Download as pdf or txt
Download as pdf or txt
You are on page 1of 279

Development of Test Methods for Assessment of

Concrete Durability for Use in


Performance-Based Specifications

By

Ahmad Shahroodi

A thesis submitted in conformity with the requirements for the degree of


Masters of Applied Science (M.A.Sc.)
Graduate Department of Civil Engineering
University of Toronto

Copyright by Ahmad Shahroodi (2010)

ABSTRACT

Development of Test Methods for Assessment of Concrete Durability for


Use in
Performance-Based Specifications

Ahmad Shahroodi
(M.A.Sc., 2010)

Department of Civil Engineering


University of Toronto

Many Ministry of Transportation of Ontario (MTO) projects consist of construction and


maintenance of reinforced concrete structures. Where appropriate test methods exist,
MTO has been moving towards use of performance-based specifications for durability
control of concrete. MTO currently uses ASTM C1202 (RCPT) coulomb values to assess
concrete durability. This test requires taking cores, so replacing this test with a faster nondestructive technique is important.
The main focus of this program was to study the Wenner probe surface resistivity as a
non-destructive test device and evaluate the potential for replacement of RCPT with the
Wenner resistivity.
This research program consists of the determination of RCPT values, water sorptivity
coefficients and electrical resistivities (bulk and surface) of nine concrete mixtures.

ii

In addition, the development of the Wenner probe instrument was studied. As well,
correlations between resistivity and ASTM C1202 and C1585 are provided followed by
technical recommendations for improving the Wenner test.

Keywords: Concrete durability; Permeability; Surface electrical resistivity; Water


sorptivity; Wenner probe; DC-cyclic bulk resistivity; RCPT coulombs passed

iii

ACK/OWLEDGEME/TS
A special note of thanks is extended to Professor R. D. Hooton for his guidance,
inspiration, and support throughout my thesis endeavour. Sincere appreciation is also due
to Ministry of Transportation of Ontario (MTO) for providing a large portion of funding
and supporting throughout the project and in particular Dr. B. Berszakiewicz.
For their help in performing all of the gruelling laboratory work at the University of
Toronto, Department of Civil Engineering, Materials Research section, recognition is also
due to Olga Perebatova and summer student Sean Hutchinson. Their kind help in this
research project is appreciated.
And last, but certainly not the least, the author wishes to thank his parents for their vital
positive energy, financial assistance, and encouragement. Also many thanks to my lovely
brother and sister. Your love and advice helped me survive the lonely life in Canada and
the long hours spent in the laboratory.

iv

TABLE OF CO/TE/TS
ABSTRACTii
ACKNOWLEGDMENTS......iv
LIST OF TABLES.........xii
LIST OF FIGURES..............xiv

1. INTRODUCTION...........1
1.1 Objective and scope of study...3

2. LITERATURE REVIEW............4
2.1 Influencing factors on concrete durability...4
2.1.1 External aggressive factors.....7
2.1.1.1 Environmental ...7
2.1.1.1.1 External chemical attack...7
2.1.1.1.1.1 Sulphates (sulphate attack)....8
2.1.1.1.1.2 Chloride corrosion.......12
2.1.1.1.1.3 Carbon dioxide and corrosion.....15
2.1.1.1.1.4 Acid Attack...16
2.1.1.1.2 External physical attack..17
2.1.1.1.2.1 Freezing and thawing damage.17
2.1.1.1.3 Environmental factors damaging internally....19
2.1.1.1.3.1 Alkali-silica reaction (ASR).....19
2.1.1.2 Operation and loading..20
2.1.1.2.1 Abrasion..20
2.1.1.2.2 Leaching..21
2.1.2 Internal structure of concrete22
2.1.2.1 Aggregate phase...22
2.1.2.2 Cement phase...23
2.1.2.2.1 W/CM ratio.23
2.1.2.2.2 Degree of hydration24
2.1.2.2.3 Curing.....24
v

2.1.2.2.4 Admixtures..24
2.1.2.2.5 Type of cement....25
2.1.2.2.6 Supplementary cementing materials (SCMs).25
2.1.3 Design and construction25
2.1.3.1 Design..26
2.1.3.2 Construction.27

2.2 Concrete durability measurement..28


2.2.1 Compressive strength test ....28
2.2.2 Rapid chloride permeability test (RCPT) ........28
2.2.2.1 First 5 minutes RCPT resistivity..31
2.2.2.2 Chloride ion penetration (chloride migration coefficient)...31
2.2.2.3 Linear extrapolation technique.33
2.2.2.4 Influencing factors on penetration resistance test33
2.2.2.4.1 Admixtures and RCPT values.33
2.2.2.4.2 Temperature and RCPT values...34
2.2.2.5 RCPT weak points ..34
2.2.3 Rate of water absorption (Sorptivity)....35
2.2.3.1 Calculation ..35
2.2.3.2 Laboratory sorptivity test ....36
2.2.3.3 Field sorptivity test..37
2.2.3.3.1 Field sorptivity apparatus overview37
2.2.3.3.1.1 Vacuum-attachment base plate38
2.2.3.3.1.2 Disc plate.39
2.2.3.4 Influencing factors on water sorptivity values ....41
2.2.3.4.1 Aggregate and the rate of water absorption....41
2.2.4 Electrical resistivity test41
2.2.4.1 Bulk electrical resistivity 44
2.2.4.2 Surface electrical resistivity.49
2.2.4.2.1 Surface disc.49
2.2.4.2.2 Four - probe line array (Wenner probe)..51

vi

2.2.4.2.3 Four - probe square array55


2.2.4.2.4 Application..57
2.2.4.3 Influencing factors on concrete electrical resistivity...58
2.2.4.3.1 Temperature and electrical resistivity.59
2.2.4.3.2 Chemical admixtures and electrical resistivity...60
2.2.4.3.3 Aggregate and electrical resistivity.60
2.2.4.3.4 Cement type and electrical resistivity.61
2.2.5 Influencing factors on durability tests...62
2.2.5.1 W/CM ratio .62
2.2.5.2 Supplementary cementitious materials64
2.2.5.3 Curing .66
2.2.5.4 Moisture content..67

3. EXPERIMENTAL PROGRAM ...69


3.1 Methodology..69
3.1.1 Compressive strength ...69
3.1.2 Rapid chloride permeability test ..69
3.1.3 Water sorptivity test......70
3.1.3.1 Laboratory sorptivity test 70
3.1.3.2 Field sorptivity test..71
3.1.3.2.1 Methodology...71
3.1.4 Electrical resistivity...75
3.1.4.1 Methodology75
3.2 Experimental project......78
3.2.1 Research project tests....78
3.2.1.1 Electrical resistivity of concrete...78
3.2.1.1.1 Surface electrical resistivity....78
3.2.1.1.1.1 RM MKII technical properties.....79
3.2.1.1.1.2 Surface electrical resistivity measurement..80
3.2.1.1.2 Cyclic-DC bulk electrical resistivity ..85
3.2.1.1.2.1 Cyclic-DC bulk resistivity of full length cylinders...85

vii

3.2.1.1.2.2 Cyclic-DC bulk resistivity of concrete discs....85


3.2.1.1.3 Rapid Chloride permeability test resistivity (first 5 minutes).86
3.2.1.2 Rapid chloride permeability test (RCPT).86
3.2.1.3 Rate of water absorption (water sorptivity test)...89
3.2.1.3.1 Laboratory sorptivity test89
3.2.1.3.2 Field sorptivity90
3.2.1.4 Compressive strength test (fc)90
3.2.2 Specimens.93
3.2.2.1 Concrete Cylinders...93
3.2.2.2 Concrete slabs..94
3.2.3 Materials...96
3.2.3.1 Cementitious materials ...96
3.2.3.2 Aggregates...97
3.2.3.2.1 Sieve analysis of fine and coarse aggregates..97
3.2.3.2.2 Physical properties of fine and coarse aggregates..99
3.2.3.3 Chemical admixtures.100
3.2.4 Mix design..102
3.2.5 Testing process....104
3.2.5.1 Concrete methodology...105
3.2.5.1.1 Material preparation..105
3.2.5.1.2 Mixing concrete....106
3.2.5.1.3 Fresh concrete properties..108
3.2.5.1.4 Casting concrete....108
3.2.5.1.5 Curing ..110

4. RESULTS AND DISCUSSION..111


4.1 Compressive strength.......111
4.1.1 Effects of changing W/CM ratio on compressive strength.....113
4.1.2 SCMs effects on compressive strength...........115
4.1.2.1 Comparison between compressive strength of SCMs concretes ..115

viii

4.2 Rapid chloride permeability test (RCPT).116


4.2.1 Total charge passed

....116

4.2.1.1 Effects of changing W/CM ratio on the RCPT coulombs.119


4.2.1.2 SCMs effects on RCPT values...120
4.2.2 RCPT electrical resistivity (first 5 min)..121
4.2.2.1 W/CM ratio effects on concrete resistivity123
4.2.2.2 Effects of adding SCMs on the RCPT electrical resistivity...125
4.2.3 Depth of chloride ion penetration (colorimetric method)...127
4.2.3.1 Effects of adding SCMs on the depth of chloride ion penetration.131
4.2.3.2 W/CM ratio effects on chloride ion penetration133
4.2.4 RCPT extrapolation....135

4.3 Rate of water absorption (sorptivity)...139


4.3.1 Laboratory sorptivity test139
4.3.1.1 SCMs effects on the lab sorptivity values..147
4.3.1.2 W/CM ratio effects on the lab sorptivity values ...148
4.3.2 Field sorptivity test.....149
4.3.3 SCMs effects on the rate of water absorption.155
4.3.4 W/CM ratio influences on the rate of water sorptivity coefficient.........156
4.3.5 Water sorptivity and degree of saturation...158
4.3.6 Calibration curves ......158

4.4 Electrical resistivity..164


4.4.1 DC-cyclic bulk resistivity (Monfore resistivity).165
4.4.1.1 DC-cyclic bulk resistivity of full length concrete cylinders..166
4.4.1.1.1 SCMs effects on the DC-cyclic bulk resistivity168
4.4.1.1.2 W/CM ratio effects on bulk resistivity..169
4.4.1.2 Bulk resistivity as an indicator for chloride penetrability..170
4.4.1.3 Bulk resistivity of concrete discs ..171
4.4.2 Surface electrical resistivity....173

ix

4.4.2.1 Surface electrical resistivity of concrete cylinders....173


4.4.2.1.1 Statistical analysis of concrete cylinders surface resistivity.177
4.4.2.2 Surface electrical resistivity of concrete slabs ..179
4.4.2.2.1 Circular slab number 1..180
4.4.2.2.2 Circular slab number 2..182
4.4.2.2.3 Statistical Analysis of concrete slabs surface resistivity...184
4.4.2.3 Unifying the surface resistivity values...188
4.4.2.4 W/CM ratio effects on surface electrical resistivity .....196
4.4.2.5 SCMs effects on surface electrical resistivity .......197
4.4.2.6 Effects of specimen geometry on the Wenner probe values......198
4.4.2.7 Surface electrical resistivity as an indicator for other properties of
concrete....202
4.4.2.7.1 Surface electrical resistivity and compressive strength....202
4.4.2.7.2 Surface electrical resistivity and total charge passed204
4.4.2.7.3 Surface electrical resistivity and the other types of electrical
resistivity..206
4.4.2.7.4 Surface resistivity and water sorptivity coefficient..209
4.4.2.7.5 Surface electrical resistivity and diffusion of chloride ion through
concrete....210
4.4.2.8 Wenner probe as a practical instrument.212

5. CONCLUSIONS AND RECOMMENDATIONS..215


5.1 Conclusions......215
5.2 Recommendations219

6. REFERENCES220

APPENDIX A: CONCRETES MIX DESIGN....232


APPENDIX B: COMPRESSIVE STRENGTH TEST................................241
APPENDIX C: LABORATORY SORPTIVITY TEST RESULTS...249
APPENDIX D: FIELD SORPTIVITY TEST RESULTS...250

APPENDIX E: BULK ELECTRICAL RESISTIVITY...251


APPENDIX F: SURFACE ELECTRICAL RESISTIVITY255

xi

LIST OF TABLES
TABLE

PAGE

Table 2.1: Adequate air content for different concretes.....19


Table 2.2: Chloride ion penetrability based on charge passed...30
Table 2.3: Electrical resistivity values for rebar corrosion rate..43
Table 2.4: Electrical resistivity of rocks.60

Table 3.1: Probe spacings used in the research project..85


Table 3.2: List of concrete cylinders for the project tests..93
Table 3.3: Durability tests for the concrete slabs...95
Table 3.4: Chemical composition of cementitious materials ....96
Table 3.5: Sieve analysis of fine aggregate ...97
Table 3.6: Sieve analysis of coarse aggregate ..98
Table 3.7: Sand sieve analysis required for the FM calculation...100
Table 3.8: Aggregates physical properties....100
Table 3.9: Basic mix design data......101
Table 3.10: Chemical admixture properties .....101
Table 3.11: Research program concrete mixes.102
Table 3.12: Concrete mixes proportions (mix design).....103
Table 3.13: Chemical admixture dosages ...104
Table 3.14: Fresh concrete properties...108

Table 4.1: Average compressive strength of different mixes at various ages..112


Table 4.2: Chloride ion penetrability based on charge passed.....116
Table 4.3: The RCP test results ...117
Table 4.4: The RCPT electrical resistivity values122
Table 4.5: Electrical resistivity values and rebar corrosion rate...122

xii

Table 4.6: Depth of chloride ion penetrated during the RCP test.128
Table 4.7: Non-steady-state migration coefficient of concrete mixes..129
Table 4.8: Extrapolated passing charges (coulombs).......135
Table 4.9: Effects of maximum anodic temperature on the RCPT coulomb values........137
Table 4.10: Average water sorptivity values of concrete discs142
Table 4.11: Water sorptivity coefficient of the circular concrete slabs....150
Table 4.12: Degree of saturation of the concrete circular slabs...152
Table 4.13: Bulk resistivity values of concrete cylinders.166
Table 4.14: The Monfore resistivity of 100 x 50 mm concrete discs....172
Table 4.15: Concrete cylinders apparent surface electrical resistivity values .173
Table 4.16: Statistical analysis of resistivity measurement of concrete cylinders (a= 50mm).177
Table 4.17: Average apparent surface electrical resistivity of circular concrete slabs.184
Table 4.18: Statistical analysis of resistivity measurement of concrete slabs (a= 50mm)...186
Table 4.19: Cell constant conversion factors for different probe spacings..190
Table 4.20: Surface resistivity of concrete cylinders with different probe spacings....191
Table 4.21: True surface electrical resistivity of concrete cylinders....192
Table 4.22: Surface resistivity of concrete slabs with different probe spacings..193
Table 4.23: True surface electrical resistivity of concrete slabs...194

xiii

LIST OF FIGURES
FIGURE

PAGE

Figure 2.1: Affecting factors on durability of concrete6


Figure 2.2: Sulfate-attacked fence posts located where a saline water table exists..9
Figure 2.3: Concrete suffering from thaumasite.....10
Figure 2.4: Physical sulphate attack mechanism11
Figure 2.5: Physical sulphate attack disintegrates concrete surface...11
Figure 2.6: Schematic description of corrosion process of a reinforcing steel in concrete...13
Figure 2.7: Cracking concrete by product of corrosion..13
Figure 2.8: Rebar corrosion decreases concrete loading capacity and strength.14
Figure 2.9: Concrete carbonation process .15
Figure 2.10: Concept of change in quantity of porosity in adequately cured carbonated concretes..16
Figure 2.11: Acid attack deteriorates concrete...17
Figure 2.12: Freeze-thaw damage in a concrete column caused rebar corrosion and more deterioration.18
Figure 2.13: ASR Mapping cracking and gel leakage....20
Figure 2.14: Surface abrasion due to the heavy traffic volume..21
Figure 2.15: Leaching and spalling under footway of a bridge..22
Figure 2.16: ASTM C1202-07 rapid chloride permeability test setup...29
Figure 2.17: Whitish chloride front migration in concrete sprayed after the RCP test..32
Figure 2.18: Field sorptivity apparatus (horizontal orientation).38
Figure 2.19: Vacuum- attachment base plate.39
Figure 2.20: Field sorptivity disc plate...40
Figure 2.21: Schematic view of reinforcing bar corrosion in concrete..43
Figure 2.22: Concrete bulk electrical resistivity test with two electrodes .45
Figure 2.23: Schematic diagram for the DC test setup ..46
Figure 2.24: Schematic diagram for the AC test setup ..48
Figure 2.25: Setup of one electrode (disc) measurement of concrete resistivity....50

xiv

Figure 2.26: Concrete cube with embedded electrodes..51


Figure 2.27: Schematic representation of four-electrode resistivity test....52
Figure 2.28: Cell constant correction factor for the centered end face configuration....54
Figure 2.29: Cell constant correction factor for the centred longitudinal measuring configuration..55
Figure 2.30: Four-probe square array principle .56
Figure 2.31: Four-probe square array schematic representation for studying crack parameters....56
Figure 2.32: Surface resistivity as a function of crack depth.58
Figure 2.33: Relationship between measured resistivity and air temperature ...59
Figure 2.34: Relation between resistivity and applied voltage of different cement concretes...61
Figure 2.35: Relation between electrical resistivity and W/CM ratio ...63
Figure 2.36: Relative reduction in diffusion coefficient with silica fume..65

Figure 3.1: Hose barbs attached to inner and outer chambers of a field sorptivity test apparatus.72
Figure 3.2: Schematic overview of the field sorptivity process.74
Figure 3.3: DC-cyclic bulk electrical resistivity test set up76
Figure 3.4: Wenner probe being used to measure surface resistivity of a concrete cylinder.77
Figure 3.5: RM MKII (surface resistivity-meter)...79
Figure 3.6: Schematic representation of four-electrode resistivity test.....79
Figure 3.7: Surface electrical resistivity measurement order for concrete cylinders.....81
Figure 3.8: Influencing factors for probe spacing in the Wenner resistivity..81
Figure 3.9: Effect of concrete section dimensions on surface resistivity measurement.82
Figure 3.10: Effect of edge and end proximity on surface resistivity measurement......83
Figure 3.11: Effect of maximum aggregate size on surface resistivity measurement84
Figure 3.12: DC-cyclic bulk resistivity of a concrete disc (test setup)...86
Figure 3.13: The RCP test setup.....87
Figure 3.14: Splitting concrete discs after The RCPT....88
Figure 3.15: AgNO3 solution appears the depth of chloride penetration...88
Figure 3.16: Laboratory water sorptivity test setup....89

xv

Figure 3.17: Field sorptivity test setup on a concrete slab.90


Figure 3.18: Research program concrete tests..91-92
Figure 3.19: Concrete cylinders testing layouts ....94
Figure 3.20: Fine aggregate grading curve....98
Figure 3.21: Coarse aggregate grading curve.99
Figure 3.22: Materials loading order107
Figure 3.23: Magnesium float for finishing the concrete slabs surface ...109
Figure 3.24: First 24 h concrete curing under plastic sheets and wet burlaps..109

Figure 4.1: Crushing a concrete cylinder during compressive strength test.111


Figure 4.2: Compressive strength of concrete mixtures at various ages..112
Figure 4.3: Compressive strength of SFSL concrete mixes.113
Figure 4.4: Compressive strength of mixes contain slag.114
Figure 4.5: Concrete strength of plain cement concretes.114
Figure 4.6: Comparison between compressive strength of ternary and binary concrete mixes ..115
Figure 4.7: The RCPT coulomb values with age..117
Figure 4.8: Total coulombs passed (silica fume and slag concrete mixes)..118
Figure 4.9: Total coulombs passed (slag concrete mixes)....118
Figure 4.10: Total coulombs passed (plain cement concrete mixes) .......119
Figure 4.11: Comparison between total charge passed through concrete mixes..120
Figure 4.12: First 5 minutes RCPT electrical resistivity .123
Figure 4.13: Effect of changing W/CM ratio on the RCPT electrical resistivity of ternary mixes..124
Figure 4.14: Effect of changing W/CM ratio on the RCPT electrical resistivity of binary mixes...125
Figure 4.15: Silica fume effects on the first 5 min. RCPT electrical resistivity...126
Figure 4.16: The effects of adding slag on the first 5 min. RCPT electrical resistivity...127
Figure 4.17: Splitting a concrete disc after the RCP test..128
Figure 4.18: Relation between the RCPT passing charges and the chloride migration coefficient.131
Figure 4.19: Depth of chloride ion penetrated into concrete specimens during the RCP test..132

xvi

Figure 4.20: Zero depth of chloride ion penetration (ternary concrete)...133


Figure 4.21: W/CM ratio effects on the depth of chloride ion penetration during the RCP test......134
Figure 4.22: RCPT recorded and extrapolated charges passed (day 7)136
Figure 4.23: RCPT recorded and extrapolated charges passed (day 28)..136
Figure 4.24: Relation between the extrapolated and recorded passing charges...138
Figure 4.25: Rate of water absorption of top slices..140
Figure 4.26: Rate of water absorption of middle slices141
Figure 4.27: Rate of water absorption of bottom slices142
Figure 4.28: Extracting cores from the concrete rectangular slabs .143
Figure 4.29: Rate of water absorption of concrete discs extracted from slabs..143-144
Figure 4.30: Comparison between water sorptivity of concrete discs sliced from concrete cylinders and
concrete cores extracted from concrete slabs...145
Figure 4.31: Water sorptivity and chloride migration coefficient146
Figure 4.32: Relation between the lab sorptivity test values and concrete compressive strength147
Figure 4.33: Average rate of water absorption of concrete mixes148
Figure 4.34: Field sorptivity apparatus (horizontal orientation)...149
Figure 4.35: Water sorptivity coefficients of concrete circular slabs ..150
Figure 4.36: Middle piece taken from a cored rectangular slab for moisture content measurement...151
Figure 4.37: Degree of saturation of the concrete slabs at different ages153
Figure 4.38: Moisture content effects on the rate of water absorption.154
Figure 4.39: SCMs effects on the field sorptivity test results..155
Figure 4.40: W/CM ratio effects on the water sorptivity coefficients of plain cement concrete.156
Figure 4.41: W/CM ratio effects on the sorptivity coefficients of concrete mixes containing slag.157
Figure 4.42: W/CM ratio effects on the water sorptivity coefficients of the ternary mixes157
Figure 4.43: Water sorptivity- saturation calibrating curve for HPC mix ...159
Figure 4.44: Water sorptivity- saturation calibrating curve for HPC+ mix .159
Figure 4.45: Water sorptivity- saturation calibrating curve for SFSL 0.40 mix ..160
Figure 4.46: Water sorptivity- saturation calibrating curve for PCSL 0.40 mix .160

xvii

Figure 4.47: Water sorptivity- saturation calibrating curve for PCSL 0.40+ mix ...161
Figure 4.48: Water sorptivity- saturation calibrating curve for PCSL 0.45 mix .161
Figure 4.49: Water sorptivity- saturation calibrating curve for PCSL 0.45+ mix ...162
Figure 4.50: Water sorptivity- saturation calibrating curve for PC 0.45 mix ..162
Figure 4.51: Water sorptivity- saturation calibrating curve for PC 0.45+ mix.....163
Figure 4.52: Drying circular concrete slabs..164
Figure 4.53: Concrete bulk resistivity test with two steel electrodes...165
Figure 4.54: DC-cyclic bulk resistivity of 100 x 200 mm concrete cylinders...167
Figure 4.55: SCMs effects on the DC-cyclic bulk resistivity values168
Figure 4.56: W/CM ratio effects on the Monfore resistivity values.....170
Figure 4.57: Relation between chloride migration coefficient and the DC-cyclic bulk resistivity with the
RCPT 5 min. resistivity171
Figure 4.58: DC-cyclic bulk resistivity of 100 x 50 mm concrete discs...172
Figure 4.59: Surface electrical resistivity of concrete cylinders (25 mm probe spacing)....175
Figure 4.60: Surface electrical resistivity of concrete cylinders (30 mm probe spacing)175
Figure 4.61: Surface electrical resistivity of concrete cylinders (40 mm probe spacing)....176
Figure 4.62: Surface electrical resistivity of concrete cylinders (50 mm probe spacing)176
Figure 4.63: Surface resistivity measuring order for concrete circular slabs...179
Figure 4.64: Surface electrical resistivity of concrete slabs labelled number 1 (a= 20 mm) ..180
Figure 4.65: Surface electrical resistivity of concrete slabs labelled number 1 (a= 30 mm)...181
Figure 4.66: Surface electrical resistivity of concrete slabs labelled number 1 (a=40 mm)181
Figure 4.67: Surface electrical resistivity of concrete slabs labelled number 1 (a=50 mm)....182
Figure 4.68: Surface electrical resistivity of concrete slabs labelled number 2 (a=20 mm)182
Figure 4.69: Surface electrical resistivity of concrete slabs labelled number 2 (a=30 mm)....183
Figure 4.70: Surface electrical resistivity of concrete slabs labelled number 2 (a=40 mm)183
Figure 4.71: Surface electrical resistivity of concrete slabs labelled number 2 (a=50 mm)....184
Figure 4.72: Cell constant correction factor for specimen used ..189
Figure 4.73: Required limitations for calculating the optimum probe spacing....195

xviii

Figure 4.74: W/CM ratio effects on the surface electrical resistivity of concrete cylinders ...196
Figure 4.75: SCMs effects on the surface electrical resistivity of concrete cylinders..197
Figure 4.76: SCMs effects on the surface electrical resistivity of circular slabs..197
Figure 4.77: Relation between different specimens resistivity with the optimum probe spacing....199
Figure 4.78: Comparison between specimens resistivity values and different probe spacings....200
Figure 4.79: Probe spacings effect on the penetration depth of the Wenner applied current ..201
Figure 4.80: Surface resistivity of concrete cylinders versus compressive strength....203
Figure 4.81: Surface resistivity of concrete slabs versus compressive strength...203
Figure 4.82: Surface resistivity of concrete cylinders versus the RCPT passing charge.204
Figure 4.83: Modified surface resistivity versus the RCPT results based on MTO specification...205
Figure 4.84: Surface resistivity of concrete slabs versus the RCPT passing charge206
Figure 4.85: Surface resistivity of concrete cylinders versus the RCPT resistivity.207
Figure 4.86: Surface resistivity of concrete slabs versus the RCPT resistivity207
Figure 4.87: Surface resistivity of fully saturated cylinders versus Monfore resistivity..208
Figure 4.88: Surface resistivity of not fully saturated slabs versus Monfore resistivity..208
Figure 4.89: Surface resistivity of concrete cylinders and water sorptivity coefficient...210
Figure 4.90: Surface resistivity of concrete cylinders versus chloride migration coefficient..211
Figure 4.91: Surface resistivity of concrete slabs versus chloride migration coefficient.211
Figure 4.92: Practical relation between the surface resistivity values and other tests values...213
Figure 4.93: Regression analysis of compressive strength and surface electrical resistivity...214

xix

CHAPTER 1
I/TRODUCTIO/
Many Ministry of Transportation of Ontario (MTO) projects consist of construction and
maintenance of reinforced concrete bridge structures. Where appropriate test methods
exist, the Ministry has been moving towards use of performance-based specifications for
construction acceptance and durability control of concrete.
Two primary characteristics of concrete, strength and durability, need to be assessed to
obtain performance of concrete structures.
Concrete strength is measured and controlled by standard testing techniques and
guidelines but the lack of durability assessment standards is an issue. A durable concrete
must retain its original form, quality, and serviceability under its working environmental
conditions.
Generally, concrete durability is affected by five factors:
1) Design: type of materials, materials conditions and proportions, concrete mix
design, and thickness of concrete cover over reinforcing steel
2) Construction practices: mixing, delivering, discharging, consolidating, finishing,
and curing conditions
3) Hardened concrete properties: compressive strength, permeability
4) Environmental exposure conditions: sulphate attack, freeze-thaw, alkali-silica
reaction
5) Loading conditions: type of loading, loading duration, crack width and depth.

Concrete design and construction practices are controlled by standard guidelines. Also
the loading conditions are considered during the design process. The major issue needed
to be studied for preparing durable concretes is the hardened concrete durability.
Permeability is the most influencing factor on the durability and service life of reinforced
concrete members because movement of aggressive fluids from the surrounding
environment into concrete is the main cause for most concrete deteriorations. In other
words, developing an impermeable pore system is necessary to produce a durable
concrete.

The permeability is currently estimated by the ASTM C1202, permeability index test.
The testing method requires the removal of cores (a destructive technique) and
preparation of samples for testing purposes which is expensive and time consuming.
Therefore providing suitable and time-saving non-destructive testing methods to assess
concrete performance as measured on hardened concrete in place would be important for
the Ministry.
Some non-destructive tests exist for concrete strength (e.g. Schmidt hammer), but there is
not any direct non-destructive test for measuring the permeability of in-place concrete.
Therefore a fast and simple non-destructive test has to be developed to evaluate the
durability of concrete. A geometry-independent test is the electrical resistivity test. The
electrical resistivity test can be used as a non-destructive alternative to the current
permeability index test. Initially the resistivity test was used on concrete cores and
cylinders taken from existing structures (destructive). Later surface electrical resistivity
measurement techniques which had been used by geologist were developed for concrete
structures. Pre-planned embedment electrodes were required for this test method, so the
method was still destructive and time-consuming. Recently surface electrical resistivity is
measured by an instrument called a Wenner probe without embedding any electrode into
concrete. This test can be done quickly on surfaces both as is and after saturation with
water, since moisture conditions affect surface resistivity results.
The Wenner probe measures the electrical properties of covercrete concrete which is
different from that of the bulk concrete due to compaction, bleeding, finishing, and
curing.
In general, durability of covercrete is improved by obtaining a discontinuous capillary
pore structure which is caused by using lower water-to-cementitious ratio (W/CM), using
supplementary cementitious materials, and applying adequate moist curing. A durable
covercrete concrete is necessary to achieve a long service life of concrete structures in a
severe environment, so surface properties were studied in this research project.

3
1.2 Objective and scope of study
The objective of this research consists of testing concrete samples with a range of mix
designs to evaluate the surface electrical resistivity test. In addition, the concrete samples
were tested in the rapid chloride permeability test (RCPT), and rate of water absorption
due to capillary suction, referred to as sorptivity (Hooton, Mesic and Beal (1993) in
DeSouza et al., 1996), in order to correlate the results.
Surface electrical resistivity and water absorption tests can be used on cores and cylinders
extracted from structures, but they also have potential as non-destructive tests for use on
constructed surfaces. Laboratory types of electrical resistivity and water absorption are
the Monfore resistivity test and the ASTM C1585 water sorptivity test (named as
laboratory sorptivity test), respectively. Non-destructive types of electrical resistivity and
water absorption are the Wenner probe and field sorptivity testing technique (DeSouza
and Hooton, 1996). Therefore both destructive and non-destructive forms of these two
tests in addition to the RCPT were used in this research program.

The study begins with a literature review considering concrete durability aspects,
concrete deterioration mechanisms, and environmental effects on surface concrete. A
brief description of testing methods, advantages and limitations, and the mechanisms of
the instruments are included. In the next chapter a detailed methodology for all tests
applied in this project followed by in-depth project plan including experimental variables
are described. Data presentation and analyses of nine series of test cores, cylinders, and
slabs are presented to quantify the applicability of the Wenner probe. Finally practical
correlations between the Wenner probe values and the other standard permeability tests,
conclusions from the study, and recommendations are presented.

CHAPTER 2
LITERATURE REVIEW
Two major properties of concrete are Strength and Durability.
Strength is the ability of concrete to resist stress (compressive, tensile, shear, or torsion).
Concrete strength is a function of many factors as well as concrete mix design, curing
process, using reinforcing steels, and etc. On the other hand, durability is another
important property of concrete.
In this research, measuring the durability of concrete was the main concern.
The American Concrete Institute in ACI 116 R defines durability of concrete as its ability
to resist weathering action, chemical attack, abrasion, and other conditions of service.
Also durability of concrete is defined as the capability of concrete by itself of keeping the
original properties for a certain period (Collepardi, 2000). Durable concrete may be
defined as concrete that keeps its original quality, form, and serviceability when it is
exposed to its environment (Savas, 1999).
Many factors affect concrete durability either directly or indirectly.

2.1 Influencing factors on concrete durability


The most important characteristic of concrete that is believed to be affecting its
durability is permeability (better to use penetrability as it is not mechanism specific) of
concrete (Baykal, 2000 in Chini, 2003). There is an approximate inverse relationship
between concrete penetrability and compressive strength as well as durability. But it is
worth noting that durability of concrete is not necessarily related to the compressive
strength of concrete.
Penetrability of concrete can be determined by measuring the rate of fluids (oxygen,
water, and chloride ions) penetration into concrete to reach a certain level, for example
level of steel bars because most of the types of deterioration are influenced by fluid
ingress (or movement) in concrete. It is not just capillary action that causes a given

5
specimen to absorb fluids, as an aggressive fluid can be transported through concrete pore
structure by various mechanisms as mentioned below:
a)

Permeability ( due to an external pressure head P, both side of the test are

saturated, so Darcys law can be applied)


b)

Sorptivity (absorbing fluid by an unsaturated pore system due to the capillary

force, much like a sponge)


c)

Vapour diffusion (due to the humidity gradient: an equilibrium state of saturation

causes liquid in a specimen to move from one saturation gradient to another).


d)

Ionic diffusion (due to the ionic such as. Cl- ion gradient)

The main purpose of this research was to evaluating and improving concrete durability,
so studying the factors governs concrete durability is necessary. All these factors
(summarized in Figure 2.1) influence concrete durability by affecting concrete
penetrability.

