577 0.378 0.249 0.133 0.028 0.020
2 1.439 0.758 0.536 0.376 0.215 0.099 0.021 0.019
0.45 3 1.371 0.652 0.509 0.353 0.206 0.090 0.020 0.016
average 1.450 0.722 0.541 0.369 0.223 0.107 0.023 0.018
1 1.846 0.622 0.304 0.137 0.101 0.104 0.125 -
2 1.608 0.591 0.296 0.081 0.099 0.097 0.124 -
0.55 3 1.600 0.564 0.286 0.134 0.093 0.094 0.119 -
average 1.685 0.592 0.295 0.117 0.098 0.098 0.122 -
1 1.855 0.871 0.685 0.570 0.460 0.424 0.405 -
2 1.827 0.866 0.651 0.561 0.426 0.377 0.401 -
0.65 3 1.818 0.828 0.579 0.558 0.422 0.367 0.371 -
average 1.833 0.855 0.638 0.563 0.436 0.389 0.392 -
128

2.0
1.8 w/c = 0.45
Chloride conc. (% by concrete weight) w/c = 0.55
1.6
w/c = 0.65
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0-6 6-13 13-19 19-25 25-32 32-38 38-45
Depth range (mm)

Fig. 6.5: Chloride profiles for different (w/c) after 200 days exposure time.

8% NaCl
Solution

150mm

300mm

Fig. 6.6: Segregation of aggregate.


129

As shown in Table (C.1) of Appendix C, the water-cement ratio and cement

content for mix 46 are 0.45 and 430 (Kg/m3); for mix 56 are 0.55 and 380 (Kg/m3);

and for mix 66 are 0.65 and 341 (Kg/m3), respectively. So, the increase in the water-

cement ratio is associated with a decrease in cement content. The review of literature

on the effect of cement content on chloride penetration revealed that chloride binding

capacity increases with the increase of the cement content (Mangat and Molloy,

1995). This was proven indirectly in this experimental study, where the increase in

the chloride penetration due to the increase in water-cement ratio was

counterbalanced by the decrease in the bound chlorides due to decrease in cement

content. This evidence becomes apparent after 200 days of exposure, where chloride

penetration is increased with the increase of water-cement ratio, as shown in Fig.

(6.5).

6.4.3 Effect of aggregate volume fraction

The chloride profiles for different aggregate volume fractions 0.55, 0.65 and

0.75 (mixes 55, 56 and 57) after 50, 100 and 200 days of exposure are shown in Figs.

(6.7) through (6.9). The figures show that the chloride concentration in a specimen

made of mix 55 is larger than the concentration in specimen made of mix 56 in the

depth range from 0-19 mm and smaller in the depth range from 19-45 mm. Due to

large particle size and high specific gravity aggregate tends to segregate from the

concrete mix and settles down in the lower portion of the specimen. The proportions

of mix 55 (see Table C.1 of Appendix C) shows that the mix has a large amount of

mixing water and cement. This resulted in segregation of aggregates and decreasing

the local aggregate volume fraction near the exposed surface and increasing the
130

aggregate content away form the exposed surface as shown in Fig. (6.6). The non-

uniform distribution of aggregates resulted in high diffusivity near the exposed

surface and low diffusivity away from the exposed surface. This results in higher

chloride concentration near the exposed surface (0-19 mm), and lower concentration

away from the exposed surface (19-45 mm).

On the other hand, mix 57 (see Table C.1 of Appendix C) has high aggregate

content, which results in poor compaction of the concrete and thus increasing the

porosity and air voids. The increase of the porosity resulted in high penetration rates

of chloride into concrete.

Thus the increase in the volume fraction of aggregate resulted in a decrease of

chloride penetration if the construction practices and the concrete mix are properly

designed to prevent the segregation and honeycombing. Otherwise, the segregation

of aggregate will have significant impact on the chloride resistance of the concrete at

different depths.
131

Table 6.4: Chloride concentration for different (gi) after 50 days of exposure.

Depth range (mm)


gi Sample #
0-6 6-13 13-19 19-25 25-32 32-38
1 1.808 0.709 0.129 0.053 0.038 0.023
2 2.113 0.825 0.183 0.030 0.081 0.036
0.55 3 1.901 0.791 0.176 0.020 0.017 0.040
average 1.941 0.775 0.163 0.035 0.045 0.033
1 1.153 0.593 0.267 0.074 0.057 -
2 1.169 0.576 0.277 0.086 0.051 -
0.65 3 1.158 0.582 0.305 0.070 0.048 -
average 1.160 0.584 0.283 0.077 0.052
1 1.252 0.784 0.699 0.770 0.394 0.322
2 1.132 0.844 0.732 0.471 0.390 0.334
0.75 3 1.293 0.766 0.719 0.454 0.399 0.335
average 1.226 0.798 0.716 0.565 0.394 0.330

2.5
gi = 0.55
Chloride conc. (% by concrete weight)

2.0 gi = 0.65
gi = 0.75

1.5

1.0

0.5

0.0
0-6 6-13 13-19 19-25 25-32
Depth range (mm)

Fig. 6.7: Chloride profiles for different (gi) after 50 days of exposure.
132

Table 6.5: Chloride concentration for different (gi) after 100 days of exposure.

Sample Depth range (mm)


gi
# 0-6 6-13 13-19 19-25 25-32 32-38 38-45 45-51
1 2.213 0.858 0.331 0.082 0.039 0.018 0.013 0.038
2 1.835 0.846 0.323 0.081 0.033 0.018 0.012 0.024
0.55 3 1.748 0.764 0.317 0.075 0.032 0.017 0.012 0.020
Average 1.932 0.823 0.324 0.079 0.035 0.017 0.012 0.027
1 1.081 0.578 0.299 0.188 0.099 0.072 0.048 0.047
2 1.050 0.616 0.291 0.181 0.099 0.069 0.043 0.040
0.65 3 1.035 0.543 0.282 0.176 0.096 0.063 0.036 0.035
Average 1.055 0.579 0.291 0.182 0.098 0.068 0.042 0.040
1 1.179 0.634 0.373 0.284 0.154 0.145 0.062 -
2 1.151 0.601 0.366 0.272 0.149 0.125 0.061 -
0.75
3 1.151 0.595 0.354 0.269 0.143 0.121 0.059 -
Average 1.160 0.610 0.365 0.275 0.149 0.131 0.061 -

2.5
gi = 0.55
Chloride conc. (% by concrete weight)

2.0 gi = 0.65
gi = 0.75

1.5

1.0

0.5

0.0
0-6 6-13 13-19 19-25 25-32 32-38 38-45
Depth range (mm)

Fig. 6.8: Chloride profiles for different (gi) after 100 days of exposure.
133

Table 6.6: Chloride concentration for different (gi) after 200 days of exposure.

Depth range (mm)


gi Sample #
0-6 6-13 13-19 19-25 25-32 32-38 38-45
1 2.074 0.779 0.370 0.118 0.046 0.116 0.046
2 1.832 0.775 0.340 0.107 0.043 0.071 0.045
0.55 3 1.666 0.772 0.349 0.118 0.048 0.078 0.051
Average 1.857 0.775 0.353 0.115 0.046 0.089 0.047
1 1.846 0.622 0.304 0.137 0.101 0.104 0.125
2 1.608 0.591 0.296 0.081 0.099 0.097 0.124
0.65 3 1.600 0.564 0.286 0.134 0.093 0.094 0.119
Average 1.685 0.592 0.295 0.117 0.098 0.098 0.122
1 1.359 0.639 0.472 0.383 0.318 0.233 0.195
2 1.306 0.581 0.469 0.377 0.302 0.225 0.194
0.75 3 1.287 0.572 0.453 0.373 0.296 0.222 0.187
Average 1.318 0.597 0.465 0.378 0.305 0.226 0.192

2.0
1.8 gi = 0.55
Chloride conc. (% by concrete weight)

1.6 gi = 0.65
gi = 0.75
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
0-6 6-13 13-19 19-25 25-32 32-38 38-45
Depth range (mm)

Fig. 6.9: Chloride profiles for different (gi) after 200 days of exposure.
134

6.5 Conclusions

1. Chloride penetration in concrete increases with the increase of water-cement ratio,

which is due mainly to the increase of the capillary porosity.

2. Chloride penetration in concrete decreases with the increase of aggregate volume

fraction. The additional aggregates decrease the diffusion coefficient of concrete

and thus impede chloride penetration in concrete.

3. Chloride binding in concrete increases with the increase of the cement content.

4. The experimental results of this study emphasized the importance of proper

construction practices and concrete mix design, where segregation of aggregates

and poor compaction of concrete would significantly increase chloride

penetration.
CHAPTER 7

EXPERIMENTAL STUDY ON THE COUPLING EFFECT OF MOISTURE


DIFFUSION AND CHLORIDE PENETRATION IN CONCRETE

7.1 Introduction

The interaction between moisture diffusion and chloride penetration becomes

more important in concrete structures subjected to repeated wetting and drying cycles.

Those structures include bridge decks, parking garage floors and marine structures in

the splash and tidal zones. Chloride diffusion due to concentration gradient is the

dominant penetration mechanism in fully saturated concrete, where pores inside

concrete become completely filled with solution and water movement ceased. On the

other hand, chloride penetration in partially saturated concrete involved many

mechanisms besides diffusion due to concentration gradient. Those mechanisms

include water movement (carrying chloride ions) due to diffusion, sorption and

permeability.

The review of the available literature showed that very few experimental

studies were conducted on the coupled effect of moisture and chloride penetration.

Swaddiwudhipong et al. (2000) investigated the effect of moisture transport due to

concentration gradient, capillary suction and pressure gradient on chloride ingress in

partially saturated concrete. Their experimental results showed that moisture transport

has no effect on chloride penetration for long term exposure period and chloride
136

concentration gradient is the most dominant mechanism for chloride penetration in

concrete. Buenfeld et al. (1995) studied the effect of wick action on chloride

penetration in concrete. Wick action can be defined as the transport of water or

solution through concrete from a wet front (face in contact with solution) to dry front

(face in contact with atmosphere).

The experimental study in this chapter concentrated on the mutually coupled

effects between moisture diffusion and chloride penetration. Two series of

experiments were conducted. One is for the effect of moisture diffusion on chloride

penetration, while the other for the effect of chloride penetration on moisture

diffusion. The effect of moisture diffusion on chloride was conducted by a

comparative study on saturated and non-saturated concrete. The inverse problem was

studied utilizing wick action.

7.2 Specimens Preparation

The specimen shown in Fig. (7.1) was used to study the coupled effect of

moisture diffusion and chloride penetration in concrete. The specimens were designed

as concrete cubes (20x20x15 cm) with a central hole, 10cm in diameter and extends

to the depth of 11.4 cm. The samples were casted, demolded after 24 hours and

stored in the moist room for 28 days.


137

Fig. 7.1: Concrete sample for coupled moisture diffusion and chloride penetration

7.3 Effect of Moisture Diffusion on Chloride Penetration


7.3.1 Experimental procedure

The mix proportions of the concrete used in this experiment are shown in

Table (C.3) of Appendix C. After 28 days of moist curing, 8 samples were taken from

the moisture room. Four of them were exposed to wetting by placing them in a tank

containing pure water for one month. The other four samples were exposed to drying

in the laboratory environment (temperature=27 oC and relative humidity=40%) for

the same period of time. The compression test for concrete cylinders of 10x20cm

(4”x8”) made of the same mix was performed as described in Appendix D. The

compression test results are shown in Table (D.2) of Appendix D. The table shows

that the average 28-day compressive strength of the concrete cylinders is 23.4 MPa.

After one month of exposure to drying (four samples) and wetting (the other

four samples), one sample of each set was tested for the internal relative humidity by

using the commercial HM44 concrete humidity measurement system manufactured


138

by Vaisala, Inc. This measurement system is mainly composed of relative humidity

probe connected to electronic indicator. The relative humidity can be measured by

boring a hole to the required depth as shown in Fig. (7.2). After cleaning the hole, a

plastic sleeve should be inserted in the hole. At this point, the probe can be pushed

into the sleeve and sealed. The concrete at the bottom of the hole release moisture

into the space around the probe until equilibrium is reached. The probe is then

connected to the electronic indicator, which gives the reading of the relative humidity.