Affecting
Factors on
Concrete
Durability

Internal
Structure of
Concrete

External
Aggressive
Factors

Environmental

Operation
and Loading

External
Chemical
Attacks

External
Physical
Attacks

Environmental
Factors
Damaging
Internally

Design
and
Construction

(Physical Properties)

Aggregate
Phase

Design

Construction
Practices

Concrete
Design

Pore structure
of
Cement Paste

Reinforcement
Design

Physical
Characteristic

Pore
Solution

Figure 2.1: Affecting factors on durability of concrete

Penetrability is important for the MTO because a) the Ministry builds, manages, and
maintains bridges, pavements, walls, and piers exposed to moisture and chloride, and b)
all forms of deterioration directly or indirectly result from movement of fluids in
concrete.

7
Fluid movement can be a result of permeability, sorptivity, ionic or vapour diffusion.
Appearance of any of this penetrability types results in a permeable concrete. In addition,
deteriorating factors influence concrete durability only if concrete is permeable.
Therefore, it is necessary to consider concrete penetrability during studying the effects of
durability influencing factors.

2.1.1 External aggressive factors


External factors influence concrete durability and deteriorate concrete. Generally,
external deteriorative factors may be environmental factors, or results of facility
operation and maintenance.

2.1.1.1 Environmental
When a concrete member is exposed to an environment, the surrounding environmental
factors can affect concrete durability. Environmental factors cause several types of
deterioration which can affect the durability of concrete. In most of the cases, the
concrete degradation process involves the penetration and subsequent movement of
water, air, or other fluids which are transporting aggressive agents into concrete pore
system (Bryant et al., 2009).
Environmental factors must be studied before designing, mixing, casting and operating
any concrete structure.
Environmental damaging factors can be categorized as external factors attacks concrete
chemically (e.g. sulphate), external factors attacks concrete physically (e.g. freeze-thaw
damage), and environmental factors damaging concrete internally.

2.1.1.1.1 External chemical attack


Concrete structures are exposed to many varying chemicals. These environmental
chemicals attack concrete and deteriorate the concrete physically. Common chemicals
which can attack concrete are discussed in the following sections.

8
2.1.1.1.1.1 Sulphates (sulphate attack)
The major source for sulphate is soil, sea water, sewers, and some chemical operations.
The common sulphates forms which attack concrete are calcium sulfate (CaSO4), sodium
sulfate (NaSO4), and magnesium sulfate (MgSO4), and potassium sulfate (KSO4).
In all sulphate attack, formation of large ettringite crystals (because of the large number
of available water molecules) results tensile force because concrete is hardened at the
time of sulphate attack. Concrete cracks and damages when these forces become higher
than concrete tensile strength. Beside losing strength and loading capacity, cracked
concrete is vulnerable to the other aggressive external factors.
Generally three types of sulphate attack which can influence the durability of concrete
have been found:

I) Classic sulphate attack (chemical): This type of sulphate attack forms in two steps:
Sulphate attacks calcium hydroxide which is available in concrete as a hydration product
and gypsum is formed:
Ca(OH)2 + SO4= CaSO4.2H2O (Gypsum)
Gypsum formation is a not deteriorating process; however formation of ettringite crystals
during the second step damages the concrete (Hooton, 2007):
3CaSO4.2H2O + C3A + 25H20 3CaSO4.C3A.31H2O
This type of sulphate attack is associated with sea water and sulphate sources such as soil.

Figure 2.2: Sulfate-attacked fence posts located where a saline water table exists (St. John, 1998)
(http://filer.case.edu/slr21/Bridge/sulfate.htm)

Using low C3A cements (Types MS and HS), adding supplementary cementing materials
(SCMs) such as slag, fly ash, or silica fume, using air entraining admixture, and low
water-to-cement (W/CM) ratio can help concrete resist sulphate attack.

II) Thaumasite sulphate attack: Sulphate ions, generally from groundwater, react with
calcium silicate hydrate (C-S-H) and carbonate to form thaumasite (Clark, 2007), so
thaumasite is a less common type of sulphate attack. It can be categorized as internal and
external thaumasite. The chemical reaction and required conditions for thaumasite attack
is not fully known however in both types, low temperature (15C and particularly at
approx. 0-5C (Bensted, 1999)), presence of moisture, and presence of carbon dioxide are
necessary. In external sulphate attack, these prerequisites are available in the surrounding
environment, while in internal sulfate attack; presence of a source of readily soluble
CaCO3 in cement paste causes thaumasite. In both types, a silica-bearing sulphate
compound

attacks

the

C-S-H

matrix

of

the

cement

paste

(either

CaO.SiO2.CaSO4.CaCO3.15H2O or CaSiO3.CaCO3.CaSO4.15H2O) (Neville, 1995). C-S-

10
H matrix is completely replaced by a white mushy incohesive mass as shown in Figure
2.3.

Figure 2.3: Concrete suffering from thaumasite (Bickley and Hooton, 2001)

Sulphate resisting cement does not prevent thaumasite because in this type of sulphate
attack, C-S-H is attacked, not C3A. Provision of low-permeable concrete (by low W/CM
ratio, SCMs, and adequate curing period) would control thaumasite attack (Hooton, 2007)
because in low permeable concretes, ingress of sulphate and carbonate ions is controlled.

III) Physical sulphate attack: Free water in concrete contains many ions including
sulphates (commonly dissolved Na2SO4). This water is drawn up through a continuous
capillary pores structure by wick action. If drawn water evaporates near a surface due to
arid weather, dissolved salts deposit near drying surface as shown in Figure 2.4.

11

Figure 2.4: Physical sulphate attack mechanism

The most common efflorescent material is anhydrous sodium sulfate (Na2SO4,


Thenardite) and hydrous sodium sulfate (Na2SO4.10H2O, Mirabilite or Glauber's salt)
(Thaulow and Sahu, 2004).
More salt concentration causes surface scaling similar to freezing and thawing damage as
shown in Figure 2.5.

Figure 2.5: Physical sulphate attack disintegrates concrete surface


(http://img638.imageshack.us/i/saltdeposits.jpg/)

12

There are three major explanations for efflorescent damage in concrete:


1) Volume change from Thenardite to Mirabilite may cause pressure (Hime, 2001 in
Thaulow and Sahu, 2004).
2) Mirabilite is unstable and quickly dehydrates to Thenardite in dry air (Thaulow
and Sahu, 2004). Some scientists have believed that even Thenardite hydration
causes damaging pressure (Evans, 1970 in Thaulow and Sahu, 2004).
3) Damage occurs whether or not Thenardite was produced previously by Mirabilite
decomposition. Together with recent results from the literature, these results
indicate that damage occurs because Thenardite dissolution can produce solutions
highly supersaturated with respect to Mirabilite, so that precipitation of this
mineral can lead to large crystallization pressures (Tsui et al., 2003).
It has been reported that the first two hypotheses are rejected while crystallization
pressure created by salts growing from as supersaturate solution is reported as the actual
damaging mechanism (Thaulow and Sahu, 2004).
To control the sulphate attack, lower W/CM ratio, adequate curing, use of pozzolans, and
use of air-entraining admixture (resultant ettringite crystals can grow in air bubbles) are
necessary. Among all this factors, lower W/CM ratio is the most effective item because it
can reduce the rate of moisture movement in concrete.

2.1.1.1.1.2 Chloride corrosion


Although sea water contains sulphates, it is also a source of chlorides as well as saline
groundwater and de-icing salts used on road and bridges in Canada. Due to chloride
penetration into the reinforced concrete, the de-passive layer around the reinforcing bars,
which protects the bars from corrosion, is destroyed. Steel bars are normally corrosion
protected when embedded in high pH concrete.
The area which receives Cl ion becomes an Anode and other area of the bar which
receives oxygen and water becomes a Cathode. Water and oxygen availability is
necessary for corrosion of depassivated bars.

13

Figure 2.6: Schematic description of corrosion process of a reinforcing steel in concrete


(http://www.cement.org/tech/cct_dur_corrosion.asp)

As shown in Figure 2.6, electrical current flows from anode to cathode through the rebar
and transfers electrons from anode to cathode and takes OH from cathode to anode
through the permeable concrete. As a corrosion electrical circuit forms, iron hydroxide,
Fe (OH) 2, will form as the product of corrosion at the anode.
Iron oxide and hydroxide (two major corrosion products) occupy more space as shown in
Figure 2.7. They cause tensile stress and subsequently cracking and spalling of concrete
happens at the level of reinforcement as shown schematically in Figure 2.7.

Figure 2.7: Cracking concrete by product of corrosion


(http://www.corrosion-club.com/concretecorrosion.htm)

14
More chloride, water, and oxygen get into cracked concrete to accelerate the corrosion
process and deteriorate concrete as shown in Figure 2.8.

Figure 2.8: Rebar corrosion decreases concrete loading capacity and strength
(http://www.cp-tech.co.uk/img28.jpg)

Since the hydroxide ion passes through concrete from Cathode to Anode, increasing the
electrical resistivity of concrete can break the corrosion electrical circuit. Concrete
electrical resistivity is a function of concrete moisture content, pore stricture connectivity,
and ionic content of pore solution (Savas 1999, in Chini 2003). Concretes in dry
environment are corrosion resistant because of their high electrical resistivity.
It is worth noting that due to the destructive effects of chloride on reinforced concrete
structures, accelerating admixtures containing calcium chloride (CaCl2) is no longer
allowed in reinforced or pre-stressed concrete structures.

15
2.1.1.1.1.3 Carbon dioxide and corrosion
In normal concrete, steel bars are chemically protected from corrosion by the alkaline
nature of concrete as concrete high alkalinity causes the formation of a passive oxide film
around the steel reinforcement, which is stable in high pH (Neville, 1995). When carbon
dioxide (CO2) of the air penetrates the pore structure of concrete, calcium hydroxide
(common among other hydration products because of its reactivity) carbonates if
moisture is in the pores and carbonic acid forms (Monkman and Shao, 2006). Carbonic
acid reduces the pH of concrete which was initially about 13-14 because of the cement
hydration alkali products. Calcium carbonate is the product of the reaction as shown in
Figure 2.9.

Figure 2.9: Concrete carbonation process

Carbonation reaction reduces the pH of the pore solution to 8-9, at which level the
passive film (oxide film) around the steel bars is not stable. Therefore, corrosion happens
not from chloride ion penetration, but from carbonation. Moisture is necessary to form
carbonation. It is important to mention that carbonation happens in 50% - 70% concrete
relative humidity (Neville, 1995).

16
pH reduction in concrete not only depassivates the reinforcing bars, but also breaks down
the C-S-H matrix into calcite and silica gel (Ishida and Li, 2008). It results in less
porosity (shown schematically in Figure 2.10) and higher electrical resistivity (Polder,
2000).
Before Carbonation

After Carbonation

Ca(OH)2
CaCO3
CaCO3

C-S-H
Gel

Silica Gel

Void

Void

Figure 2.10: Concept of change in quantity of porosity in adequately cured carbonated concretes
(Ishida and Li, 2008)

Although the concrete porosity reduces by carbonation (Monkman and Shao, 2006),
carbonation major negative effect (rebar corrosion) is more dominant (Bertolini et al.,
2008). Increasing the depth of covercrete, using surface coating and coated rebars,
adequate curing, and casting good quality and low permeability concrete can reduce
chloride and carbon dioxide ingress into concrete (Neville, 1995).
2.1.1.1.1.4 Acid Attack
Acid attack usually happens at the crown of sewers where the hydrogen sulfide, H2S, is
rising from anaerobic bacteria in sewage. Hydrogen sulphide reacts with aerobic bacteria
located on the inner crown surface, so sulphuric acid is produced:

H2S +2 O2 H2SO4
The produced sulphuric acid dissolves the C-S-H matrix of concrete (Kosmatka et al.,
2002). Concrete is an alkaline substance, so many of its components readily react with
acids.

17

Figure 2.11: Acid attack deteriorates concrete (http://filer.case.edu/slr21/Bridge/acid1.jpg)

Keeping a low W/CM ratio will increase the resistance of the concrete to deterioration by
acid because low permeability, a result of the low water/cement ratio keeps the acidic
solution out of the concrete pore structure (Kosmatka et al., 2002). Surface coating by an
acid proof layer is a common industrial acid proofing.
2.1.1.1.2 External physical attack
Concrete structures are exposed to many physical deteriorating factors. The most
important physical environmental factors are temperature and humidity. The common
physical damage in concrete is caused by cold temperatures:

2.1.1.1.2.1 Freezing and thawing damage


This type of damage happens all over Ontario because of its cold weather. Upon freezing
of the internal water, it expands by 9% in volume (Neville, 1995) and if there is no room
for the expansion, tensile stress will be applied to the solid surrounding the freezing
water. Therefore, the critical saturation level of pores is about 91% if the pores are
equally saturated. The generated stress by ice formation breaks the paste results in
aggregate-paste separation and cracks; then ice melts and contracts as it thaws (Neville,
1995). Many freeze-thaw cycles crack internal concrete and concrete surface resulting in

18
more concrete deterioration. Cracked concrete has lower strength and higher permeability
resulting in more penetration of chloride. Therefore, other deteriorating damages start to
form in concrete as shown in Figure 2.12. This figure was taken from a deteriorated
bridge column located at Finch Ave. East, Toronto, Ontario, Canada.

Figure 2.12: Freeze-thaw damage in a concrete column caused rebar corrosion and more
deterioration

The damaged region is covered by snow shovelled to the road side during the winter.
Permeable concrete froze and deteriorated after several freeze-thaw cycles. Chloride ion
from the melting salt used by the Toronto city penetrated into the cracked concrete and
destroyed the rebars passive layer. Steel corrosion started so concrete was more damaged
by the corrosion expanding products.
If there is enough room for ice expansion, tensile forces generated by ice formation are
not applied to the solid body of cement paste. This additional room can be provided by
air-entraining admixture. This admixture creates spherical air bubbles (with diameters
about 50 m (Neville, 1995)) in concrete during mixing. The air-bubbles should be well
distributed and have a distance between each other of less than 200 m in the cement

19
paste (Neville, 1995) after the concrete has hardened. During ice passage into the air
bubbles, some stresses are developed, so concrete needs to have a minimum strength
before exposed to the first freeze-thaw cycle. Neville (1995) recommended that concrete
should not be exposed to the cold weather until its strength has reached 24 MPa.
The total air content and the spacing factors are important properties of air-entrainment.
Table 2.1 contains the total air content and average spacing factors for different
aggregates (CSA A23.1, 2009).
Table 2.1: Adequate air content for different concretes (CSA A23.1, 2009)
Maximum Aggregate Size (mm)

Air Content (%)

Spacing Factor, L (m)

10

6-9

<230

14 - 20

5-8

<230

Also concretes with low W/CM ratio are more freeze-thaw resistant because of their
lower permeability which is enough to prevent critical saturation of the pores, less
internal freezable water, and smaller volume of capillary pores (Neville, 1995).

2.1.1.1.3 Environmental factors damaging internally


In some cases, environmental factors do not attack concrete directly, but their presence in
concrete mixture can cause later physical deterioration caused by internal chemical
reactions. These factors get into concrete during the mixing period. The most common
type of concrete damaging, caused by the presence of reactive environmental factors, is
alkali-silica attack.

2.1.1.1.3.1 Alkali-silica reaction (ASR)


The main alkali sources in cement paste are calcium, sodium, and potassium hydroxide
(from Ca(OH)2, NaOH and KOH). These alkalis (mainly hydroxide ion, OH) attack
various types of micro-crystalline and the silica bonds of active aggregates provided that
the required alkali content, 300 mM/L (Hooton, 2007), is available. . Finally alkali-silica
gel forms which invites water and then swells mainly around aggregates. The swelling
gel creates tensile stress to hardened concrete, so concrete cracks. Sign of this type of
deterioration is known as mapping cracks (because the alkalies and reactive silica

20
particles are well distributed through concrete), leaking of the gel from cracks and joints,
and joint closure due to the gel swelling.

Figure 2.13: ASR Mapping cracking and gel leakage


(http://www.todaysconcretetechnology.com/wp-content/uploads/2009/10/ASR1.jpg)

The major solutions for ASR are a) Using low-alkali cement, b) Limit the cement content
to restrict the alkali loading of concrete (Neville, 1995), c) Avoid using of reactive
aggregate (can be recognised by spectrographic analysis or physical long-term and shortterm tests), and d) Since bonding larger quantities of alkalies improves ASR resistance,
using SCMs causing more C-S-H formation is another practical solution (Neville, 1995).

2.1.1.2 Operation and loading


The maintenance loads are considered during designing a structure, so they do not affect
concrete durability unless concrete has deteriorated or cracked. In case of loading,
unexpected impact loads are the most common affecting factor. The latest standards have
considered the loads might be applied during a structures service life. Therefore, loading
is not a major factor affecting concrete durability in modern structures. On the other
hand, the most common operational stresses, affecting the durability of concrete are
surface acid attack (Section 2.1.1.1.1.4), abrasion, and leaching.

21
2.1.1.2.1 Abrasion
Generally friction between a substance and concrete surface as well as constant rolling of
machinery and equipment, traffic, or water flow causes surface abrasion and cover loss.
Due to abrasion, concrete cover above the reinforcing bars is reduced, so concrete
durability is reduced as shown in Figure 2.14 taken on the University of Toronto campus.

Figure 2.14: Surface abrasion due to the heavy traffic volume

As cover concrete thickness reduces, aggressive ions can reach the rebar, so concrete
durability decreases. Surface abrasion resistance is directly proportional to the concrete
strength, strength of aggregate, and density of concrete (Forster, 2000), so factor
improving strength and density such as adequate curing, long water curing, and using
SCMs improve abrasion wear resistance.

2.1.1.2.2 Leaching
In case of water leaching from structural joints, or structure exposure to water flow
cement hydration products (Ca(OH)2 and C-S-H) may start to dissolve, so concrete
becomes soft and etched away. In this type of operation deterioration, lime dissolves in
water at the surface exposed to water and in some cases calcium carbonate is seen on the
other side of concrete as shown in Figure 2.15.

22

Figure 2.15: Leaching and spalling under footway of a bridge (water on top of the footway)
(http://www.rta.nsw.gov.au/cgi-bin/index.cgi?action=heritage.show&id=4301049)

Salt crystallization deteriorates concrete due to the crystallization pressure as mentioned


in Section 2.1.1.1.1.1 (part III). Beside adequate water drainage system, reducing the
permeability of concrete protects concrete against leaching.

2.1.2 Internal structure of concrete


Cement paste phase and aggregate phase (fine and coarse aggregates) make the concrete
internal structure. Properties of both aggregate and cement part of concrete affect
concrete durability.

2.1.2.1 Aggregate phase


Aggregate properties are physical (such as density, water absorption, moisture content,
and grading) and chemical (e.g. silica content). These properties which can be measured
by standard tests must be controlled before the designing process. They change concrete
mix design, setting time, packing, and fresh and hardened concrete behaviours.
Therefore, aggregate properties must be measured and controlled in any concrete project.

23
Since aggregates are usually provided by suppliers, its properties have been standardized,
so cement paste properties are the most dominant factors affecting concrete durability.
However some concrete properties such as electrical resistivity may be influenced by
aggregate properties as described in Sections 2.2.3.4.1 and 2.2.4.3.3.

2.1.2.2 Cement phase


The important part of cement paste in durability studies, is the pore system. Pore system
behaviour is defined as the physical characteristics of pore structure (magnitude, size, and
connectivity of pores) and water in pore system (ionic content and concentration in pore
solution) that can influence the fluid transportation properties through concrete (McCarter
et al., 2009). Therefore, both pore structure physical characteristics and pore solution
must be studied in case of concrete durability improvement.
Penetrability of the concrete is affected by both pore structure and voids between
aggregates as described in Section 2.2.5. Pore structure (voids in cement paste) properties
as well as size, distribution, and internal connection of pores can influence the rate of
deterioration of concrete as pores volume increases; the apparent chloride diffusion
coefficient increases (Savas, 1999 in Chini et al., 2003).
The following factors can improve pore structure properties:

2.1.2.2.1 W/CM ratio


The W/CM ratio represents the amount of water in concrete. As this ratio increases,
concrete porosity increases and the pore structure becomes more continuous (Chini,
2003). Therefore, in a low W/CM ratio concrete, penetrability is low while electrical
resistivity is high because of less continuous pore structure (Neville, 1995).
It is worth mentioning that in order to obtain an acceptable workability and surface finish,
mixes with the water-to-cement ratios below 0.45 required relatively higher amount of
cement and/or utilization of chemical admixtures (Al-khaiat and Fatuuhi, 2002).
The W/CM ratio effects on standard durability tests results are described briefly in
Section 2.2.5.1.

24
2.1.2.2.2 Degree of hydration
More hydration (cement hydration or secondary hydration of added SCMs), results in
more produced C-S-H. Therefore, cement paste voids are filled with more hydration
products and porosity and pore connectivity is reduced (Neville, 1995).
Hydration effects on standard durability tests results are described briefly in Section
2.2.5.3.

2.1.2.2.3 Curing
Proper curing of concrete has a very important influence on the final properties of the
concrete. Length of curing period and curing temperature are the most important factors
defining concrete curing procedure.
Since concrete penetrability (and durability) is related to physical characteristics of the
pore system, any action increases cementing materials hydration, reduces concrete
penetrability. Longer curing period improves concrete durability because the volume of
permeable voids is decreases with longer curing while poor curing results in high
absorptivity near the surface (usually the first 30 mm from the surface) (Saricimen et al.,
2000 in Chini et al., 2003).
Higher curing temperature causes continuity pores; this decreases concrete electrical
resistivity and negatively affects concrete durability (Chini et al., 2003) as described in
Section 2.2.5.3. Also curing scenario such as type of curing (steam or moist curing) and
curing temperature affect concrete durability.

2.1.2.2.4 Admixtures
Any admixture causing discontinuous pore system or finer microstructure, improves
concrete durability. Using water reducing admixtures leads in lower required water
content, so free water decreases which results in less and smaller pores, so concrete
becomes more durable. Adding an air-entraining admixture results in millions of small air
bubbles being produced in the concrete, so concrete becomes freeze-thaw resistant. In
spite of the higher setting time, more plastic shrinkage, in case of using retarding
admixtures, later age compressive strength is higher (Neville, 1995).
Admixtures effects on concrete electrical resistivity are described in Section 2.2.4.3.2.

25
2.1.2.2.5 Type of cement
The difference in cement chemical compositions (C3S, C2S, C3A, and C4AF) contents and
cement particles finest are the leading factors in categorizing different cement types.
Cements with low C3S content has lower early age strength and higher early age
penetrability while their later age properties (more than 90 days) are better than high C3S
cements (Neville, 1995). Cement particles size governs cement paste early strength, as
finer cement particles has higher cement paste 28-day strength due to the more surface
area hydrated and more hydrated cement content (Neville, 1995). Also it has been
reported that, using different types of cement results in different electrical resistivity
values as described in Section 2.2.4.3.4.

2.1.2.2.6 Supplementary cementing materials (SCMs)


The use of supplementary cementitious materials has become one of the best solutions to
make durable concrete mixes. These materials are used to replace some cement (or in
some cases in conjunction with the required cement) in concrete to improve concrete
strength and durability. Also using SCMs is an environmentally acceptable course of
action because of the reduced amount of cement used in concrete industry.
By using SCMs into concrete mixture a denser concrete (improved particle packing)
which has finer and discontinuous pore structure is made (Neville, 1995). Therefore, in
concrete containing SCMs, penetrability coefficient is lower and concrete strength and
electrical resistivity is higher than plain cement concrete mixes (Vieira et al., 2000 in
Chini et al., 2003).
Also pore solution in concrete mixes containing SCMs is different than that in plain
cement concretes because of the SCMs secondary hydration which causes reduction in
pore solution alkali content (Neville, 1995).
SCMs effects on standard durability tests results are described briefly in Section 2.2.5.2.

2.1.3 Design and construction


Generally concrete is a porous material and its durability is influenced by its porosity.
Concrete contains two types of pores: voids between aggregates and pores in cement
paste: voids are formed in concrete between aggregates and cement paste while pore

26
structure is formed in cement paste. Both types of porosity must be minimized with
proper designing methods and standard construction practices to make concrete durable.

2.1.3.1 Design
Both components of a reinforced concrete member, concrete and reinforcing bars, must
be designed properly in case of having a durable structure. Reinforcing bars is designed
according to standards such as CSA A23, Design of concrete structures. As the steel
component of a concrete structure can be designed properly by following the standards,
designing the concrete components of the structure is the major issue.
Concrete is made of three major components: cementitious materials, aggregate, and
water (mineral admixtures can be categorized as the forth components used in recent
concretes). Durability of concrete is affected by either individual or combination
characteristics of all these components (will be explained briefly in Section 2.2).
The purpose of designing concrete is to establish a proper proportion for each component
but all concrete will experience the following three phases (Savas 1999, in Chini et al.,
2003):
a) Aggregate (macro)
b) Hydrated cement paste (macro)
c) Interfacial transition zone, ITZ (micro)

ITZ is a low density region around the aggregate between aggregate and bulk hydrated
cement paste. Concrete permeability is influenced by this regions permeability which
contains smaller cement particles.
Since durability of concrete largely depends on the ease with which fluids can enter and
move through (Neville, 1995), concrete durability is more affected by the last two phases
than the aggregate phase. The property or hydrated cement paste and the transition zone
changed with time as such measurements have shown that the higher porosity present
initially in ITZ is significantly diminished by the migration of ions during cement
hydration (Scrivener et al., 2004). Also cementitious materials can improve the durability
of these two phases: adding silica fume improves concrete properties in two ways: (1) by
reducing the ITZ porosity because silica fume is a super fine material. (2) Silica fume

27
generates more C-S-H which results in more discontinuity in pore structure (Neville,
1995).
Finally it can be concluded that high strength concrete may or may not be a durable
concrete (Hooton, 1993 in DeSouza et al., 1998). In general, durability will result if the
concrete has a low W/C ratio, has achieved adequate thermal and moisture curing, and
has achieved a discontinuous capillary pore structure free of micro and micro defects
(DeSouza et al., 1998).

2.1.3.2 Construction
Concrete construction methods and practices influence the final quality of in-place
concrete (Rasheeduzzafar, 1989 in Bryant et al., 2009) although the mix design dictates
the proportion and types of concrete components. These constructional practices which
affect concrete durability, serviceability, and integrity are:
a)

Batching materials regarding to the mix design proportions in correct order (ACI

Guide 304R, 2000).


b)

Mixing to coat all aggregate particles with cement paste (ASTM C685)

c)

Transportation (ACI Guide 304R, 2000)

d)

Placement into frameworks without distributing concrete uniformity

e)

Consolidation to reduce amount of voids in concrete and achieve the highest

possible density (ACI Guide 309R, 2005)


f)

Finishing (ACI Guide 302R, 2004)

g)

Curing the surface concrete to protect concrete from drying which has to begin

immediately after finishing (Kosmatka et al., 2002) explained in Section 2.1.2.2.3.

In other words, if the concrete mixture is not batched, mixed, transported placed,
finished, and cured properly, it will not exhibit the desired performance qualities and
concrete does not become durable (Bryant et al., 2009).

28

2.2 Concrete durability measurement


Durability is another factor, besides strength, that defines performance of concrete.
Measuring concrete durability is more difficult and complicated than measuring
compressive strength. As mentioned before, the most important factor affecting concrete
durability is penetrability. Penetrability is defined in general as the ease with which
fluids, both liquids and gases, can enter into or move through the concrete (Savas, 1999).
It is a function of W/CM ratio, aggregate size, pore size, and pore distribution (Savas,
1999). The key to concrete durability and therefore its performance is to enable concrete
to attain a highly impermeable pore structure (Swamy, 1996 in Bryant et al., 2009).
Concrete permeability can be measured by standard testing techniques. These tests are
destructive for concrete strictures and time-consuming. The ultimate goal of this research
project was to provide reliable correlations between standard tests and non-destructive
and fast electrical tests.
Consequently, standard required tests (e.g. chloride migration under electrical gradient)
and electrical tests must be studied.

2.2.1 Compressive strength test


Although compressive strength test was one of the tests done in this research program, it
was not a durability test. Compressive strength test, methodology, and influencing factors
are briefly described in Appendix B.

2.2.2 Rapid chloride permeability test (RCPT)


Since the ability of concrete to resist chloride penetration is an essential factor in
determining concrete performance, chloride permeability of concrete must be measured
in any concrete durability study. This property of concrete can be measured by a standard
test method for electrical indication of concretes ability to resist chloride ion penetration
named as the rapid chloride permeability test (RCPT). The test, designated as AASHTO
T277 in 1983 by the American Association of State Highway and Transportation
Officials (AASHTO), was the first-ever test proposed for rapid qualitative assessment of
chloride permeability of concrete (Chini et al., 2003).

29
The RCP test monitors the amount of electrical current passed through a 50 3 mm thick
concrete slice of actual diameter ranging from 95mm to 102 mm nominal diameter core
or cylinder over 6 h. testing period (ASTM C1202, 2007). A potential difference of 60 V
induces a direct current (Elkey, 1995 in Chini et al., 2003: may cause polarization in the
pore water and a transport of ions) between two cells containing sodium chloride (NaCl)
and sodium hydroxide (NaOH) solutions. The setup is shown in Figure 2.16.

Figure 2.16: ASTM C1202-07 rapid chloride permeability test setup (Stanish et al., 1997)

The RCP testing methodology is described in Section 3.1.2.


Electric charges travel a tortuous path because of obstructing particles, so the effective
path length is longer than the dimensional of the concrete in the direction of the current
(Monfore, 1968). Concrete tortuosity is defined as:
T= (

Le 2
) ,
L

where, Le is the effective path length and L is the apparent path length of specimen.
This test method evaluates the electrical conductance of concrete (by calculating the total
passing charges in coulombs) over 6 h. to provide a rapid indication of its resistance to
chloride ion penetration as the total charge passed (the area under the electrical current

30
versus time curve), in Coulombs, has been found to be related to resistance of specimen
to chloride ion penetration (ASTM C1202, 2007) as shown in Table 2.2

Table 2.2: Chloride ion penetrability based on charge passed (ASTM C1202, 2007)
Charged Passed (Coulombs)

Chloride Ion Penetrability

> 4000

High

2000 - 4000

Moderate

1000 - 2000

Low

100 - 1000

Very Low

< 100

Negligible

These values have been shown to be representative of chloride ion permeability which is
an indirect indication of the permeability of concrete (Baykel, 2000 in Chini et al., 2003).
It can be concluded from Table 2.2 that a low passed charge means relatively
impermeable concrete; However, values may be inaccurate if concrete is atypical (Stanish
et al., 1997).
Since slow hydration of some SCMs delay their impacts by about one or two month and
the RCPT results depend on pore solution chemistry and pore structure characteristics,
Table 2.2 limits would be unreasonable to apply on the 28 day concretes but would be
sufficient for 91 days concrete (Shi, 2004).
Since the test result is a function of electrical resistance of the concrete specimen, the
presence of reinforcing steel or other embedded electrically conductive materials may
have a significant effect. Therefore, the test is not valid for specimen with longitudinal
reinforcing bars that can provide continuous electrical path between the two ends of the
specimen (ASTM C1202, 2007).
This testing method can be used in applications such as quality control and acceptance
testing in practice.
In addition, it is known that RCPT data reflects the electrical resistance of concrete rather
than the resistance to chloride penetration (Wee et al., 2000 in Chini, 2003).

31

2.2.2.1 First 5 minutes RCPT resistivity


After the first 5 minutes of testing in the RCPT, electrical resistivity, , can be calculated
from the current using the following equations:
R=
R=

L
A
=R
A
L

V
(Ohms law)
I

d 2 V
4L

where, d is the average of four diameters measured on the specimens cross-sections, L is


the thickness of the specimen, I is the first 5 min. electrical current passing through the
concrete disc, and V is the applied voltage, which is 60 V.
Since the solution temperatures remain constant during early minutes, the RCPT
electrical resistivity value is independent from heat effects.
Since the durability of salt-exposed concrete structures is influenced by the ability of
concrete to resist the penetration of chloride ion, chloride ion penetration, measured after
the RCP test, is used for a non-steady-state migration similar to Nordtest NT build 492.

2.2.2.2 Chloride ion penetration (chloride migration coefficient)


The ability of concrete to resist chloride ion penetration is a key component for durability
analysis of concrete structures especially in salt-exposed structures. The RCPT results
have been used as an indicator for concrete durability for many years. However, a
number of scientists have criticized this test for its bias to supplementary cementing
materials, especially silica fume, because cement replacement with silica fume can lead
to an order of magnitude reduction in Na+, K+, Ca++, and OH- ion concentration in pore
solution (Shi et al., 1998 in Ahmed et al., 2009). They have reported that it is not
appropriate to use the RCPT charge passed alone to evaluate chloride ion penetrability in
concretes incorporating materials affecting the ionic concentration of pore solution
especially silica fume (Bassuoni et al., 2006). However, others have found that the effect
is small (Nokken and Hooton, 2006).