The setup, which was used in the measurement of relative humidity, is shown in Fig.

(7.2).

Figure 7.2: Setup for concrete humidity measurement.

The rest of the specimens (6 specimens) were immersed in a bath containing

sodium chloride solution with the concentration of 5% NaCl (by weight) as shown in
139

Fig. (7.3). The specimens were exposed to chloride solution from outside, and the

initially fully saturated specimens were exposed to pure water in the central hole.

Fig. 7.3: Immersion bath of the concrete samples.

After certain exposure periods (30 days, 90 days and 180 days), samples of

concrete powder were collected at various depths from the outside surface of the wall

of specimens. The powder samples were analyzed for total chloride content by using

the rapid chloride test instrument manufactured by Germann Instruments, Inc. as

described in Appendix E.

7.3.2 Experimental results and discussion

The inside relative humidity, measured by concrete humidity measurement

system, in the set of specimens which was exposed to drying found to be 40%. In the
140

other set, which was exposed to wetting found to be 100%.

The results of the electrochemical analysis (mille Volt readings) of the two sets of

the specimens at different depths from the exposed surface and different time of

exposure (30, 90, and 180 days) are presented in Tables (E.7) through (E.9) in

Appendix E. The mV readings of the known chloride concentration solutions

(0.005%, 0.02%, 0.05% and 0.5%) are shown in Table (E.10). The total chloride

concentrations converted from mV-readings using the calibration curves shown in

Fig. (E.6) are presented in Tables (7.1) through (7.3).

The chloride penetration profiles obtained from the two different sets of

specimens after 30, 90 and 180 days of exposure are shown in Figs. (7.4) through

(7.6). One can see clearly from the figures that the chloride concentrations are higher

in the partially saturated concrete than the fully saturated concrete, which means that

the moisture diffusion accelerates the chloride diffusion when the two diffusion

processes occur in the same direction. Mathematically, the chloride diffusion in

saturated concrete can be expressed by a standard partial differential equation, Eq.

(7.2), based on Fick’s law, Eq. (7.1). However, due to the simultaneous occurrence

of the moisture diffusion and chloride diffusion in non-saturated concrete, an

additional term in the Fick’s law must be included to take into account the effect of

moisture diffusion, as shown in Eq. (7.3). As a result, the governing equation

becomes Eq. (7.4)

J Cl = − DCl grad ( C f ) (7.1)


141

∂ Ct
= − div ( J Cl ) = div[ DCl grad (C f )] (7.2)
∂t

J Cl = − DCl grad (C f ) − DCl − H grad ( H ) (7.3)

∂ Ct
= − div ( J Cl ) = div[ DCl grad ( C f )] + div[ DCl − H grad ( H )] (7.4)
∂t

in which, JCl = chloride flux, Ct = total chloride concentration, Cf = free chloride

concentration, H = pore relative humidity in concrete, DCl = chloride diffusivity, DCl-

H = chloride diffusivity due to moisture diffusion, t = time. The first term in the right

hand side of Eqs. (7.3) and (7.4) represents the driving force of chloride concentration

gradient, and the second term reflects the effect of moisture diffusion on chloride

penetration.

Table 7.1: Total chloride concentration after 30 days of exposure.

Depth Initially partially saturated Initially fully saturated


range sample number sample number
(mm) average average
1 2 3 1 2 3
0-6 0.241 0.248 0.256 0.248 0.217 0.203 0.155 0.192
6-13 0.161 0.146 0.162 0.156 0.109 0.076 0.076 0.087
13-19 0.102 0.081 0.092 0.091 0.036 0.035 0.020 0.031
19-25 0.024 0.034 0.025 0.028 0.017 0.012 0.011 0.014
25-32 0.008 0.011 0.008 0.009 0.012 0.011 0.006 0.010
142

0.30

Initially Fully Saturated

chloride conc. (% by concrete weight)


0.25
Initially Partially Saturated

0.20

0.15

0.10

0.05

0.00
0-6 6-13 13-19 19-25 25-32
Depth range (mm)
Fig. 7.4: chloride profiles after 30 days of exposure.

Table 7.2: Total chloride concentration after 90 days of exposure.

Depth Initially partially saturated Initially fully saturated


range sample number sample number
(mm) average average
1 2 3 1 2 3
0-6 0.319 0.294 0.289 0.301 0.277 0.258 0.226 0.253
6-13 0.203 0.197 0.191 0.197 0.131 0.130 0.130 0.130
13-19 0.116 0.101 0.100 0.106 0.066 0.060 0.056 0.061
19-25 0.053 0.053 0.051 0.052 0.017 0.016 0.016 0.016
25-32 0.041 0.039 0.036 0.039 0.009 0.009 0.009 0.009
143

0.35

Initially Fully Saturated


0.30

chloride conc. (% by concrete weight)


Initially Partially Saturated

0.25

0.20

0.15

0.10

0.05

0.00
0-6 6-13 13-19 19-25 25-32
Depth range (mm)
Fig. 7.5: chloride profilest after 90 days of exposure.

Table 7.3: Total chloride concentration after 180 days of exposure.

Depth Initially partially saturated Initially fully saturated


range sample number sample number
(mm) average average
1 2 3 1 2 3
0-6 0.363 0.360 0.353 0.359 0.294 0.277 0.269 0.280
6-13 0.251 0.238 0.228 0.239 0.183 0.165 0.165 0.171
13-19 0.164 0.157 0.140 0.153 0.099 0.095 0.094 0.096
19-25 0.126 0.122 0.117 0.122 0.046 0.046 0.045 0.046
25-32 0.111 0.085 0.082 0.093 0.024 0.022 0.020 0.022
144

0.40
Initially Fully Saturated
0.35
chloride conc. (% by concrete weight)
Initially Partially Saturated
0.30

0.25

0.20

0.15

0.10

0.05

0.00
0-6 6-13 13-19 19-25 25-32

Depth range (mm)


Fig. 7.6: chloride profiles after 180 days of exposure.

7.4 The Effect of Chloride Penetration on Moisture Diffusion

As shown in the previous section, the moisture diffusion accelerates chloride

penetration. Another interesting issue is the effect of chloride penetration on

moisture diffusion. This is the inverse problem to the first one. The two problems

are equally important, but the inverse problem has not been investigated before. A

seepage test was designed specifically to study the inverse effect.

7.4.1 Experimental setup and procedures

Concrete mix 56 was used in this experiment. This mix had a water-cement

ratio of 0.55, aggregate volume fraction of 0.65 and average 28-day compressive

strength of 34.1 MPa. The mix proportions and compression test results (see mix 56)
145

are shown in Table (C.2) of Appendix C, and Table (D.1) of Appendix D.

The experimental setup shown in Fig. (7.7) was used to study the effect of

chloride diffusion on moisture diffusion. In this experiment, three sets of concrete

specimens were immersed in three different bathes containing three different

solutions: pure water in Bath 1, sodium chloride (NaCl) solutions with concentrations

of 0.5M in Bath 2 and 1.0M in Bath 3, respectively. Each set had three concrete

specimens. The concrete specimens were initially saturated by water (Relative

Humidity =100%). The external surface of the center hole was exposed to the

atmosphere (40% relative humidity) during the test. In Bathes 2 and 3, chloride ions

penetrate into concrete specimens by diffusion due to concentration gradient and

permeation due to pressure gradient. Water (including chloride in specimens

immersed in bathes 2 and 3) accumulated in the center hole due to flow and wicking

action. The accumulated solution was collected and used as an indicator for the effect

of chloride penetration on moisture transfer. The accumulated solution in the center

holes was collected after 10, 20, 60 and 90 days of ponding. Except at the time of

collection for the accumulated solution, the ponding containers were covered all the

time to reduce the evaporation of the accumulated solution in the hole.


146

Fig. 7.7: Experimental setup for the effect of chloride diffusion on moisture
diffusion.

7.4.2 Results and discussion

The test results shown in Table (7.4) and Fig. (7.8), represent the weight of

the accumulated solution (in grams) in the center hole after 10, 20, 30 and 90 days.

One can see that the amount of the accumulated solution increases with the increase

of chloride concentration in the ponding solution. This means that chloride

penetration accelerates water flow and wicking.

Two important observations can be made from the test data shown in Fig.

(7.8). The first one is that the seepage test can obtain meaningful data in a rather short

time period, less than 30 days. In general, concrete has been considered as a solid

material with very low permeability, thus, direct seepage test is not viable for testing

concrete permeability. But, one can see from Fig. (7.8), that the test data of 10 days
147

shows the same trend as of 90 days, which implies that 10-day testing period may be

sufficient for studying the coupling effect. The second observation is that the

chloride penetration accelerates the moisture diffusion greatly! The amount of

seepage water is about tripled from zero chloride concentration to 0.5M, and then

doubled from 0.5M to 1.0M. It is very clear that the effect of chloride penetration on

moisture diffusion cannot be neglected.

Table 7.4: Weight of the accumulated solution.

Elapsed time (days)


Solution type Sample #
10 20 30 60 90
1 8.252 16.868 27.059 54.017 72.506
2 3.536 5.744 8.746 17.090 21.758
Pure water
3 0.287 0.444 0.797 2.023 2.923
Average 4.025 7.685 12.201 24.377 32.396
1 6.484 10.779 15.197 24.344 30.595
2 31.646 53.315 72.309 116.535 148.817
0.5M NaCl
3 11.873 21.498 30.559 55.062 73.459
Average 16.668 28.531 39.355 65.314 84.290
1 37.350 57.358 86.288 157.521 209.078

2 15.760 43.526 70.457 131.205 175.961


1.0M NaCl
3 32.340 44.796 59.071 95.566 120.817

Average 28.483 48.560 71.939 128.097 168.619


148

180

160 exposed to water


0.5M NaCl solution

weight of collected solution (g)


140 1.0M NaCl solution
120

100

80

60

40

20

0
10 20 30 60 90
elapsed time (days)

Fig. 7.8: The weight of accumulated solution in the center hole for the three
different solutions.

Similar to Eqs. (7.1) and (7.2) for the chloride penetration in saturated

concrete, the standard partial differential equation for moisture diffusion without

chloride ions is Eq. (7.6), which is based on Fick’s law as shown in Eq. (7.5). To take

into account for the effect of chloride penetration, an additional term in the Fick’s law

must be included, as shown in Eq. (7.7). As a result, the governing equation can be

modified as Eq. (7.8)

J w = − DH grad ( H ) (7.5)

∂w
= − div ( J w ) = div[ DH grad ( H )] (7.6)
∂t
149

J w = − DH grad ( H ) − DH − Cl grad ( C f ) (7.7)

∂w
= − div ( J w ) = div[ DH grad ( H )] + div[ DH − Cl grad (C f )] (7.8)
∂t

where Jw = moisture flux, H = pore relative humidity in concrete, w = moisture

content of concrete, DH = moisture diffusivity, DH-Cl = moisture diffusivity due to

chloride diffusion. The two governing equations for chloride penetration and

moisture diffusion, equations (7.4) and (7.8), are two fully coupled equations (two-

way coupled). The results of the seepage test showed that the second term in the right

hand side of equation (7.8) plays a very important role in the moisture diffusion in

non-saturated concrete, which has not be carefully evaluated so far.

7.5 Conclusions

1) In this chapter an experimental study was conducted on the coupled effect of

moisture diffusion and chloride penetration in concrete.

2) The experimental results show that the moisture diffusion accelerated the chloride

penetration (when the two processes are in the same direction). On the other hand,

the chloride penetration accelerated the moisture diffusion.

3) The effect of chloride diffusion on moisture diffusion is substantial and has not

been investigated systematically.

4) The moisture diffusion and chloride penetration are two mutually coupled

processes (two-way coupled), and the coupling effects should be considered in


150

theoretical modeling on the chloride penetration in non-saturated concrete.

5) The seepage test proposed in the study can be used to estimate the chloride

permeability of concrete in a relatively short period of time.


CHAPTER 8

EXPERIMENTAL STUDY ON THE EFFECT OF SHRINKAGE-


INDUCED DAMAGE ON CHLORIDE PENETRATION

8.1 Introduction

Drying shrinkage is one of the major problems that affect the durability of

concrete structures. The contraction of concrete due to moisture loss is usually

hindered by either internal or external restraints and thus tensile hygral stresses are

induced. Those stresses may exceed locally the tensile strength of concrete and thus,

generates microcracks (Bissonnette et al. 1999; Beddoe and Lippok, 1999;

Ramamurthy and Narayanan, 2000). The microcracks change transport properties of

concrete adversely, accelerating the moisture and chloride diffusion processes and

reduce service life of concrete structures. The external restraints include joints,

supports, connections and the bond between the newly cast concrete and the old

concrete in the case of overlays on top of existing concrete pavement. The internal

restraints include the differential shrinkage resulted from the non-uniform drying

along the depth of the concrete from the exposed surface and the non-shrinking

phases such as aggregates, unhydrated cement particles and calcium hydroxide

crystals (CH).

Pioneering experimental work by Powers et al. (1954) studied the effect of

drying on permeability of cement paste. Their results showed that the resistance to
152

water flow through the intact specimens was 70 times greater than those of

companion specimens that underwent a cycle of drying to 79% relative humidity and

resaturation. Since then, many experimental studies and theoretical analyses have

conducted (Aldea et al., 2000; Aldea et al., 1999; Gerard et al., 1996; Daian and

Saliba 1993; Francois and Maso 1988; Bazant et al., 1987). All results showed that

the concrete permeability is considerably increased with the increase of internal

damage in concrete. Bazant et al. (1987) studied the effect of cracking resulted from

drying on the permeability and diffusivity of concrete. Two types of C-shaped

reinforced concrete beams (cracked and uncracked) were exposed to controlled

drying environment and the loss in the weight of the specimen was reported. It was

observed that the cracked specimens had more weight loss, which means that the

drying process is faster in the cracked specimens. Samaha and Hover (1992)

investigated the effect of air-drying (similar to moderate environmental drying

condition) and oven-drying on the transport properties of concrete. They observed

that in air-dried concrete the shrinkage cracks are limited to a small depth from the

exposed surface. From this observation, they concluded that those cracks did not

affect greatly the transport properties of concrete. On the other hand, they noticed a

severe cracking in the oven dried concrete samples. Those cracks are uniformly

distributed through the sample and resulted in a great internal damage. And thus, they

contributed largely to the deterioration of transport properties of concrete. Gerard et

al. (1996) investigated the effect of mechanically induced cracking on the

permeability of normal and high strength concrete. The cracks were produced by

uniaxial tension loading. The water permeability was measured in a direction


153

perpendicular to the axis of loading. Two series of specimens having an average

strength of 32 and 50 (MPa) were tested to better understand the effect of the strength

on the crack pattern and thus the diffusion properties of concrete. It was found that

the damage is controlled by the length and aperture of the cracks. Based on the

observations from the cracking patterns a simple micro-macro model was suggested.

This model can be used to predict the effect of the damage evolution on the diffusion

properties of concrete.

In this chapter, a comparative study on the chloride penetration in cracked (due

to restrained drying shrinkage) and uncracked concrete was conducted. The concrete

was exposed to air drying for a certain period of time then to chloride penetration.

During the drying period the development of the shrinkage strains and cracking were

monitored. Also, chloride profiles for the cracked and uncracked concrete were

obtained.

8.2 Experimental Program

8.2.1 Material

The concrete mix used in this experiment has a water-cement ratio of 0.55 and

aggregate volume fraction of 0.55. The mix proportions are shown in Table (C.4) of

Appendix C. One can see from the proportions, that the mix has high water-cement

ratio, high cement content and low aggregate volume fraction. This was done in order

to have high shrinkage strains and thus significant damage in the concrete specimens.

The average 28-day compressive strength of concrete cylinders prepared from the

same mix is 34 (MPa). The compression test results are shown in Table (D.3) of
154

Appendix D.

8.2.2 Test specimens preparation

Two sets of concrete specimens were used in this experiment. One set was a

full concrete ring restrained by inner steel ring and the other set was half ring free to

shrink as shown in Fig. (8.1). The concrete rings had an outside diameter of 457 mm

(18”), inside diameter of 305 mm (12”) and height of 152 mm (6”). The full ring was

restrained by an inner steel ring having a thickness of 25.4 mm (1”). Total of four

specimens (two of each set) were casted, demolded after 24 hours and kept in the

laboratory environment for another day to dry the exterior surfaces for the purpose of

application of strain gages.

Steel ring

152mm

Side view
45° 45°

45° 45°
Extensometers

254 mm
305 mm 305 mm
457 mm 457 mm

a) Restrained ring b) Unrestrained half ring

Fig. 8.1: Restrained and free shrinkage concrete specimens.

The full ring is restrained against movement in the radial direction, which

develops internal pressure, p when the ring shrinks. This pressure creates tensile
155

stresses in the circumferential direction, σθ and compression stresses in the radial

directions, σr as shown in Fig. (8.2). Cracks in the concrete ring are initiated from the

inner boundaries because of the coexistence of the maximum tensile stress and the

biaxial compression-tension condition (Shah et al., 1998).

σr

internal
σθ pressure
p

σθ σr

σr σθ

Fig. 8.2: Stresses in the restrained shrinkage concrete sample.

8.2.3 Experimental procedure


Measurement of the shrinkage strains

After drying of the surface, a total of 6 strain gages were attached to the

exterior surface of the specimens in order to monitor the development of shrinkage

strains. Three gages were applied on one of the restrained rings while the others were

applied on one of the free shrinkage half rings as shown by Fig. (8.1). All the strain

gages were connected to a data acquisition system as shown in Fig. (8.3). The data

acquisition system was connected to personal computer used to display and process
156

the obtained strains in microstrains. The measurement of the shrinkage strains started

after two days of concrete casting. One day in the moist room before demolding and

the other day to dry the surface of the specimens for installation of the strain gages.

Fig. 8.3: Data acquisition system and personal computer.

Ponding by chloride solution

After 107 days of casting, two specimens (one restrained full ring and one free

half ring) were placed in 8% sodium chloride (NaCl) solution bath as shown in Fig.

(8.4). Before the placement, the specimens were insulated by aluminum foil on all

surfaces expect the exterior surface facing the chloride solution. After 50 days of

continuous exposure to chloride solution the chloride profiles are obtained for both

specimens (restrained and free shrinkage) following the procedure described in

Appendix E. The m-Volt readings resulted from the electrochemical analysis are

shown in Table (E.11) of Appendix E. The mV readings of the known chloride


157

concentration solutions (0.005%, 0.02%, 0.05% and 0.5%) were taken and shown in

Table (E.12). The calibration curve for standard concentration solutions is shown in

Fig. (E.7).

Fig. 8.4: Ponding with sodium chloride solution.

8.3 Experimental Results and Discussion

8.3.1 Shrinkage strain

The shrinkage strains due to drying from initially saturated state were measured

and recorded from four out of the six strain gages placed on the two specimens. This

is mainly due to the limitation by the data acquisition system. The readings of the

shrinkage strain from the restrained and free shrinkage specimens are shown in Table

(8.1). As shown in the table, five readings were taken in the first day, then one

reading per day up to 30 days, and finally weekly reading up to 15 weeks.


158

Table 8.1: Measured shrinkage strains (in microstrains).

Elapsed* Restrained shrinkage Free shrinkage


Time (days) strain gage # 1 strain gage # 2 strain gage # 3 strain gage # 4
0.000 0 0 0 0
0.563 7 18 32 23
0.604 10 18 31 22
0.615 12 22 33 26
0.780 10 22 42 30
1.481 22 43 63 47
1.892 25 47 72 58
3.940 66 92 128 111
4.857 86 110 152 133
5.504 98 123 168 147
5.961 115 128 178 155
6.531 122 132 187 160
7.504 145 153 204 175
10.594 182 198 253 221
10.760 185 200 255 224
11.459 194 204 258 229
12.439 204 217 270 240
13.453 215 229 281 253
14.465 232 239 289 263
17.565 261 280 322 295
18.682 271 288 330 300
19.583 276 297 338 314
20.565 282 303 343 318
21.504 291 313 355 329
24.555 314 337 375 347
25.562 321 344 383 353
26.569 330 352 392 358
27.767 339 359 398 366
31.674 356 379 418 384
33.510 373 394 430 393
39.628 401 420 456 415
45.511 416 439 474 430
46.944 419 445 477 433
52.558 427 442 493 443
56.649 425 454 497 446
62.055 432 466 507 453
66.680 439 473 514 461
75.496 443 479 524 469
79.817 450 487 531 478
83.667 452 481 522 472
94.522 465 505 545 498
105.000 470 505 544 499
* Elapsed time measured after 2 day of concrete casting.
159

Fig. (8.4) shows the development of the shrinkage strains with time, obtained

from the free and the restrained rings. The figure shows that the shrinkage in both free

and restrained specimens developed at higher rate in the first ten days of drying, then

at slower rate up to 60 days and almost leveled after that.

600
Shrinkage strain (microstains)

500

400

300

Restrained (gage #1)


200
Restrained (gage #2)
Free (gage #3)
100
Free (gage #4)
0
0 10 20 30 40 50 60 70 80 90 100 110
Elapsed time (days)

Fig. 8.5: Shrinkage strains development in restrained and free shrinkage concrete

8.3.2 Restrained shrinkage cracking

Microcracks as shown in Figs. (8.6) through (8.8), became visible on the outer

circumferential surface of restrained ring after 90 days of exposure. No cracks were

noticed on the free shrinkage specimens. Most of the time, the cracks started from the

edge (stress concentration location) and extend to the interior on the surface of the

specimens (see Fig. 8.8).


160

Fig. 8.6: One of the microcracks in the restrained shrinkage ring.

Fig. 8.7: Microcrack on the surface of the restrained ring.


161

Fig. 8.8: Microcracking of the surface of the restrained ring.

8.3.3 Chloride profiles

The calibration curve shown in Fig. (E.7) of Appendix E, was used to convert

the mVolt readings to total chloride concentration as shown in Table (8.2). The table

presents a comparison between the chloride concentrations at different depth ranges

for restrained and unrestrained specimens. The specimens were exposed to 105 days

of air drying, then 50 days to chloride penetration. The chloride profiles for damaged

(due to air drying) and undamaged specimens are shown in Fig. (8.9). The figure

shows that chloride penetrate restrained concrete ring in a higher rate compared to the

unrestrained half ring, which means that the shrinkage-induced damage increases the

chloride penetration. One can see that the concentration of chloride in the free

shrinkage specimen is higher in the depth range of 0-13 mm. Since cracks were
162

initiated and propagated from the inner part of the restrained ring, the figure shows

that chloride concentration was higher for restrained ring in depth range from 13-38

mm, but the difference is not significant. This is partly due to nonlinear inelastic

behavior due to the degradation of the initial elastic modulus and the unrecoverable

part of concrete creep.

Table 8.2: Total chloride concentration after 50 days of exposure.

Restrained shrinkage Free shrinkage


Depth
range sample number sample number
(mm) average average
1 2 3 1 2 3

0-6 0.871 0.784 0.767 0.807 0.952 0.847 0.750 0.850

6-13 0.489 0.261 0.230 0.327 0.440 0.412 0.373 0.408

13-19 0.316 0.309 0.204 0.276 0.276 0.258 0.212 0.249

19-25 0.199 0.198 0.197 0.198 0.154 0.150 0.149 0.151

25-32 0.131 0.116 0.101 0.116 0.113 0.109 0.073 0.099

32-38 0.119 0.089 0.070 0.093 0.089 0.082 0.073 0.081


163

0.90

0.80 Free Shrinkage

chloride conc. (% by concrete weight)


Restrained Shrinkage
0.70

0.60

0.50

0.40

0.30

0.20

0.10

0.00
0-6 6-13 13-19 19-25 25-32 32-38
Depth range (mm)

Fig. 8.9: chloride profiles for restrained and free shrinkage concrete samples after 50
days of exposure.