32
Therefore, additional property of concrete which is less dependent on the electrical
conductance of concrete and pore solution is requires to be measured. This property
(depth of chloride ion penetration) represents concrete resistance against chloride ion
penetration forced into concrete over 6 h. RCP test (Bassuoni et al., 2006). This property
is mainly related to continuity of the concrete porosity because the ions can not travel
through the solid pore walls and are only free to travel through the liquid-filled pore
structure (Stanish et al., 2004).
Tested samples can be split open after the RCP test. Fracture surfaces of split samples are
sprayed with 0.1 M silver nitrate solution, AgNO3, to determine the depth of chloride
penetration. Fifteen minutes later the area containing chloride remains white, silver
chloride, while the color of the area free of chloride ion remains natural or brown as
shown in Figure 2.17.

Figure 2.17: Whitish chloride front migration in concrete sprayed after the RCP test
(Bassuoni et al. 2006)

The depth of chloride ion penetration is used to calculate a non-steady-state diffusion


coefficient using the equation in Nordtest NT Build 492 (Bassuoni et al., 2006):
D=

(273 + T ) Lx d
0.0239(273 + T ) L
( xd - 0.0238
),
V 2
(V 2)t

where, D is non-steady-state migration coefficient (x 10 12 m 2 /s), V is the applied voltage


(V), T is the average value of initial and final temperature in the anolyte solution (C), L

33
is the thickness of the specimen (mm), xd is the average depth of chloride penetration
(mm), and t is time (h).
It is important to mention that although the electrical potential gradient remains constant
during the 6 h. test, chloride binding and direction of pores resistance decrease the rate of
chloride ion penetration (Stanish et al., 2004) which has not been considered in the
migration coefficient relation.

2.2.2.3 Linear extrapolation technique


The RCPT test is continued for either 6 h. or until the NaCl solutions temperature
reaches 80C. This temperature is the upper limit to avoid any damage of test cell
components (Bassuoni, 2006). To avoid high temperature effects on the RCPT values, the
total charge passed through concrete discs over 6 h. can be estimated by multiplying the
30-min. values by 12 (Hooton et al., 1997 in Bassuoni et al., 2006).
By calculating the extrapolated charge passed, the heating effects on the RCPT charge
passed and electrical resistivity results are discounted. The extrapolated charge passed is
always lower than the actual passing charged measured by 6 h. standard procedure
(Bassuoni et al., 2006) because the heat effect is eliminated.

2.2.2.4 Influencing factors on penetration resistance test


Electrical conductivity of concrete depends on both its pore structure characteristics
(charges in pore structure) and pore solution chemistry (Monfore, 1968). Both mentioned
factors are functions of the W/CM ratio, admixture, mix design, supplementary
cementing materials, and etc. (Savas, 1999). Most of them such as W/CM ratio, adding
SCMs, and length of curing not only influence the RCPT results but affect other
durability tests results as described briefly in Section 2.2.5.

2.2.2.4.1 Admixtures and RCPT values


Addition of chemical admixtures, such as calcium nitrite ,Ca(NO2)2 , which is found in
corrosion inhibitor admixtures, affects concrete pore solution chemistry (Wee et al., 2000
in Chini et al., 2003). This may have effects on the accuracy of the Coulomb value.
Calcium nitrite (which is found in corrosion inhibitor admixtures) reduces electrical

34
resistivity (increases the RCPT charge passed of a given concrete), but it does not
increase the rate of chloride ingress (Savas, 1999).
Other chemical admixtures may affect results similarly, so electrical properties of mixes
containing chemical admixtures should be compared with control mixes with caution.

2.2.2.4.2 Temperature and RCPT values


Temperature of the solution should be limited between 20 to 25 C during measurements
according to ASTM C1202-07. As temperature increases, the reported RCPT level of
permeability will be higher than the actual permeability level measured under the normal
temperature (Bassouni et al., 2006). It can be avoided by linear extrapolation estimation
technique. This disadvantage does not affect first 5 min. electrical resistivity.

2.2.2.5 RCPT weak points


As indicated in the literature, there are four main disadvantages for the RCP test:
1) Applied electrical charges are taken by all chemical ions presented in pore
solution, not only the chloride ions. The RCP test does not distinguish between
the current carried by chloride ion and that carried by the other ions available in
pore solution (Suryavanashi et al., 2002 in Ahmed et al., 2009).
2) The heat generated during the RCP test is the other problem which increases the
total reported charge passed in case of high solution temperature (Bassouni et al.,
2006).
3) Pore fluid characteristics and chemical composition can affect the RCPT results
especially in mixes with pozzolanic materials such as silica fume (Hale et al.,
2002 in Ahmed et al., 2009).
4) The RCPT does not measure the permeability of concrete directly, but it measures
the total charge passed through concrete over 6 h. of testing period which is
related to chloride permeability. This relation is criticized in concrete mixture
containing SCMs especially silica fume since it reduces ionic concentration in
pore system. Therefore, some scientist have claimed that the reduction in total
RCPT charge passed was due to the reduced ionic concentration not discontinuous
and small volume of pore system. This claim is rejected in Section 4.2.3.

35
On the other hand, this test can evaluate concrete electrical resistivity. Also its
convenience and short-time duration are other advantages for this standard durability test.

2.2.3 Rate of water absorption (Sorptivity)


As water penetration is the major cause of steel corrosion and freeze-thaw damage, water
absorption into hardened concrete is an important factor in determination of concrete
durability. Moreover, studying the water absorption of concrete surfaces is more essential
than that of the concrete core. Rate of water absorption of a concrete surface is different
from the rate of absorption of a sample taken from the interior because the exterior
surface is often subjected to less curing, and more exposed to potentially adverse
conditions (ASTM C1585, 2004).
Therefore, surface concrete penetrability is studied by measuring the rate of water
penetrability due to the capillary rise (sorptivity).

2.2.3.1 Calculation
The sorptivity test measures the rate of absorption of water when one surface of concrete
specimen is exposed to water. Capillary suction is the reason for water absorption into the
concrete specimen. DeSouza presented the rate of absorption, I, using the sorptivity
relation expressed by Hall (1989) if a specimen is in contact with water from one of its
surfaces (DeSouza et al., 1997):
I=

mass
,
A.

where, I is cumulative water absorption (mm), mass is the change in the mass of the
specimen which is in contact with water (g) , A is the cross-section area of the specimen
(mm2), and is water density (

g
).
mm 3

The mass represents the amount of water absorbed by the specimen.

36
According to the principle of mass conservation:
Q=where, q is volume flow rate per unit area (

,
t

m
), t is the flow time (s), and is degree of
s

hydration.
This equation can be combined to yield the one-dimensional non-linear diffusion
equation (Nokken et al., 2002):

=
[D ( )
],
t z
z

where, z is elevation head (mm).


Philip (1957) solved the last equation for the boundary conditions (0,t)=1 and (x,0)=0
(Nokken et al., 2002). Therefore, the sorptivity equation was derived:

I=S t,
where, S is water sorptivity (

mm
s

) and t is time of absorption (s).

In other words, if I against square root of time data is represented by a straight line,
water sorptivity in mm/min0.5 should be determined as the slope of the least-squares
linear regression line.
The I-t0.5 plot has to be linear with a regression coefficient, r2, of less than an arbitrary
value of 0.98 (ASTM C1585, 2004), otherwise the sorptivity can not be derived
accurately with this linear relationship.
At the beginning of the test, the graph is not linear due to the skin layer of concrete and
saturation of skin concrete, but later, the plot becomes linear (Desouza, 1996).
Water absorption is strongly affected by the moisture condition of the concrete at the time
of testing, so standard amounts of concrete moisture must be assigned for the test.
Previous works (Hall, 1989 in DeSouza et al., 1996 and Parrott, 1994) have indicated that
certain pre-conditioning regime must be applied to obtain a uniform moisture distribution
in specimens (explained briefly in the methodology section, Section 3.1.3.1).

37
Sorptivity is a term used for water ingress into pores of concrete under unsaturated
conditions (50 to 70% internal relative humidity which is similar to the RH found near
the surface in some field structures according to ASTM C1585-04) due to capillary
suction. Rate of water absorption or sorptivity of a concrete mixture at each age can be
measured by both a laboratory sorptivity test and a field sorptivity test.

2.2.3.2 Laboratory sorptivity test


As mentioned in ASTM C1585-04, the average of test results on at least two 100 6
mm diameter with a length of 50 3 mm specimens shall be used as the test results.
Specimens are obtained from either molded cylinders according to practices or drilled
cores according to test method. Concrete top surfaces are in contact with tap water for
eight days and the amount of water absorbed by the specimen is measured at time
intervals specified in the standard (ASTM C1585, 2004).

2.2.3.3 Field sorptivity test


Since concrete deterioration processes (e.g. rebar corrosion) are influenced by the
concrete fluid penetrability especially in covercrete, the ability of concrete to absorb
water on site provides information about its durability. Concrete long term durability in a
severe environment is achieved by the quality of cover concrete between reinforcing bars
and the exterior surface of the member. This cover layer concrete contains three sublayers; cement skin which is about 0.1 mm thick, mortar skin which is about 5 mm thick,
and the concrete skin about 30 mm (DeSouza, 1998).
For measuring the fluid penetrability of covercrete non-destructively, the field sorptivity
test is used. Capillary rise, or sorptivity, is the case of one-dimensional absorption, in
which flow is normal to the inflow face throughout the wetted region (test methodology
is briefly described in Section 3.1.3.2.1).

2.2.3.3.1 Field sorptivity apparatus overview


The field sorptivity instrument has two major parts; a separate vacuum base plate and
plexiglass test disc which can be easily removed. Figure 2.18 shows the field sorptivity
apparatus in use on a concrete slab.

38

Figure 2.18: Field sorptivity apparatus (horizontal orientation)

This test configuration causes neutral pressure between the test area and the vacuum
region, so water can not wick into the surrounding vacuum under the base plate (DeSouza
et al., 1998).

2.2.3.3.1.1 Vacuum-attachment base plate


The base plate was fabricated from a 6 mm aluminum plate, 300 mm in diameter with a
125mm central hole. It was equipped with three Delta 150 mm vide clamps (Quick
Setting Drill Press Work Hold-Downs), mounted on stainless steel blocks spaced 120
apart around the diameter of the plate.
To make a proper seal, the aluminum base plate contains a double gasket. These gaskets
were made of a 6 mm square Neoprene, closed-cell gasket within the inner regions and a
10 mm Neoprene sponge gasket around the outer regions. A 3 mm NPT vacuum port was
tapped into the plate in inner vacuum chamber as shown in Figure 2.19.

39

Figure 2.19: Vacuum- attachment base plate (DeSouza et al., 1996)

All gaskets were glued in place with contact cement (DeSouza et al., 1996)

2.2.3.3.1.2 Disc plate


The sorptivity disc plate consists of a 150 mm diameter; 25 mm thick cast Acrylic plate
which is sealed to the base plate using three bench clamps as shown in Figure 2.20.

40

Figure 2.20: Field sorptivity disc plate (DeSouza et al., 1996)

The plexiglass has two chambers. The chambers were sealed with 12mm thick roomtemperature-vulcanized (RTV) silicone rings to utilize lateral confining forces to properly
seal the test surface. To maintain a high point, the inner 91 mm diameter region was
machined to have a concave domed interior profile as shown in Figure 2.23. Therefore,
air can get out.
By placing the instrument on a concrete surface, water in the domed part of the sorptivity
disc is absorbed due to the capillary pore suction. The volume of water absorbed by the
cover layer concrete over 16 minutes can be read from the graduated pipette connected to
the top part of the instrument. The water absorption is the volume of absorbed water

41
divided by the affected area. The slope of the water absorption versus square root of time
plot is the water sorptivity.
This type of water sorptivity test is inexpensive, quick (about 20 minutes), and
nondestructive.

2.2.3.4 Influencing factors on water sorptivity values


The water penetrability of concrete only depends on the pore structure of concrete (Shi,
2004). Water permeability of concrete is not only a function of porosity, but also size,
distribution, shape, tortuosity, and continuity of the pores (Neville, 1995). Therefore, any
factors influencing the physical characteristic of the pore system, affects the rate of water
absorption of concrete. Most of these factors (e.g. W/CM ratio, adding SCMs, curing
period, and moisture content) influence other durability test results which are described in
Section 2.2.5.

2.2.3.4.1 Aggregate and the rate of water absorption


Because the flow path has to circumvent the aggregate particles, the effective path
becomes longer so that the effect of aggregate in reducing the permeability may be
considerable (Neville, 1995). Presence of aggregate in concrete mixtures influence the
permeability coefficient of concrete if aggregate is low-permeable. On the other hand, if
aggregate is more permeable than cement paste, aggregate introduce will increase the
permeability of concrete (Shi, 2004). However, for a given W/CM ratio and degree of
hydration, sorptivity of concrete made with low-permeability aggregate is about one to
two orders lower than that of cement paste due to the ITZ between aggregate and cement
paste (Neville, 1995 in Shi, 2004).

2.2.4 Electrical resistivity test


In addition to ASTM C1202-07, electrical resistivity has been used to characterize the
electrical properties of concrete. Electrical resistivity of concrete is a measurement of its
ability to resist electron transfer. This electrical property (ionic transport property) of
concrete has become an indicator for assessment the physical properties of the pore

42
structure and microstructure and the chemistry of pore solution for more than forty years
(Monfore, 1968).
Electrical resistivity is a geometry-independent material property that describes the
electrical resistance of concrete (Gowers and Millard, 1999). Resistance is the ratio
between applied voltage and resulting current in a unit cell according to Ohms law. The
dimension of resistivity is resistance multiplied by length, so the unit is usually .m or
K.cm. This physical parameter is different for every material and can be found from
electro-magnetism tables. On the other hand, it is difficult to find a unique number as
resistivity for composite materials because of different properties of their components.
Concrete is a composite material whose compounds can be described as: (a) a solid phase
purely resistive (aggregates), (b) a solid phase which participates to conduction through
its porous structure (cement matrix) and being the source of ions found in the third phase,
(c) the liquid phase, i.e. interstitial solution (Lastaste, 2003). Therefore, an electrical
circuit can be created in this composite material, which can be linked to the circulation of
fluids through the pore network. Generally, the current flows through the pore liquid in
the cement paste and aggregate can be considered inert (Ferreira and Jalali, 2010).
Electrical resistivity of concrete varies over very broad ranges: from 1011 .cm for ovendried concretes to less than 103 .cm for saturated concretes (Lopez and Gonzalez,
1993).
Electrical property of concrete is important for civil engineers because of two reasons:
1) Since the major types of fluid transportation through concrete (permeability and
diffusion) are analogous to the flow a current under a potential difference (hence,
electrical resistance), electrical properties of concrete can serve as a simple and
effective assessing indicator of fluid transport processes and hence, durability
(McCarter et al., 2009).
2) During rebar corrosion in concrete (one of the main causes of deterioration of
reinforced concrete), an electrical circuit is formed as shown in Figure 2.21;
corrosion current passes thought concrete from cathode to anode. Therefore, to
make a corrosion resistant concrete, electrical resistivity of concrete must be
increased. Increasing the electrical resistivity of concrete breaks the electrical

43
circuit which is required for the steel corrosion. Therefore, the concrete electrical
property can control the magnitude of the corrosion current.

Figure 2.21: Schematic view of reinforcing bar corrosion in concrete (Millard and Gowers, 1991)

The steel corrosion is inversely proportional over a wide electrical resistivity range, so an
important application for the electrical resistivity values is to indicate the corrosion rate
of steel bars in concrete (Lopez and Gonzalez, 1993).
There is not a general agreement about the resistivity level above which corrosion risks
will be negligible, but a relation between resistivity and rate of steel corrosion is shown in
Table 2.3. In general, a low resistivity is related to a high risk of corrosion (Ferreira and
Jalali, 2010).

Table 2.3: Electrical resistivity values for rebar corrosion rate (Millard and Gowers, 1991)

Electrical Resistivity (K. Cm)

Corrosion Risk Level

> 20

Low Rate

10 - 20

Moderate Rate

5 -10

High Rate

<5

Very High Rate

44
These levels have been confirmed from correlations with linear polarisation an AC
impedance corrosion measurements (Millard et al., 1989).
The electrical resistivity of concrete is dependent on the degree of pore saturation
(humidity), pore structure continuity, ionic concentration in pore solution, and to a lesser
extent, on the degree of paste hydration (Lopez and Gonzalez, 1993).
The resistivity of a concrete specimen can be measured non-destructively using
electrodes placed on a specimen surface. This requires at least two electrodes, one of
which may be a reinforcing bar in case of in-situ measurements (Polder, 2000). A voltage
is applied between the electrodes and the resulting current is measured or vice versa
(Morris et al., 1996). From Ohms law, the ratio of voltage to electrical current results in
resistance of the sample (R=V/I). The resistivity is obtained by multiplying the measured
resistance by a conversion factor, called the cell constant, m (Polder, 2000). For a given
cell arrangement, the cell constant can be obtained either from theoretical considerations
or from calibration using standard concrete samples or electrolytes of known resistivity
(Polder, 2000). In order to have the cell constant, different cell arrangements must be
studied.
Generally, electrical resistivity tests can be categorized as test measuring the bulk
resistivity of concrete and tests measuring the surface electrical resistibility of concrete.

2.2.4.1 Bulk electrical resistivity


In principle, the bulk electrical resistivity of concrete can be measured using two
electrodes placed on concrete opposite surfaces (Morris et al., 1996).
One electrode induces the electrical current and the other electrode receives the current.
Voltmeter can calculate the potential drop, P, during this electric circuit or the external
plate-electrodes provide a uniform electrical field within concrete specimen and the
voltage drop is measured by the instrument (McCarter et al., 2009) as shown in Figure
2.22.

45

Figure 2.22: Concrete bulk electrical resistivity test with two electrodes

The ratio of the potential drop to the current is the member electrical resistance according
to the Ohms law. Electrical resistivity, , can be calculated from the following equations:

P=RI (Ohms Law)


R= (L/A) =P/I (A/L)

d 2

P
=

4L I ,
where, d is the section diameter (mm), L is the length of the cylinder (mm), I is the
current passed through the specimen (A), and P is the voltage drop (V).
This measuring technique can be complicated by the need for effective and uniform
contacts between the end electrodes and concrete specimen end surfaces, so flat cylinder
end surface, flexible electrodes, and resistance free connections between electrodes and
concrete are required (Morris et al., 1996).
Depending on the type of applied electrical current, this type of resistivity measurement
can be classified into two methods: DC method (using cyclic direct current, Monfore) and
AC method (using alternating current).
A schematic view of a DC resistivity meter is shown in Figure 2.23.

46

Figure 2.23: Schematic diagram for the DC test setup (El-Dieb et al., unpublished)

The DC measurement is carried out using a cyclic DC potential similar to that of the
Monfore. The resistivity meter used at the University of Toronto applies a cyclic voltage
across the specimen between 3 and 5 volts every 5 seconds, so the derived resistivity
equation is modified into:
Resistivity =

(V5 V3 ) A
(K.cm),
(I5 I3 ) L

where, V3 and V5 are average applied voltage for 3 and 5 volts respectively, I3 and I5 are
average applied electrical current for 3 and 5 volts in Amperes respectively, A is the area

47
of the core face (cm2) and L is specimens thickness (cm). In DC resistivity-meters, high
contact-persistence between the current electrodes and the concrete surface desensitize a
resistivity measuring instrument by reducing the amount of applied current flowing for a
fixed voltage drive (Ewins, 1990). In other words, if the conductor is an electrolyte (e.g.
cement paste) the passage of direct current will cause polarization and the establishment
of a potential at the electrodes that opposes the applied potential (Monfore, 1968). In this
case the polarization potential reduces the current passed through the specimen:

I=

Ea E p
R

where, Ea is the applied potential (V) and Ep is the polarization potential (or back emf)
(V).
Polarization potential which depends on the ions present and the materials of the
electrodes results from reactions that take place at the electrodes. During this reaction,
thin films of oxygen, hydrogen, or other gases may be formed on the electrodes resulting
lower potential created (Monfore, 1968).
On the other hand, resistivity-meter used in this research program worked with the cyclic
direct current. The cyclic DC avoids polarization effects.
In AC resistivity-meters, electrode/concrete interface polarization and capacitive effects
(formation of a thin gas film between electrodes and electrolyte) are not seen, resulting in
real (and lower than DC instrument) resistivity values (El- Dieb et al., unpublished).
Schematic diagram of an AC resistivity-meter is similar to a DC resistivity meter as
shown in Figure 2.24.

48

Figure 2.24: Schematic diagram for the AC test setup (El-Dieb et al., unpublished)

The DC method overestimated by about 58% the concrete resistivity than the AC method
due to the electrode polarization potential or back emf at the electrode surfaces (El-Dieb
et al., unpublished). This polarization effect can be avoided at 50 Hz frequency and more,
as Hammond and Robson interpreted this to mean that the capacitative reactance of
concrete is much larger than its electrical resistivity (Neville, 1995). On the other hand,
El-Dieb et al. reported that the AC resistivity values are similar to the RCPT resistivity
values measured by direct current (instantaneous measurement of the resistivity (first 5
min.) under the application of a direct current).
Electrical resistivity of stainless steel electrodes is near zero, but in case of using
saturated sponge as a connection between electrode and specimen; concrete resistance is
calculated as (McCarter et al., 2009):

Rconc.= R measured - Rspone ,


where, Rspone is the electrical resistance of the saturated sponge (). To obtain the true
resistivity, the use of a high frequency instrument is necessary, so the sponge-specimen
interface region is short-circuited and not considered as measurement (McCarter et al.,
2009).

49
In conclusion, DC measurements are more appropriate than AC if the current (or
potential required to overcome polarization) due to polarization is subtracted from the
total value of current or potential, respectively (Monfore, 1968 and Hausmann, 1964 in
Hansson, 1983). This potential drop due to polarization effects is the intercept on the
voltage axis of the linear portion of the Vcell vs. I curve (Hansson, 1983).
Bulk resistivity measurement techniques are useful for cast specimens or cores taken
from an existing structure. To apply this testing method to a concrete member on site;
both electrodes should be installed on un-sealed surface(s) of the member, if installing in
two opposite surfaces is possible.
In practice, this method is expected to be less accurate and poorly reproducible because
the electrode size has an important effect on the measured values (Polder, 2000).

2.2.4.2 Surface electrical resistivity


There are two major reasons for evaluating the surface electrical resistivity of concrete:
1) Concrete long term durability in a severe environment is achieved by the quality
of concrete between reinforcing bars and the exterior surface of the member
because all deteriorating factors attack concrete and rebar from covercrete.
Therefore, studying the covercrete electrical resistivity (and its correlation with
penetrability) is necessary in case of durability assessment.
2) Bulk electrical resistivity tests are destructive (cores must be taken from an
existing structure), while surface electrical resistivity tests are non-destructively
(can be performed on as constructed surface). Non-destructive evaluations appear
more and more important in civil engineering projects especially in high sensitive
structures such as nuclear plants where coring is not allowed.
The surface electrical resistivity of concrete can be measured non-destructively by several
techniques:

2.2.4.2.1 Surface disc


One of surface electrical assessment testing methods is a disc-electrode method. This
method involves an electrode-disc placed on concrete surface over a rebar and measures

50
the resistance between the disc and the rebar. This method requires a connection to the
reinforcement cage and full steel continuity (Polder, 2000) as shown in Figure 2.25.

disc

rebar

Figure 2.25: Setup of one electrode (disc) measurement of concrete resistivity (Polder, 2000)

By applying this test method to several concrete slabs of known resistivity, the cell
constant can be determined. For cover depths, disc and bar diameters being 10-50 mm,
the cell constant is approximately 0.1 m (Polder, 2000). Therefore, the resistivity
measured using a disc electrode is approximately:

p(disc)= 0.1*R(disc bar) = 0.1V / I ,


where, V (or V is used sometimes) is the potential drop (V), I is the current passed
through covercrete (A), and is surface electrical resistivity (.cm).
The other type of surface resistivity measurement technique for measuring resistivity
which was first used by geologists for investing soil strata (Wenner, 1980 in Millard et
al., 1989) is a four-probe technique. In this technique four electrodes are located on
concrete surface. Two electrodes are for current insertion and the other two are the
potential measurement points (Morris et al., 1996). Primary measurements have shown
that direct current can not be used to measure the resistivity of concrete because of the
polarization effect on electrode/concrete interface (Millard et al., 1989). If a sine-wave
current source is used, impedance is likely to be measured between the voltage probes

51
(probe/concrete interface problems) rather than a resistance, and this may vary with
frequency, so the results are doubtful (Ewins, 1990).
Two common types of four-probe array technique are different due to their electrode
configuration.

2.2.4.2.2 Four - probe line array (Wenner probe)


This convenient technique measures the in-place surface electrical resistivity of concrete
non- destructively. Four equi-spaced, a, electrodes, placed on the surface of concrete in a
linear array were used in Wenner probe.
In this method a low-frequency alternating current (AC) passes between the two outer
probes. The two inner electrodes serve as the potential drop electrodes. Electrodes can be
embedded in concrete during casting as shown in Figure 2.26.

Figure 2.26: Concrete cube with embedded electrodes (McCarter et al., 2009)

If the concrete specimen be approximated as a semi-infinite medium, resistivity of


concrete is given as (Tagg, 1964 in McCarter et al., 2009):

4aRc
1+

2a
a 2 + 4d 2

2a
4a 2 + 4d 2

52
where, a is the electrode spacing, d is the depth of embedment of the electrode tip, and Rc
is the concrete resistance which is V/I according to the Ohms law.

In another type of surface electrical resistivity measurement, electrodes are pressed


against constructed surfaces of concrete (Wenner probe). Alternating current (AC) is
passing between two other electrodes while the voltage drop is measured by two inner
electrodes as shown in Figure 2.27.

Figure 2.27 Schematic representation of four-electrode resistivity test (Gowers and Millard, 1999)

Probe spacing, a, must be determined carefully. If the probe spacing is too small, the
presence or absence of aggregate particles (very high resistivity) will lead to a high
degree of scatter while in case of too large spacing, inaccuracies due to the current field
being constricted by the specimens edges is seen (Millard et al., 1989).

53
The Wenner resistivity can be calculated from the previous equation, but the depth of
embedment of the electrode tip (d) is zero:
4 aRc

=
1+

2a
2

a + 4d

2a
2

4a + 4d 2

d=0 (electrodes on located on the surface) and

= 2a

V
,
I

where, a is the probe spacing (mm), V is the voltage drop measured by two inner probes
(V), and I is the applied current by the two outer probe (A). 2a is the cell constant for
four linear probe array.
Four major difficulties must be overcome when the four electrode method is applied to
concrete (Millard et al., 1989):
a) Steel bars must be kept away from the depth affected by the applied electrical
current flow (see Figure 2.27), otherwise the apparent resistivity is being recorded
significantly lower than the true resistivity of concrete (Millard and Gowers,
1991).
b) Probe spacing must be chosen as specimen becomes semi-infinite (will be
explained in Section 3.2.1.1.1.2).
c) When measuring the resistivity it is important to eliminate any resistance between
the probes and the body of concrete. High probe resistance causes significant
errors especially in high frequency instruments (Ewins, 1990). This resistance can
be avoided by using saturated wooden bars (or sponges) or contact gel as the
connection between the electrodes and concrete surface.
d) Dramatic error occurs when there are two surface layers with different resistivities
such as might be caused by a recent wetting of concrete already having surface
carbonation which causes a increase in the apparent resistivity but also salt ingress
into the surface zone (Millard and Gowers, 1991).

54
All surface electrical resistivity relations are based on the assumption that the probe is in
contact with the face of a semi-infinite uniform body (Morris et al., 1996). In practice, the
electrodes are in contact with a body of finite dimensions unless the optimum probe
spacing is figured out based on the specimens geometry, aggregate size, and rebar
location (Section 3.2.1.1.1.2).
Therefore, concrete resistivity is calculated from a relation between apparent resistivity
measured by the resistivity-meter and the cell constant correction, K, a function of interprobe distance, a, and the geometry of concrete body (Morris et al., 1996):

app
K

Generally, two types of Wenner probe configuration for concrete cylinders (specimens
used in this research program) are common:

Probe configuration type 1) The first configuration is considered the metallic electrodes
centre on one of the end faces of a concrete cylinder.
Figure 2.28 presents appropriate cell constant correction, K, for this type of Wenner
probe configuration.

Figure 2.28: Cell constant correction factor for the centered end face configuration
(Morris et al., 1996)

55
Usually for the case that removing the sample from its mould is not wanted or enough
room is not available to install the probes on the side of the cylinder (Morris et al., 1996).

Probe configuration type 2) The second situation is consisted in placing the resistivitymeter electrodes longitudinally on the side of a concrete cylinder. This is the most
common type of electrode configuration for Wenner probe technique in civil engineering
projects. By using the graph shown in Figure 2.29, the cell constant correction for this
type of probe installation can be found.

Figure 2.29: Cell constant correction factor for the centred longitudinal measuring configuration
(Morris et al., 1996)

The cell constant correction factors found for concrete cylinders based on Figures 2.28
and 2.29 can be used for other types of specimen if the geometrical dimensions, required
in those graphs, are adapted.

2.2.4.2.3 Four - probe square array


This convenient testing technique measures the in-place electrical resistivity of concrete
non-destructively. In this type of surface electrical resistivity measurement, four
electrodes spaced out 5 or 10 cm arranged in a square (Lataste et al., 2003) is used as
shown in Figure 2.30.

56

Figure 2.30: Four-probe square array principle (Lataste et al., 2003)

Two neighboring electrodes (A and B) inject a known electrical intensity while the
potential difference, V created by the passage of the current in the material is measured
between the two remaining electrodes (M and N) (Lataste et al., 2003). 3D current and
voltage flow pattern for four-probe square array is shown in Figure 2.31.

Figure 2.31: Four-probe square array schematic representation for studying crack parameters
(Lataste et al., 2003)

57
The equation presented on Figure 2.30 is the apparent resistivity of concrete sample and
2a
is the cell constant for four-probe square array instrument.
2 2
To calculate the resistivity of concrete, the following equation should be followed:

app
K

Cell constant correction, K, is a function of the probe distance, a, and the geometry of
concrete body test as described in Section 2.2.4.2.2. In this type of the surface resistivity
measurement, both sides length of concrete member should be considered to find the
correct cell constant correction (Lataste et al., 2003). This is the main strategy to find the
correct cell constant correction, in four-probe square array.

2.2.4.2.4 Application
Electrical resistivity measured by the Wenner probe can be related to permeability
properties and moisture content of concrete, which is the ultimate goal of this project. As
well, the Wenner probe can be used for resistivity mapping (Polder, 2001). Resistivity
mapping is useful in case of actively corroding steel. In resistivity mapping, surface
electrical resistivity values of each electrical zone can distinguish the high corrosion risk
locations by the corrosion prediction levels as shown in Table 2.3 (Millard and Gowers,
1991). This potential mapping technique identifies regions where corrosion in occurring,
but does not give any information about the rate of corrosion activity. But the corrosion
rate of steel rebar can be estimated by electrical resistivity when concrete resistivity
exceeds 70 .m (Lopez and Gonzalez, 1993):

CR =

1000

concrete

where, C R is corrosion rate (m Fe/yr) and concrete is concrete electrical resistivity


(.m).
Nevertheless, in some cases steel corrosion can not be truly detected by electrical
resistivity; saturated carbonated concretes (supports high corrosion rate once the steel has
lost the passivity layer) have been found 3 to 4 times more resistant than saturated uncarbonated concretes (Millard and Gowers, 1991).

58
Resistivity mapping by surface electrical resistivity is also beneficial for local
electrochemical repair methods which are based on the locations with different corrosion
rates (Polder, 2001).
Also the Wenner probe can be used as a crack detector within concrete members.
Average resistivity value of an un-cracked concrete is around 800 10% .m while the
average resistivity in the delamination zones is between 1700 and 3000 .m (Lataste et
al., 2003). In addition, depth of cracks can be estimated by the apparent resistivity values
as shown in Figure 2.32.

Figure 2.32: Surface resistivity as a function of crack depth (Lataste et al., 2003)

It can be concluded from Figure 2.32 that the resistivity value increases if the applied
electrical current is perpendicular to the crack. If cracks are conductive (i.e. full of water)
the top line in Figure 2.32 remains relatively 800 .m constant (un-cracked concrete).

2.2.4.3 Influencing factors on concrete electrical resistivity


The electrical resistivity of a concrete can be defined as the resistance of concrete for
flow of electrical current through concrete, so any factor affecting passing electrical
current, influences concrete electrical resistivity (Sengul and Jjorv, 2009).
As conduction of electricity through moist concrete is visualized as the ion movement in
the evaporable water (in paste matrix and in some cases in aggregates pores), any factors
affects amount of liquid (moisture content), pore solution ionic concentration, or

59
continuity of pores will influence the resistivity of concrete (Monfore, 1968). Electrical
conductivity of concrete depends on its pore structure (volume of porosity and porous
connectivity) and pore solution chemistry (Lataste et al., 2003 and Savas, 1999).
Therefore, factors influencing physical characteristic of pore system (e.g. porosity and
pore-size distribution), pore solution chemistry, and ionic mobility in the pore solution
must affect electrical resistivity values of concrete. Among of all these influencing
factors, W/CM ratio, adding SCMs, moisture content, and curing affect other durability
tests as described in Section 2.2.5.