8.4 Conclusions

1) Concrete drying resulted in high shrinkage strains in both free and restrained

shrinkage specimens. The free shrinkage strain was 450-500x10-6 after 105 days

of drying. In restrained shrinkage it was 500-550x10-6 after 105 days of drying.

2) Drying shrinkage resulted in cracking and substantial damage when the concrete

was restrained.

3) The surface cracks resulted from the shrinkage-induced damage increase the

permeability of concrete and thus increase the moisture diffusion and chloride

penetration. However, the effect of shrinkage-induced damage on chloride

penetration is not very strong comparing with other effects (such as mechanical

loading).
CHAPTER 9

CONCLUSIONS AND FUTURE RESEARCH RECOMMENDATIONS

The main goal of this thesis was to investigate some of the processes related to

durability of concrete structures by means of theoretical modeling and experimental

study. The processes include moisture diffusion, chloride penetration, freezing-

thawing and the coupling effects between them. The objectives were achieved by

three consecutive steps. The first step was achieved by reviewing the available

literature and following up the new publications on the researched topics. This step

was necessary for strengthening the background information and understanding the

different mechanisms for each process involved in this study. Also, it paved the road

to the following steps. The second step was the theoretical modeling, which involved

the formulation of the governing equations and boundary conditions, identification of

the material parameters and the numerical solutions. This step resulted in numerical

models for the coupled effect of moisture diffusion and drying shrinkage and the

effect of moisture diffusion on chloride penetration. A systematic description of the

proposed models and parametric studies on the effects of different influential

parameters are given in chapters three and four. The third or the final step was the

experimental study, and it covered four series of experiments. The first series was

designed to study the coupling effect of freezing-thawing and chloride penetration,

which is described in chapter five. The second series was focused on the effect of
165

concrete material parameters on chloride penetration, which is presented in chapter

six. The third series was concentrated on the coupling effect of moisture diffusion

and chloride penetration, which is presented in chapter seven. And the last series was

for the effect of shrinkage-induced cracking on chloride penetration, which is

described in chapter eight. The experimental results obtained in this study can be used

to validate and demonstrate the accuracy of the numerical models developed in this

study and the future studies.

9.1 Conclusions

1) Concrete deteriorations mechanisms are attributed to several factors including

physical, chemical and biological attacks. The common feature of those factors is

that they are all diffusion-controlled processes. Those processes include moisture

diffusion, chloride penetration and freezing-thawing.

2) The diffusion properties of concrete are related to its permeability, which is

affected by many factors including material parameters and proportions,

construction practices, exposure conditions and the features of cracking and air

voids.

3) Durability of concrete structures can be enhanced by improving their resistance to

moisture diffusion, chloride penetration and frost action. This can be achieved by

lowering the permeability, lowering the water-cement ratio, appropriate selection

of materials, good compaction, prevent premature damage and cracking from

shrinkage and thermal effects, and good curing conditions.

4) The Multiscale modeling of the coupled effect of moisture diffusion and drying

shrinkage predicted well the basic trends and the effects of concrete material
166

parameters, such as the water-cement ratio and the aggregate volume fraction. In

this model, the effect of damage on moisture diffusion was characterized by the

concept of continuum damage mechanics using the isotropic damage parameter

proposed by Kachanov (1958). This is an acceptable approach, since microcracks

induced by drying shrinkage are randomly distributed without a preferential

orientation. The effect of drying shrinkage on the moisture capacity was modeled

based thermodynamics laws and minimum potential energy principle.

5) The numerical study of chapter four resulted in a rigorous model for chloride

penetration into non-saturated concrete. The model predicts very well the effect of

moisture diffusion on the chloride ion penetration in partially saturated concrete.

The numerical results show that the moisture diffusion accelerates the chloride

penetration. Also, it incorporates the effects of concrete mix parameters; such as

volume fraction of aggregate and water-cement ratio.

6) The numerical study also presented the stability and convergence of the

Alternating-Direction Implicit method for solving the governing equations of

moisture diffusion and chloride penetration.

7) The model can be used in the durability design of reinforced concrete structures to

predict the onset of steel corrosion, which is important for predicting the service

life of concrete structures and for scheduling inspection and rehabilitation of the

structures.

8) The experimental study of chapter 5 emphasized the following:

o The large deteriorating action of alternating freezing and thawing on

durability of concrete. The progressive damage and cracking of concrete cover


167

increases the permeability and accelerate the rate of chloride penetration

towards the steel bars, which increases the chloride concentration in concrete

and triggers the corrosion process.

o The consistency of the pulse velocity and the resonant frequency for

predicting the concrete damage.

o The permeability is not only related to the level of damage and crack width,

but also to the crack density and connectivity.

9) The experimental study of chapter six, presented the effect of concrete material

parameters as well as the construction practices on the resistance to chloride

penetration.

10) The experimental work in chapter seven emphasized the mutually effect of

moisture diffusion and chloride penetration in concrete. The experimental results

show that the moisture diffusion accelerated chloride penetration (when the two

processes are in the same direction). On the other hand, chloride penetration

accelerated the moisture diffusion.

11) The experimental study of chapter eight presented the effect shrinkage-induced

damage on the chloride penetration. In which, the surface cracks increase the

permeability of concrete and thus increase the moisture diffusion and chloride

penetration in concrete.

9.2 Future Research Recommendations

1) This study can be further extended to include the effect of the mechanical loading
168

on the diffusion processes.

2) Since the theoretical modeling was done before the experimental study, a lot of

new findings have not been included in the current theoretical models. i) the

effect of chloride penetration on moisture diffusion needs to be included in the

model for moisture diffusion, and in this way, the moisture diffusion and chloride

penetration become a fully two-way coupled processes; ii) a thermal conduction

model must be established in order to evaluate the damage due to the freezing-

thawing cycles, and furthermore to evaluate the chloride penetration under the

coupled thermal effect; iii) in the material models, such as binding capacity and

chloride permeability, one more material parameter, cement content, must be

incorporated; and finally iv) once the coupled diffusion models and material

models are established, they must be validated by using the experimental data

from Chapters 5, 6, 7, and 8.

3) Further experimental and numerical studies are needed on the effect of mineral

additives on chloride penetration, such silica fume (SF), ground blast furnace slag

(GBFS) and fly ash (FA) on the chloride penetration.


169

BIBLIOGRAPHY

Ababneh, A., Xi, Y., and Willam, K. (2001). “Multiscale Modeling of the Coupled
Moisture Diffusion and Drying Shrinkage of Concrete.” Proc. of 6th International
Conference on Creep, Shrinkage & Durability Mechanics of Concrete and other
Quasi-Brittle Materials, Aug. 20-22, MIT, Cambridge, USA, pp. 159-164.

Aldea, C.M., Shah S.P., and Karr, A. (1999), “Effect of Cracking on Water and
Chloride Permeability of Concrete.” Journal of Materials in Civil Engineering,
11(3), 181-187.

Aldea, C.M., Shah, S.P., and Karr, A. (2000). “Effect of Microcracking on Durability
of High-Strength Concrete.” Transportation Research Record, (1668),
Transportation Research Board, Washington DC, 86-90.

Al-Khaja, W.A. (1997). “Influence of temperature, cement type and level of concrete
consolidation on chloride ingress in conventional and high-strength concrete.”
Construction and Building Materials, 11(1), 9-13.

Altoubat, S.A. and Lange, D.A. (1997). “Early Age Shrinkage and Creep of Fiber
Reinforced Concrete For Airfield Pavement.” Proc. Of Aircraft/Pavement
Technology: In the Midst of Change, ASCE, Seatle, WA, 229-243.

Amey, S.L., Johnson, D.A., Miltenberger, M.A., and Farzam, H. (1998). “Predicting
the service life of concrete marine structures: an environmental methodology.”
ACI Structural Journal, 95(2), 205-214.

Andrade, C. (1993). “Calculation of Chloride Diffusion Coefficients in Concrete from


Ionic Migration Measurement.” Cement and Concrete Research,23,724-742.

Andrade, C., and Whiting, D. (1996). “A comparison of chloride ion diffusion


coefficients derived from concentration gradients and non-steady state
accelerated ionic migration.” Material and Structures, 29, 476-484.

Argyris, J.H., Vaz, L.E., and Willam, K.J. (1977). “Higher Order Methods for
Transient Diffusion Analysis.” Computer Methods in Applied Mechanics and
Engineering, 12, 243-278.

Arya, C., Buenfeld, N.R., and Newman, J.B. (1990). “Factors influencing chloride-
binding in concrete.” Cement and Concrete Research, 20(2), 291-300.

Baluch, M.H., Qureshy, A.B., and Azad, A.K. (1989). “Fatigue Crack Propagation in
Plain Concrete.” Proc. Of Fracture of Concrete and Rock, Ed. Shah, S.P., and
Swartz, S.E., Springer-Verlag, New York, 80-87.
170

Bazant, Z.P. and X., Y. (1994). “Drying creep of concrete: constitutive model and
new experiments seprating its mechanisms.” Materials and Structures, 27, 3-14.

Bazant, Z.P., and Chern, J.C. (1985). “Concrete creep at variable humidity:
Constitutive law and mechanism.” Materials and Structures (RILEM), 18, 1-20.

Bazant, Z.P., and Najjar, L.J. (1972). “Nonlinear water diffusion of nonsaturated
concrete.” Matériaux et Constructions, 5(25).

Bazant, Z.P., and Raftshol, W.J. (1982) “Effect of Cracking in Drying and Shrinkage
Specimens.”Cement and Concrete Research, 12, 209-226.

Bazant, Z.P., and Thonguthai, W. (1978). “Pore Pressure and Drying of Concrete at
High Tempearture.” Journal of Engineering Mechanics-ASCE, 104, 1058-1080.

Bazant, Z.P., Chern, J.C., Rosenberg, A. M., and Gaidis, J. M. (1988). “Mathematical
Model for Freeze-Thaw Durability of Concrete.” J. of Ame. Ceram. Soc., 71(9),
776-783.

Bazant, Z.P., Sener, S., and Kim J.K. (1987). “Effect of Cracking on Drying
Permeability and Diffusivity of Concrete.” ACI Materials Journal, 84, 351-358.

Beddoe, R.E., and Lippok, R. (1999). “Hygral stress in hrdenend cement paste.”
Materials and Structures, 32, 627-634.

Bentz, D.P., Clifton, J.R., Ferraris, C.F., and Garboczi, E.J. (1999). “Diffusion
Properties and Durability of Concrete: Literature Review and Research Plan.”
Building and Fire Research Laboratory, NIST, Gaithersburg, Maryland.

Bentz, D.P., Garboczi, E.J. and Lagergren E.S. (1998). “Multi-Scale Microstructural
Modeling of Concrete Diffusivity: Identification of Significant Variables.”
Journal of Cement, Concrete & Aggregates, 20(1), 129-139.

Bentz, D.P., Quenard, D.A., Veronique, B.B., Garboczi, E.J. and Jennings, H.M.
(1995). “Modelling during shrinkage of cement paste and mortar, Part 1.
Structral models from nanometers to millimeters.” Materials and
Stuctures,28,450-458.

Bentz, D.P.,(2000). “Influence of silica fume on diffusivity in cement-based materials


II. Multi-scale modeling of concrete diffusivity.” Cement and Concrete
Research, 30, 1121-1129

Berke, N.S. (1986). “Corrosion Rates of Steel in Concrete.” ASTM Standardization


News, 14(3), 57-61.

Bissonnette, B., Pierre, P., and Pigeon, M. (1999). “Influence of Key Parameters on
171

Drying Shrinkage of Cementitious Materials.” Cement and Concrete Research,


29(10), 1655-1662.