2.2.4.3.1 Temperature and electrical resistivity


Temperature has an important effect on concrete electrical resistivity. Overall, a
temperature increase causes a decrease of electrical resistivity and vice versa because
concrete has electrolytic properties and temperature influences ion mobility, ion-ion and
ion-solid interactions (Millard et al., 1989) as shown in Figure 2.33.

Figure 2.33: Relationship between measured resistivity and air temperature


(Gowers and Millard, 1999)

60
From laboratory work, it appears that the temperature effect may vary with moisture
content; electrical resistivity changes with 3% for saturated and 5% for dry concrete for
each degree K temperature change (Bertolini and Polder, 1997, Polder, 2000, and Elkey,
1995 in Chini, 2003).
However, temperature effects are far more significant than small changes in moisture
content; from site experience, electrical resistivity is significantly higher during the
winter (although the moisture content is high) than the summer period (Millard et al.,
1989).

2.2.4.3.2 Chemical admixtures and electrical resistivity


As mentioned before, electrical properties of concrete are influenced by the ionic
concentration and mobility in pore solution. Therefore, any chemical admixture affecting
pore solution chemical composition will influence concrete electrical resistivity:
1) The use of ammonium phosphate in concrete has been reported to increase the
electrical resistance of concrete (Freitag, 1961 in Monfore 1968).
2) Also it has been reported that concretes with nitrate based corrosion inhibiting
admixtures, such as calcium nitrite, Ca(NO2)2, would not be reflected properly using
a resistivity test (Stanish et al., 2004).

2.2.4.3.3 Aggregate and electrical resistivity


Electrical resistivity of aggregates is larger than resistivity of the other components of
concrete (Neville, 1995). Table 2.4 contains electrical resistivity of common aggregates
used in concrete industry.

Table 2.4: Electrical resistivity of rocks (Monfore, 1968)

Type of Aggregate

Resistivity
(.cm)

Sandstone

18,000

Limestone

30,000

Marble

290,000

Granite

880,000

61
In contrast with the numbers presented in Table 2.4, resistivity of rocks embedded in
concrete (exposed to alkalies) is lower, hence resistivity of concrete is considerably
dependant upon the resistivity of its cement paste matrix (Monfore, 1968).
In addition, it has been derived that the electrical conductivity of concrete at a certain
degree of hydration is inversely proportional to the volume fraction of aggregate in
concrete (Xie et al., 1991 in Shi, 2004).
Some aggregates may release alkalis into the pore solution (reported for limestone
aggregate) which will have a significant effect on electrical resistivity and the RCPT
results of concrete (Grattan, 1994 in Shi, 2004)

2.2.4.3.4 Cement type and electrical resistivity


As the chemical composition of cement controls the quantity of ions present in the
evaporable water, electrical resistivity becomes more dependent on the cement used
(Neville, 1995). Figure 2.34 has shown that concretes made of different types of cement
have shown different resistivity values.

Figure 2.34: Relation between resistivity and applied voltage of different cement concretes with
W/CM= 0.49 (/eville, 1995)

62
Also is can be seen that electrical resistivity values and applied voltages are directly
proportional.

2.2.5 Influencing factors on durability tests


Among all affecting factors on durability test results, four of them are common for all
tests.

2.2.5.1 W/CM ratio


W/CM ratio represents the evaporable water and gel porosity in concrete (Neville, 1995).
Higher W/CM ratio concretes have more continuous pore systems and larger pore size
distributions resulting in higher total charge passed (Ahmed et al., 2009).
In the case of depth of chloride penetration, the apparent chloride diffusion coefficient
increases, as the pore volume increases. The transport of chloride ions has little to do with
the chemistry of pore solution, but it is influenced by pore structure characteristics (Shi,
2004) as there is an inverse relationship between penetrability and diffusion coefficients
as well (Savas, 1999).
Bassouni et al., (2006) have concluded that since chloride migration coefficient is only
influenced by the physical characteristics of the pore system, W/CM ratio is the most
governing factor for chloride ion penetration.
Since W/CM ratio represents the porosity and pore size distribution of concrete as higher
the ratio, higher the amount of porosity in concrete (Neville, 1995) water sorptivity of
concrete mixes with higher W/CM ratio is higher than that in mixes with lower ratio
(Nokken et al., 2002). In addition, for cement pastes hydrated to the same degree, the
permeability is lower the higher the cement content, i.e. the lower the W/CM ratio
(Neville, 1995).
It has been reported that the concrete resistivity trend over time is similar to that of
mechanical strength of concrete (significantly influenced by the W/CM ratio), so W/CM
ratio is an influencing factor on concrete electrical resistivity (Hansson, 1983). It affects
the electrical resistivity value of concrete in two ways:

63
a) Higher W/CM results in increases in the volume of evaporable water (conductive
material) in concrete. Therefore resistivity of cement paste decreases (Neville,
1995).
b) Electrical resistivity of hardened concrete is sensitive to the volume of porosity
and to the porous connectivity degree which are increased in higher W/CM ratio
concretes (Andrade, 2010).
These effects can be seen in Figure 2.35.

Figure 2.35: Relation between electrical resistivity and W/CM ratio at 28 days with different cement
contents (/eville, 1995)

64
Therefore, to provide a high electrical resistivity concrete, the W/CM ratio must be
minimized.
Also it can be concluded from Figure 2.35 that at a constant W/CM ratio, concrete with
lower cement content has higher electrical resistivity because of the less available
electrolyte for the current to pass. In case of cement content, for a constant W/C ratio,
increasing the paste volume creates more channels, pore water, for electrolytic
movement, so electrical resistivity is decreased (Elkey et al., 1995).

2.2.5.2 Supplementary cementitious materials


By adding SCMs to the concrete mixture, a denser concrete (improved particle packing)
which has finer and discontinuous pore structure is made (Neville, 1995) due to SCMs
secondary hydration products which blocks the pore system and makes it discontinuous.
It results in lower total charge passed and the level of permeability of concrete (Chini et
al., 2003).
The dependence of RCPT results on pore fluid conductivity has little relevance to the
chloride permeability of the concrete; therefore, for concrete contains SCMs, the
interpretation of chloride permeability based on the RCPT results becomes unrealistic
(Wee et al., 2000 in Chini et al., 2003). Concrete with ground granulated blast furnace
slag (GGBFS) as well as all SCMs has significantly lower coulomb values than a plain
concrete without SCMs because SCMs refines pore size; it also decreases alkali of the
concrete by decreasing the Na+ and OH ion from the solution (Ahmed et al., 2009).
Among all SCMs, applying the RCPT to concretes containing silica fume has been
criticized. The replacement of Portland cement with silica fume can reduce the electrical
conductivity of concrete more than 90% due to the changes in the chemical composition
of pore solution, which have little to do with the transport of chloride ions in concrete
(Shi, 2004). On the other hand, Pun et al. (1997) showed that using silica fume in
concrete mixes did not have any significant effect on ASTM C1202 chloride penetration
resistance results since there was not any significant difference between the concrete
behaviour from the RCP test and other standard tests measuring chloride diffusion
coefficient as shown in Figure 2.36.

65

Figure 2.36: Relative reduction in diffusion coefficient with silica fume (W/CM= 0.35)
(Pun et al., 1997)

Therefore, it has been recommended to measure the chloride migration coefficient similar
to Nordtest NT build 492 beside the total charge passed at the end of the RCPT in cases
of binary or ternary mixes containing silica fume (Bassouni et al., 2006).
Using SCMs results in lower pores tortuosity and sorptivity value (Neville, 1995). Silica
fume concretes has a much more substantial effect on permeability than normal Portland
cement (Johnston, 1992 in DeSouza, 2996).
Electrical resistivity of concretes containing SCMs is higher than for plain cement
concretes because the pore system becomes discontinuous and blocked as results
secondary hydration of SCMs (Chini et al., 2003). Discontinuity of the pore system
blocks the ionic movement in pore water, so electrons can not easily pass through pores
and must transfer across the high resistivity parts of concrete such as hydrated cement gel
(Hansson, 1983).
Electrical charges applied by the resistivity-meters electrodes must be taken by the ions
presented in pore solution, so concrete electrical resistivity decreases with increasing the
alkalinity of pore solution (Monfore, 1968). Therefore, adding SCMs results in a lower
alkali concentration or lower pH value in the pore solution because SCMs incorporate
more alkali into hydration products than they release (Shehata et al., 1999), so electrical
resistivity increases due to the lower ionic concentration (Elkey et al., 1995).

66
In analysing electrical resistivity, knowledge of pore solution is necessary because if a
concrete exhibits a high resistivity value, this may be due to a reduction in ionic
concentration rather than fine, tortuous pore structure (McCarter et al., 2009). This debate
is common especially when silica fume is used within the pore fluid compared to
Portland cement. a replacement of 5% Portland cement with silica fume decreases the
specific conductivity of pore solution to approximately 75% of that of Portland cement at
day 7 (Shi, 2004). Silica fume can reduce the ITZ porosity because silica fume is a super
fine material witch results in higher electrical resistivity (Neville, 1995).

2.2.5.3 Curing
Sarkar (1987) in Ahmed et al. (2009) reported that, normal concrete is characterized by
more open microstructure, increased presence and well crystallized formation Ca(OH)2,
and higher Ca/Si ratio compared to high performance concrete mixes. These
characteristics of C-S-H and CH in normal concrete result in very high charge passed.
As cement hydrates, gel porosity decreases in addition to lowering continuity of the pore
system due to cement hydration (Neville, 1995), resulting in lower charge passed, higher
electrical resistivity, and lower water sorptivity values. Mature concrete has more filled
pores which results in lower coulombs at later ages than early ages. In addition, the
reduced charge in later age can be attributed to the finer pore-size distribution which is
affected by the more cement hydration products (Bassouni et al., 2006).
Therefore, any action that increases cement hydration is beneficial; the longer the moist
curing period, the higher the degree of hydration, so the lower chloride permeability level
and higher electrical resistivity (Savas, 1999). Curing temperature is another influencing
factor on pore system. Standard cured concrete has lower 91 day coulomb values than
accelerated cured specimens (Savas, 1999). Normally cured concrete has lower chloride
ion diffusion than the high-temperature cured concrete at later ages (Stanish et al., 1997).
With the progress of cement hydration, water permeability values decrease rapidly
because the gross volume of gel increases, so the gel gradually fills some of the original
water-filled spaces (Neville, 1995). Therefore, longer curing period will reduce concrete
capillary pores water sorptivity.

67
In case of fixed water saturation level, electrical conductivity (inverse of resistivity)
changes as the solution in the pores changes (Kessler et al., 2004), so any factors
affecting the pore solution and pore system such as curing condition and type of
cementing materials change electrical resistivity values.
Although electrical resistivity of concrete increases with age due to more cement
hydration, the rate of resistivity increase is proportional to the length of curing period
(Monfore, 1968). In other words, for constant moisture content, electrical resistivity
increases with longer curing and more hydration.

2.2.5.4 Moisture content


Although the specimens are vacuum saturated prior to the RCP test according to ASTM
C1202-07, some scientists have tested semi-dry specimens. Moisture content between 4060% saturation (that pore water begins to gain or lose continuity) results in a 3%
fluctuation in the RCPT values per each 1% change in degree of saturation (Elkey, 1995
in Chini et al., 2003).
Concrete moisture content influences its water permeability: the higher the moisture
content of the concrete, the lower the measured sorptivity values (Neville, 1995). This
conclusion is seen in both types of sorptivity test results:
1) Filed water sorptivity values increase with decreasing level of concrete saturation
(DeSouza, 1996).
2) Although the relative humidity of the specimens used for laboratory sorptivity test
was constant, the same conclusion has been discussed by other researchers:
Nokken et al. (2002) have shown that the sorptivity decreases with increasing
degree of saturation and also decreasing W/CM ratio.

Concrete resistivity is sensitive to moisture content of concrete; a value of less than 1


K.cm for water-saturated concrete can increase to over 100 K.cm for the same
concrete when oven dried (Millard et al., 1989).

68
There are two main reasons for high electrical resistivity in dry concretes:
1) Water is conductive: Electrical resistivity increases when concrete dries out.
Generally, dried materials have more resistivity than wet materials, because of the
electrical conductivity of water filled the pore system (Monfore, 1968).
2) Discontinuity of Pore water system: Pore water begins to be discontinuous, at
moisture contents less than 80% which causes an increase in electrical resistivity
values (Neville, 1995).

It is recommended that for on-site measurements, a quick study of moisture effects is


performed before proper measurements (Lataste et al., 2003).

CHAPTER 3
EXPERIME/TAL PROGRAM
The main focus of this study was to develop non-destructive testing methods to evaluate
the durability of covercrete. Most of the report is discussing concrete surface electrical
resistivity measured by four-electrode instrument (Wenner probe). To improve reliability
of data measured by Wenner probe, it was crucial to calibrate the testing device with
other standard laboratory testing methods. In addition, standard time-consuming
durability tests such as ASTM C1202-07 were evaluated for comparison.
Therefore, testing methodology must be studied at the beginning point of the project.

3.1 Methodology
This methodology is to read as a narrative. The reason is to describe the testing methods,
required in this research program, in detail to maintain the consistency of the testing
procedures.
The testing methodology is based on available standard testing methods (e.g. ASTM
C1202, 2007). In case of the field sorptivity test and electrical resistivity test which have
not been standardized, the testing methodology is based on previous works presented in
the literature.

3.1.1 Compressive strength (ASTM C39, 2005)


The compressive strength test of moist cured concrete cylinders was done in accordance
to ASTM C39-05. The detailed methodology is described in Appendix B.

3.1.2 Rapid chloride permeability test (ASTM C1202, 2007)


According to ASTM C1202-07 the following steps should be taken during the rapid
chloride permeability test:
a) Sample preparation
b) The RCP test
Both sample preparation and the RCPT process were done according to ASTM C120207.
69

70

3.1.3 Water sorptivity test


Sorptivity test measures the rate of water absorption when only one surface of concrete
specimen is exposed to water. Capillary suction under unsaturated conditions is the
reasons for water absorption of concrete specimens.
Rate of water absorption or sorptivity of a concrete mixture at each age can be measured
by two types of sorptivity test; laboratory sorptivity test (destructive) and field sorptivity
test (non-destructive).

3.1.3.1 Laboratory sorptivity test (ASTM C1585, 2004)


As mentioned in ASTM C1585-04, the average of test results on at least two 100 6
mm diameter with a length of 50 3 mm specimens were used.
Specimens are obtained from either molded cylinders according to practices or drilled
cores according to test method. The cross sectional area of a specimen should not vary
more than 1 % from the top to the bottom of the specimen (ASTM C1585, 2004).
Laboratory sorptivity test was done according to ASTM C1585-04. The only difference
was sample conditioning procedure.
Since the amount of water absorbed is affected by the amount of water present in
concrete, the most important part of the lab sorptivity test is to find a best way for
conditioning concrete discs. There are two types of sample conditioning (to reach 50 to
70% internal relative humidity which is similar to the RH found near the surface in some
field structures according to ASTM C1585-04) recommended by the standard and other
researchers:

1) As standard mentioned test specimens should be placed in the environmental


chamber at a temperature of 50 2C and RH of 80 3 % for three days.
Alternatively, test specimens can be placed in a dessicator inside an oven at a
temperature of 50 2C for three days. The relative humidity in the dessicator is
controlled with a saturated solution of potassium bromide (KBr). The solution
should be placed in the bottom of the dessicator to ensure the largest surface of
evaporation possible, so specimen is not in contact with the solution. After 3
days, each specimen must be placed inside a sealable container. Precautions must

71
be taken to allow free flow of air around the specimen by ensuring minimal
contact of the specimen with the walls of the container. The container is stored at
23 2C for at least 15 days before the start of the absorption procedure. At the
time of testing, internal relative humidity is about 50 to 70% which is similar to
the relative humidity found near the surface in the some field structures (ASTM
C1585, 2004).
2) Previous works have found that three days oven drying at 50C followed by
sealed drying for four days at 50C provides a rapid and convenient method of
obtaining a uniform moisture distribution and a surface relative humidity of 50 to
60% and moisture contents above 1.0% (Parrott, 1994). This amount of relative
humidity represents in-situ concrete condition. It is worth mentioning that
ettringite destroys at 70C (Neville, 1995), so 50C is the best conditioning
temperature because micro-structural cracking is minimized in this temperature.

The second type of sample conditioning was used in this research program.
The testing process and calculation were done according to ASTM C1585-04.

3.1.3.2 Field sorptivity test


The permeability of outer zone of concrete can be different from that of the bulk concrete
due to compaction, bleeding, finishing and curing, as well as the choice of constituent
materials (DeSouza et al., 1997), so studying the outer layer of skin penetrability is
necessary to analyse the durability of concrete. An updated and non-destructive sorptivity
testing technique was created at the University of Toronto for measuring the water
penetrability of concrete surface. The apparatus used during this sorptivity measurement
was described in Section 2.2.3.3.1.

3.1.3.2.1 Methodology
There was not any standard methodology for the field sorptivity test. The main concept
beyond this measuring technique was to attain the best approximation of unidirectional
flow within the inner chamber. Therefore, the outer chamber needed to be full off water
before the inner domed area. The outer chamber saturated the concrete beneath, so the

72
water penetrating into the concrete under inner chamber flowed uni-directionally. Two 6
mm hose barbs were housed on both the interior and outer chamber as shown in Figure
3.1.

Figure 3.1: Hose barbs attached to inner and outer chambers of a field sorptivity test apparatus

The hose barb at the higher point of the dome was attached to the water column
(graduated pipette), while the lower hose barb was attached to the water reservoir to fill
the domed area. The lower hose barb in the outer chamber was attached to the water
reservoir, while the upper hose barb was open to the atmosphere (air vent). All
connections were made of 6 mm Quick Disconnect Shutoff valves, to allow for easy
connection and removal.

73
In order to produce accurate and reliable results, the following procedures were
implemented (DeSouza et al., 1996):

Step I) Specimen preparation: Specimens preparation for the filed sorptivity test had
two major parts:
1) If it is not possible to pre-condition concrete in situ to standard moisture content,
moisture effect on the rate of absorption must be considered. In this project concrete
samples were weighed each time prior to testing. At the end of the test the moisture
content was calculated from the fully dried mass of the specimens.
2) In order to achieve a maximum vacuum pressure, the section of the specimen
underneath the outer guard ring had a smooth flat finished surface, without cracks.
This surface area could be either ground or epoxy.

Step II) The sorptivity base plate was placed on the specimen surface with the smooth
surfaces of the specimen between two circular Neoprene sponge gaskets. To ensure the
gasket did not traverse any large voids or cracks, the base plate was pressed manually
against the surface.

Step III) The vacuum pump was engaged while the sorptivity base plate was pushed
downward to obtain a vacuum between two the circular neoprene sponge gaskets. The
vacuum pump had to be adjusted to approximately 380 mm HG (DeSouza et al., 1996).
The pump should reach a steady-state condition in approximately 30-60 s (vacuum was
applied during the full test duration).

Step IV) After 2 minutes vacuuming, the water reservoir (tap water 23 2C) was
attached to the lower hose bard of the outer chamber. Outer chamber was filled while the
other outer hose barb acted as an air vent. The timer was started with the first contact of
water to concrete surface. Now the area underneath the outer chamber was going to be
saturated which caused unidirectional flow in the inner chamber as shown in Figure 3.2.

74

Figure 3.2: Schematic overview of the field sorptivity process

Step V) After 2 minutes and 30 seconds, saturation period of the outer ring, water
reservoir was connected to the lower hose barb of the inner domed area until water came
up the full length of graduated pipette tube and started dripping from it. After removing
all air bubbles from the tube, the hose was detached from the water reservoir.

Step VI) A volume of water was placed in the apparatus. The initial height off the water
above the concrete slab was recorded immediately after the last air bubble was removed.
Concrete started to absorb water. The height of water was being measured as water was
absorbed into the surface of the concrete and drawn along the measuring pipette.
Readings were taken every minute (after initial contact of water with the test surface) for
the first 10 minutes, then at 12 and 16 minutes (the tail sorptivity curve). During all
readings, vacuum pump needed to be monitored to ensure that it remained constant.

Step VII) After the 16 minute reading, the vacuum pump was turn off, remained water
was drain from the system, and the apparatus was removed from concrete surface.

75

Step VIII) Calculation


The rate of absorption, I (mm), was more generally calculated using the following
relationship as expressed by Hall (1989) in DeSouza et al. (1998):

I=

mass
,
Area

where, mass is the change in mass of the sample (g), Area is the cross sectional area of
test specimen (mm), and is fluid density (g/mm3). If the relationship between I and the
square root of time is well represented by a straight line, the sorptivity, S (mm/min)
should be determined by the slope of the least-squares linear regression line of I against
t .
As discussed before, it was not possible to condition concrete in-situ to standard moisture
content, so it was necessary to correlate the effect of moisture content on the rate of
absorption. Moisture content ([test mass- dry mass at 110c] / dry mass at 110c) of a
concrete sample was important to calibrate the instrument and crate a calibration curve
for the sorptivity test. Therefore, every time concrete specimens w weighed before the
field sorptivity test.

3.1.4 Electrical resistivity


Three types of electrical resistivity were measured in this research program: the RCPT
resistivity (explained in Section 2.2.2.1), DC-cyclic bulk resistivity (Monfore resistivity),
and surface electrical resistivity (Wenner probe). There was no any standard test
methodology for electrical resistivity, so the testing method (surface and bulk electrical
resistivity) is briefly explained.

3.1.4.1 Methodology
DC-cyclic resistivity of concrete specimens was measured by taking the following steps:

Step I) Concrete specimen end faces were ground flat prior to the test.

76

Step II) Concrete specimen was placed between two stainless steel electrodes. Since
electrode/concrete interface resistance was the major cause of data scatter, salt free water
based contact gel named as Spectra 360 was used as shown in Figure 3.3.

Figure 3.3: DC-cyclic bulk electrical resistivity test set up (concrete disc between electrodes)

Step III) By turning on the instrument, a cyclic voltage was applied across the specimen
between 3 and 5 volts every 5 seconds. Electrical current passed trough specimen was
measured and reported by the instrument.

Step IV) Based on the voltage and current reported by the instrument after 15 min.,
electrical resistivity was calculated by the following equation:
=

(V5 V3 ) A
(K.cm),
(I5 I3 ) L

where, V3 and V5 are average applied voltage for 3 and 5 volts respectively, I3 and I5 are
average applied electrical current (A) for 3 and 5 volts respectively, A is specimens
cross-section area (cm2) and L is specimens thickness (cm).

77
For measuring surface electrical resistivity, a four-probe line array (Wenner probe) was
used in all testes presented and analysed in this report.

Following steps were taken during surface electrical resistivity measurement of a


concrete specimen during this research project:

Step I) The equally spaced probes was adjusted to that required spacing by sliding the
probes along the guide rails and connect the probe to the instrument using the cable
provided.

Step II) The spacing control value (cm) was set to that of the probe spacing from step I.
Step III) The plastic covers was removed from the four electrodes tips (All wooden
electrodes should have been wet at the time of testing as well as during the measurement;
otherwise measured data were not reliable).

Step IV) Any surface water was removed by a damp rag from concrete sample surface to
avoid any short connection.

Step V) Probes were placed in contact with the surface of the concrete specimen and a
firm steady pressure downwards was maintained on the probes as shown in Figure 3.4.

Figure 3.4: Wenner probe being used to measure surface resistivity of a concrete cylinder

78

Step VI) The surface electrical resistivity of the concrete sample displayed on the meter
screen in K.cm was recorded.

Step VII) The probes were left off the surface of the concrete and moved them to the
next position.

If the screen reading was unsteady of drifting slowly there would be a contact problem.
For bulk resistivity test, more contact gel must be used and specimen end faces must be
completely ground. In case of the Wenner resistivity the probe tips should be soaked in
water for a few more minutes. Contact problems will be prevented by using fully water
saturated probe tips.
Concrete electrical resistivity which is required to limit the rate of steel corrosion remains
virtually unchanged in a saturated pore network (Lopez and Gonzalez, 1993). In this
situation, resistivity values represent pore system continuity and ionic concentration in
pore solution, so corrosion rate decreases as a result of chloride diffusion control
(Monfore, 1968). All concrete specimens tested for electrical resistivity were fully
saturated in this research program.

3.2 Experimental project


Before starting the project, different concrete mix design specimens were used in order to
understand the operation of the instrument and to make modification to it.
All required tests, required materials, mix design, curing regime, and testing schedule for
this research program are presented in this chapter.

3.2.1 Research project tests


Three durability tests were conducted in this research project: Wenner probe and bulk
electrical resistivity, rapid chloride permeability, and rate of water absorption.
Besides, compressive strength of concrete cylinders was measured.

3.2.1.1 Electrical resistivity of concrete


Three types of electrical resistivity were measured at each age in this research program:

79

3.2.1.1.1 Surface electrical resistivity


The CNS Farnell RM MKII, shown in Figure 3.5, made by CNS FARNELL Ltd.,
England, was used in this research project for measuring concrete electrical surface
resistivity. The resistivity meter was developed by engineers from Taylor Woodrow.

Figure 3.5: RM MKII (surface resistivity-meter)

By using a standard Wenner linear four- probe array, water-saturated wood contact
points, a flat-topped AC wave form for the current source and sophisticated electronic
circuity, a true measure of the DC component of resistivity is measured.

3.2.1.1.1.1 RM MKII technical properties


Two outer probes apply low-frequency alternating current to concrete while the
instrument measures voltage drop between its two inner probes as shown in Figure 3.6.

Figure 3.6: Schematic representation of four-electrode resistivity test (Wenner method)


(http://www.canin-concrete-corrosion.com/analyzing-methods.html)

80
Probe spacing can be adjusted between 0-100 mm. Spacing between probes can be
influenced by several factors as well as the thickness of concrete member, location of the
probe array from edges of concrete specimen, and the thickness of concrete cover.
Sinusoidal electrical current applied can be adjusted from 20A to 2mA. Alternating with
a flat topped, trapezoidal waveform at a frequency of about 13 Hz. A low frequency
should be used to eliminate capacitive effects (Millard et al., 1989). The amount of
electrical current passed through concrete remains constant and limited to a certain value
recommended by the manufacturer. If the electrical current showed by the instrument was
less than the recommended amount in the manual, the electrical current had to be changed
to an acceptable range. By adjusting the applied electrical current, sufficient current flows
between the outer probes, so the concrete surface electrical resistivity measured is
reliable.
There were four possible basic ranges of full-scale resistivity measurement: 0-2 K.cm,
0-20 K.cm, 0-200 K.cm, and 0-2 M.cm.
The basic accuracy was reported to be 2% of reading, + 3 digits, for current drives
down to 1/20th of the nominal constant value assuming equal contact resistances on the
current probe. The Wenner probe was calibrated, so it was not necessary to calibrate the
instrument by a metal sheet before the project.

3.2.1.1.1.2 Surface electrical resistivity measurement


Three 100 x 200 mm water saturated concrete cylinders were tested at each age. Four
longitudinal readings (90 degrees apart) along the cylinder height from each concrete
cylinder were measured as shown in Figure 3.7.
Hence, reported surface resistivity at each time was an average of twelve readings (4 x 3
cylinders).

81

Figure 3.7: Surface electrical resistivity measurement order for concrete cylinders

Eight measurements on the finished surface of each concrete slab were read. Therefore,
reported surface resistivity at each time was an average of sixteen readings (8 x 2 slabs).
Data from previously studies were used to find the optimum probe spacing in the first
trial. Previous work has shown that the factors shown in Figure 3.8, scattering the
resistivity values measured by the Wenner probe, were needed to be considered in order
to find the optimum probe spacing.

Influencing
Factors on
Probe Spacing

Specimen

Specimen
Thickness

Edge Effects

Concrete

Rebar Locations
(Covercrete Thickness)

Figure 3.8: Influencing factors for probe spacing in the Wenner resistivity

Maximum
Aggregate Size

82
To have a homogeneous current flow, electrodes must be far from the reinforcing bars.
All concrete samples used in this project were non-reinforced specimens, so steel location
was not an affecting factor on determining the probe spacing.
In practice it has been found that a significant error occurs if resistivity measurements
are taken on a thin concrete section or near to an edge (Gowers and Millard, 1999). To
avoid this problem, Figure 3.9 provides a relation between probe spacing and resistivity.

Figure 3.9: Effect of concrete section dimensions on surface resistivity measurement


(Gowers and Millard, 1999)

If the dimensions of a concrete element are relatively small, the current is constricted to
flow into a different field pattern to that shown in Figure 3.6. This will result in an overestimation of the evaluation of surface electrical resistivity of concrete (Gowers and
Millard, 1999). The optimum probe spacing used in this research project was selected
with regard to the specimens geometry.

83
As mentioned before, all resistivity equations had been derived from semi-infinite
samples. In case of infinite specimens either a constant correction, K, should be
considered or geometrical effects should be eliminated. Therefore, the edge effect must
be considered. Figure 3.10 provides a correlation between electrode distance from the
specimen edges and concrete surface electrical resistivity.

Figure 3.10: Effect of edge and end proximity on surface resistivity measurement
(Gowers and Millard, 1999)

It can be concluded from Figure 3.10 that the distance of probe contact from any element
edge should be at least twice the probe spacing.
Maximum aggregate size can also influence the space between electrodes. The influence
of individual aggregate particles on resistivity measurements is not significant if particle
the size is smaller than the Wenner contact spacing (Gowers and Millard, 1999).

84
Figure 3.11 illustrates the effect of maximum aggregate size on concrete surface electrical
resistivity.

Figure 3.11: Effect of maximum aggregate size on surface resistivity measurement


(Gowers and Millard, 1999)

It can be concluded from Figure 3.11 that contact spacing is 1.5 times (or greater) as large
as the maximum size of the aggregate, 10 mm in this research program, to obtain a
standard deviation less than 5%.
Regarding to the last three correlations, the optimum probe spacing for the
100 x 200 mm concrete cylinders was 25 mm and for the 406 x 75 mm concrete slabs
was 15 mm. Due to the Wenner probes instrumental limitations, the optimum probe
spacing for circular slabs was changed to 20 mm because it was the minimum probe
spacing for Wenner probe.
For all surface resistivity measurements, a 50 mm probe spacing was necessary because
of the MTO instrument probe spacing. Besides, different probe spacing between the

85
optimum spacing and the common spacing, 50 mm, were chosen to study different probe
spacing effects on electrical resistivity.
Probe spacings used in this research program are summarized in Table 3.1.
Table 3.1: Probe spacings used in the research project

Probe Spacing (mm)*

Types of Specimen
Concrete Cylinders
(100 x 200 mm)

25

(4 readings at each age on each

(optimum)

30

40

30

40

50
(MTO)

concrete cylinder)
Concrete Circular Slabs
(Thickness = 75 mm)

20

(8 readings :Diagonally + Horizontally +

(optimum)

50
(MTO)

Vertically readings)
*Probe spacing should be more than 15.9 mm according to the Farnell probes manual

3.2.1.1.2 Cyclic-DC bulk electrical resistivity (Monfore resistivity)


Two bulk resistivity tests were performed in this program:

3.2.1.1.2.1 Cyclic-DC bulk electrical resistivity of full length cylinders


Three 100 x 200 mm water saturated concrete cylinders, used for surface electrical
resistivity measurement, were tested at each age. The average of three Cyclic-DC
resistivity readings was measured for calculating the electrical resistivity of concrete.

3.2.1.1.2.2 Cyclic-DC bulk electrical resistivity of concrete discs


A saturated concrete cylinder was sliced into three discs with thicknesses of 50 3 mm.

The bottom and middle slices were tested for the cyclic-DC bulk resistivity at each age
as shown in Figure 3.12.

86

Figure 3.12: DC-cyclic bulk resistivity of a concrete disc (test setup)

This test takes 15 minutes for each specimen.

3.2.1.1.3 Rapid Chloride permeability test resistivity (first 5 minutes)


This type of resistivity will be explained in Section 3.2.1.2.

3.2.1.2 Rapid chloride permeability test (RCPT)


In this test, chloride ions are forced into concrete by an external DC voltage on concrete
surface. For a concrete mix at each age, two 50 3 mm thick concrete discs sliced from
the bottom and middle part of a concrete cylinder were tested (100 x 50mm). The discs
were sliced by a water-cooled diamond saw and conditioned one day prior to the test time
under vacuum tap water for 3 hours according to ASTM C1202-07. After 18 2h, the
soaked specimens under water were removed, surface dried, measured for the average
section diameter and thickness, and tightly edges-sealed with vinyl electrical tape. The
sealed specimens were placed between two half cells. The half cell contacting the top
surface of the specimen was filled out with 3.0% NaCl (the catholyte) and the other cell

87
was filled out with 0.3N NaOH. Temperature of the cell contained NaCl was monitored
during the test. This temperature has to be less than 90C to avoid boiling of the solution
and damaging the cell.