Bouny, V.B., Perrin, B., and Chemloul, L. (1997). “Experimental Determination of


Moisture Properties of Hardened Cement Pastes, Showing Hysteresis Effects.”
Materials and Structures (RELEM), 30, July, 340-348.

Browne, R. (1982). “Design Prediction of the Life for Reinforced Concrete in a


Marine and Other Chloride Environment.” Durability of Building Materials,
1(2), 113-125.

Brunauer S., Skalny J., and Bodor E.E. (1969). “Adsorption on nonporous solids.”
Journal of Colloid and Interface Science, 30 (4), 546-552

Buenfeld, N.R., Shurfa-Daoudi, M.T., and McLoughlin, I.M. (1995). “Chloride


Transport Due to Wick Action in Concrete.” Proc. Of International RILEM
workshop on chloride penetration in concrete, France, 315-324.

Cady, P.D., and Weyers, R.E. (1992). “Predicting Service Life of Concrete Bridge
Decks Subject to Reinforcement Corrosion.” Corrosion Forms & Control for
Infrastructure, San Diego, CA.

Carnahan, B., Luther, H.A. and Wilkes, J.O. (1969). Applied numerical methods.
John Wiley & Sons, Inc., New York.

Chatterji, S. (1994). “Transportation of Ions through Cement Based Materials, Part I:


Fundamental Equations and Basic Measurement Techniques.” Cement and
Concrete Research, 24(5), 907-912.

Christensen, R.M. (1979). Mechanics of composite materials. Wiley Interscience,


New York.

Coggins, F.B., and French, C.W. (1990). “Chloride Ion Penetration in Twenty-Year-
Old Prestressed Bridge Girders.” ACI Materials Journal,87(5),479-488.

Collepardi, M., Marciall, A., and Turriziani, R. (1972). “Penetration of chloride ions
in cement pastes and concrete.” Journal of American Ceramic Society, 55(10),
534-535.

Costa, A., and Appleton, J. (1999). “Chloride penetration into concrete in marine
environment–Part II: Prediction of long term chloride penetration.” Material and
Structures (RILEM), 32, 354-359.

Daian, J.F., Saliba, J. (1993). “Transient Moisture Transport in a Cracked Porous


Medium.” Transport in Porous Media, 13, 239-260.
172

Derucher, K. N., and Heins, C. P. (1981). Materials for Civil and Highway
Engineers. Prentice-Hall, New Jersey.

Diamond, S., and Bonen, D. (1993). “Microstructure of Hardened Cement Paste—A


New Interpretation.” J. Am. Ceram. Soc., 76(12), 2993-2999.

Digler, W.H., Ghali, A., Chan, M., Cheung, M.S., and Maes, M. A. (1983). “
Temparature-Induced Stresses in Composite Box-Girder Bridges.” Journal of
Structural Engineering, ASCE, 109(6), 1460-1478.

Elbadry, M., and Ghali, A. (1983). “Temperature Variations in Concrete Bridges”,


Journal of Structural Engineering, 109(10), 2355-2374.

Emanuel, J. H., and Hulsey, J. L. (1978). “Temperature Distribution in Composite


Bridges.” Journal of Structural Division, ASCE, 104, 65-78.

Fagerlund, G. (1992). “Effect of the Freezing Rate on the Frost Resistance f


Concrete.” Nordic Concrete Research, Publication No. 11, Oslo.

Fagerlund, G. (1994). “Predicting the Service Life of Concrete Exposed to rust


Action Through a Modeling of the Water Absorption Process in the Air Pore
syste.” Proc. of NATO/RILEM workshop on the Modeling of Microstructure and
its Potential for Studying Transport Properties and Durability, St. Remy-les-
Chevreuse, France, July 10-13.

Francois, R., Maso, J.C. (1988), "Effect of Damage in Reinforced Concrete on


Carbonation or Chloride Penetration", Cement and Concrete Research, 18, 961-
970.

Frey, R., Balogh, T., and Balazs, G. L. (1994). “Kinetic Method to Analyze Chloride
Diffusion in Various Concretes.” Cement and Concrete Research, 24(5), 863-
873.

Gerard, B., and Marchand, J. (2000). ”Influence of cracking on the diffusion


properties of cement-based materials, Part I: Influence of continuous cracks on
the steady-state regime.” Cement and Concrete Research, 30, 37-43.

Gerard, B., Breysse, D., Ammouche, A., Houdusse, O., and Didry, O. (1996).
“Cracking and permeability of concrete under tension.” Materials and Structures
(RILEM), 29, 141-151.

Gerard, B., Breysse, D., Ammouche, A., Houdusse, O., and Didry, O. (1996) “
Cracking and permeability of concrete under tension”, Materials and Structures,
29, 141-151.

Gontar, W.A., Martin, J.P., Popovics, J.S. (2000), “ Effect of Cyclic Loading on
173

Chloride Permeability of Plain Concrete”, Special Publication of ASCE


“Condition Monitoring of Materials and Structures”, F. Ansari, Editor, 84-94.

Hale, D.K. (1976). “Review the Physical Properties of Composite Materials.” J. of


Materials Science, 11, 2105-2141.

Hashin, Z. (1983). “Analysis of Composite Materials - A Survey.”, Journal of


Applied Mechanics, 50, 481-505.

Helmuth, R. A. (1960). “Capillary Size Restrictions on Ice Formation in Hardened


Portland Cement Pastes.” Proc. of the 4th Int. Sym. on the Chemistry of Cement,
Washington, D.C., Paper VI-S2, 855-869.

Herve, E., and Zaoui, A. (1990). “Modeling of Effective Behavior of Nonlinear


Matrix-Inclusion Composites.” Eur. J. Mech. And Solids, 9(6), 505-515.

Hilsdorf, H.K., and Kesler, C.E. (1966). “Fatigue Strength of Concrete under Varying
Flexural Stresses.” ACI Journal , 63(10), 1059-1075.

Jennings, H. M., and Tennis, P. D. (1994). “A model for the Developing


Microstructure in Portland Cement Paste.” Journal of American Ceramic Society,
77(2), 3161-3172.

Jennings, H.M., and Xi, Y. (1993) “Microstructurally based Mechanisms for


Modeling Shrinkage of Cement Paste at Multiple Levels.” Proc. of the 5th Int.
Symp. on Creep and Shrinkage of Concrete, Barcelona, Spain.

Kachanov, L. M. (1958). “Time rupture process under creep conditions.” Izv. Akad.
Nauk, USSR, Otd. Tech. Nauk. (8), 26-31.

Kachanov, L. M. (1986). Introduction to Continuum Damage Mechanics. Martinus


Nijhoff Dordrecht, The Netherlands.

Kawai, K., Kohno, K., Morinaga, T., and Tazawa, E. (1996). “Effects of Organic and
Carbonic Acids on Concrete Deterioration Caused by Aerobic Microorganisms.”
Journal of Materials, Concrete Structures and Pavements, 550, 105-115.

Kermani, A. (1991), “Stressed Concrete: Permeability of Stressed Concrete”,


Building Research and Information, 19(6), 360-366.

Kong, J., Ababneh, A., Frangopol, D.M., and Xi, Y. (2001). “Reliability Analysis of
Chloride Penetration in Saturated Concrete.” submitted to Probabilistic
Engineering Mechanics.

Konin, A., francois, R., Arliguie, G. (1998), “Penetration of Chloride in Relation to


the Microcracking State into Reinforced ordinary and High Strength Concrete”,
174

Materials and Structures (RILEM), 31, 310-316.

Li, Z., Qi, M., Li, Z., and Ma, B. (1999). “Crack Width of High-Perfromance
Concrete Due to Restrained Shrinkage.” Journal of Materials in Civil
Engineering, 11(3), 214-223.

Litvan, G.G. (1973). ”Pore Structure and Frost Susceptibility of Building Materials."
Proc., RILEM/IUPAC Int. Sym. on Pore Structure and Properties of Materials,
Sept. 18-21, Prague, F17-F30.

Luping, T., and Nillsson, L.O. (1993). “Chloride binding capacity and binding
isotherms of OPC pastes and mortars.” Cement and Concrete Research, 23(2),
247-253.

Majorana, C., and Mazars, J. (1997). “Thermohygrometric and Mechanical Behavior


of Concrete Using Damage Models.” Materials and Structures (RILEM), 30,
349-354.

Mangat, P.S., and Molloy, B.T. (1995). “Chloride binding in concrete containing
PFA, GBS or silica fume under sea water exposure.” Magazine of Concrete
Research, 47(171), 129-141.

Martys, N. S. (1995). Survey of Concrete Diffusion Properties and Their


Measurement. Building and Fire Research Laboratory, NIST, Gaithersburg,
Maryland.

Martys, N. S., Torquato, S., and Bentz, D. P. (1994). “Universal Scaling of Fluid
Permeability for Sphere Packing.” Phys. Rev., E, 50(1), 403-408.

Mehta, P. K. (1991). “Durability of Concrete -- Fifty Years of Progress?” Proc.,


Second International Conference on Durability of Concrete, V. M. Malhorta, ed.,
Montreal, Canada, v. 1, 1-31.

Mehta, P.K., and Monteiro, P.J.M. (1993). Concrete, Structures, Properties, and
Materials. Prentice Hall, Englewood Cliffs, New Jersey, p.120.

Meziani, H., Skoczylas, F. (1999), “An Experimental Study of the Mechanical


Behavior of a Mortar and of its Permeability under deviatoric Loading”,
Materials and Structures (RILEM), 32, 403-409.

Mier, J. G. van., (1997). Fracture processes of concrete: assessment of material


parameters for fracture models. CRC Press, Inc., Boca Raton, Florida.

Mohamed, O. A., Rens, K. L. and Stalnaker, J. J. (2000). “Factors Affecting


Resistance of Concrete to Freezing and Thawing Damage.” Journal of Materials
in Civil Engineering,12(1),26-32.
175

Neubauer, C. M., Bergstrom, T. B., Sujata, K., Xi, Y., Garboczi, E.J., and Jennings,
H.M. (1997). “Drying Shrinkage of Cement Paste as Measured in an ESEM and
Comparison with Microstructural Models.” Journal of Materials Science, 32,
6415-6427.

Neville, A. M. (1996). Properties of Concrete, John Wiley & Sons, Inc., New York.

Nordby, G.M. (1958). “Fatigue of Concrete-A Review of Research.” Journal of the


ACI Journal, 30(2), 191-219.

Olsen, M.P.J. (1984). “Mathematical Modeling of the Freezing Process of Concrete


and Aggregates", Cement and Concrete Research, 14, 113-122.

Page, C.L., Short, N.R., and El Tarras, A. (1981). “Diffusion of Chloride Ions in
Hardened Cement Paste”. Cement and Concrete Research, 11(3), 395-406.

Philleo, R.E. (1986) “Freezing and Thawing Resistance of High-Strength Concrete”,


NCHRP Synthesis 129, Transportation Research Board, NRC, Washington D.C.

Popovics, S. (1987). “A Classification of the Deterioration of Concrete Based on


Mechanism.” Proc., Katharine and Bryant Mather International Conference on
Concrete Durability, Atlanta, Georgia, 1, 131-142.

Powers, T.C., Copeland, L.E., Hayes, J.C., and Mann, H.M. (1954). “Permeability of
Portland Cement Paste.” ACI Journal, 51,285-298.

Ramamurthy, K. and Narayanan, N. (2000). “Influence of composition and curing on


drying shrinkage of aerated concrete.” Material and Structures, 33, 243-250.

Reinhardt, H.W., Sosoro, M., Zhu, X.-F. (1996), “Cracked and Repaired Concrete
Subject to Fluid Penetration”, Materials and Structures (RILEM), 31, 74-83.

Rosen, B. W., and Hashin, Z. (1970). “Effective Thermal Expansion Coefficients and
Specific Heats of Composite Materials.” Int. J. Engng. Sci., 8, 157-173.

Saetta, A.V., Scotta, R. V., and Vitaliani, R. V. (1993). “Analysis of Chloride


Diffusion into Partially Saturated Concrete.” ACI Materials Journal, 90(5), 441-
451.

Saito M., Ishimori, H. (1995) “Chloride Permeability of Concrete Under Static and
Repeated Loading”, Cement and Concrete Research, 25(5), 803-808.

Samaha, H., and Hover, K. C. (1992). “Influence of Microcracking on the Mass


Transport Properties of Concrete.” ACI Materials Journal, 89(4), 416-424.

Scherer, G. W. (1999). “Crystallization in Pores.” Cement and Concrete Research,


176

29,1347-1358.

Selleck, S. F., Landis, E. N., Peterson, M. L., Shah, S. P., and Achenbach J. D.
(1998). “Ultrasonic Investigation of Concrete with Distributed Damage.” ACI
Materials Journal, 95(1), 27-36.

Shah, S.P., Ouyang, C., Marikunte, S., Yang, W. and Becq-Giraduon, E. (1998). “A
Method to Predict Shrinkage Cracking of Concrete.” ACI Materials Journal,
95(4), 339-346.

Shah, S.P., Weiss, W.J., and Yang, W. (1997). “Shrinkage Cracking in High
Performance Concrete.” Proc. of the PCI/FHWA International Symposium on
High Performance Concrete, New Qrleans, LA, 148-158.

Sugiyama, T., Bremner, T.W., Holm, T.A. (1996), “Effect of Stress on Gas
Permeability in Concrete”, ACI Materials Journal, 93(5), 443-50.

Suryavanshi, A.K., Swamy, R.N., and McHugh, S. (1998). “Chloride penetration into
reinfrorced concrete slabs.” Canadian Journal of Civil Engineering, 25, 87-95.

Swaddiwudhinpong, S., Wong, S.F., Wee, T.H., and Lee, S.L. (2000). “Chloride
ingress in partially and fully saturated concretes.” Concrete Science and
Engineering, 2, 17-31.

Tang, L., and Nilson, L.O. (1993). “Chloride binding capacity and binding isotherms
of OPC pastes and mortars.” Cement and Concrete Research, 23, 247-253.

Taylor, H. F. W. (1987). “A method for Predicting Alkali Ion Concentration in


Cement Pore Solutions.” Advanced Cement Research, 1(1), 5.

Tritthart, J. (1989). “Chloride binding in cement. II. The influence of hydroxide


concentration in the pore solution of hardned cement paste on chloride binding.”
Cement and Concrete Research, 19, 683-691.

Troxell, G. E., Davis, H. E. and Kelly, J. W. (1968). Composition and Properties of


Concrete. McGraw-Hill, Inc., Second Edition.

Wang, K, Jansen, D.C., Shah, S.P. (1997), “Permeability Study of Cracked


Concrete”, Cement and Concrete Research, 27(3), 381-393.

Wee, T.H., Wong, S.F., Swadiwudhipong, S., and Lee, S.L. (1997). “A Prediction
Method for Long-Term Chloride Concentration Profiles in Hardened Cement
Matrix Materials.” ACI Materials Journal, 94(6), 565-576.

Xi, Y. (1991) Analysis of Concrete Creep, Shrinkage and Fracture by Deterministic


and Probabilistic Methods. Ph.D. Dissertation, Northwestern University,
177

Evanston, Il.

Xi, Y. (1995a). “A model for moisture capacities of composite materials, Part I:


formulation.” Computational Materials Science, 4, 65-77.

Xi, Y. (1995b). “A model for moisture capacities of composite materials, Part II:
application to concrete.” Computational Materials Science, 4, 78-92.

Xi, Y. and Ababneh, A. (2000) “Prediction of the Onset of Steel Corrosion in


Concrete by Multiscale Chloride Diffusion.” Proc. of the International
Symposium on High Performance Concrete: Workability, Strength and
Durability, December 10-15, Hong Kong & Shenzhen, China, pp. 181-186.

Xi, Y. and Jennings, M. (1997). “Shrinkage of Cement Paste and Concrete Modeled
by a Multi-scale Effective Homogeneous Theory.” Materials and Structures
(RILEM), 30, 329-334.

Xi, Y., and Bazant, Z.P. (1999). “Modeling Chloride Penetration in Saturated
Concrete.” Journal of Materials in Civil Engineering, 11(1), 58-65.

Xi, Y., and Jennings, H. M. (1992). “Relationships Between Microstructure and


Creep and Shrinkage of Cement Paste.” Material Science of Concrete, III, Ed. J.
Skalny, The Amer. Cer. Soc., Westerville, OH, 37-69.

Xi, Y., Bazant, Z.P., Molina, L. and Jennings, H.M. (1994b). “Moisture Diffusion in
Cementitious Materials: Moisture Capacity and Diffusivity.” Advanced Cement
Based Materials, 1, 258-266.

Xi, Y., Bazant, Z.P., Molina, L., and Jennings, H.M. (1994a). “Moisture Diffusion in
Cementitious Materials: Adsorption isotherm.” Advanced Cement Based
Materials, 1, 248-257.

Xi, Y., Willam, K., and Frangopol, D. M. (2000). “Multiscale Modeling of Interactive
Diffusion Processes in Concrete.” Journal of Engineering Mechanics, ASCE,
126(3), 258-265.

Zhen-hai, G. and Xiu-qin, Z. (1987). “Investigation of complete stress-deformation


curves for concrete in tension.” ACI Materials Journal, 84(4), 278-285.

Zuber, B., Marchand, J., Delagrave, A. and Bournazel, J. P. (2000). “Ice Formation
Mechanisms in Normal and High-Perfrormanc Concrete Mixtures.” Journal of
Materials in Civil Engineering,12(1),16-23.
APPENDIX A

ADSORPTION ISOTHERMS AND MOISTURE CAPACITIES OF

CEMENT PASTE AND AGGREGATE

The adsorption isotherm of porous media (the relationship between moisture

content and relative humidity at a constant temperature) can be described by the BSB

theory (Brunauer et al. 1969) as follows:

C k Vm H
w = (A.1)
(1 − kH )[1 + (C − 1 )kH ]

The moisture capacity can be obtained by deriving the adsorption isotherm with

respect to H as follows:

dw C k V m + wk [1 + (C − 1 )kH ] − wk [(C − 1 )(1 − kH ) ]


= (A.2)
dH (1 − kH )[1 + (C − 1 )kH ]

where, Vm is the monolayer capacity which is the mass of adsorbate required to cover

the adsorbent with a single molecular layer, C is the net heat of adsorption, which is a

function of temperature, and k determines the degree of saturation of the pores. The

three material parameters Vm, C and k can be evaluated by using adsorption test data

of the material.
179

Adsorption Isotherm and Moisture Capacity of Cement Paste

The adsorption isotherm and moisture capacity of cement paste can be

calculated using Eq. (A.1) and (A.2) and the following formulas for the Vm, C and k

parameters (Xi 1995b):

⎛ 0.22 ⎞⎛ w⎞ w
Vm = ⎜ 0.068 − ⎟⎜ 0.85 + 0.45 ⎟Vct t > 5 days, 0.3 < < 0.7 (A.3)
⎝ t ⎠⎝ c⎠ c

in which t is the time in days; w/c is the water-cement ratio; and Vct is for the type of

cement, Vct = 0.9, 1, 0.85 and 0.6 for type I, II, III and IV cements, respectively. The

model for parameter C is

⎛C ⎞
C = exp⎜ 0 ⎟ (A.4)
⎝T ⎠

in which T is the temperature in Kelvin; and C0 = 855. The model for parameter k is

(1 − 1 / n) C − 1
k= (A.5)
C −1

in which n is the number of the adsorbed layers in the pore and it is a function of t (in

days), water cement ratio, w/c, and type of cement (Nct = 1.1, 1.0, 1.15 and 1.5 for

type I, II, III and IV cements, respectively):

⎛ 15 ⎞⎛ w⎞ w
n = ⎜ 2.5 + ⎟⎜ 0.33 + 2.2 ⎟ N ct t > 5 days, 0.3 < < 0.7 (A.6)
⎝ t ⎠⎝ c⎠ c
180

Adsorption Isotherm and Moisture Capacity of Aggregate

The adsorption isotherm and moisture capacity of aggregates can be

calculated using the same BSB theory, and thus, Eqs. (A.1), (A.2), (A.4) and (A.5)

are still valid, but the monolayer capacity, Vm and the number of the adsorbed layers,

n, are different. The following formulas may be used for parameters Vm and n of

aggregate (Xi 1995b):

Vm = 0.0647Vagg (A.7)

n = 4.063nagg (A.8)

in which Vagg and nagg depend on the pore structures of the aggregates. For dense

aggregates, Vagg ranges from 0.05 to 0.1 and nagg from 1.0 to 1.5. For porous

aggregates, Vagg ranges from 0.01 to 0.04 and nagg from 1.7 to 2.0. It should be noted

that dense aggregates have larger Vagg and smaller nagg because of their small size

pores and high surface area. Dolomite, granite, crushed limestone and river gravel

are examples of dense aggregates. Lightweight aggregates and river sand are

examples of porous aggregates.


APPENDIX B

VOLUME PERCENTAGES OF DIFFERENT HYDRATION PRODUCTS

Jenning and Tennis model (1994) considers five different phases of cement

paste: anhydrous cement particles, inner hydration products, outer hydration products,

capillary pores and calcium hydroxide, CH (include AFm phase). The volume of C-S-

H gel is the sum of the volumes of inner and outer hydration products.

VC − S − H = Vinner + Vouter (B.1)

in which Vinner and Vouter are the volumes of inner and outer hydration products. The

volume fraction of C-S-H gel can be written as:

Vtotal − V1 − V pores
f C −S −H = = 1 − f1 − f pores (B.2)
Vtotal

in which Vpores and fpores are the volume and volume fraction of capillary pores and V1

and f1 are the volume and volume fraction of anhydrous cores of cement particles and

all the crystalline phases such as calcium hydroxide (CH) and ettringite.

The volume fraction of capillary pores can be written as follows:

f pores = ((1 − c) − cΩ ) p (B.3)

in which, c and p parameters depend on the water-cement ratio, w/c and the densities
182

of cement, ρc and water, ρw as follows:

1
c= (B.4)
1 + ( w / c)

(1 + (w / c))ρ c
p= (B.5)
1 + ( w / c) * ( ρ c / ρ w )

And Ω is a parameter depends on the degree of hydration of cement and can be

written as follows:

Ω = 0.437α C 3 S W C 3 S + 0.503α C 2 S W C 2 S + 0.397α C 3 AW C 3 A + 0.136α C 4 AF W C 4 AF (B.6)

in which WC3S, WC2S, WC3A, and WC4AF are the initial weight percentages of the

compounds of the cement. Also, αC3S, αC2S, αC3A, and αC4AF are the degree of

hydration of the compounds of the cement. The initial weight percentages of the

cement compounds, Wi (i = C3S, C2S, C3A and C4AF) depends on the type of the

cement. The degree of hydration of the cement compounds, αi (i = C3S, C2S, C3A and

C4AF) can be determined as follows:

αi = 1 − e − a ( t − b )
ci
i i
(B.7)

in which ai, bi and ci (i = C3S, C2S, C3A and C4AF) are constants determined

experimentally (Taylor et. al 1987) and shown in Table (B.1).


183

Table (B.1): Parameters related to the hydration of different compounds of cement.