Figure 3.13: The RCP test setup


(the black wire connected the /aCl cell to the negative terminal of the power supply while the other
wire was connected between the /aOH cell and the positive terminal)

The purpose of this test was not only to measure the total charge passed in coulombs but
also to measure the RCPT first 5 min. electrical resistivity (explained in Section 2.2.2.1).
As chloride ingress can be the most dominant factor that affects the rebar corrosion
process concrete, the depth of chloride ion penetration over 6 h. RCP test was measured.
At the end of the RCP test, samples were split open as shown in Figure 3.14.

88

Figure 3.14: Splitting concrete discs after The RCPT

Split samples were then sprayed with a 0.01 M silver nitrate, Ag/O3, solution to
determine the depth of the chloride penetration (colorimetric method).

Figure 3.15: Ag/O3 solution appears the depth of chloride penetration (AgCl2 white color)

This was used to calculate a non-steady-state diffusion coefficient using the equation in
Nordtest NT Build 492:

89

D=

(273 + T ) Lx d
0.0239(273 + T ) L
( xd - 0.0238
),
V 2
(V 2)t

where, D is non-steady-state migration coefficient (x 10 12 m 2 /s), V is applied voltage


(V), T is average value of initial and final temperature in the anolyte solution (C), L is
thickness of the specimen (mm), xd is average depth of chloride penetration (mm), t is
time (h).

3.2.1.3 Rate of water absorption (water sorptivity test)


Generally two types of water sorptivity test are common: laboratory and field sorptivity
test.

3.2.1.3.1 Laboratory sorptivity test


In addition to a disc extracted from a rectangular slab, top, middle, and bottom slices of a
concrete cylinder were tested according to ASTM C1585-04. Samples were conditioned
for seven days, three days oven dry 50C followed by sealed storage for four days at 50C
to obtain a uniform moisture distribution and obtain a relative humidity of 50-70%
(suggested by Parrott ,1994).

Figure 3.16: Laboratory water sorptivity test setup


(Exposed faces were covered with plastic sheets)

90
According to ASTM C1585-04, the maximum water level was 5 mm from the plastic
mesh located under concrete discs.

3.2.1.3.2 Field sorptivity


Lab sorptivity results were compared to the surface sorptivity test performed on two
406 x 75 mm circular concrete slabs at each time.

Figure 3.17: Field sorptivity test setup on a concrete slab

The average of two measurements was used to calculate the water sorptivity of concrete.

3.2.1.4 Compressive strength test (fc)


Three 100 x 200 mm concrete cylinders removed from the moist room prior to the test,
were used to obtain the compressive strength at each age: 3, 7, 28, 56, 91 days according
to ASTM C39-05. The concrete cylinders were end ground.

91
It is important to mention that concrete temperature effect was not studied in this research
program and all concrete specimens were tested in the room temperature, 23 2C.

All tests required in this research program are summarized in Figure 3.18.

Tests

Rate of
Water
Absorption
(Sorptivity)

Electrical
Resistivity

Surface
Electrical
Resistivity
(Wenner
Probe)

Concrete
cylinders
100 x 200
mm

Bulk
Electrical
Resistivity
(Monfore
Cyclic DC)

Concrete
circular
slabs
406 x 75
mm

Field
Sorptivity
Concrete
slabs
406 x 75
mm

Full length
specimens

Concrete
discs

100 x 200
mm

100 x 50
mm

Figure 3.18 (a): Research program concrete tests

Lab.
Sorptivity
Concrete discs
100 x 50mm

ASTM C1585

92

Tests

Rapid Chloride
Penetrability
Test
(RCPT)

RCPT
(6 hours)
100 x 50 mm
ASTM C1202

RCPT
Electrical
Resistivity

Depth of Cl ion
penetration
with spraying
AgNO3

Total charged
passed trough
concrete discs over 6
hours
(Coulombs)

Extrapolated
RCPT
Coulombs
(30 min. x 12)

Figure 3.18 (b): Research program concrete tests

Compressive
Strength
ASTM C39

93

3.2.2 Specimens
The sample types for this project include 100 x 200 mm concrete cylinders, 406 x
75mm circular concrete slabs, and 300 x 400 x 75 mm rectangular concrete slabs.

3.2.2.1 Concrete Cylinders


According to Table 3.2, thirty one cylinders were required for each concrete mixture.

Table 3.2: List of concrete cylinders for the project tests

Concrete Cylinders (100 x 200 mm)


TEST

3 day

7 day

28 day

56 day

91 day

Total

15

16

31

Compressive
Strength
and

Surface
Resistivity
RCPT
and

Bulk
Resistivity

2(Continuously

and

Cured)

Sorptivity
Total

The testing layouts for the concrete cylinder tests are shown in Figure 3.19.

94

Figure 3.19: Concrete cylinders testing layouts

3.2.2.2 Concrete slabs


Two 406 x 75 mm circular slabs and one 300 x 400 x 75 mm rectangular slab were cast
for each concrete mix. The concrete slabs were de-moulded after 24 h of casting and
cured for a further period of seven days. After seven days, slabs edge was sealed with
electrical tape and exposed in an environmental room at 23C and 50% relative humidity.

95
The circular slabs were used for field sorptivity tests at 14, 28, 56, 91 days and surface
electrical resistivity tests at 3, 7, 14, 28, 56, 91 days. Since a slabs moisture content was
an influencing factor on surface electrical resistivity (explained in Section 2.2.5), each
circular slab was weighed before the resistivity measurement (scale maximum capacity
was 30,000 5 g). After 91 days, the mass of all concrete slabs did not change which
means the moisture content of the slabs were in equal condition with the 50% RH room.
Moisture content at each age was calculated by the following equation

Moisture content at each age (%) =

Mass at each age - Mass at day 91


x100
Mass at day 91

At 28, 56, and 91 days, a core was extracted from the rectangular slab for the laboratory
sorptivity test.
Table 3.3 summarizes all the durability tests for the concrete slabs required in the
research program.
Table 3.3: Durability tests for the concrete slabs

Type of the concrete


slab

Size

Circular

406 x 75 mm

Rectangular

300 x 400x75mm

Durability Tests
I) Surface electrical resistivity
II) Field sorptivity test
Laboratory sorptivity test on extracted
cores

According to Tables 3.2 and 3.3, the amount of concrete required for each concrete
mixture was

Concrete Cylinders: 31 x (0.100 x 0.100 x 3.14/4) x (0.200) = 0.04867 m = 48.67 L.


Concrete Circular Slabs: 2 x (0.406 x 0.406 x 3.14/4) x 0.075= 0.01940 m = 19.40 L.
Concert Rectangular Slabs: 1 x 0.35 x 0.25 x 0.075= 0.009 m = 6.50 L.
Total Volume= 74.57 x 1.1 (allowing for waste) 82 L.

96

3.2.3 Materials
Concrete contains cementitious materials, aggregate, chemical admixtures, water, and air.
Now that physical and chemical properties or concrete components affect fresh and
hardened concrete properties, they must be measured before designing the mix design.

3.2.3.1 Cementitious materials


The main cementing material included an ordinary Portland cement (GU) meeting CSA
A3001. Supplementary cementitious materials included the ground granulated blast
furnace slag (GGBFS), and blended silica fume cement meeting CSA A3001.
Two types of cement were used as concrete paste; Ordinary Portland cement (OPC) and
silica fume blended Cement (Type GUb-8SF). 7-8 % silica fume provides a dramatic
improvement in chloride penetration resistance regardless of the test procedure used (Pun
et al., 1997). Silica fume made up of free silica has a large specific surface area, so it is a
fast in reacting material. Silica fume has two major limitations: it has a big demand of
water which causes a viscous or even dry concrete and silica fume is expensive.
The chemical composition (oxide analysis) and physical properties of cementitious
materials as measured by Holcim Canada Inc. are given in Table 3.4.

Chemical Composition (%)

Table 3.4: Chemical composition of cementitious materials


Holcim

Blended

Ground

GU

silica fume

granulated

Cement

cement

blast

(OPC)

GUb-8SF

furnace slag

SiO2

19.24

25.28

34.1

Al2O3

5.43

5.02

Fe2O3

2.36

1.98

0.7

CaO

60.94

55.97

41.8

MgO

2.34

2.30

6.3

SO3

4.11

3.71

2.4

K2O

1.11

1.10

0.34

3134

3080

2854

350

337

425

Density (kg/m )
2

Blaine Fineness (m /kg)

13.2

97

3.2.3.2 Aggregates
Crushed lime stone with a maximum nominal size of 10 mm meeting ASTM C33-07
were used as coarse aggregate. Washed natural siliceous river bed sand was used as fine
aggregate meeting ASTM C33-07. Coarse and fine aggregates used in this research
program were supplied by Dufferin Concrete.

3.2.3.2.1 Sieve analysis of fine and coarse aggregates


Sieve analysis of aggregates is necessary for control of the production of concrete. Sieve
analysis data is useful in developing relationships concerning concrete porosity and
packing.
Sieve analysis of fine aggregate was measured according to ASTM C136-06 as reported
in Table 3.5.
Table 3.5: Sieve analysis of fine aggregate
Sieve
Size
9.50 mm
4.75 mm
2.36 mm
1.18 mm
600 m
300 m
150 m
75 m
Pan

Mass Retained (g)


Sieve Cumulative
0
0
1.90
1.90
49.47
51.37
119.28
170.65
123.18
293.83
123.35
417.18
55.29
472.47
15.76
488.23
5.80
494.03

Percent Retained
Sieve Cumulative
0%
0%
0%
0%
10%
10%
24%
35%
25%
59%
25%
84%
11%
96%
3%
99%
1%
100%

Percent
Passing
100%
100%
90%
65%
41%
16%
4%
1%
0%

OPSS 1002
Minimum Maximum
100%
95%
100%
80%
100%
50%
85%
25%
60%
10%
30%
0%
10%
0%
3%
0%

To compare the results to the grading requirements of CSA A23.1, data have been plotted
in Figure 3.20.

98

100%

Percent Passing (%)

90%
80%
70%
60%
50%
40%
30%
20%
10%
0%
0.01

0.1

10

Sieve Size (mm)


Figure 3.20: Fine aggregate grading curve

Sieve analysis of coarse aggregates was measured according ASTM C136-06 as shown in
Table 3.6.

Sieve Size
19.00 mm
12.50 mm
9.50 mm
4.75 mm
2.36 mm
Pan

Table 3.6: Sieve analysis of coarse aggregate


Percent
Mass Retained (g)
Percent Retained
Sieve Cumulative Sieve Cumulative Passing
0
0
0%
0%
100%
421.0
421.0
8%
8%
92%
1525.4
1946.4
30%
38%
62%
2937.0
4883.5
58%
97%
3%
125.7
5009.2
2%
99%
1%
47.6
5056.8
1%
100%
0%

ASTM C33-07
Minimum Maximum
100%
90%
100%
40%
70%
0%
15%
0%
5%
0%

99
Data have been plotted along with CSA A23.1 grading requirements in Figure 3.21.

100%

Percent Passing (%)

90%
80%
70%
60%
50%
40%
30%
20%
10%
0%
1

10

100

Sieve Size (mm)


Figure 3.21: Coarse aggregate grading curve

Both fine and coarse aggregates grading graphs are within the standard grading limits.

3.2.3.2.2 Physical properties of fine and coarse aggregates


Coarse and fine aggregate properties such as density, sieve analysis, water absorption and
the fineness modulus were measured according to ASTM C127-07 and ASTM C128-04,
respectively.

100
The fineness modulus (FM) of sand was calculated from the sieve analysis of 160m,
315 m, 630 m, 1.25mm, 2.5mm, and 10 mm sieves.

Table 3.7: Sand sieve analysis required for the FM calculation


Sieve
Size
9.50 mm
4.75 mm
2.36 mm
1.18 mm
600 m
300 m
150 m
75 m
Pan
Total

Mass Retained (g)


Sieve Cumulative
0
0
1.90
1.90
49.47
51.37
119.28
170.65
123.18
293.83
123.35
417.18
55.29
472.47
15.76
488.23
5.80
494.03

Percent Retained
Sieve Cumulative
0%
0%
0%
0%
10%
10%
24%
35%
25%
59%
25%
84%
11%
96%
3%
------1%
--------285

F.M= (Total cumulative mass retained) / 100 = 285 / 100 = 2.85

In addition, aggregates physical properties were measured as tabulated in Table 3.8.

Table 3.8: Aggregates physical properties

SSD
Aggregate

Density

Fineness
Modulus

Kg
( 3)
m

Dry-Rodded
Density

Water
Absorption

Kg
( 3)
m

(%)

Fine Aggregate

2720

2.85

-----------

1.16

Coarse Aggregate

2702

----------

1582

1.38

3.2.3.3 Chemical admixtures


Water reducing admixture (WR) and Superplasticizer (SP), high range water reducer,
were added to fresh concrete to maintain its fresh properties such as workability
especially in low W/CM mixes. Also air entraining admixture (AE) was required. In
mixes containing water reducing admixture, water content was limited to 160 kg/m3
while in mixes with superplasticizer this limitation was 140~150 kg/m3. The water

101
content and the W/CM ratio were used to calculate the required cement content, as shown
in Table 3.9.
Table 3.9: Basic mix design basic data

W/CM

Cement
Content
(Kg/m)

Water
(Kg/m)

Admixture

Target Air
Volume
(m/1 m)

0.35

410

143.5

SP-AE

0.06-0.08

0.40

375

150

WR/SP-AE

0.06-0.08

0.45

355

160

WR-AE

0.06-0.08

All chemical admixtures were produced by BASF Construction Chemicals LLC. The
admixtures properties reported by the supplier are listed in Table 3.10.

Table 3.10: Chemical admixture properties


Recommended
Dosage by BASF

Density
Chemical

Commercial

Admixture

Name

Water Reducing
Admixture
(WR)

POZZOLITH
210

gr
)
cm 3

pH

1.138

Colour

Odour

Required

Co.

Standard
(

Dark Brown Odourless

ASTM
C494-08

mL
)
100 Kg
130-390

Air-Entraining

11.8

Admixture
Vinsol-Resin

MB-VR

1.030

(AE)

Dark Brown

Soapy

12.8

ASTM
C260-06

16-260

Superplasticizer
Based on
Polycarboxylate
(SP)

GLENIUM
7700

1.064

Purple
Brown

Mild

ASTM
C494-08

260-975

102

3.2.4 Mix design


The concretes cast in this research program represented a wide range of concrete
mixtures covering those typically used by Ministry of Transportation of Ontario. The
number of concrete mixtures was a combination of five mix designs and four re-tempered
mixes. For four of the mixes, additional batches were cast and left in the mixer for
approximately one hour, and then water was added to maintain the original slump. In this
way, the sensitivity of electrical resistivity and sorptivity values to re-tempering could be
assessed.
It is important to mention that concrete slump, set at 90 25 mm, was measured
according to ASTM C143-08 at the end of the mixing period.
The concrete mix variables for the project are presented in Table 3.11.
Table 3.11: Research program concrete mixes

Cement

w/cm

Cementitious content

0.35

0.40

0.45

410 kg/m3

375 kg/m3

355 kg/m3
X+

Portland cement

X+

25% slag

Silica fume cement and 25% slag

X+

X+

+ Additional samples were cast with re-tempering water added after a delay period (one hour)

Nine concrete mixtures were cast according to concrete mixes proposed in Table 3.11.
The required volume for each mixture was about 82 L which was enough for casting all
specimens. The W/CM ratios of 0.35, 0.40, and 0.45 were selected. The W/CM ratios and
aggregate content were corrected to SSD conditions. All cement used in this program
was normal Portland cement (GU) while for two mixes named high performance concrete

103
(HPC) and one water-to-cement ratio of 0.40, silica fume-cement (Type GUb-8SF) was
used.
Concrete mix design and other pertinent information used in this program are presented
in Table 3.12. Concrete mixtures labelled + were re-tempered. Detailed mix design is
presented in Appendix A.

Table 3.12: Concrete mixes proportions for 82 litre concrete (mix design)

Concrete
Mixture

HPC

Cement
(GU)
(kg)

Silica
fume

Sand

Cement

(kg)

Stone
10 mm

(kg)

Water

Slag

Added Water

(kg)

(kg)

re-tempering

(kg)

(L)

-------

24.6

57.9

87.2

9.6

8.2

-------

25.2

59.3

89.1

10.1

8.4

SFSL 0.40

-------

22.5

59.2

87.5

9.7

7.5

PCSL 0.40

23.1

-------

60.6

89.3

10.7

7.7

PCSL 0.40+

23.1

-------

61.2

89.7

9.7

7.7

PCSL 0.45

21.8

-------

60.0

88.8

11.9

7.3

PCSL 0.45+

21.8

-------

60.3

88.8

11.5

7.3

PC 0.45

29.1

-------

60.6

88.7

12.0

-------

PC 0.45+

29.1

-------

60.4

89.9

11.0

-------

(SFSL 0.35)
HPC+
(SFSL 0.35+)

0.95

0.71

The mixtures were selected to characterize a wide range of concrete mixtures from highperformance concrete (HPC) used in Ontario bridges to concrete that similar to mixtures
used in a residential application.
For each concrete mixture, a 10 L trial mix was used to find the optimal admixture
dosage. The optimum dosage is the admixture content which added to the concrete
mixture to maintain the expected fresh properties such as slump and air content.
The finalized admixtures dosages (mL per100 kg cement) are provided in Table 3.13.

104
Table 3.13: Chemical admixture dosages (mL/ 100 kg cement)

Air

Water

Entraining

Reducing

Admixture

Admixture

510

--------------

600

510

--------------

600

SFSL 0.40

400

--------------

450

PCSL 0.40

105

--------------

375

PCSL 0.40+

90

340

146

PCSL 0.45

75

340

--------------

PCSL 0.45+

75

340

--------------

PC 0.45

70

315

--------------

PC 0.45+

70

315

--------------

Concrete
Mixture

Superplasticizer

HPC
(SFSL 0.35)
HPC+
(SFSL0.35+)

3.2.5 Testing process


The testing process in this program depends on concrete age and type of concrete
specimens.

Concrete cylinders: Concrete cylinders were cast for the compressive strength, the DCCyclic bulk resistivity, and the surface electrical resistivity measurements. Besides, the
total charge passed through the concrete samples was measured by the RCPT method at
each age except day 3. Therefore, two concrete cylinders were sliced and conditioned
according to ASTM C1202-07 a day prior to testing.
At 28, 56 and 91 days one concrete cylinder was sliced into three 100 x 50 mm discs in
addition to a core extracted from the rectangular slab for the laboratory water sorptivity
test. All concrete slices were prepared according to the method described before (three
days oven drying at 50C following by four days sealed in 50C). It was assumed that
cement hydration slowed to near zero during the seven days conditioning period.

105
Therefore, the laboratory sorptivity was measured seven days from the day that cylinders
were removed from the fog room.

Circular slabs: Circular slabs were cast for two tests: surface electrical resistivity and
field water sorptivity.
The surface electrical resistivity was the average of two measurements on both circular
slabs at 3, 7, 14, 28, 56, and 91 days by the 4-electrode Wenner probe.
Since a saturated concrete slab did not absorb water, the field sorptivity test was
performed on the concrete slab aged more than 7 days (removing day from the moist
room). Therefore, the testing time for the field sorptivity test was 14, 28, 56, 91 days.

3.2.5.1 Concrete methodology


In order to have consistency in all concrete mixes to allow comparison of results, all
material preparation, mixing, casting, and curing steps were performed in the exact same
manner meeting ASTM C192-07, as described in the following steps.

3.2.5.1.1 Material preparation


Two days prior casting, coarse aggregate, twice as needed ,was washed to remove dust
and excessive fines, and to avoid the intrusion of chloride in the ingredients of the mixes.
The coarse aggregate was placed on a big pan over night to drain. A day later, the fine
aggregate was placed on a pan and mixed to maintain constant moisture content. The
washed coarse aggregate was mixed to maintain constant moisture content. Finally the
aggregates were sealed in pails to prevent any moisture changes. Two samples were taken
from coarse and fine aggregate containers.

After weighing the samples, they were dried in 110 C for 24 hour; moisture content was
calculated as:

Moisture content (%) =

Initial Mass - Dried Mass at 110C


x100
Dried Mass at 110C

106

3.2.5.1.2 Mixing concrete


Immediately prior to the mixing, the masses of coarse aggregate, fine aggregate, and
water were adjusted to compensate for the respective moisture content.
A horizontal rotating cum flow-pan type 150 litres capacity EIRICH model EAG21 mixer
was used for preparing concrete mixes. All materials were mixed in the EIRICH
mechanical mixer in accordance with the ASTM C192-07 standard procedure.
To minimize the water absorption by the mixer, the inner wall and blades were damped
with wet burlap. The weighed materials were added in the mixer drum in the following
order:
I- Coarse aggregate (stone)
II- Cement + SCMs
III- Fine aggregate (sand)
IV-Air entraining admixture (placed on top of the fine aggregate prior to mixing to avoid
absorption by the stone) as shown in Figure 3.22.

107

Figure 3.22: Materials loading order

According to the ASTM C192-07, alternating periods of mixing and rest were used:
(a) Mixing dry materials for 30 seconds prior to adding water and the water reducing
admixture (or superplasticizer).
(b) Mixer was turned off after 3 minutes mixing, 3 minutes rest and 2 more minutes
mixing.
Total mixing period = 3 minutes mixing (30 seconds dry materials, then add water and
mix for 2.5 minutes) + 3 minutes waiting + 2 minutes mixing.
Water reducing admixture was combined with the mix water prior to application into the
mixer. On the other hand, the optimum superplasticizer effect is generally obtained with a
delayed addition (Neville, 1995), so the superplasticizer was added to the mixer, 30
seconds after adding water.

108

3.2.5.1.3 Fresh concrete properties


At the end of the mixing period, fresh concrete properties were measured as shown in
Table 3.14. Fresh properties of re-tempered concrete mixes were changed after adding
water as shown in the table.
Table 3.14: Fresh concrete properties

Concrete
Mixture

Design
W/CM

Air

Actual
W/CM

SCMs

Content

Slump

Concrete

(ASTM C143-08)

Temperature

(mm)

(C)

(ASTM C173-10)

Density
(

Kg
)
m3

Ratio

Ratio

0.35

0.35

SF-SL

7.4

95

25

2349

0.35

0.38

SF-SL

7.8 8.0

85 100

26 26

2390 2277

SFSL 0.40

0.40

0.40

SF-SL

5.8

85

25

2410

PCSL 0.40

0.40

0.40

SL

5.9

85

24.5

2405

PCSL 0.40+

0.40

0.43

SL

8.4 8.0

85 105

24 24

2424 2324

PCSL 0.45

0.45

0.45

SL

6.9

105

25

2313

PCSL 0.45+

0.45

0.47

SL

8.2 8.0

115 110

24 23.5

2291 2272

PC 0.45

0.45

0.45

-------

6.1

80

23.5

2410

PC 0.45+

0.45

0.52

-------

6.0 5.8

85 110

23 23

24032359

HPC
(SFSL 0.35)
HPC+
(SFSL 0.35+)

(%)

3.2.5.1.4 Casting concrete


After measuring the fresh concrete properties, concrete was cast into polyethylene
cylindrical and steel rectangular and circular moulds.
Concrete slabs were cast in two layers. Each layer was consolidated by placing the slab
on the vibrating table for up to 10 seconds.
After casting all specimens, cylindrical moulds were capped with polyethylene caps and
covered by wet burlap. Concrete slabs were finished with a magnesium float after initial
set of the concrete as shown in Figure 3.23.

109

Figure 3.23: Magnesium float for finishing the concrete slabs surface

All three (two circular and one rectangular) concrete slabs, were covered with plastic
sheets followed by wet burlap for 24 hours. The temperature in the laboratory varied
from 21 to 26C.

Figure 3.24: First 24 h concrete curing under plastic sheets and wet burlaps

110
24 hours later, all specimens were demolded and labelled.

3.2.5.1.5 Curing
The curing regime in this research program had three parts:

1- First 24 hours before demolding: Concrete samples were sealed or covered to


prevent moisture loss during the first 24 hours.

2- First seven days: Since cement hydrates at a much reduced rate when the internal
relative humidity drops below 100% and ceases at approximately 80% (Neville, 1995),
initial curing is important. Therefore, concrete specimens were kept in the moist room for
seven days. It was recommended not to use saturated lime water for the following reason:
The specimens should not be cured in a saturated lime water tank, as this curing
condition decreases the resistivity of the concrete (Florida Method of Test for Concrete
Resistivity, FM 5-578).

3- After day seven: After seven days moist curing, concrete slabs were removed from
the moist room. The edges were sealed, so they dry from the top and the bottom surfaces.
Concrete slabs were exposed in an environmental room at 50 4 % relative humidity and
23 1 C according to ASTM C157-08. Concrete cylinders were kept in the fog room
until the test time. Therefore, concrete cylinders were fully saturated while concrete slabs
were partially dried.
All the slabs were weighed after sealing the edges. These masses, masses of the fully
saturated concrete, were required for calculating the moisture content of the specimens at
different ages.

CHAPTER 4
RESULTS A/D DISCUSSIO/
Tables of results included in this chapter depict final results or averages from testing
procedures.

4.1 Compressive strength


The average compressive strength of three 100 x 200 mm water saturated concrete
cylinders is shown in Table 4.1. The rate of loading was recommended to be within the
range of 1.14-2.76 KN/S. The rate of loading applied during this research project was
1.9-2.1 KN/S. Most of the types of fracture were cone while the rest was shear. If a
crushed cylinder had an unusual type of fracture, the results would be ignored according
to ASTM C39-05.

Figure 4.1: Crushing a concrete cylinder during compressive strength test (shear fracture)

The compressive strengths of concrete mixes at different ages are tabulated in Table 4.1.
It is important to mention that each value presented in Table 4.1 is the average of three
strengths obtained from three cylinders.

111

112
Table 4.1: Average compressive strength of different mixes at various ages

Concrete
Mixtures

Actual
W/CM

HPC
HPC +
SFSL 0.40

0.35
0.38
0.40

PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

0.40
0.43
0.45
0.47
0.45
0.52

SCMS
Day 3
SF, SL 45.67
SF, SL 44.57
SF, SL 42.84
SL
SL
SL
SL
-

35.79
32.48
30.26
27.00
31.54
30.45

Compressive Strength
(MPa)
Day 7 Day 28 Day 56
57.41 68.58 73.73
54.02 61.87 65.72
52.97 59.84 61.89
39.18
34.87
31.54
29.84
33.87
31.94

49.34
44.43
40.28
38.08
37.28
34.11

52.02
47.32
44.22
41.73
39.18
37.43

Day 91
76.46
68.22
63.28
55.49
49.38
45.82
44.41
43.13
40.13

The compressive strength test results of all concrete mixes are plotted in Figure 4.2.

S ilic a fu m e + S la g

C o m p re s s iv e S tre n g th (M P a )

70

S la g

60
50
40
No SCMs

HPC
(SFSL 0.35)

W /C M
+<===================

80

HPC+
(SFSL 0.38)
SFSL 0.40

PCSL 0.40

PCSL 0.40+
(PCSL 0.43)
PCSL 0.45

PCSL 0.45+
(PCSL 0.47)

30

PC 0.45

20
03 7

28

Age (day)

56

91

PC 0.45+
(PC 0.52)

Figure 4.2: Compressive strength of concrete mixtures at various ages

As shown in Figure 4.2, compressive strength increased rapidly during the first 28 days.
In concrete mixes with silica fume (HPC, HPC+, and SFSL 0.40 which are three top
graphs in Figure 4.2) rate of strength gain was significantly high within concrete early-

113
ages even in the first 7 days because silica fume made a significant contribution to earlyage strength. In concrete mixes without any supplementary cementitious material (PC
0.45 and PC 0.45+) compressive strength increased but not so much at later ages. In early
ages, compressive strengths of 25% slag replacement mixes were lower than plain
cement concretes (similar W/CM ratio) since slag contributed in later ages. Rate of
strength gain reduced after 28 days in all mixes especially in mixes without slag because
slag reacted later.
Factors affecting compressive strength are W/CM ratio, the presence of SCMs, concrete
age, degree of consolidation, air-void system and type of curing.

4.1.1 Effects of changing W/CM ratio on compressive strength


Changing W/CM ratio influences concrete porosity. Figures 4.3, 4.4, and 4.5 present

C om p ressive Stren gth (M Pa)

compressive strength of concrete mixes at different ages.

80
75
70
65
60
55
50
45
40
35
30
25
20

W/CM=0.35

W/CM=0.38
W/CM=0.40

HPC
(SFSL 0.35)

HPC+
(SFSL 0.38)

SFSL 0.40

03 7

28

Age (day)

56

91

Figure 4.3: Compressive strength of SFSL concrete mixes

Compressive Strength (MPa)

114

60
W/CM=0.40

55
50

W/CM=0.43
W/CM=0.45

45

W/CM=0.47

40

PCSL 0.40

35

PCSL 0.40+
(PCSL 0.43)

30

PCSL 0.45

25

PCSL 0.45+
(PCSL 0.47)

20
03 7

28

Age (day)

56

91

Figure 4.4: Compressive strength of mixes contain slag

Compressive Strength (MPa)

50
45

W/CM=0.45

40

W/CM=0.52

35
30

PC 0.45

25
PC 0.45+
(PC 0.52)

20
03 7

28

Age (day)

56

91

Figure 4.5: Concrete strength of plain cement concretes

It can be concluded from the figures that in all concrete mixes, higher W/CM ratio results
in less compressive strength. More water in concrete mix causes more porosity and
lowers compressive strength.

115

4.1.2 SCMs effects on compressive strength


Three different types of concrete mixture were cast in this research project: Plain cement
concrete mixes, concrete mixes with 25% slag replacement, and concrete mixes with
25% slag and 8% silica fume replacement.
It is concluded from Figure 4.2 that concrete mixes with slag and silica fume had higher
early and later age compressive strength. As the PCSL 0.45 mix had less Portland cement
content than the PC 0.45 mix, its 7 days compressive strength is lower while its strength
is higher at later ages as seen in Figure 4.2. It can be explained as the pozzolanic
reaction between the SiO2 in slag and the Ca(OH)2 is a slower reaction than the clinker
hydration reaction (Hansson, 1983).
The addition of silica fume may accelerate the growth of compressive strength compared
with addition of slag.

4.1.2.1 Comparison between compressive strength of SCMs concretes


Both silica fume and slag improve concrete strength because of their secondary
hydration. Figure 4.6 compares compressive strength of two concrete mixes with the

Compressive Strength (MPa)

same W/CM ratio.

70
60
SFSL 0.40

50
40

PCSL 0.40

30
20
03 7

28

Age (day)

56

91

Figure 4.6: Comparison between compressive strength of ternary and binary concrete mixes
(W/CM= 0.40)

116
Since silica fume is finer than slag, it hydrates earlier and makes a significant
contribution to early-age strength of concrete while slag affects later-age strength. In
conclusion a concrete mixture containing silica fume and slag shows higher strength than
a concrete mixture with slag and both concrete mixes are stronger than a plain cement
concrete mix.

4.2 Rapid chloride permeability test (RCPT)


Three different properties were measured from the ASTM C1202-07, RCP testing: (a)
total charged passed through concrete over 6 hours, (b) the RCPT first 5 minutes
electrical resistivity, and (c) chloride migration coefficient from depth of chloride
penetrated during the test.

4.2.1 Total charge passed


The RCPT methodology was briefly described in Section 3.1.2. The ASTM C1202-07
standard presents a table to guide the interpretation of results. This table is presented here
as Table 4.2.

Table 4.2: Chloride ion penetrability based on charge passed (ASTM C1202, 2007)

Charged Passed (Coulombs) Chloride Ion Penetrability


> 4000

High

2000 - 4000

Moderate

1000 - 2000

Low

100 - 1000

Very Low

< 100

Negligible

Total charged passed through two 100 x 50 mm concrete discs, simulating concrete
above the rebar level are presented in Table 4.3. Three 50 mm thick discs were cut from a
100 x 200 mm cylinder. The middle and bottom discs were used for the RCP test. The
total charge of each mixture at each age is the average of charge passed through middle
disc and bottom disc.

117
Table 4.3: The RCP test results

Age (day)
Mix
Design

7
Charged
Passed

28

Penetrability
Level

(Coulombs)

Charged
Passed
(Coulombs)

91

56

Penetrability
Level

Charged
Passed

Charged

Penetrability
Level

(Coulombs)

Passed

Penetrability
Level

(Coulombs)

HPC

754

Very Low

231

Very Low

219

Very Low

125

Very Low

HPC+

822

Very Low

241

Very Low

229

Very Low

165

Very Low

SFSL 0.40

914

Very Low

292

Very Low

247

Very Low

186

Very Low

PCSL 0.40

1621

Low

847

Very Low

640

Very Low

403

Very Low

PCSL 0.40+

2585

Moderate

1044

Low

808

Very Low

636

Very Low

PCSL 0.45

2938

Moderate

1272

Low

810

Very Low

705

Very Low

PCSL 0.45+

3256

Moderate

1258

Low

1004

Low

744

Very Low

PC 0.45

5154

High

2665

Moderate

1906

Low

1565

Low

PC 0.45+

6014

High

3922

Moderate

2423

Moderate

1918

Low

RCPT results are plotted in Figure 4.7. The bottom three lines represent the total charge
passed through specimens which contained silica fume. The two top lines represent

6500
6000
5500
5000
4500
4000
3500
3000
2500
2000
1500
1000
500
0

PC 0.45+
(PC 0.52)
PC 0.45
PCSL 0.45+
(PCSL 0.47)
PCSL 0.45

W /CM

<================

A verag e P assed C h arg es (C oulombs)

specimens without any SCMs and with the highest W/CM ratio.

28

Age (day)

56

Figure 4.7: The RCPT coulomb values with age

91

PCSL 0.40+
(PCSL 0.43)
PCSL 0.40
SFSL 0.40
HPC+
(SFSL 0.38)
HPC
(SFSL 0.35)

118
RCPT results for different types of mix design are shown in Figures 4.8, 4.9, and 4.10.

Figure 4.8: Total coulombs passed (silica fume and slag concrete mixes)

Figure 4.9: Total coulombs passed (slag concrete mixes)

119

Figure 4.10: Total coulombs passed (plain cement concrete mixes)

A comparison of the charge passed for different mixtures in Figures 4.8, 4.9, and 4.10 has
shown that all ternary mixes (8% silica fume and 25% slag replacement) were ranked
very low with a total passed charges less than 1000 coulombs. Binary mixes were ranked
moderate at 7 days (except W/CM= 0.40 mix). The reduction in coulombs passed is
significant for the slag mixes from approximately 2700 at 7 days to less than 700 at 91
days. Plain cement concrete mixes (GU cement concretes without any SCMs) were
ranked high at 7 days, but this rank is improved over 91 days.
The mobility of pore water ions through the cement paste porosity influences the total
charge passed, so any factors affecting the ionic quantity and ionic mobility, influences
the RCPT coulombs.

4.2.1.1 Effects of changing W/CM ratio on the RCPT coulombs


As mention in Section 2.2.5, higher W/CM ratio results in higher porosity, so the total
charge passed is higher in lower W/CM ratio concrete as shown in Figure 4.7. In high
W/CM ratio concrete mixes, the pore structure is more connected than the pore structure

120
in low W/CM ratio concretes (Neville, 1995) which results in higher electrical charge
passed over the 6 h test. Also Table 4.3 indicates that as W/CM ratio increases, chloride
ion penetrability ranking increases from very low to high.

4.2.1.2 SCMs effects on RCPT values


It can be seen in Figure 4.8 that the ternary silica fume blend with 25% slag (three lower
graphs) had shown lower charge passed when compared with the PC 0.45 control or
binary blend of Portland cement and slag (PCSL 0.45).
Figure 4.11 indicates that concrete mixes containing slag has lower penetrability level
than plain cement concrete mixes.

Total Charge Passed (Coulombs)

7000
6000
PCSL 0.40

5000

PCSL 0.40+
(PCSL 0.43)

4000

PCSL 0.45

3000

PCSL 0.45+
(PCSL 0.47)

2000

PC 0.45

PC 0.45+
(PC 0.52)

1000
0
7

28

56

91

Age (Day)
Figure 4.11: Comparison between total charge passed through concrete mixes containing slag and
plain cement concrete mixes

As mentioned in the literature review, researchers have believed that because of the
discontinuous pore structure in concrete mixes containing silica fume and slag, the total
charged passed through SFSL mixes is lower than SL mixes and plain cement mixes at

121
all ages (Section 2.2.2.4.2). Silica fume which is significantly fine reduces concrete
porosity which results in lower ion penetrability.
In addition the RCPT passing charge is sensitive to the pore solution chemistry since
electrons are transported by ions through the specimen pore system. Cement replacement
with silica fume can lead to an order of magnitude reduction in Na+, K+, Ca++, and OHion concentration in pore solution (Shi et al., 1998 in Ahmed et al., 2009), so lower
coulombs in silica fume concretes can be a result of either discontinuous pore system or
reduced ionic concentration. In other words, in concrete mixes containing silica fume
ASTM C1202 is not an appropriate method to evaluate chloride ion penetrability since
the results are affected by ionic concentration in pore solution (explained in Section
2.2.2.2). To make this standard test appropriate, Section 4.2.3 provides an additional
testing method to modify ASTM C1202-07.

4.2.2 RCPT electrical resistivity (first 5 min)


Another use for the RCP test is the measurement of bulk electrical resistivity. Bulk
electrical resistivity of a 100 x 50 mm concrete disc was calculated as,

R=
R=

L
A
=R
A
L

V
(Ohms law)
I

d 2 V
4L

where, d is taken as the average of four diameters measured on the specimen crosssections; L is the thickness of the specimen; I is the electrical current passed through the
specimen over the first 5 minutes; and V is the applied voltage which is 60 Volts.
The RCPT electrical resistivity values for concrete mixes cast in this research program
are presented in Table 4.4. Electrical resistivity of a concrete mixture at each age was
taken as the average of resistivity of two tests.

122
Table 4.4: The RCPT electrical resistivity values (K. cm)

Mix Design

W/CM

HPC
HPC+
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

0.35
0.38
0.40
0.40
0.43
0.45
0.47
0.45
0.52

Age (days)
7
23.4
20.9
18.9
10.5
7.6
6.4
6.1
4.2
3.7

28
71.6
66.9
58.7
19.3
15.7
13.0
13.8
7.5
5.5

56
76.8
73.5
70.3
26.5
20.2
20.1
18.7
9.6
8.1

91
83.9
80.7
77.5
36.7
26.7
24.5
23.6
11.4
10.2

In general, a low electrical resistivity is related to a high level of rebar corrosion risk.
Previous research has related the bulk electrical resistivity and severity of rebar corrosion
as indicated in Table 4.5.
Table 4.5: Electrical resistivity values and rebar corrosion rate (Feliu et al., 1996)

Electrical Resistivity (K. Cm)

Corrosion Risk Level

> 20

Low Rate

10-20

Moderate Rate

5-10

High Rate

<5

Very High Rate

The levels shown in Table 4.5 will be changed in presence of de-icing salts. The RCPT
electrical resistivity of concrete mixtures is shown in Figure 4.12. Now that the main
cause for deterioration in reinforced concrete structures is rebar corrosion, the RCPT
resistivity, which is related to the rebar corrosion risk level, can be a reliable indicator for
concrete durability.

123

90

80

First 5 min. Resistivity (K.cm)

70

HPC
(SFSL 0.35)
HPC+
(SFSL 0.38)
SFSL 0.40

60

50

PCSL 0.40
PCSL 0.40+
(PCSL 0.43)
PCSL 0.45

40

30

PCSL 0.45+
(PCSL 0.47)
PC 0.45

20

PC 0.45+
(PC 0.52)

10

0
0

28

56

Age

91

(day)

Figure 4.12: First 5 minutes RCPT electrical resistivity

It can be seen that adding silica fume to the concrete mixture influences the early age
electrical resistivity (three top lines in Figure 4.12).
Changing the W/CM ratio and adding SCMs affect the RCPT electrical resistivity as
described with following:

4.2.2.1 W/CM ratio effects on concrete resistivity


W/CM ratio influences concrete porosity. The RCPT electrical resistivity values of
different W/CM ratio mixes, shown in Table 4.4, have shown that higher W/CM ratio
concrete has lower electrical resistivity.
Relations presented in Figures 4.13 and 4.14 have shown that reducing the W/CM ratio
increases electrical resistivity of concrete. The difference between electrical resistivity of
higher and lower W/CM ratio concrete remains constant in plain cement concrete and

124
25% slag replacement cement concrete. In concrete mixes containing silica fume and
slag, this difference amount is significant between days 7 and 28.

8% Silica Fume and 25% Slag Replacement


First 5 min. RCPT Electrical Resistivity (K.cm)

90

0.35
0.38

80

0.35
W/CM = 0.35
0.38

70
60

0.40

0.38
0.40

0.40

HPC
(SFSL 0.35)

50
HPC+
(SFSL 0.38)

40
30
20

0.35

SFSL 0.40

0.38

0.40

10
0
7

28

56

91

Age (Day)
Figure 4.13: Effect of changing W/CM ratio on the RCPT electrical resistivity of ternary mixes

125

25% Slag Replacement

First 5 min. RCPT Electrical Resistivity (K.cm)

40

0.40

35
30
W/CM =0.40

0.45

25
0.43

15

0.47

0.43
0.45
0.47

0.40

20

0.45

PCSL 0.40

0.43

PCSL 0.45

0.47
PCSL 0.45+
(PCSL 0.47)

0.40

10

PCSL 0.40+
(PCSL 0.43)

0.43
0.45 0.47

5
0
7

28

56
Age

91

(Day)

Figure 4.14: Effect of changing W/CM ratio on the RCPT electrical resistivity of binary mixes

Increasing the W/CM ratio from 0.40 to 0.47 reduced the RCPT electrical resistivity at 7,
28, and 56 days by 42 %, 28%, and 29%, respectively. In concrete mixes with silica fume
and slag, increasing the W/CM ratio from 0.35 to 0.40 reduced the RCPT electrical
resistivity at 7, 28, and 56 days by 19 %, 18%, and 9%, respectively.
It is worth mentioning that the difference between the RCPT resistivity values of all
mixes decreased with concrete age.

4.2.2.2 Effects of adding SCMs on the RCPT electrical resistivity


The RCPT electrical resistivity of a binary (25 % slag replacement and GU cement) and a
ternary (8%silica fume, 25% slag, and GU cement) concrete mixes, W/CM ratio is 0.40,
are shown in Figure 4.15.

126

RCPT Electrical Resistivity


(W/CM= 0.40)
77.5

80
70.3

70
58.7

60
(K .cm)

First 5 min. R C PT Electrical R esistivity

90

SFSL 0.40

50
36.7

40

PCSL 0.40

26.5

30
19.3

18.9

20

10.5

10
0

28

56

91

Age (Day)
Figure 4.15: Silica fume effects on the first 5 min. RCPT electrical resistivity

As shown Figure 4.15, concrete mixes with silica fume have higher resistivity than
concrete mixes without silica fume, so adding SCMs increases the first 5 min. RCPT
electrical resistivity of concrete significantly. Silica fume improves electrical resistivity
of concrete by factor of 3 at early ages (e.g. at 28 days) in comparison with similar mixes
not containing silica fume.
Although slag contributes to later age properties of concrete, the difference between the
RCPT resistivity of mixes with silica fume and without silica fume remains constant from
day 7 to day 56 because the mixes with silica fume, contain slag as well. In other words,
the ternary blends of slag and silica fume behave much better than binary slag mixes.
Figure 4.16 illustrates that concrete mixes containing slag as a SCM have higher
electrical resistivity than plain cement concrete mixes (the W/CM ratio was the same in
both mixes).

127

RCPT: Electrical Resistivity


(W/CM= 0.45)

35
30
24.5

PCSL 0.45

25
(K.cm)

First 5 min. RCPT Electrical Resistivity

40

20.1

20

PC 0.45

15

13.0

11.4
9.6

10
5

6.4

7.5
4.2

0
7

28

56

91

Age (Day)
Figure 4.16: The effects of adding slag on the first 5 min. RCPT electrical resistivity

As slag improves later age properties of concrete, the difference between resistivity of
Portland cement concrete and 25% slag concrete, is higher at later age.

4.2.3 Depth of chloride ion penetration (colorimetric method)


Concrete discs used for the RCP test were split open along the central line by Carver
manual hydraulic pump at the end of the RCP testing as shown in Figure 4.17.

128

Figure 4.17: Splitting a concrete disc after the RCP test

The fracture surfaces of split samples were sprayed with 0.1 M silver nitrate solution,
AgNO3, to illustrate the depth of the chloride penetration.
From these, the average depth of chloride ion penetration at eight different locations
across the width of the specimens is tabulated in Table 4.6.

Table 4.6: Depth of chloride ion penetrated during the RCP test

Depth of Chloride Ion Penetration (mm)

MIXTURE
HPC
HPC+
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

Concrete Age
Day 7

Day 28

Day 56

Day 91

5.5
4.0
8.0
10.5
16.0
17.0
19.5
32.0
34.0

1.5
1.5
2
5.5
6.0
7.5
9.0
16.0
21.5

0.50
1.0
1.5
5.0
6.0
7.0
7.5
14.0
17.0

0.0**
0.0*
0.0*
3.5
5.5
6.0
6.5
10.5
16.0

* The depth of chloride penetration was not visible and uniform enough to be measured
** /o chloride penetration as shown in Figure 4.20

129
The values presented in Table 4.6 were used to calculate a non-steady-state diffusion
coefficient using the equation in Nordtest NT Build 492:

D=

(273 + T ) Lx d
0.0239(273 + T ) L
( xd - 0.0238
),
V 2
(V 2)t

where, D is the non-steady-state migration coefficient (x 10 12 m 2 /s), V is the applied


voltage (60 Volts), T is the average value of initial and final temperature in the anolyte
solution (C), L is the thickness of the specimen (mm), xd is the average depth of
chloride penetration (mm), and t is time (6 h).

The non-steady-state migration coefficients calculated based on the chloride ion


penetration values, presented in Table 4.6, are tabulated in Table 4.7.

Table 4.7: /on-steady-state migration coefficient of concrete mixes

Age
MIXTURE

HPC

HPC+

SFSL 0.40

Day 7

Day 28

Day 56

Day 91

5.53

1.66

0.50

0.00

Cl Penetration (mm)

26.75

24.25

24.5

23.1

Temperature (C)

52.94

51.12

50.81

50.69

Thickness (mm)

5.0

1.2

0.2

0.0

3.69

1.63

1.13

0.00

Cl Penetration (mm)

25.75

24.5

24.75

23.5

Temperature (C)

53.13

51.07

50.15

50.27

Thickness (mm)

3.2

1.2

0.7

0.0

8.24

1.85

1.23

0.00

Cl Penetration (mm)

25.5

24.75

24.5

24

Temperature (C)

53.08

51.17

50.73

50.12

Thickness (mm)

7.7

1.4

0.8

0.0

Migration coefficient
(10-12 m2/s)

Migration coefficient
(10-12 m2/s)

Migration coefficient
(10-12 m2/s)

130

PCSL 0.40

PCSL 0.40+

PCSL 0.45

PCSL 0.45+

PC 0.45

PC 0.45+

10.56

5.51

4.97

3.50

Cl Penetration (mm)

28

26.25

27

24.25

Temperature (C)

53.06

51.2

50.89

49.86

Thickness (mm)

10.2

4.8

4.3

2.8

16.19

6.29

5.81

5.60

Cl Penetration (mm)

42.75

27

26

24.5

Temperature (C)

52.33

52.95

52.8

50.82

Thickness (mm)

16.5

5.8

5.3

4.9

17.13

7.31

6.91

6.05

Cl Penetration (mm)

32

27

26.5

25

Temperature (C)

52.36

53.04

50.91

49.66

Thickness (mm)

17.0

6.8

6.2

5.2

19.45

8.69

7.50

6.42

Cl Penetration (mm)

41.75

27.25

26.5

26

Temperature (C)

53.07

53

51.34

48.92

Thickness (mm)

20.3

8.2

6.8

5.5

31.75

15.94

13.75

10.50

Cl Penetration (mm)

39.5

33

31

27.75

Temperature (C)

52.6

52.72

52.6

52.6

Thickness (mm)

33.3

15.9

13.5

10.0

34.25

21.40

17.03

16.05

Cl Penetration (mm)

42

35.25

30.75

30

Temperature (C)

53.25

53.01

52.3

52.99

Thickness (mm)

36.7

21.9

16.8

15.9

Migration coefficient
(10-12 m2/s)

Migration coefficient
(10-12 m2/s)

Migration coefficient
(10-12 m2/s)

Migration coefficient
(10-12 m2/s)

Migration coefficient
(10-12 m2/s)

Migration coefficient
(10-12 m2/s)

As expected, the migration coefficient decreased with a reduction of the W/CM ratio and
adding SCMs. The addition of SCMs had a significant effect on reducing the migration
coefficient by approximately 48% at 28 days of age due to the pozzolanic and micro-filler
(silica fume) effects.

131
A linear relation represents the correlation between the total charge passed and the RCPT
migration coefficient as shown in Figure 4.18.

RCPT Chloride Migration Coefficient


-12
2
(10 m /s)

40.0
35.0
30.0

D = 0.006 q- 0.004
2
R = 0.98

25.0
20.0
15.0
10.0
5.0
0.0
0

1000

2000

3000

4000

5000

6000

7000

RCPT Charge (Coulombs)

Figure 4.18: Relation between the RCPT passing charges and the chloride migration coefficient

The coefficient of determination (R2) between the total charge and the migration
coefficient data is 0.98. It reflects a significant association between these two datasets.
Consequently, it can be concluded that diffusion migration coefficient is marginally
affected by conductivity of pore solution which is in agreement with previous research
done by Stanish et al., 2004.

4.2.3.1 Effects of adding SCMs on the depth of chloride ion penetration


Figure 4.19 illustrates the depth of chloride penetration based on the values presented in
Table 4.6.

132

Rapid Chloride Permability Test


(Chloride Ion Penetration)

35

Cl
Depth of Penetration (mm)

30

25
PC 0.45
20

PCSL 0.45

15

PCSL 0.40
No SCM
SFSL 0.40

10

Slag

Silica fume + Slag


7

28

56

Age

91

(day)

Figure 4.19: Depth of chloride ion penetrated into concrete specimens during the RCP test

Concrete mixes containing SCMs had lower penetrability levels and lower chloride
penetration depths. Adding silica fume reduced concrete penetrability especially in early
ages in comparison with the other mixes. The depth of chloride ion penetrated into silica
fume specimens over 6 hours RCP test is the lowest. This value became zero at 91 days.
These results have proved that the RCPT is a reliable technique for concretes with SCMs
because concretes containing silica fume had the lowest (and zero at 91 days) depths of
chloride penetration in addition to their low passing coulombs.
Adding silica fume decreased concrete penetrability level as shown in Table 4.3. Also
depth of chloride ion penetration is affected by addition of silica fume as the depth of
chloride penetration is zero at 91 days (Figure 4.20).

133

Figure 4.20: Zero depth of chloride ion penetration from the top surface during the RCP test
(silica fume and slag concrete)

Therefore, the criticizes of some scientist (explained in Section 2.2.2.5) have be rejected
as silica fume mixes showed less chloride penetration coefficient which is independent
from ionic concentration and connectivity than mixes not containing silica fume.

4.2.3.2 W/CM ratio effects on chloride ion penetration


As mentioned in the literature, the W/CM ratio is the governing factor for chloride ion
penetration because the chloride diffusion is mostly dependent on concrete pore structure
than pore solution.
Figure 4.21 shows that by increasing the W/CM ratio, the depth of chloride ion
penetration is increased because higher W/CM ratio results in more porosity and higher
pore size distribution. Therefore, the pore structure is more connected in high W/CM
ratio concretes.

134

Figure 4.21: W/CM ratio effects on the depth of chloride ion penetration during the RCP test

In conclusion, the chloride penetrability depth is more sensitive to the W/CM ratio than
by adding SCMs because the migration coefficient is sensitive to the physical
characteristics of the pore structure. Therefore, it is not appropriate to use the RCPT
passing charge alone to evaluate chloride penetrability because it is more sensitive to the
pore solution chemistry while chloride penetrability is more affected by the pore structure
which is influenced by the W/CM ratio and the degree of hydration. Therefore, another
property which is related to the physical characteristics of pore system must be
investigated. A relation between the RCP tests Dcl and coulombs has to be used to
modify the ASTM C1202.

135

4.2.4 RCPT extrapolation


High current flow can generate heat in solutions which results in higher reported
coulombs. Therefore, the RCPT test continued for either 6 hours or until NaCl solutions
temperature reaches 80C. To discount heat effects, the extrapolated RCPT values are
calculated. As described by Hooton et al. (1997) in Bassuoni et al., 2006, linear
extrapolation to 6 h was done by multiplying the first 30 minutes coulombs by 12. In
such cases, extrapolated data may provide better estimation than the ultimate recorded
charge.
The calculated extrapolated data for concrete mixes are shown in Table 4.8.

Table 4.8: Extrapolated passing charges (Coulombs)

Concrete

Day 7

Day 28

Day 56

Day 91

HPC

734

232

233

143

HPC+

796
884
1593
2314
2627
2885
4264
4859

237
285
821
1011
1221
1204
2398
3269

239
255
631
778
779
951
1764
2104

179
198
434
674
776
755
1469
1724

Mixtures

SFSL 0.35
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

The extrapolated charges were compared with the recorded values as shown in Figures
4.22 and 4.23.

136

Figure 4.22: RCPT recorded and extrapolated charges passed (day 7)

The extrapolated passing charges were lower than the recorded charges because of the
heat effect which increased the total charges passed over 6 hours.
Replacing Portland cement with silica fume by 8% led to 43% reduction in both the
extrapolated and recorded passing charge at 7 days of age.

Figure 4.23: RCPT recorded and extrapolated charges passed (day 28)

137
As the total charge increased, the difference between the extrapolated and recorded
values (q) increased due to the heat effect during the standard RCP test; e.g. from 2.7%
in HPC to 19% in PC 0.45+ at day 7 and from 0% in HPC to 17% in PC 0.45+ at day 28
.
Heat generated during the RCP testing increased the recorded coulombs. This increase is
significant in concrete mixes with a high level of penetrability. Also there is a relation
between the maximum anode temperature and the difference between the extrapolated
and recorded passing charge as shown in Table 4.9.

Table 4.9: Effects of maximum anodic temperature on the RCPT coulomb values

Day 7

Mixtures

Day 28

Day 56

Tmax

Tmax

Tmax

(c)

(Coulombs)

(c)

(Coulombs)

(c)

(Coulombs)

HPC

28

2.7

25

0.0

25

-6.0

HPC+

28

3.3

26

1.9

25

-4.1

SFSL 0.40

28.5

3.4

26

2.5

25

-3.2

PCSL 0.40

33

1.6

29

2.7

28

1.6

PCSL 0.40+

49.5

10.5

29

3.2

28

3.8

PCSL 0.45

41

10.6

31

4.0

29

3.9

PCSL 0.45+

50

11.4

32

4.3

30

5.3

PC 0.45

54.5

17.3

40

10.0

37

7.5

PC 0.45+

59

19.2

47

16.6

39

13.2

Since there is a linear relation between extrapolated and recorded values (Figure 4.24),
the extrapolating passing charge values are reliable indicators for chloride penetrability in
highly permeable concrete mixtures. Also the linear extrapolation results are not
influenced by conductivity biases caused by SCMs especially silica fume.

Extrapolated Charge (Coulombs)

138

7000
6000
5000
4000
3000

y = 0.82 x + 118.76
2

R = 0.99

2000
1000
0
0

1000

2000

3000

4000

5000

6000

7000

Recorded Charge (Coulombs)


Figure 4.24: Relation between the extrapolated and recorded passing charges

The generated heat during the test is directly proportional to the amount of passing
charge which is affected by the level; of penetrability. High performance concretes at 91
days, had the lowest level of penetrability, so the heat generated during the test is the
least.
In conclusion, extrapolating the 6 h. coulombs from 30 minute readings is a reliable
technique to avoid the heat effects on the values, but just for concretes with level of
penetrability higher than very low, 1000 coulombs, according to ASTM C1202-07.
Concretes with the total charge passed lower than 1000 coulombs showed higher 6 h.
charges than calculated extrapolated charges.
As a practical recommendation, extrapolated values can be replaced with the recorded
values if the extrapolated passing charge is less than 800 coulombs (or the maximum
anodic temperature is less than 26 C. This conclusion is based on data presented in
Tables 4.8 and 4.9.

139

4.3 Rate of water absorption (sorptivity)


Water sorptivity is a term used to describe water ingress into pores of unsaturated
concrete due to capillary suction.
Two types of water sorptivity were studied in this research program:
a) Laboratory water sorptivity
b) Field water sorptivity

In both types of sorptivity test capillary sorptivity is the case of one-dimensional


absorption. The rate of absorption, I, is calculated using sorptivity relation:

I=

mass
,
A.

where, I is cumulative water absorption (mm), A is cross-section area of the specimen


which is in contact with water (mm2), and is density of water (

g
).
mm 3

The rate of water absorption, sorptivity (S), is the slope of I- t graph (mm/sec1/2).

4.3.1 Laboratory sorptivity test


The rate of water absorption was measured according to ASTM C1585-04 as described
briefly in Section 3.1.3. Four types of concrete discs (ASTM C1585-04: 100 6 mm
diameter with a length of 50 3 mm) were tested at any age: a bottom disc, a middle
disc, and a top disc sliced from a concrete cylinder in addition to a disc cored from the
rectangular concrete slabs.
Three days oven drying at 50C followed by sealed drying for four days at 50C was the
conditioning regime to obtain a uniform moisture distribution and a surface relative
humidity of 50 to 60% as described by Parrott (1994) and DeSouza et al. (1997).
After 7 days conditioning, the side of the specimens were sealed with electrical tape to
simulate one-dimensional flow. The laboratory sorptivity test was done according to
methodology described in Section 3.1.3.1.

140
The initial and secondary water sorptivity values of concrete specimens are shown in
Figures 4.25, 4.26, 4.27, and 4.29.
Since the water absorption is the slope of the I- t graph with a correlation of coefficient
more than 0.98 (ASTM C1585, 2004), the initial rate of absorption (1 min to the first 6
hours) of some mixes with correlation of coefficient less than 0.98 were calculated based
on data measured from1 to 60 min or 2h to 6h.

Figure 4.25: Rate of water absorption of top slices

The amount of water absorption is reduced as concrete ages. Since mature concrete has a
more discontinuous pore system, water absorption due to the capillary pores suction is
reduced. Concrete mixes containing silica fume (three bottom lines in Figure 4.25) had
lower initial water sorptivity values because of the silica fume effects on concrete pose
structure (described in the literature review, Section 2.2.3.4.2).

141

Figure 4.26: Rate of water absorption of middle slices

Since higher W/CM ratios result in higher porosity and high continuity of the pore
structure, the water sorptivity of PCSL 0.45 mix is higher than water sorptivity value of
PCSL 0.40+ (0.43) and PCSL 0.40 mixes. But the secondary rate of water absorption of
the middle disc from PCSL0.45 mix at 56 days is lower than the same mix with the lower
W/CM ratio (as shown in Figure 4.26). This anomalous result happened only at 56 days
for PCSL 0.45 mixture.

142

Figure 4.27: Rate of water absorption of bottom slices

In most cases, the rate of water absorption of the bottom discs is the lower than that of the
middle and the top discs because the bottom slices were denser due to the gravity.
The average rate of water absorption (initial and secondary absorption values) of three
slices cut from a concrete cylinder was calculated to the nearest 0.1 x 10-4 mm/sec as
tabulated in Table 4.10.

Table 4.10: Average water sorptivity values of concrete discs (x 10-4 mm/sec)

Age
Mixtures

Day 28

Day 56

Day 91

Initial

Secondary

Initial

Secondary

Initial

Secondary

HPC

17.0

3.4

13.1

2.6

11.2

1.2

HPC+

19.5
21.8
25.3
27.5
31.6
34.9
39.5
43.3

3.9
4.9
6.1
6.8
8.0
8.9
12.6
16.1

16.0
17.1
20.1
24.1
26.4
30.6
32.6
37.3

3.2
3.6
5.1
5.6
5.6
6.6
8.0
12.0

12.6
14.8
16.2
19.1
20.9
25.2
31.6
35.8

2.0
2.8
3.7
4.2
4.7
5.5
6.9
10.1

SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

143
In addition, a 100 6 mm diameter and 50 3 mm thick core was taken at each age
from the concrete rectangular slabs as shown in Figure 4.28.

Figure 4.28: Extracting cores from the concrete rectangular slabs

After conditioning, the finished surface of the concrete core was tested by water
sorptivity test to simulate the situation on site.
The secondary water absorption values of concrete cores taken at 28, 56, and 91 days are
shown in Figure 4.29.
90.0
PC 0.45+
PC 0.45
PCSL 0.45+
PCSL 0.45
PCSL 0.40+

70.0

No SCM

PCSL 0.40
SFSL 0.40

60.0

HPC+ (SFSL 0.37)


HPC (SFSL 0.35)

Slag

40.0
Silica fume + Slag
30.0
20.0
10.0
0.0
0

28

56
Age (day)

91

=============>+

50.0

W/CM

Water Sorptivity (x10-4 mm/sec)

80.0

Figure 4.29 (a): Initial rate of water absorption of the finished surface of concrete discs extracted
from rectangular concrete slabs

144

16.0
PC 0.45+
PC 0.45

14.0

PCSL 0.45+
PCSL 0.45
PCSL 0.40+

12.0

PCSL 0.40
SFSL 0.40

10.0

HPC+ (SFSL 0.37)


HPC (SFSL 0.35)

8.0
Slag
6.0
4.0

Silica fume + Slag

2.0
0.0
0

28

56
Age (day)

91

W/CM
=============>+

Water Sorptivity (x10-4 mm/sec)

No SCM

Figure 4.29 (b): Secondary rate of water absorption of the finished surface of concrete discs extracted
from rectangular concrete slabs

The only differences between concrete discs sliced from a concrete cylinder and a
concrete disc extracted from the rectangular slabs were the edge effect, curing regime,
and the finished surface. Concrete cylinders were kept in the fog room till the testing time
while concrete slabs were kept in the 50% RH room after first seven days moist curing.

To study the effects of curing regime on the rate of water absorption, the average
secondary water sorptivity of concrete cylinders (average values for the top, middle, and
bottom slices) were compared with the secondary water sorptivity values of concrete
cores as shown in Figure 4.30.

145

16.0

y = 0.98 x
2

R = 0.98

14.0

-4

Water Sorptivity of Sliced Discs (x10 mm/sec )

18.0

12.0
10.0
8.0
6.0
4.0
2.0
0.0
0.0

2.0

4.0

6.0

8.0

10.0

12.0
-4

14.0

16.0

18.0

Water Sorptivity of Extracted Cores (x10 mm/sec )

Figure 4.30: Comparison between the secondary water sorptivity values of concrete discs sliced from
concrete cylinders and concrete cores extracted from concrete slabs

The secondary rate of water absorption of concrete discs and cores are almost identical,
but concrete cores absorbed more water. Water sorptivity values of concrete cores were
higher than that in concrete slices by an average of 5% at 91 days. Since the moisture
content of concrete slabs kept in 50% RH room reduced after they had been removed
from the moist room, the degree of hydration was reduced. Consequently, at later ages
the tortuosity of cores was higher than concrete cylinders resulted in higher rate of water
absorption. It resulted in higher water sorptivity coefficients although the differences
between sorptivity values of cores and slices were not significant since mid-depth of
slabs, where represented by cores, had more than 70% RH during most of the time as
presented in Table 4.12. Therefore hydration did not stop. Also paste skin of the cylinders
may influence the rate of absorption.

146
As mentioned, volume and tortuosity of the pore system which affects the rate of water
absorption can be represented by the RCPT chloride migration coefficient (Table 4.7),
the RCPT results and water sorptivity can be related as shown in Figure 4.31.

25
m /s)

y = 1.64 x - 3.82
2

RCPT Chloride Migration Coefficient (x10

-12

R = 0.92
20

15

10

0
0.0

2.0

4.0

6.0

8.0

10.0

12.0
-4

14.0

16.0

18.0

Water Sorptivity of Sliced Discs (x10 mm/sec )

Figure 4.31: Water sorptivity and chloride migration coefficient

The linear relation indicates that both water sorptivity coefficient and chloride ion
penetration coefficient are influenced by similar factors.
There is a relation (non-linear) between the water sorptivity and compressive strength
(Figure 4.32).

147

16

y = 49030 x

W ater S orptivity (x 10 -4 m m /s ec )

14

-2

R = 0.89
12
10
8
6
4
2
0
30

35

40

45

50

55

60

65

70

75

80

Compressive Strength (MPa)

Figure 4.32: Relation between the lab sorptivity test values and concrete compressive strength

It can be concluded that the water sorptivity is influenced by factors affecting capillary
pore system and its continuity such as the W/CM ratio, the addition of SCMs and the
degree of hydration.
Although the relative humidity of the specimens used for laboratory sorptivity test was
constant, it has been discussed by other researchers (Nokken et al., 2002) that concrete
sorptivity decreases with increasing degree of saturation and also decreasing W/CM ratio.

4.3.1.1 SCMs effects on the lab sorptivity values


Silica fume, an ultrafine material, strengthens interfacial transition zone (ITZ) by better
particle packing and providing nucleation by its pozzolanic reaction with portlandite.
Therefore, the microstructure of concrete becomes denser resulting in lower water
permeability. As shown in Figure 4.33, concrete mixes containing slag had lower water
sorptivity values than plain cement concrete mixes at all ages.

148

18.0

Water Sorptivity (x10-4 mm/sec )

16.0

No SCMs

14.0

PC 0.45+

12.0

PC 0.45
PCSL 0.45+

10.0

PCSL 0.45

Slag

PCSL 0.40+

8.0

PCSL 0.40
6.0

SFSL 0.40
HPC+ (SFSL 0.37)

4.0

HPC (SFSL 0.35)

Silica fume +
Slag

2.0
0.0
0

28

56
Age (day)

91

Figure 4.33: Average rate of water absorption of concrete mixes

Although not measured, it can be predicted that at 7 days water sorptivity of plain cement
concretes would lower than slag concrete mixes with similar W/CM ratio because of
slags later age contribution to concrete properties. In contrast, it can be anticipated that
at later ages, ternary mixes have lower sorptivity coefficients than any other mixes.