Compound ai bi ci
C3S 0.25 0.90 0.70
C2S 0.46 0.00 0.12
C3A 0.28 0.9 0.77
C4AF 0.26 0.90 0.55

Also, f1 is the volume percentage of the anhydrous cores of cement particles, calcium

hydroxide (CH) crystals and ettringite as follows:

f 1 = f core + f CH + f AFm (B.8)

The volume fraction of the anhydrous cores of cement particles can be determined by

the following formula:

⎛ 1 −ψ ⎞
f core = c⎜⎜ ⎟⎟ p (B.9)
⎝ ρc ⎠

in which c and p are determined by Eqs. (B.4) and (B.5), ρc is the density of cement

and ψ is a factor depends on the degree of hydration of the compounds of the cement

and can calculated based weight average as follows:

ψ = α C 3SWC 3S + α C 2 SWC 2 S + α C 3 AWC 3 A + α C 4 AFWC 4 AF (B.10)

The volume fraction of CH crystals and AFm phase can be determined as follows:

f CH = c(0.189α C 3 S WC 3 S + 0.058α C 2 S WC 2 S ) p (B.11)


184

f AFm = c(0.849α C 3 AWC 3 A + 0.472α C 4 AF WC 4 AF ) p (B.12)

The volume fraction of the other twp phases can be calculated as follows:

f 3 = [0.706( w / c) + 0.131](1 − f1 − f pores ) + f pores (B.13)

f 2 = 1 − f1 − f 3 (B.14)
APPENDIX C

MIX DESIGN FOR THE EXPERIMENTAL PROGRAM

The description of the materials used in the concrete mixes, are as follows:

Coarse Aggregate (CA) or Gravel

Size, No. 4 to ¾ in., angular aggregate.

Absorption (%) = 0.384

SG (bulk dry) = 2.8

SG (bulk saturated surface dry) = 2.811

Moisture content (%) = 0.242

Dry-rodded unit weight (lb/ft3) = 96.81 [1550 Kg/m3]

Fine Aggregate (FA) or Sand

Fineness modulus (FM) = 2.93

Absorption (%) = 0.8

SG (bulk dry) = 2.153

SG (bulk saturated surface dry) = 2.305

Moisture content (%) = 0.351

Concrete Mix Design

This mix was designed according to Goldbeck and Gray Method (see the book

Materials for Civil and Highway Engineers, Kenneth N. Derucher and Conrad P.

Heins, 1981, pp. 153).

b: solid volume of coarse aggregate in a unit volume of concrete.

bo: solid volume of coarse aggregate in a unit volume of dry-rodded coarse aggregate.
186

Solid weight of CA = SG (bulk dry) *unit weight of water

= 2.8*62.4 = 174.72 lb/ft3 [2800 Kg/m3]

So,

dry − rodded unit weight 96.81


b0 = = = 0.554
solid weight 174.72

• From table 6-4 in page 155 (Derucher et al. 1981) and by using the size of coarse

aggregates (No. 4 to ¾ in) and fineness modulus of fine aggregates (2.93), the

value of b/ bo is 0.61.

• b can be calculated as follows:

b
b= * bo = 0.61 * 0.554 = 0.338
bo

• From table 6-5 page 156 in the above reference, and by using the size of CA

aggregate (No. 4 to ¾ in), angular in shape, slump of 3” and 28-day compressive

strength, ( f c' ) of 4333 psi [30 MPa],

¾ The required mixing water = 40 gal/yd3 (≈333lb/yd3) [198 Kg/m3]

333
Volume of mixing water = = 5.337 ft3/yd3
62.4

¾ The required cement = 6.4 sacks/ yd3 (≈604 lb/yd3) [358 Kg/m3]

604
Volume of cement = = 3.073 ft3/yd3
3.15 * 62.4

Volume of the entrapped air ≈ 2 % *27 = 0.54 ft3/yd3

333
water-cement ratio, w / c = ≅ 0.55
604

The volume of the coarse aggregate (CA) = 0.338*27 = 9.126 ft3/yd3


187

¾ The weight of the coarse aggregate (CA) = 9.126*174.72 ≈ 1595 lb/yd3

[946 Kg/m3]

The volume of fine aggregate (FA) = 27-volume of cement-volume of water-

volume of air-volume of coarse aggregate

= 27-6.4*0.48-40/7.5-0.02*27-9.126

= 8.929 ft3/yd3

¾ The weight of the fine aggregate (FA) = 8.929*2.153*62.4

≈ 1200 lb/yd3 [712 Kg/m3]

9.126 + 8.929
The aggregate volume fraction, g i = = 0.67
27

Design for certain water-cement ratio and different aggregate volume fractions

(w/c = 0.55 and gi = 0.55, 0.65 and 0.75)

The total volume of concrete, VT:

VT = Vw + Vc + VCA + VFA + Vair


123 1424 3 {
Vcp Vagg 2%

VT = Vcp + Vi + V air

dividing by VT :

1 = g cp + g i + 0.02

for gi = 0.55,

g cp = 0.43

Also, the volume of cement paste, Vcp :

Vcp = g cp *VT = 0.43 * 27 = 11.61 ft3/yd3


188

The volume of cement paste, Vcp:

Ww Wc W *w/c Wc W 1
Vcp = V w + Vc = + = c + = c (w / c + )
ρw 3.15 ρ w ρw 3.15 ρ w ρ w 3.15

And so the weight of cement, Wc:

Vcp * ρ w
Wc =
( w / c + 1 / 3.15)

For w/c = 0.55, the weight of cement:

Wc = 71.934 * Vcp

For gi = 0.55, the weight of cement :

¾ Wc = 71.934 *11.61 = 835 lb/yd3 [495 Kg/m3]

The volume of cement:

Wc 835
Vc = = = 4.248 ft3/yd3
3.15 * ρ w 3.15 * 62.4

The weight of water:

¾ Ww = w / c *Wc = 0.55 * 835 = 459 lb/yd3 [272 Kg/m3]

The volume of water:

Ww 459
Vw = = = 7.356 ft3/yd3
ρw 62.4

The volume of aggregate (Coarse and fine):

Vagg = 27 − Vcp − Vair = 27 − 11.61 − 0.02 * 27 = 14.85 ft3/yd3

Assuming that the VCA = VFA = 14.85/2 = 7.425 ft3/yd3:

The weight of gravel:

¾ WCA = VCA * 2.8 * 62.4 = 1297 lb/yd3 [769 Kg/m3]


189

The weight of sand:

¾ W FA = V FA * 2.153 * 62.4 = 998 lb/yd3 [592 Kg/m3]

By following the same procedure mentioned previously the mix design for w/c= 0.55,

gi= 0.65 and 0.75 can be obtained. The results of the mix design for w/c = 0.55, gi=

0.55, 0.65 and 0.75 are shown in the Table (C.1) below.

Design for certain aggregate volume fraction and different water-cement ratios

(gi = 0.65 and w/c = 0.45, 0.55 and 0.65)

for gi = 0.65:

g cp = 0.33

Also, the volume of cement paste, Vcp :

Vcp = g cp * VT = 0.33 * 27 = 8.91 ft3/yd3

The volume of aggregate (Coarse and fine):

V agg = 27 − V cp − V air = 27 − 8.91 − 0.02 * 27 = 17.55 ft3/yd3

Assuming that the VCA = VFA = 17.55/2 = 8.775 ft3/yd3:

The weight of gravel:

¾ WCA = VCA * 2.8 * 62.4 = 1533 lb/yd3 [909 Kg/m3]

The weight of sand:

¾ W FA = V FA * 2.153 * 62.4 = 1179 lb/yd3 [699 Kg/m3]

And the weight of cement, Wc:


190

Vcp * ρ w Vcp * 62.4


Wc = =
1 1
(w / c + ) (w / c + )
3.15 3.15

The weight of cement for w/c = 0.55 :

8.91 * 62.4
Wc = = 641 lb/yd3 [380 Kg/m3]
(0.55 + 0.317)

The weight of water for w/c = 0.55 :

Ww = w / c * Wc = 0.55 * 641 = 353 lb/yd3 [209 Kg/m3]

By following the same procedure mentioned previously the mix design for gi = 0.65,

w/c = 0.45 and 0.65 can be obtained. The results of the mix design for gi = 0.65, w/c

= 0.45, 0.55 and 0.65 are shown in Table (C.1) below.

Table C.1: Concrete mix design proportions.

Mix 46 56 66 55 57
water-cement ratio (w/c) 0.45 0.55 0.65 0.55 0.55

aggregate volume fraction (gi) 0.65 0.65 0.65 0.55 0.75

weight of water (Kg/m3) 193* 209* 222* 272* 146*

weight of cement (Kg/m3) 430 380 341 495 265

weight of sand (Kg/m3) 699* 699* 699* 592* 806*

weight of gravel (Kg/m3) 909* 909* 909* 769* 1049*

* Values need to be corrected to account for the moisture condition of aggregates.

Correction for moisture content of aggregates

The proportions of water, coarse and fine aggregates are based on their dry

weights. In reality aggregates are not dry (unless dried) and have certain moisture
191

content. The amount of water calculated based on the mix design should be

completely available for cement hydration during the hardening of concrete. Thus, the

amount of water absorbed by aggregates after the deduction of moisture content must

be added. This is correct if the moisture content of aggregate is less the amount of

water required to reach saturated surface dry condition (SSD). If the aggregates are

wet the amount of mixing water need to be decreased by the amount of free additional

water. Also, the proportions of coarse and fine aggregates must be adjusted to

compensate for the water content.

As an example for mix 46 (w/c = 0.45 and gi = 0.55) the additional amount of

water need to be added is:

From Coarse Aggregates (CA) = (0.00384 − 0.00242 ) * 909 = 1.25 Kg

From the Fine Aggregate (FA) = (0.008 − 0.00351) * 699 = 3.14 Kg

The amounts of aggregates need to be added are:

Coarse Aggregate (CA) = 0.00242*909 = 2.2 Kg

Fine Aggregate (FA) = 0.00351*699 = 2.5 Kg

For mix 46, the new mix proportions are, (193+1.25+3.14≈197 Kg/m3) for the

water, (699+2.5≈702 Kg/m3) for sand (FA) and (909+2.2≈911 Kg/m3) for gravel

(CA). All mix proportions for the five mixes are corrected and tabulated again in

Table (C.2).
192

Table C.2: Mix proportions used in the experimental study of Chapters 5, 6 and 7.

Mix 46 56 66 55 57
water-cement ratio (w/c) 0.45 0.55 0.65 0.55 0.55
aggregate volume fraction (gi) 0.65 0.65 0.65 0.55 0.75
weight of water (Kg/m3) 197 213 226 276 151
weight of cement (Kg/m3) 430 380 341 495 265
weight of sand (Kg/m3) 702 702 702 594 809
weight of gravel (Kg/m3) 911 911 911 771 1052

Table C.3: Concrete mix proportions used in experimental study of the effect of
moisture diffusion on chloride penetration (Chapter 7)

water-cement ratio (w/c) 0.65


3
weight of water (Kg/m ) 232
weight of cement (Kg/m3) 356
weight of sand (Kg/m3) 847
weight of gravel (Kg/m3) 1031

Table C.4: Proportions of the concrete mix used in the experiment of the effect of
induced-shrinkage cracking on chloride penetration (Chapter 8)

water-cement ratio (w/c) 0.55


aggregate volume fraction (gi) 0.55
weight of water (Kg/m3) 244
weight of cement (Kg/m3) 448
weight of sand (Kg/m3) 593
weight of gravel (Kg/m3) 771
APPENDIX D

COMPRESSION TEST RESULTS

Concrete cylinders of 10x20cm (4”x8”) were casted, remolded after 24 hours

and then cured in a moisture room for 28 days. After curing, all specimens were

capped using a suitable capping material. The compression test of the concrete

cylinders was performed using a universal MTS loading machine, with maximum

capacity of 484 KN (110 Kips). In this test was, a compressive axial load was applied

on the concrete cylinder at rate of 2 (Kn/sec) until failure. The maximum loads were

recorded and the compressive strength was calculated by dividing the average of

maximum loads by the cross-sectional area of the specimen.

The following are the compression test results of concrete specimens prepared

from the mixes used in the experimental studies of Chapters 5 through 8.

Table D.1: Compression test results for the concrete mixes used in the experimental
study of Chapters 5, 6 and 7.

Sample # Maximum load (KN)


Mix-46 Mix-56 Mix-66 Mix-55 Mix-57
1 303 265 226 226 245
2 307 267 233 268 265
3 315 280 235 276 279
4 323 283 248 287 286
5 325 297 255 310 306
Average 314.6 278.7 239.5 273.5 276.4
f c' (MPa) 38.5 34.1 29.3 33.5 33.8
194

Table D.2: Compression test results for the concrete used in the experiment of the
effect of moisture diffusion on chloride penetration (Chapter 7).