4.3.1.2 W/CM ratio effects on the lab sorptivity values


Since the ability of concrete to resist water penetration is influenced by the connectivity
of its capillary pore structure (explained in Section 2.2.5), lower W/CM ratio mixes had
lower sorptivity values as shown in Figures 4.29 and 4.33. Concrete mixes with low
W/CM ratio have lower porosity and the pore system is less continuous. They result in
lower amount of water absorbed by the capillary suction.
It is important to mention that the HPC mixes containing silica fume and slag had fast
initial sorptivity due to their finer pore structure, but sorptivity decreased rapidly due to
discontinuity.

149

4.3.2 Field sorptivity test


Since concrete deterioration processes (e.g. rebar corrosion) are influenced by fluid
penetrability especially of covercrete, the ability of concrete to absorb water on site can
provide information about its durability.
For each concrete mixture, two 406 x 75 mm circular slabs were tested at ages 14, 28,
56, and 91 days as described briefly in Section 3.1.3.2.

Figure 4.34: Field sorptivity apparatus (horizontal orientation)

The water sorptivity was measured in naturally occurring conditions (concrete slabs were
tested as they removed from the 50% RH room) and moisture content of slabs were
measured by weighting slabs. Therefore, no pre-conditioning regime was applied to the
concrete slabs. The rate of water absorption is calculated as shown in Table 4.11.

150
Table 4.11: Water sorptivity coefficient of the circular concrete slabs (mm/min1/2)

Mixture
HPC
HPC +
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

Day 14

Day 28

Day 56

Day 91

CS1

CS2

Ave.

CS1

CS2 Ave. CS1

CS2 Ave. CS1

CS2 Ave.

0.009
0.010
0.010
0.012
0.015
0.028
0.030
0.006
0.025

0.006
0.008
0.009
0.013
0.016
0.029
0.029
0.016
0.030

0.008
0.009
0.009
0.013
0.016
0.028
0.029
0.011
0.028

0.009
0.015
0.015
0.017
0.032
0.042
0.054
0.017
0.025

0.012
0.008
0.012
0.019
0.031
0.053
0.060
0.018
0.037

0.013
0.019
0.023
0.032
0.052
0.057
0.063
0.030
0.046

0.019
0.024
0.026
0.050
0.072
0.100
0.116
0.048
0.078

0.010
0.011
0.013
0.018
0.031
0.047
0.057
0.018
0.031

0.020
0.022
0.027
0.031
0.047
0.060
0.074
0.031
0.038

0.017
0.021
0.025
0.031
0.050
0.058
0.068
0.030
0.042

0.022
0.028
0.030
0.049
0.071
0.089
0.115
0.044
0.057

It is seen that the amount of water absorption is increased with a decrease in the relative
humidity in the pore system of the concrete specimens due to continued drying with age
as shown in Figure 4.35.

0.120

0.100

Water Sorptivity (mm/min

0.5

PCSL 0.45+
(PCSL 0.47)
PCSL 0.45

0.080

PCSL 0.40+
(PCSL 0.43)
PC 0.45+
(PC 0.52)
PC 0.45

0.060

PCSL 0.40
SFSL 0.40

0.040

HPC+
(SFSL 0.38)
HPC
(SFSL 0.35)

0.020

0.000
0

14

28

56

91

Age (day)

Figure 4.35: Water sorptivity coefficients of circular concrete slabs

0.020
0.026
0.028
0.049
0.072
0.094
0.115
0.046
0.067

151
Since day 14 is important for the Ministry of Transportation of Ontario (bridge joints
installation), the field sorptivity test was started at this day. As shown in Figure 4.35, rate
of water absorption increased as concrete becomes more mature. This unusual finding
was caused by the reduction in the concrete moisture level. Concrete slabs were not preconditioned before the test, so moisture content at each testing time was less than that in
the previous time. Therefore, drier concrete absorbed more water although the pores
tortuosity was reduced due to the cement hydration and SCMs secondary hydrations.
Despite the test results can not provide useful data about the physical characteristics of
the pore structure and the degree of hydration, it can be a useful technique measuring the
moisture content of in situ concretes.

It was not possible to precondition concrete slabs to a standard relative humidity prior to
testing as can be done in the laboratory sorptivity test. Therefore, the rate of water
absorption must be correlated with the slabs moisture content. Since it was not possible
to place a 406 x 75 mm circular slab in an oven to dry, a concrete piece was taken from
the cored rectangular slabs (Figure 4.36), which were kept in the same condition as the
circular slabs, weighed, and dried at 110C.

Figure 4.36: Middle piece taken from a cored rectangular slab for moisture content measurement

152
It is important to mention that these concrete pieces were taken approximately seven
months from the casting time. At this time the weight of slabs remained constant, so there
was a humidity balance between the room conditions and the concrete (50% RH). The
amount of water (a percentage of the dry mass) in 50% RH concrete pieces was used to
calculate the dry mass of circular slabs.
Table 4.12: Degree of saturation of circular concrete slabs

Concrete Age

Mixture

7 days 14 days 28 days

56 days

91 days

7 months

HPC

100%

92%

89%

86%

84%

50%

HPC +
SFSL 0.40

100%
100%

90%
90%

85%
86%

80%
81%

77%
77%

50%
50%

PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+

100%
100%
100%
100%

91%
88%
82%
83%

84%
81%
77%
78%

78%
74%
72%
71%

73%
69%
64%
64%

50%
50%
50%
50%

PC 0.45
PC 0.45+

100%
100%

83%
85%

77%
78%

70%
71%

64%
63%

50%
50%

Data presented in Table 4.12 are visualized in Figure 4.37.

153

100

95
RCPT Chloride
Penetration Depth

Degree of Saturation (%)

90

HPC
85

HPC +
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

80

75

70

65

60
0

14

28

56

91

Drying Time (day)


Figure 4.37: Degree of saturation of the concrete slabs at different ages

It can be seen that the physical characteristic of concrete which affects the depth of
chloride ion penetrated into concrete can be considered as an influencing factor for the
rate of moisture loss and degree of saturation. Concrete mixes with higher volume and
connectivity of pores, so higher chloride penetration coefficient (e.g. PC 0.45+) lost their
moisture more and faster than concretes with less porosity and pore system continuity
such as HPC. In other words, moisture evaporation is less and slower in concretes with
less porosity and pore continuity.
If water sorptivity test results are calibrated, the field sorptivity test can be used as a rapid
quality assurance procedure. The calibration curve is plotted as shown in Figure 4.38
based on data presented in Tables 4.11 and 4.12.

154

0.120

0.5

Water Sorptivity Coefficient (mm/min )

0.100

PCSL 0.45+
0.080

PCSL 0.45
PCSL 0.40+
PCSL 0.40
PC 0.45+

0.060

SFSL 0.40
HPC
HPC+

0.040

PC 0.45

0.020

0.000
60

65

70

75

80

85

90

95

100

Degree of Saturation (%)

Figure 4.38: Moisture content effects on the rate of water absorption

The need to adjust the sorptivity values for moisture content can be minimized if all tests
were performed after the same conditioning procedure (e.g. ASTM C 1202-07).
The rate of absorption increased as the moisture content decreased with the smallest rate
of absorption occurring at 14 days with degrees of saturation close to 100%. The largest
rate of water absorption occurs at the lowest moisture content and zero rate at the
saturated condition (after removing from the fog room after 7 days moist curing).
Water sorptivity is influenced by many factors. Among these factors, cementitious
materials ratio, the water-to-cement ratio, and concrete moisture content are the most
important factors.
To plot the calibration curves for each concrete mix (Figures 4.43-4.51), water sorptivity
coefficient at 50% RH (when slabs weight located in 50% RH room became constant

155
which meant slabs were in humidity equilibrium condition with the room) were
measured.

4.3.3 SCMs effects on the rate of water absorption


Both laboratory and field sorptivity test data have shown that adding SCMs to a concrete
mixture, results in less water absorption, but it depends on the age of the concrete. As
shown in Figure 4.39, concrete mixes containing silica fume and slag had the lowest
early age sorptivity coefficient because of silica fume rapid pozzolanic reactivity.

0.100

0.5

Water Sorptivity (mm/min )

0.080
PCSL 0.45

0.060
PCSL 0.40
PC 0.45

0.040
SFSL 0.40

0.020

0.000
0

14

28

56
Age (day)

91

Figure 4.39: SCMs effects on the field sorptivity test results

The ternary mix with both silica fume and slag had shown a low rate of water absorption
in later ages because of the secondary hydration products of silica fume and mostly slag.
Cementing materials hydration reduced the moisture content of the concrete slabs in
addition to decreasing the tortuosity of pores. Concrete mix containing GGFS absorbed
more water than plain cement concrete with similar W/CM ratio because slag effects are
significant in concrete later ages. At later ages, binary mixes (slag and Portland cement)

156
had even higher sorptivity coefficients than plain cement concrete mixes because slag
absorbed more water during its secondary hydration to form more C-S-H. Therefore, the
moisture content of concrete slab containing slag (PCSL 0.45) is lower than plain cement
concrete (PC 0.45) which results in higher water sorptivity coefficients.
In other words, the field sorptivity test with the procedure applied during this research
program is not an applicable test to analyse the characteristics of the pore structure while
it can be a useful non-destructive technique to distinguish between different mix designs
especially in mixes containing silica fume.

4.3.4 W/CM ratio influences on the rate of water sorptivity coefficient


Since pore size and pore connectivity decreased with decreasing the W/CM ratio (Ahmed
et al., 2009), an increase of water sorptivity is observed with an increase of water-tocementitious materials ratio. In other words, high W/CM concretes have greater pore
volume with connected pores which results in faster water absorption. It can be seen in
all concrete mixes as shown in Figures 4.40, 4.41, and 4.42.

0.050

==========>

W/CM

0.060

0.5

Sorptivity Coefficient (mm/min )

0.070

0.040
PC 0.45+
(PC 0.52)

0.030

PC 0.45

0.020
0.010
0.000
0

14

28

56

91

Age (day)

Figure 4.40: W/CM ratio effects on the water sorptivity coefficients of plain cement concrete

157

0.080

0.060

==========>

0.100

W/CM

0.5

Sorptivity Coefficient (mm/min )

0.120

PCSL 0.45+
(PCSL 0.47)

0.040

PCSL 0.45

0.020

PCSL 0.40+
(PCSL 0.43)

0.000
0

14

28

56

91
PCSL 0.40

Age (day)

Figure 4.41: W/CM ratio effects on the sorptivity coefficients of concrete mixes containing slag
(25 % replacement)

0.025

0.020

==========>

W/CM

0.5

Sorptivity Coefficient (mm/min )

0.030

0.015
SFSL 0.40

0.010

HPC+
(SFSL 0.38)
HPC
(SFSL 0.35)

0.005

0.000
0

14

28

56
Age (day)

91

Figure 4.42: W/CM ratio effects on the water sorptivity coefficients of the ternary mixes

158
In all concrete mixes, higher W/CM ratio mixes had higher water sorptivity coefficient.
But the amount of increase in water sorptivity coefficient is dependant to the concrete
mix design. It is seen that increasing the W/CM ratio for 0.05, increased the water
permeability values by 26 % in ternary mixes and 38% in binary and 32% in plain
cement mixes at 91 days. It can be concluded that the binary mixes absorbed more water
because of their lower moisture content in contrast with their lower porosity than plain
cement mixes.

4.3.5 Water sorptivity and degree of saturation


Relative humidity of concrete specimens used for laboratory sorptivity test was about
80% while concrete slabs used for field sorptivity test had different moisture contents
because of the slow conditioning process of 406 x 75 mm circular slabs. Field sorptivity
test results have supported DeSouzas results (DeSouza, 1998) which showed that water
sorptivity increases with decreasing levels of concrete saturation.
In conclusion, the field sorptivity test can be a useful test for similar mix design concretes
with different W/CM ratios. To compare the change in physical characteristics of the
pore system of different mix designs, the field sorptivity test can be applicable provided
that all specimens are conditioned to a similar level of relative humidity.
It is recommended for future works to pre-condition the slabs to a fully dried level by
vacuum and silica gel. In this case, the effects of the change in physical properties of the
pore structure will be studied by the water absorption values regardless of the moisture
history of concrete.

4.3.6 Calibration curves and prediction of the later age sorptivity coefficient
Since each concrete has a unique calibration curve, it is necessary to plot it as shown in
Figures 4.43 to 4.51 based on water sorptivity coefficient at first 91 days and at 50% RH.
It is important to mention that the water sorptivity coefficients after 50% RH can be
calculated regarding to the best fit representing the measured values. 50% RH point
represented the day when the weight of concrete slabs remained constant in 50% RH
curing room. The days when the slabs weight became constant, were varied in mixes as
rates of absorption and evaporation of mixes were different.

159

HPC (SFSL 0.35)


day 7
100

day 14

Degree of Saturation (%)

90

day 28
day 56

y = 102.3 e

-12x

R = 0.98

day 91
80

70

60

0.08

0.06

0.04

0.02

0.00

50

0.5
Water Sorptivity Coefficient (mm/min )

Figure 4.43: Water sorptivity- saturation calibrating curve for HPC mix

HPC+ (SFSL 0.38)


100

y = 100.0 e

-11.3 x

R = 0.99
80

70

60

0.5
Water Sorptivity Coefficient (mm/min )

Figure 4.44: Water sorptivity- saturation calibrating curve for HPC+ mix

0.08

0.06

0.04

0.02

50
0.00

Degree of Saturation (%)

90

160

SFSL 0.40
100

90

Degree of Saturation (%)

-11.4 x

y = 102.0 e
2

R = 0.98
80

70

60

50
0.00

0.02

0.04

0.06

0.08

0.5

Water Sorptivity Coefficient (mm/min

Figure 4.45: Water sorptivity- saturation calibrating curve for SFSL 0.40 mix

PCSL 0.40
100

-5.07 x

y = 95.0 e
2

R = 0.98
80

70

60

0.16

0.14

0.12

0.10

0.08

0.06

0.04

0.02

50
0.00

Degree of Saturation (%)

90

0.5

Water Sorptivity Coefficient (mm/min

Figure 4.46: Water sorptivity- saturation calibrating curve for PCSL 0.40 mix

161

PCSL 0.40+ (PCSL 0.43)


100

Degree of Saturation (%)

90
-5.9 x

y = 99.1 e
2

R = 0.97
80

70

60

0.14

0.12

0.10

0.08

0.06

0.04

0.02

0.00

50

0.5
Water Sorptivity Coefficient (mm/min )

Figure 4.47: Water sorptivity- saturation calibrating curve for PCSL 0.40+ mix

PCSL 0.45
100

y = 96.7 e

-4.7 x

R = 0.98
80

70

60

0.16

0.13

0.10

0.06

0.03

50
0.00

Degree of Saturation (%)

90

0.5

Water Sorptivity Coefficient (mm/min

Figure 4.48: Water sorptivity- saturation calibrating curve for PCSL 0.45 mix

162

PCSL 0.45+ (PCSL 0.47)


100

Degree of Saturation (%)

90
y = 98.5 e

-4.5 x

R = 0.95
80

70

60

0.16

0.14

0.12

0.10

0.08

0.06

0.04

0.02

0.00

50

0.5

Water Sorptivity Coefficient (mm/min

Figure 4.49: Water sorptivity- saturation calibrating curve for PCSL 0.45+ mix

PC 0.45
100

Degree of Saturation (%)

90
y = 93.1 e

-8.3 x

R = 0.97
80

70

60

50
0.00

0.02

0.04

0.06

0.08

0.10

0.5

Water Sorptivity Coefficient (mm/min

Figure 4.50: Water sorptivity- saturation calibrating curve for PC 0.45 mix

163

PC 0.45+ (PC 0.52)


100

90

Degree of Saturation (%)

-5.7 x

y = 95.4 e
2

R = 0.96
80

70

60

50
0.00

0.02

0.04

0.06

0.08

0.10

0.12

0.14

0.5

Water Sorptivityy Coefficient (mm/min

Figure 4.51: Water sorptivity- saturation calibrating curve for PC 0.45+ mix

As the rest of the calibration curves can be plotted (not measured) based on the first 91
days water sorptivity values and sorptivity coefficient at 50% RH, some water sorptivity
coefficients of dry samples were against the basic concepts of the concrete technology;
for example, although the PCSL 0.45 mix had 25% slag replaced with the Portland
cement (higher electrical resistivity), its permeability coefficient at 0% degree of
saturation was higher than that in PC 0.45 (lower resistivity). In case of W/CM ratio
effects, the water sorptivity coefficient at 0% RH of higher a W/CM ratio concrete was
higher than that in a lower W/CM ratio specimen (similar mix design) which proved that
this procedure can be useful for analysing the W/CM ratio effects on the pore system (as
concluded in Section 4.3.4).
If a similar pre-conditioning regime was applied to all concrete specimens and specimens
were in a similar RH at the time of testing, the later age (0% moisture content) sorptivity
coefficients would represent the change in the physical characteristics of the pore system
caused by cementing materials hydration and the W/CM ratio.

164
To check the calculated sorptivity coefficient at 0% MC, one concrete slab from every
mix designed was fully dried as shown in Figure 4.52.

Figure 4.52: Drying circular concrete slabs

Water sorptivity of fully dried (0% MC) circular concrete slabs were measured and
compared with the calculated coefficients (not shown in Figures 4.43-4.51). The
difference between the measured sorptivity coefficients and calculated coefficients was
not significant (except in PC 0.45 mix), so estimating 0% MC sorptivity values based on
the calibration curves is a reliable method. In this case, it can be predicted that, the water
sorptivity coefficient of higher resistivity concrete is lower and vice versa.

4.4 Electrical resistivity


Measurement of concrete electrical resistivity has been proposed as a possible method for
assessment of the physical characteristics of pore structure and microstructure and pore
solution chemistry. Electrical current passes through the pore structure because saturated
pores in the porous cement paste contain ions although cement hydrates in solutions.

Beside the RCPT resistivity (Section 4.2.2), DC-cyclic bulk resistivity and surface
electrical resistivity were measured in this research program.
Measuring the electrical bulk resistivity is the most common and reliable method for
measuring electrical resistivity. The measurement of surface electrical resistivity by the

165
Wenner technique is a time saving, but less reliable method if adequate technical
recommendations have not been considered as critics believed: Millard and Gowers
(1991) and Morris et al. (1996). But this research has indicated that the Wenner technique
is a reliable test method.
As mentioned in the literature, electrical resistivity of concrete is a function of the pore
solution chemistry, the pore volume, its connectivity, and the degree of saturation of
concrete (moisture content). All the specimens tested for electrical resistivity were fully
water saturated, so the effect of moisture content on concrete conductivity was not
studied in this research program.

4.4.1 DC-cyclic bulk resistivity (Monfore resistivity)


The bulk electrical resistivity of a concrete cylinder was measured using two stainless
steel electrodes placed on opposite surfaces of a specimen as shown in Figure 4.53.

Figure 4.53: Concrete bulk resistivity test with two steel electrodes (DC-cyclic bulk resistivity)

166
One electrode induced the cyclic direct current and the other electrode received the
current. The DC measurement was carried out using a cyclic DC potential similar to that
of Monfore (Monfore, 1968 and Polder, 2001) but switching between 3 volts and 5 volts
every 5 seconds. The bulk electrical resistivity in this research program was calculated as:

(V5 V3 ) d 2
=
(K.cm),
(I5 I3 ) L

where, V5 is 5 volts, V3 is 3 volts, I5 is current at V5 (A), I3 is current at V3, d is specimen


section diameter (mm), and L is the length of specimen (mm).
In this research program, the Monfore resistivity of three 100 x 200 mm saturated
cylinders and two 100 x 50 mm discs sliced from middle and bottom parts of concrete
cylinders were measured at each age. In all specimens, the top faces were in contact with
the negative electrode while the bottom surfaces were in contact with the positive
electrodes.

4.4.1.1 DC-cyclic bulk resistivity of full length concrete cylinders


According to the last equation, the DC-cyclic resistivity of three water saturated 100 x
200 mm concrete cylinders, was measured in 15 minutes at each age. Table 4.13 presents
the average of three resistivity values.

Table 4.13: Bulk resistivity values of concrete cylinders

Mixture

DC-Cyclic Bulk Resistivity (K.cm)

Day 3
HPC
17.5
HPC +
16.9
SFSL 0.40
9.1
PCSL 0.40
6.3
PCSL 0.40+ 3.8
PCSL 0.45
3.5
PCSL 0.45+ 3.7
PC 0.45
3.4
PC 0.45+
3.5

Day 7
26.3
24.8
20.5
9.8
5.7
3.8
3.8
4.9
4.6

Day 28 Day 56
82.4
94.5
79.2
88.1
72.5
78.2
23.3
29.4
18.1
25.0
15.7
22.1
15.1
19.7
6.8
9.8
8.2
8.6

Day 91
102.4
92.7
82.3
42.0
32.6
29.8
28.8
12.6
10.4

167
Figure 4.54 in plotted based on the average electrical resistivity presented in Table 4.13.

Silice fume + Slag

110
100

80
70

HPC (SFSL
0.35)
HPC+ (SFSL
0.38)

SFSL 0.40
PCSL 0.40

50

PCSL 0.40+
(PCSL 0.43)

40

PCSL 0.45

Slag

60

30

PCSL 0.45+
(PCSL 0.47)

PC 0.45

20

No SCM

DC Cyclic Bulk Resistivity (K.cm)

90

10

PC 0.45+
(PC 0.52)

0
03 7

28

Age (day)

56

91

Figure 4.54: DC-cyclic bulk resistivity of 100 x 200 mm concrete cylinders

It can be seen that the addition of slag and silica fume to concrete mixtures, affects the
electrical resistivity significantly. Also any change in W/CM ratio influences electrical
resistivity of concrete.
The ionic concentration in pore solution increases with age because of the dissolution of
calcium and alkali ions (Nokken et al., 2006). It causes higher electrical conductivity
(lower resistivity). On the other hand, the connectivity of the pore structure and pore size
distribution decrease as concrete ages.
As shown in Figure 4.54, the Monfore resistivity of all concrete mixes increased as
concrete aged. It shows that the change in the physical characteristics of the pore system
is more dominant than the change in pore solution. Mature concrete has less connected
pores which results in higher electrical resistance.

168
Electrical properties of concrete are dependant on the amount of free water (moisture
content which was always 100% in this study), the type of supplementary cementitious
materials, the W/CM ratio, and the degree of hydration.

4.4.1.1.1 SCMs effects on the DC-cyclic bulk resistivity


The influence of adding silica fume to concrete mixes can be appreciated by comparing
data for mixes with 8% silica fume and without silica fume (with similar W/CM ratio) as
shown in Figure 4.55.

W/CM

110
100
DC Cyclic Bulk Resistivity (K.cm)

90

Silica fume +Slag


0.40

80
70

SFSL 0.40

60
PCSL 0.40

50
Slag

0.40

40
Slag

30

PCSL 0.45

0.45
PC 0.45

20

No SCM
0.45

10
0
03 7

28

Age (day)

56

91

Figure 4.55: SCMs effects on the DC-cyclic bulk resistivity values

The W/CM ratio in the two lower lines is 0.45. The only difference in their mix design is
the presence of slag. Concrete mixtures containing slag had higher electrical resistivity
after 10 days than plain cement concrete. This difference is more significant in later ages
because of the later age contribution of slag. In contrast, electrical resistivity of plain
cement concretes is higher than that in similar concrete mixes with 25% slag replacement

169
during the first 10 days although no one is interested in 10 days resistivity (Figure 4.55).
All these facts have proved that concrete mixes containing GGBS have lower durability
than plain cement concrete in early ages and higher durability in later ages significantly
after 28 days.

It is known that silica fume will increase the bulk resistivity of concrete in three ways:

1) Silica fume reacts with calcium hydroxides, Ca(OH)2, in the pore solution to form
secondary hydrates. Silica fume reduces the ionic concentration of the pore
solution resulting in less electrical charge passing trough the pore system
(Bassuoni et al., 2006), so this is a critic for the bulk resistivity measurements of
silica fume concretes,
2) Silica fume is a super fine material, so it increases the density of cement paste.
The ITZ penetrability is decreased significantly by adding silica fume to the
mixture (Neville, 1995).
3) Because of the silica fume secondary hydrates (C-S-H), the volume and
connectivity of the pore structure is decreased significantly which results in less
pores tortuosity (Hansson, 1983).
Concrete mixes with 8% silica fume and 25 % slag replaced with cement had the highest
electrical resistivity in comparison with the other mixes. Silica fume mostly affects early
age properties of concrete and the difference between early age resistivity of ternary
mixes and other mixes is significantly higher than that in later ages.

4.4.1.1.2 W/CM ratio effects on bulk resistivity


W/CM ratio fundamentally affects concrete porosity. More porosity and connected pores,
caused by higher W/CM ratio, results in lower electrical resistivity because the applied
electrical current can easily pass through the pore structure. It can be concluded from
Figure 4.56 that concrete mixes with similar mix design but different W/CM ratio had
different electrical properties.

170

W/CM

DC Cyclic Bulk Resistivity (K.cm)

110
100

0.35

90

0.38

80

0.40

HPC
(SFSL0.35)
HPC+
(SFSL 0.38)

70

SFSL 0.40

60

PCSL 0.40

50
40
30

0.40

PCSL 0.40+
(PCSL 0.43)

0.43
0.45
0.47

PCSL 0.45

20

PCSL 0.45+
(PCSL 0.47)

PC 0.45

0.45

10

0.52

PC 0.45+
(PC 0.52)

0
03 7

28

Age (day)

56

91

Figure 4.56: W/CM ratio effects on the Monfore resistivity values

In all mixes tested in the research program, as W/CM ratio increased, the Monfore
resistivity decreased. On the other hand, low W/CM ratio specimens had higher electrical
resistivity.
The W/CM ratio effects were significant in concrete specimen containing both slag and
silica fume (top three lines in Figure 4.56). In contrast, the W/CM ratio did not
significantly affect electrical properties of plain cement concrete mixtures (two lower
lines). Also it can be seen that effects of any charge in the W/CM ratio is significant in
lower W/CM ratios.

4.4.1.2 Bulk resistivity as an indicator for chloride penetrability


The diffusion of chloride ion through concrete can be correlated with the DC-cyclic bulk
resistivity. Also the total RCPT passing charge represents the ranking of the concrete
with respect to its chloride permeability. Consequently, the Monfore resistivity can be
used as an indicator for chloride ion penetrability (Section 4.2.3) as shown in Figure 4.57.

171

Figure 4.57: Relation between chloride migration coefficient and the DC-cyclic bulk resistivity with
the RCPT 5 min. resistivity

4.4.1.3 Bulk resistivity of concrete discs


Two 100 x 50 mm concrete discs were sliced from a concrete cylinder at each age. All
specimens were vacuum-saturated one day prior to the testing time according to the
process recommended in ASTM C1202-07. The DC-cyclic bulk resistivity of concrete
mixtures (average of electrical resistivity of the bottom and middle slices) were measured
as shown in Table 4.14.

172
Table 4.14: The Monfore resistivity of 100 x 50 mm concrete discs

Mixture

DC-Cyclic Bulk Resistivity (K.cm)

Day 3
HPC
4.9
HPC +
4.5
SFSL 0.40
4.5
PCSL 0.40
4.0
PCSL 0.40+ 4.3
PCSL 0.45
3.9
PCSL 0.45+ 3.6
PC 0.45
4.2
PC 0.45+
4.1

Day 7
22.6
24.8
24.3
6.6
5.3
4.6
4.6
4.2
3.8

Day 28 Day 56
77.4
91.3
74.9
89.6
64.3
78.3
17.3
30.1
15.3
24.4
13.3
21.6
11.6
22.4
6.6
10.9
6.4
9.8

Day 91
99.6
96.8
84.5
42.6
32.1
29.1
27.4
13.0
11.3

DC Cyclic Bulk Resistivity (K.cm)

Figure 4.58 is based on data presented in Table 4.14.


110
100

HPC
(SFSL 0.35)

90
80
70

HPC+
(SFSL 0.35+)
SFSL 0.40
PCSL 0.40

60
50

PCSL 0.40+
PCSL 0.45

40
30
20

PCSL 0.45+
PC 0.45
PC 0.45+

10
0
03 7

28

56

91

Age (day)
Figure 4.58: DC-cyclic bulk resistivity of 100 x 50 mm concrete discs

It can be seen that there is a logical relation between the Monfore resistivity and concrete
age for all mixes; as concrete ages, electrical resistivity increases. It is important to
mention that electrical resistivities of bottom slices, more compacted, were usually higher
than middle slices resistivity due to gravity effects on aggregates.
In most of the cases, the difference between the bulk resistivity of concrete slices and
resistivity of full length cylinders from a similar mix was less than 20%. In other words,

173
DC-cyclic bulk resistivity test was independent from geometry of specimens. Therefore,
the Monfore resistivity test values represent the resistivity of concrete.

4.4.2 Surface electrical resistivity


Surface electrical resistivity was measured by four-electrode, Wenner, method. In this
method a low-frequency alternating current (AC) passes between the two outer probes.
The two inner probes measure the voltage drop (V). As explained in Section 2.2.4.2.2,
surface electrical resistivity was calculated as:

=2a

V
,
I

where, a is the probe spacing (mm) which can be adjusted in this research program, I is
the electrical current (mA) and V is the voltage drop (V). The probe spacings used in this
research program were 20, 30, 40 and 50mm for concrete slabs and 25, 30, 40, and 50mm
for concrete cylinders.

4.4.2.1 Surface electrical resistivity of concrete cylinders


Despite the fact that the Wenner method is a simple and convenient test method, probe
spacing should be adjusted carefully. Surface electrical resistivity is directly proportional
to the probe spacing (a).
Three water saturated 100 x 200 mm concrete cylinders were tested at each age.
Surface electrical resistivity of a concrete cylinder was the average of four measurements
done longitudinally on the cylinder surface along its length as shown in Figure 3.7.
The average of twelve resistivity values at each time is reported in Table 4.15.

Table 4.15: Concrete cylinders apparent surface electrical resistivity values

Mixture

HPC

Probe

Surface Electrical Resistivity

Spacing

(K.cm)

(mm)

Day 3

Day 7

Day 28

Day 56

Day 91

25

15.5

36.1

106.7

112.8

121.7

30

17.1

42.9

125.7

133.8

144.4

40

24.5

54.2

163.7

167.8

174.7

50

28.1

71.5

207.3

211.9

220.6

174

HPC+

SFSL 0.40

PCSL 0.40

PCSL 0.40+

PCSL 0.45

PCSL 0.45+

PC 0.45

PC 0.45+

25

12.3

37.0

105.7

109.3

114.7

30

13.5

40.7

121.7

125.7

132.3

40

17.7

51.8

153.1

156.0

160.3

50

22.0

66.7

198.8

200.7

208.3

25

17.8

26.0

88.3

97.6

109.6

30

19.8

37.3

108.5

113.3

122.0

40

24.6

46.6

137.8

146.7

156.4

50

32.5

60.6

180.4

189.3

196.3

25

8.4

17.4

31.0

33.9

48.3

30

9.0

17.5

37.4

41.3

57.3

40

12.4

22.6

47.0

51.7

65.3

50

16.1

29.6

61.4

69.2

88.6

25

8.2

10.6

21.2

27.8

40.1

30

9.2

11.5

23.7

35.0

45.7

40

11.6

15.2

29.5

43.9

58.1

50

15.3

20.1

39.3

57.3

79.7

25

6.0

8.4

21.0

21.6

40.1

30

6.6

10.0

23.4

24.5

45.0

40

8.3

12.7

28.8

32.5

56.7

50

11.8

16.8

39.3

47.4

74.2

25

4.7

8.2

19.3

27.1

30.3

30

5.0

9.0

21.1

29.3

35.4

40

6.6

11.3

27.4

35.9

45.8

50

9.0

15.2

36.6

46.3

60.9

25

4.0

5.0

9.9

12.3

15.3

30

4.4

5.8

10.5

13.4

17.4

40

5.6

6.9

14.3

17.0

21.9

50

7.2

9.4

19.1

22.8

28.4

25

4.1

4.8

7.8

10.2

11.7

30

4.5

5.3

8.7

11.5

13.3

40

5.9

6.7

10.5

14.6

17.1

50

7.7

8.6

13.2

19.7

23.2

175
Figures 4.59- 4.62 represent the average electrical resistivity of concrete mixtures (three
cylinders) with different probe spacings. These graphs are plotted based on the values
presented in Table 4.15.

a= 25 mm

S urface Resistivity (K .c m )

250
HPC

200

HPC+
SFSL 0.40

150

PCSL 0.40
PCSL 0.40+

100

PCSL 0.45
PCSL 0.45+

50

PC 0.45
PC 0.45+

0
03 7

28

Age (day) 56

91

Figure 4.59: Surface electrical resistivity of concrete cylinders (25 mm probe spacing)

a= 30 mm

Surface Resistivity (K.cm)

250
HPC

200

HPC+
SFSL 0.40

150

PCSL 0.40
PCSL 0.40+

100

PCSL 0.45
PCSL 0.45+

50

PC 0.45
PC 0.45+

0
03 7

28

Age (day) 56

91

Figure 4.60: Surface electrical resistivity of concrete cylinders (30 mm probe spacing)

176

a= 40 mm

S urface Resistivity (K .c m )

250
HPC

200

HPC+
SFSL 0.40

150

PCSL 0.40
PCSL 0.40+

100

PCSL 0.45
PCSL 0.45+

50

PC 0.45
PC 0.45+

0
03 7

28

Age (day) 56

91

Figure 4.61: Surface electrical resistivity of concrete cylinders (40 mm probe spacing)

a= 50 mm

Surface Resistivity (K.cm )

250
HPC

200

HPC+
SFSL 0.40

150

PCSL 0.40
PCSL 0.40+

100

PCSL 0.45
PCSL 0.45+

50

PC 0.45
PC 0.45+

0
03 7

28

Age (day) 56

91

Figure 4.62: Surface electrical resistivity of concrete cylinders (50 mm probe spacing)

In all cases, surface electrical resistivity increased with concrete age because of the
continued hydration. This improvement of surface resistivity was rapid at early ages of
the silica fume mixes. Slag increased the resistivity values significantly at later ages.