Sample number Maximum load (KN)


1 156
2 178
3 200
4 200
5 214
Average 189.6
'
f (MPa)
c
23.40

Table D.3: Compression test results for the concrete used in the experiment of the
effect induced-shrinkage cracking on chloride penetration (Chapter 8).

Sample number Maximum load (KN)


1 267
2 298
3 276
4 260
Average 275
' 34
f (MPa)
c
APPENDIX E

EXPERIMENTAL PROCEDURE FOR OBTAINING CHLORID PROFILES

After certain exposure period, the specimens were cleaned by removing the

insulation material and salts and then dried at room conditions for one day. After

drying, the surfaces of specimens were cleaned from salt deposits by using a wire

brush.

Collecting Dust Samples

Concrete dust samples were collected from certain depth ranges by using the

drill press machine shown in Fig. (E.1). Drilling started from the side exposed to

chloride penetration and extends at 1/4” ( ≈ 6 mm) depth increments.

Fig. E.1: Drilling of concrete sample.


196

Electro-chemical Analysis

The electro-chemical analysis of powder samples collected from the concrete

specimens resulted in a profile of total chloride concentration as weight percentage of

concrete weight. The test kit for electro-chemical analysis (manufactured by

Germann Instruments, Inc.) is shown in Fig. (E.2). It mainly consists of chloride

electrode and electrometer for mV (mille Volt), calibration liquids 0.005%, 0.02%,

0.05% and 0.50% Cl-, and chloride ions extraction liquid.

Fig. E.2: Chloride concentration measurement kit.

The electro-chemical analysis method procedure is described as follows:

1. A 1.5 gram of the collected dust sample was added to a vial containing 10

ml of chloride extraction liquid as shown in Fig. (E.3).


197

Fig. E.3: Sampling of 1.5 gram of concrete dust and the extraction liquid vial.

2. The vials containing the concrete dust and the extraction liquid were

secured by a lid and shaken for 5 minutes.

3. The electrode was submerged into a vial containing calibration solution

(know chloride concentration solution) and the mille Volt (mV) reading

displayed by the electrometer was recorded.

4. Four solutions of 0.005%, 0.02%, 0.05% and 0.5% chloride concentration

were used to plot the calibration curve.

5. The electrode need to be cleaned by spraying with distilled water and

wiped after each use.

6. The electrode was submerged into the vials containing the concrete dust

and extraction liquid and the measured mV-readings were recorded.

7. The measured mV-readings were converted to total chloride concentration

by using the calibration curve.


198

The following are the results of the electro-chemical analysis of the

experimental studies in chapters 5, 6, 7 and 8.

Table E.1: The mV readings after 400 freezing-thawing cycles.

Mix Sample # Depth range (mm)


0-6 6-13 13-19 19-25 25-32 32-38 38-45
1 -28.8 -5 0.7 -4.5 -1.7 0.8 6.9
2 -26.1 -4.7 1.3 -3.8 -0.8 0.9 6.9
46
3 -25.9 -4.6 2 -3.6 -0.7 1.1 8.3
1 -28.9 -19 -13 -8.6 -5.3 -4 -3.6
2 -27.8 -18.4 -12.8 -8.1 -4.9 -3.3 -3.2
56
3 -27.8 -17.6 -12.7 -7.9 -4.7 -3.1 -2.8
1 -29.9 -15.3 -10 -5.2 - - -
2 -29.4 -14.3 -9.9 -3.4 - - -
66
3 -28.3 -13.6 -8.4 -2.8 - - -
1 -30.4 -16.9 -13.5 -11.3 -9.9 -5.1 -
2 -30.1 -16.5 -12.1 -10.7 -8.7 -4.7 -
55
3 -29.3 -16.3 -11.7 -9.6 -7.4 -4.6 -
1 -27.7 -11.2 0.6 - - - -
2 -27 -7.4 0.8 - - - -
57
3 -26.3 -6.9 1.3 - - - -

Table E.2: The mV readings for the standard solutions.

% Cl- mV reading

0.005 77.4
0.02 51.3
0.05 42.5
0.5 -5.2
199

1
Chloride conc. (% by weight)

0.1

0.01

0.001
-10 0 10 20 30 40 50 60 70 80 90 100
mV

Fig. E.4: Calibration curve for obtaining chloride concentrations.


200

Table E.3: The mV readings after 50 days of exposure.

Depth range (mm)


Mix Sample #
0-6 6-13 13-19 19-25 25-32 32-38
1 -21.4 2.8 22.4 30.4 66.1 -

46 2 -20.2 -5.5 15.2 36.6 72.1 -

3 -17.3 -3.4 7.9 36.7 41.9 -


1 -20.7 -6.2 11.2 39 45 -

56 2 -21 -5.6 10.4 35.9 47.4 -

3 -20.8 -5.8 8.3 40.3 48.5 -


1 -19.2 -1.2 8.9 23.3 26.2 31.8

66 2 -20.4 -2.7 17.2 30.9 31.3 29.3

3 -18.3 -2.8 21.5 33.4 35 32


1 -30.5 -10.1 27.1 46.2 53.9 64.3

55 2 -33.9 -13.4 19.4 58.5 37.1 54.9

3 -31.6 -12.5 20.2 67.6 70.7 52.8


1 -22.5 -12.3 -9.8 -11.9 2.7 7.1

57 2 -20.3 -13.9 -10.8 -1.2 2.9 6.3

3 -23.2 -11.8 -10.4 -0.4 2.4 6.2


201

Table E.4: The mV readings after 100 days of exposure.

Sample Depth range (mm)


Mix
#
0-6 6-13 13-19 19-25 25-32 32-38 38-45 45-51

1 -22.8 -1.3 17 32.2 51.6 48.5 65.3 57.7

46 2 -22.3 -1.2 18 34.1 54.2 49.4 68.4 59.8

3 -20.4 -1.1 16 35.3 57.3 49.6 68.5 59.9

1 -18.5 -5.5 8.2 17.9 31.2 37.9 46.1 46.8

56 2 -17.9 -6.8 8.8 18.6 31.3 38.8 48.5 50.1

3 -17.6 -4.2 9.4 19.2 31.9 40.5 52.4 53

1 -21.9 -3.6 13 20 22.7 33 25.3 26.9

66 2 -21.6 -3.3 13 20.4 26.1 33.4 26.2 27.2

3 -21.1 -2.9 13.7 21.8 26.9 33.8 26.4 28.3

1 -33.4 -13.7 6.1 35.2 50.5 66.7 73.3 51.3

55 2 -29.5 -13.4 6.6 35.4 54 67 74.5 61

3 -28.5 -11.3 7 37 54.7 68 75.7 64.5

1 -20.3 -7.4 3.6 9.3 22 23.2 41 -

57 2 -19.8 -6.3 4 10.2 22.7 26.3 41.2 -

3 -19.8 -6.1 4.7 10.4 23.6 27 41.8 -


202

Table E.5: The mV readings after 200 days of exposure.

Depth range (mm)


Mix Sample #
0-6 6-13 13-19 19-25 25-32 32-38 38-45 45-51

1 -25.7 -11.1 -5.5 3.2 11.8 24.7 56.6 63.3


46
2 -24.3 -11.1 -4 3.3 14.8 30.7 62.6 65.2

3 -23.3 -8 -2.9 4.6 15.7 32.8 64.1 68.2

1 -31.3 -10 4 19.6 25.5 25 21.4 -


56
2 -28.6 -9 4.5 30 26 26.4 21.6 -

3 -28.5 -8.1 5.2 20.1 27.1 26.9 22.4 -

1 -31.4 -16.6 -11.9 -8.3 -4.1 -2.5 -1.6 -


66
2 -31.1 -16.5 -10.9 -8 -2.6 -0.2 -1.4 -

3 -31 -15.6 -8.6 -7.9 -2.4 0.3 0.1 -

1 -32.3 -12.6 2.4 25.3 44.4 25.7 44.3 -


55
2 -29.8 -12.5 4.1 27.3 45.7 35.5 44.8 -

3 -27.5 -11.5 5 27.5 46.2 36.1 44.9 -

1 -23.8 -8.6 -2.5 1.7 5.4 11.7 15.3 -


57
2 -23 -6.7 -2.4 2 6.5 12.4 15.4 -

3 -22.7 -6.4 -1.7 2.2 6.9 12.7 16.1 -


203

Table E.6: The mV readings for the standard solutions.

mV reading
% Cl-
50-days 100-days 200-days
0.005 99.4 91.3 82.5
0.02 60.3 60.8 54.4
0.05 53.3 52.8 45.85
0.5 -2.2 -4.3 -7.5

1
Chloride conc. (% by weight)

0.1
50 days
100 days
200 days
0.01 200 days
100 days
50 days

0.001
-10 0 10 20 30 40 50 60 70 80 90 100
mV

Fig. E.5: Calibration curves for obtaining chloride profiles.


204

Table E.7: The mV readings after 30 days of exposure.

Initially dry Initially saturated


Depth range
(mm) sample number sample number
1 2 3 1 2 3

0-6 15.7 15.1 14.3 18.2 19.7 25.9

6-13 25.1 27.3 24.9 34.1 42.3 42.4

13-19 35.7 41.1 38 59.6 60.2 73.3

19-25 69.4 60.8 68.5 77.5 84.5 86.2

25-32 94.4 87.6 93.3 84.6 87.2 102.7

Table E.8: The mV readings after 90 days of exposure.

Initially dry Initially saturated


Depth range
(mm) sample number sample number
1 2 3 1 2 3

0-6 14.9 16.7 17.1 18.1 19.7 22.7

6-13 25.1 25.8 26.4 35 35.2 35.2

13-19 37.7 40.8 41 50.5 52.7 54.1

19-25 55.4 55.6 56.4 80.7 82.6 82.7

25-32 61.4 62.1 64.1 94.3 95.9 96.5


205

Table E.9: The mV readings after 180 days of exposure.

Initially dry Initially saturated


Depth range
(mm) sample number sample number
1 2 3 1 2 3

0-6 2.9 3.1 3.5 9.4 10.6 11.2

6-13 10.3 11.4 12.3 18.8 20.8 20.8

13-19 18.9 19.8 22.1 31 31.8 32.1

19-25 24.1 24.8 25.6 46.1 46.1 46.4

25-32 26.8 32 32.7 59.4 60.8 62.3

Table E.10: The mV readings for the standard solutions.

mV reading
% Cl-
30-days 90-days 180-days
0.005 103.4 104.2 87.1
0.02 76 81.2 57.6
0.05 52.5 59.3 50.1
0.5 -2 2.7 -5.3
206

Chloride conc. (% by weight)

0.1

30 days
90 days
180 days
0.01
180 days
90 days
30 days

0.001
-10 0 10 20 30 40 50 60 70 80 90 100
mV
Fig. E.6: Calibration curves for obtaining chloride concentrations.

Table E.11: The mV readings after 50 days of exposure.

Depth range Restrained shrinkage Free shrinkage

(mm) sample number sample number


1 2 3 1 2 3
0-6 -10.1 -8.2 -7.8 -11.7 -9.6 -7.4
6-13 0.3 11.6 13.9 2.2 3.4 5.2
13-19 8.2 8.6 16.1 10.6 11.8 15.4
19-25 16.5 16.6 16.7 21.1 21.6 21.7
25-32 24.1 26.3 28.8 26.7 27.3 34.5
32-38 25.7 31.1 35.3 31.1 32.4 34.5
207

Table E.12: The mV readings for the standard solutions.

% Cl- mV reading

0.005 71.9
0.02 60.3
0.05 53.4
0.5 -3.4

1
Chloride conc. (% by weight)

0.1

0.01

0.001
-10 0 10 20 30 40 50 60 70 80 90 100
mV

Fig. E.7: Calibration curve for obtaining chloride concentrations.

View publication stats

You might also like