177
It is important to mention that three cylinders were tested each time (four measurements
on each cylinder). Therefore, the values presented in Figures 4.59, 4.60, 4.61, and 4.62
were the average of twelve measurements at each time.

4.4.2.1.1 Statistical analysis of concrete cylinders surface resistivity


Twelve measurements were taken for each concrete mixture at each time throughout the
cylinders surfaces. Table 4.16 presents the resistivity measurement results for the
concrete cylinders for the standard probe spacing (50 mm). Similar data can be presented
for the other probe spacings.

Table 4.16: Statistical analysis of resistivity measurement of concrete cylinders (a= 50mm)

Mixture

Statistics

Statistical Summary
Day 3

Day 7

Day 28

Day 56

Day 91

12

12

12

12

12

28.17

71.53

207.28

211.91

220.58

5.06

12.10

6.50

8.10

8.65

COV (%)

17.96

16.92

3.14

3.82

3.92

12

12

12

12

12

21.98

66.69

198.75

200.67

208.26

2.30

8.97

5.97

7.55

7.89

COV (%)

10.48

13.45

3.00

3.76

3.79

12

12

12

12

12

32.50

60.62

180.41

189.27

196.25

2.27

2.95

5.28

7.53

6.89

COV (%)

6.99

4.87

2.93

3.98

3.51

12

12

12

12

12

16.11

29.57

61.37

69.22

88.58

0.96

1.05

4.58

2.07

2.15

COV (%)

5.94

3.54

7.46

2.99

2.43

n
Mean

HPC

(K.cm)

Mean

HPC+

(K.cm)

Mean

SFSL 0.40

(K.cm)

Mean

PCSL 0.40

(K.cm)

178
n

12

12

12

12

12

15.33

20.13

39.29

57.28

79.70

0.47

0.88

2.11

1.92

4.08

COV (%)

3.09

4.38

5.37

3.35

5.12

12

12

12

12

12

11.77

16.83

39.25

47.43

74.17

0.79

1.23

1.94

2.32

2.94

COV (%)

6.74

7.32

4.94

4.88

3.97

12

12

12

12

12

8.97

15.23

36.60

46.32

60.88

0.52

0.59

0.94

3.19

3.32

COV (%)

5.82

3.90

2.57

6.89

5.46

12

12

12

12

12

7.20

9.40

19.15

22.78

28.37

0.64

0.87

1.05

0.99

1.38

COV (%)

8.91

9.20

5.46

4.35

4.86

12

12

12

12

12

7.67

8.63

13.20

19.68

23.22

0.75

0.77

0.62

1.54

1.41

COV (%)

9.76

8.91

4.67

7.81

6.09

Mean

PCSL 0.40+

(K.cm)

Mean

PCSL 0.45

(K.cm)

Mean

PCSL 0.45+

(K.cm)

Mean

PC 0.45

(K.cm)

Mean

PC 0.45+

*=

(x

) 2

n 1

(K.cm)

(sample standard deviation)

Coefficient of variation (

100% ) represents the variability of a set of numbers. The

variability of a set of resistivity measurements is low (lower than 10%) in most of the
mixtures. Both high performance concretes had a COV higher than 10 % at 3 and 7 days.
Therefore, there was not a notable difference between the twelve measurements in all
mixes and the average resistivity was confidentially the electrical resistivity of the
mixture.

179

4.4.2.2 Surface electrical resistivity of concrete slabs


Two 406 x 75 mm circular slabs were used for surface electrical resistivity at each age.
The moisture content of concrete slabs was an affecting factor on resistivity values, so
electrical resistivity was measured after the field sorptivity test. Hence, the slab top
surface was wetted for ~20 minutes prior to the surface electrical resistivity test. The
amount of water absorbed during the sorptivity test was enough to make the pore
structure of the affected depth by the resistivity equi-potential lines saturated. This
assumption is acceptable because in some slabs, electrical resistivity was re-measured
after 10 more minutes water contact and the re-measured resistivity values was not
statistically different from the initial resistivity values.
The average of eight measurements on each concrete slab, measuring pattern is shown in
Figure 4.63, was used.

Figure 4.63: Surface resistivity measuring order for circular concrete slabs (top view)

180
Five different probes spacing (20, 30, 40, and 50 mm) were used. The electrical
resistivity of each slab with different probe spacings is presented in Figures 70 to 78.

4.4.2.2.1 Circular slab number 1


Surface electrical resistivity of concrete slabs which is the average of eight measurements
at each time is graphed in Figures 4.64, 4.65, 4.66, and 4.67.
Sign a represents the resistivity meters probe spacing.

Figure 4.64: Surface electrical resistivity of concrete slabs labelled number 1 (20 mm probe spacing)

181

Figure 4.65: Surface electrical resistivity of concrete slabs labelled number 1 (30 mm probe spacing)

Figure 4.66: Surface electrical resistivity of concrete slabs labelled number 1 (40 mm probe spacing)

182

Figure 4.67: Surface electrical resistivity of concrete slabs labelled number 1 (50 mm probe spacing)

4.4.2.2.2 Circular slab number 2


Surface electrical resistivity of concrete slabs which is the average of eight measurements
at each time is graphed in Figures 4.68, 4.69, 4.70, and 4.71.
Sign a represents the resistivity meters probe spacing.

Figure 4.68: Surface electrical resistivity of concrete slabs labelled number 2 (20 mm probe spacing)

183

Figure 4.69: Surface electrical resistivity of concrete slabs labelled number 2 (30 mm probe spacing)

Figure 4.70: Surface electrical resistivity of concrete slabs labelled number 2 (40 mm probe spacing)

184

Figure 4.71: Surface electrical resistivity of concrete slabs labelled number 2 (50 mm probe spacing)

In can be seen that electrical resistivity were influenced by the addition of the
supplementary cementing materials, the W/CM ratio, and age of concrete. Due to less
connected pores and smaller pore size distribution caused by cement and SCMs
secondary hydration, electrical resistivity was attributed to increasing age of the concrete.

4.4.2.2.3 Statistical Analysis of concrete slabs surface resistivity


Two concrete slabs were tested at each age; the average surface electrical resistivities are
shown in Table 4.17.

Table 4.17: Average apparent surface electrical resistivity of circular concrete slabs
Probe

Mixture

HPC

Surface Electrical Resistivity (K.cm)

Spacing/
Age

a= 20 mm

a= 30 mm

a= 40 mm

a= 50 mm

Day 3

7.0

8.9

9.8

11.3

Day 7

21.8

24.0

26.7

29.6

Day 14

71.2

83.3

88.0

98.1

Day 28

109.8

153.1

151.1

166.5

Day 56

197.7

250.5

261.2

264.9

Day 91

230.4

295.3

300.2

311.3

185

HPC+

SFSL 0.40

PCSL 0.40

PCSL 0.40+

PCSL 0.45

PCSL 0.45+

Day 3

5.5

6.9

7.3

8.5

Day 7

17.0

21.1

22.9

26.3

Day 14

66.4

77.5

80.5

89.9

Day 28

99.6

135.5

135.9

157.7

Day 56

142.6

182.8

199.4

226.4

Day 91

154.3

194.3

230.7

254.6

Day 3

10.0

12.7

14.1

15.2

Day 7

16.3

19.5

21.8

25.2

Day 14

50.9

58.5

61.5

70.6

Day 28

90.1

120.7

120.4

134.2

Day 56

122.5

166.8

181.1

183.6

Day 91

134.9

180.4

194.8

215.0

Day 3

3.4

4.5

5.1

6.0

Day 7

8.3

10.3

11.0

12.5

Day 14

17.2

20.7

22.5

26.0

Day 28

28.3

31.7

33.9

35.8

Day 56

43.5

47.0

49.3

52.4

Day 91

54.8

75.0

81.4

99.0

Day 3

3.4

4.5

5.1

6.0

Day 7

5.9

7.9

9.0

10.6

Day 14

16.1

18.1

19.2

22.6

Day 28

22.7

28.1

30.3

33.0

Day 56

31.0

35.6

41.7

49.8

Day 91

46.3

57.2

76.4

90.5

Day 3

3.0

3.9

4.4

5.1

Day 7

4.5

6.0

6.9

7.9

Day 14

13.6

17.5

17.7

20.1

Day 28

17.7

22.4

24.8

28.6

Day 56

22.0

30.6

35.1

40.1

Day 91

33.2

49.6

60.1

77.9

Day 3

2.2

3.0

3.3

3.8

Day 7

4.5

5.5

6.2

7.3

Day 14

11.6

14.5

15.7

17.6

Day 28

14.5

20.2

22.0

24.6

Day 56

20.4

31.1

34.9

43.4

Day 91

28.4

43.1

54.6

65.9

186

PC 0.45

PC 0.45+

Day 3

2.8

3.2

3.7

4.2

Day 7

3.5

4.0

4.5

5.0

Day 14

6.7

7.7

8.2

8.7

Day 28

9.5

11.2

11.2

12.4

Day 56

13.4

16.8

16.9

18.4

Day 91

22.5

31.5

33.4

34.6

Day 3

2.3

2.8

2.9

3.7

Day 7

3.2

3.5

4.0

4.6

Day 14

6.4

8.0

7.8

8.4

Day 28

8.6

10.2

9.9

10.7

Day 56

11.5

15.0

14.5

15.8

Day 91

19.7

25.3

27.9

26.7

Sixteen measurements were taken for each concrete mixture at each time throughout the
circular slabs finished surfaces (eight measurements on each pair of slab). Table 4.18
presents the statistical analysis for resistivity measurement values obtained from the
concrete slabs for the standard probe spacing (50 mm).

Table 4.18: Statistical analysis of resistivity measurement of concrete slabs (a= 50mm)

Mixture

Statistics

Statistical Summary
Day 3

Day 7

Day 14

Day 28

Day 56

Day 91

16

16

16

16

16

16

11.28

29.61

98.07

166.53

264.88

311.26

1.40

2.53

7.23

8.22

10.86

9.99

COV (%)

12.45

8.56

7.38

4.94

4.10

3.21

16

16

16

16

16

16

8.49

26.30

89.85

157.71

226.38

254.61

0.79

2.78

7.84

6.25

6.00

4.86

COV (%)

9.30

10.58

8.72

3.96

2.65

1.91

16

16

16

16

16

16

15.16

25.19

70.63

134.20

183.57

214.97

1.39

2.77

6.29

6.51

7.54

7.98

COV (%)

9.20

11.00

8.90

4.85

4.11

3.71

n
Mean

HPC

(K.cm)

Mean

HPC+

(K.cm)

Mean

SFSL 0.40

(K.cm)

187
n

16

16

16

16

16

16

5.98

12.53

25.96

35.84

52.39

99.04

0.88

1.77

4.09

2.37

3.43

6.06

COV (%)

14.70

14.15

15.76

6.60

6.54

6.12

16

16

16

16

16

16

6.08

10.56

22.64

32.96

49.76

90.53

0.98

1.64

3.63

0.70

2.16

3.76

COV (%)

16.08

15.51

16.03

2.12

4.35

4.15

16

16

16

16

16

16

5.06

7.93

20.10

28.60

40.11

77.87

0.52

1.03

3.16

2.09

3.12

5.13

COV (%)

10.20

13.01

15.74

7.31

7.78

6.59

16

16

16

16

16

16

3.84

7.25

17.63

24.59

43.44

65.86

0.53

0.85

2.03

1.32

1.77

2.56

COV (%)

13.82

11.71

11.52

5.38

4.08

3.89

16

16

16

16

16

16

4.23

4.99

8.66

12.36

18.35

34.56

0.25

0.45

0.74

0.82

1.09

17.87

COV (%)

6.02

9.08

8.51

6.61

5.95

5.17

16

16

16

16

16

16

3.65

4.56

8.43

10.65

15.84

26.74

0.44

0.42

1.12

0.66

0.46

0.91

COV (%)

11.96

9.17

13.34

6.17

2.90

3.42

Mean

PCSL 0.40

(K.cm)

Mean

PCSL 0.40+

(K.cm)

Mean

PCSL 0.45

(K.cm)

Mean

PCSL 0.45+

(K.cm)

Mean

PC 0.45

(K.cm)

Mean

PC 0.45+

(K.cm)

The coefficient of a set of resistivity measurements is low (less than 20%) in most of the
mixtures. Binary mixes (25% slag replacement) had higher COV in early ages (between
10% and 20%).

188
Now that all coefficients of variations were generally low, the average resistivity of two
slabs was confidentially the electrical resistivity of the mixture.

4.4.2.3 Unifying the surface resistivity values


The electrical resistivity is a number for each material (at a given age, temperature and
moisture content) while the Wenner resistivity values are influenced by the probe spacing
as shown in Tables 4.15 and 4.17. In other words, different resistivity values were
obtained by using different probe spacings.
The Wenner probe has been design based on an assumption that the electrodes are in
contact with the face of a semi-infinite uniform body which can be approximated by
using the optimum probe spacing (Gowers and Millard, 1999). Therefore, using any other
probe spacing higher than the optimum probe spacing (25 mm for cylinders and 20 mm
for slabs as calculated in Section 3.2.1.1.1.2) must be calibrated. As explained in Section
2.2.4.2.2, concrete resistivity is calculated from a relation between apparent resistivity
measured by the resistivity-meter with probe spacings other than the optimum spacing
and the cell constant correction factor, K, a function of inter-probe distance, a, and the
geometry of concrete body (Morris et al., 1996). The cell constant correction factor can
be calculated based on the graph recommended by Morris et al. (1996) as shown in
Figures 2.28 and 2.29.
In contrast, the concept of optimum probe spacing for achieving the semi-infinite body
assumption was not considered in the relations shown in Figures 2.28 and 2.29. Electrical
resistivity values measured by the optimum probe spacing are the actual resistivity of
concrete and do not have to be reduced by the correction factor (K must be 1), while the
cell constant correction factor obtained from those figures for the optimum probe spacing
was more than 1. Therefore, the cell constant factors recommended by Morris et al.
(1996) must be modified.
The optimum cell constant factor (KOptimum) obtained for the optimum probe spacing can
be divided by the cell constant correction factor (Kprobe) to obtain the actual resistivity. It
results in a cell constant conversion factor, .

189
Hence,

K Optimum
K Pr obe

actual = app.
The cell constant conversion factor not only unifies the electrical resistivity values
obtained from different probe spacings, but also converts the apparent resistivity into the
actual electrical resistivity representing the resistivity of a semi-infinite concrete body. In
this research program, Figure 2.28 was used for calculating the cell constant correction
factor for both cylindrical and slab specimens. This Figure was originally designed for
concrete cylinders, but it has been modified for concrete slabs as shown in Figure 4.72.
In this research program, the relation between the length and diameter of cylinders (L/d)
is 2 and Figure 2.28 was designed based on this relation. The L/d ratio in circular
concrete slabs used in this project was 5. Therefore, for calculating the cell constant
factor, a factor of

5
must be multiplied by the thickness of the concrete slabs as shown in
2

Figure 4.72.

Figure 4.72: Cell constant correction factor for specimen used


(Modified for circular concrete slabs after Morris et al., 1996)

190
The cell constant conversion factors for the probe spacings used in this research program
are tabulated in Table 4.19. These numbers are based on relations presented in Figure
4.72.
Table 4.19: Cell constant conversion factors for different probe spacings

Specimen

Probe Spacing
(mm)

Cell Constant
Conversion Factor

a=25

100 x 200 mm
Cylinders

(Optimum)

based on relations on

1.40
= 0.88
1.60

a=40

1.40
= 0.69
2.05

a=50

1.40
= 0.53
2.65

Figure 4.72)

a= 20

406 x 75 mm
Circular Slabs

(Optimum)

based on relations on

1.00
= 1.00
1.00

a= 30

1.00
= 0.80
1.25

a= 40

1.00
= 0.74
1.35

a= 50

1.00
= 0.65
1.55

(The optimum Cell


Constant factor = 1.00

1.40
= 1.00
1.40

a=30

(The optimum Cell


Constant factor = 1.40

()

Figure 4.72)

Therefore, the electrical resistivity values of concrete cylinders were recalculated based
on the measured apparent resistivity values presented in Table 4.15 and the cell constant
conversion factors from Table 4.19 as shown in Table 4.20.

191
Table 4.20: Surface resistivity of concrete cylinders with different probe spacings
Probe

Mixture

Spacing

Surface Electrical Resistivity (actual


(K.cm)

(mm)

HPC

HPC+

SFSL 0.40

PCSL 0.40

PCSL 0.40+

PCSL 0.45

PCSL 0.45+

PC 0.45

= app.)

Day 3

Day 7

Day 28

Day 56

Day 91

25

15.5

36.1

106.7

112.8

121.7

30

0.88

15.1

37.7

110.6

117.8

127.1

40

0.69

16.9

37.4

112.9

115.7

120.5

50

0.53

14.9

37.9

109.9

112.3

116.9

25

12.3

37.0

105.7

109.3

114.7

30

0.88

11.9

35.8

107.1

110.6

116.4

40

0.69

12.2

35.7

105.7

107.6

110.6

50

0.53

11.7

35.3

105.3

106.4

110.4

25

17.8

26.0

88.3

97.6

109.6

30

0.88

17.4

32.8

95.5

99.7

107.4

40

0.69

17.0

32.2

95.1

101.2

107.9

50

0.53

17.2

32.1

95.6

100.3

104.0

25

8.4

17.4

31.0

33.9

48.3

30

0.88

7.9

15.4

32.9

36.3

50.4

40

0.69

8.5

15.6

32.4

35.7

45.0

50

0.53

8.5

15.7

32.5

36.7

46.9

25

8.2

10.6

21.2

27.8

40.1

30

0.88

8.1

10.1

20.8

30.8

40.2

40

0.69

8.0

10.5

20.4

30.3

40.1

50

0.53

8.1

10.7

20.8

30.4

42.2

25

6.0

8.4

21.0

21.6

40.1

30

0.88

5.8

8.8

20.6

21.6

39.6

40

0.69

5.7

8.8

19.9

22.4

39.1

50

0.53

6.2

8.9

20.8

25.1

39.3

25

4.7

8.2

19.3

27.1

30.3

30

0.88

4.4

7.9

18.6

25.8

31.2

40

0.69

4.6

7.8

18.9

24.8

31.6

50

0.53

4.8

8.1

19.4

24.6

32.3

25

4.0

5.0

9.9

12.3

15.3

30

0.88

3.8

5.1

9.2

11.8

15.3

192

PC 0.45+

40

0.69

3.9

4.7

9.8

11.7

15.1

50

0.53

3.8

5.0

10.1

12.1

15.0

25

4.1

4.8

7.8

10.2

11.7

30

0.88

4.0

4.7

7.7

10.1

11.7

40

0.69

4.1

4.6

7.2

10.1

11.8

50

0.53

4.1

4.6

7.0

10.4

12.3

The corrected surface electrical resistivity values of concrete which were the average of
four united and modified resistivity values obtained by different probe spacings was
independent of the Wenner probe electrode spacings as shown in Table 4.21.

Table 4.21: True surface electrical resistivity of concrete cylinders

Surface Electrical Resistivity


(K.cm)

Mixture
Day 3

Day 7

Day 28

Day 56

Day 91

HPC

15.6

37.3

110.0

114.7

121.6

HPC+

12.0

36.0

106.0

108.5

113.0

SFSL 0.40

17.4

30.8

93.6

99.7

107.2

PCSL 0.40

8.3

16.0

32.2

35.7

47.7

PCSL 0.40+

8.1

10.5

20.8

29.8

40.7

PCSL 0.45

5.9

8.7

20.6

22.7

39.5

PCSL 0.45+

4.6

8.0

19.1

25.6

31.3

PC 0.45

3.9

5.0

9.8

12.0

15.2

PC 0.45+

4.1
4.7
7.4
10.2
11.9
It can be seen that the average resistivity values unified by the cell constant conversion
factor were approximately the same as the resistivity values obtained by using the
optimum probe spacing. In other words, the surface resistivity values presented in Table
4.21 can be taken as the surface electrical resistivities of semi-infinite bodies of concrete
at different ages.
Surface electrical resistivity values of circular concrete slabs can be unified based on the
calculated cell constant conversion factors presented in Table 4.19. The converted surface
electrical resistivities of concrete slabs which represent the resistivity of a semi-infinite
concrete body are tabulated in Table 4.22.

193
Table 4.22: Surface resistivity of concrete slabs with different probe spacings
Surface Electrical Resistivity (actual

Probe

Mixture

Spacing

(K.cm)

(mm)

HPC

HPC+

SFSL 0.40

PCSL 0.40

PCSL 0.40+

PCSL 0.45

PCSL 0.45+

PC 0.45

= app.)

Day 3

Day 7

Day 14

Day 28

Day 56

Day 91

20

7.0

21.8

71.2

109.8

197.7

230.4

30

0.80

7.1

19.2

66.6

122.5

200.4

236.2

40

0.74

7.3

19.7

65.1

111.9

193.3

222.1

50

0.65

7.3

19.2

63.7

108.2

172.2

202.3

20

5.5

17.0

66.4

99.6

142.6

154.3

30

0.80

5.5

16.9

62.0

108.4

146.2

155.4

40

0.74

5.4

16.9

59.5

100.5

147.6

170.7

50

0.65

5.5

17.1

58.4

102.5

147.1

165.5

20

10.0

16.3

50.9

90.1

122.5

134.9

30

0.80

10.2

15.6

46.8

96.6

133.4

144.3

40

0.74

10.4

16.1

45.5

89.1

134.0

144.1

50

0.65

9.9

16.4

45.9

87.2

119.3

139.7

20

3.4

8.3

17.2

28.3

43.5

54.8

30

0.80

3.6

8.2

16.5

25.4

37.6

60.0

40

0.74

3.8

8.1

16.6

25.1

36.5

60.2

50

0.65

3.9

8.1

16.9

23.3

34.1

64.4

20

4.1

5.9

16.1

22.7

31.0

46.3

30

0.80

4.0

6.3

14.5

22.5

28.5

45.8

40

0.74

3.8

6.7

14.2

22.4

30.8

56.5

50

0.65

3.9

6.9

14.7

21.4

32.3

58.8

20

3.0

4.5

13.6

17.7

22.0

33.2

30

0.80

3.1

4.8

14.0

17.9

24.5

39.7

40

0.74

3.3

5.1

13.1

18.3

26.0

44.5

50

0.65

3.3

5.2

13.1

18.6

26.1

50.6

20

2.2

4.5

11.6

14.5

20.4

28.4

30

0.80

2.4

4.4

11.6

16.1

24.9

34.4

40

0.74

2.4

4.6

11.6

16.3

25.8

40.4

50

0.65

2.5

4.7

11.5

16.0

28.2

42.8

20

2.8

3.5

6.7

9.5

13.4

22.4

30

0.80

2.6

3.2

6.2

9.0

13.4

25.2

194

PC 0.45+

40

0.74

2.7

3.3

6.1

8.3

12.5

24.7

50

0.65

2.8

3.2

5.6

8.0

11.9

22.5

20

2.3

3.2

6.4

8.6

11.5

19.7

30

0.80

2.2

2.8

6.4

8.2

12.0

20.3

40

0.74

2.1

2.9

5.8

7.3

10.8

20.6

50

0.65

2.4

3.0

5.5

6.9

10.3

17.4

Surface electrical resistivity of concrete which is the average of four united and modified
resistivity values obtained by different probe spacings is independent from the Wenner
probe electrode spacing as shown in Table 4.23.

Table 4.23: True surface electrical resistivity of concrete slabs


Surface Electrical Resistivity

Mixture

(K.cm)
Day 3

Day 7

Day 14

Day 28

Day 56

Day 91

HPC

7.2

20.0

66.7

113.1

190.9

222.8

HPC+

5.5

17.0

61.6

102.8

145.9

161.5

SFSL 0.40

10.1

16.1

47.3

90.8

127.3

140.8

PCSL 0.40

3.7

8.2

16.8

25.5

37.9

59.8

PCSL 0.40+

4.0

6.4

14.9

22.3

30.7

51.9

PCSL 0.45

3.2

4.9

13.4

18.1

24.6

42.0

PCSL 0.45+

2.4

4.5

11.6

15.7

24.8

36.5

PC 0.45

2.7

3.3

6.1

8.7

12.8

23.7

PC 0.45+

2.3

3.0

6.0

7.7

11.1

19.5

Since the true surface resistivity values for specimens is the average of unified resistivity
values obtained by different probe spacings, true resistivity of slabs and cylinders are not
exactly equal. The reasons will be explained briefly in Section 4.4.2.6.
Five practical steps are recommended for use of the Wenner probe for measuring the
surface electrical resistivity of a concrete structure:

195

Step I) Surface electrical resistivity is measured by the Wenner probe on site without
limiting the probe spacing (any probe spacing can be used). The cell constant
correction factor (Kprobe) for the applied probe spacing is found based on Figure 4.72.

Step II) The optimum probe spacing which represents the electrical resistivity of a
semi-infinite body is calculated based on the technical graphs presented in Section
3.2.1.1.1.2 or from Figure 4.73.

Figure 4.73: Required limitations for calculating the optimum probe spacing
(Gowers and Millard, 1999)

MTO bridge decks are typically 225 mm thick with 70 20 mm cover above rebars
and maximum aggregate size of 19 mm, so the optimum probe spacing for most of
the MTO projects is 50 mm.

Step III) The optimum cell constant correction factor (KOptimum) is calculated based
on the optimum probe spacing (Step II) and the relations illustrated in Figure 4.72.

196

Step IV) The cell constant conversion factor () is calculated as:

K Optimum
K Pr obe

Step V) The actual electrical resistivity of concrete which is independent from the
probe spacing and a unique number for a mix design is calculated by multiplying the
conversion factor and apparent resistivity measured in Step I.

actual = app.
4.4.2.4 W/CM ratio effects on surface electrical resistivity
As Figure 4.74 shows, the lower the W/CM ratio, the higher the surface electrical
resistivity. This statement is true for concrete cylinders and slabs, as a similar order was
seen for the concrete slabs electrical resistivity (Sections 4.4.2.2.1 and 4.4.2.2.2).

W/CM
<======
Increasing

140

100
Silica fume + slag

HPC
HPC+
SFSL 0.40

80

PCSL 0.40

60

<====== <======
Increasing Increasing

Surface Resistivity (K.cm)

120

Slag
40
20
No SCM

0
0 3 7

28

Age (day) 56

PCSL
0.40+
PCSL 0.45
PCSL
0.45+
PC 0.45
PC 0.45+

91

Figure 4.74: W/CM ratio effects on the surface electrical resistivity of concrete cylinders

Lower W/CM ratio resulted in lower porosity and connected pores. In this case, the
induced electrical current by the two outer electrodes could not pass through concrete
pores system easily, so higher electrical resistivity was measured. Discontinuous pore
structure results in high resistivity paths available for ion movement.

197

4.4.2.5 SCMs effects on surface electrical resistivity


Figures 4.75 and 4.76 show the slag and silica fume effects on the surface electrical
resistivity of concrete cylinders and slabs.

120

W/CM
Silica fume + Slag
0.40

Surface Resistivity (K.cm)

100

SFSL 0.40

80
PCSL 0.40

60
Slag

0.40

40

PCSL 0.45

0.45

Slag

PC 0.45

20

0.45
No SCM

0
0 3 7

28

Age (day) 56

91

Figure 4.75: SCMs effects on the surface electrical resistivity of concrete cylinders

W/CM

Surface Resistivity (K .cm)

150

Silica fume +
Slag

0.40

SFSL 0.40

100
PCSL 0.40

0.40

Slag

50

PCSL 0.45

0.45
Slag

0.45

No SCM

03 7 14

28

56

Age (day)

91

Figure 4.76: SCMs effects on the surface electrical resistivity of circular slabs

PC 0.45

198
As mentioned in the literature, adding silica fume decreases the ionic concentration of the
pore solution because of its secondary hydration process which reduces the free alkalis
(Na+, K+ and mostly OH-) content in cement paste. The significant influence of adding
silica fume to concrete mixes can be appreciated by comparing data for mixes with 8%
silica fume and without silica fume (the two upper line in Figures 4.75 and 4.76). 8%
silica fume replacement increased the electrical resistivity of concrete significantly, but
not because of the reduced ionic concentration. Used SCMs reduced the pore system
connectivity. Also using silica fume resulted in smaller pore size distribution. In addition,
concrete mix which contained slag had higher surface electrical resistivity than plain
cement concrete mix (W/CM= 0.45 in Figures 4.75 and 4.76).
At later ages, hydration of slag made the pore system denser and less porous.
Consequently, the electrical resistivity of concrete ternary mixes containing silica fume
and slag were higher than binary mixes. Plain cement concrete mixes had the lowest
resistivity at both early and later ages.

4.4.2.6 Effects of specimen geometry on the Wenner probe values


Two types of concrete specimens were tested in this research program: 200 x 100 mm
concrete cylinders and 406 x 75 mm circular concrete slabs. It was expected that the
surface electrical resistivity values measured by the Wenner probe were geometryindependent because the electrical resistivity is influenced only by the pore system
properties and the specimens moisture content. Both types of specimens were water
saturated during the test. Figure 4.77 compares the surface resistivity of the specimens
with the optimum probe spacing calculated in Section 3.2.1.1.1.2 according to the
technical relations recommended by Gowers and Millar (1999).

199

150

Concrete Slab (a= 20 mm)

125

100

75

y = 0.97 x - 2.33
2

R = 0.98
50

25

0
0

25

50

75

100

125

150

Concrete Cylinder (a= 25 mm)

Figure 4.77: Relation between different specimens resistivity with the optimum probe spacing

Previous work has concluded that surface resistivity values are dependent to specimen
geometry as larger concrete volume of a specimen results in lower resistivity value
(Sengul and Gjorv, 2008). However it has been found that geometry effects can be
neglected by applying the optimum probe spacing as shown in Figure 4.77. Hence, the
Wenner probe can measure the actual resistivity of concrete independent of the specimen
geometry if the optimum probe spacing is applied. Surface resistivity values of concrete
cylinders were al little higher than that in slabs since cylinders were kept in the fog room
prior to the testing time. Therefore, cementing materials were fully hydrated in cylinders
while the rate of hydration in slabs (kept in 50% RH room) were lower which resulted in
less discontinuous pore system and lower Wenner resistivity values (the moisture
contents of the influenced depths were the same since the least probe spacings were used
and cylinders and slabs surfaces were saturated).

200
In contrast, the other probe spacings give scattered resistivity values as shown in Figure
4.78.

Figure 4.78: Comparison between specimens resistivity values and different probe spacings

It can be seen in Figure 4.78 that as probe spacing increased, surface resistivity measured
by the Wenner probe became more independent from specimen geometry (difference
between slabs and cylinders became near zero).
Two conclusions can be derived based on the last two figures:
a) If the optimum probe spacing, which can be found based on the specimens
thickness, maximum aggregate size, rebars location, and distance between the
probe array and specimen edges is applied, the measured resistivity values are
independent from the specimens geometry. Electrical resistivity measured by the
optimum probe spacing represents resistivity of a semi-infinite concrete body
(Figure 4.77).

201
b) Resistivity measurement tends to be representative of the concrete region located
between the electrodes current density distribution. Concrete slabs were removed
from the moist room after seven days wet curing. Prior to the Wenner test, slabs
were soaked for approximately 16 minutes (because of the field sorptivity test)
and then kept soaked till concrete surface became saturated. The surface layers of
concrete slabs were saturated, so the applied current between two outer probes of
the optimum probe spacing (a= 20 mm) passed thought the saturated volume. In
this case, the situation in concrete slabs was the same as in saturated concrete
cylinders which was proved in Figure 4.77. As the probe spacing increased, the
penetration depth of the applied current was increased, so current may have
passed through a partially wet concrete volume which resulted in higher apparent
resistivity (Figure 4.78). A probe spacing of 50 mm resulted in the highest current
penetration through the specimens depth. In this case applied current path may
pass through the core of the circular slab where the last drying level was, as
shown schematically in Figure 4.79.

Figure 4.79: Probe spacings effect on the penetration depth of the Wenner applied current

Therefore, the applied current passed through a saturated (or near saturation) volume, so
the measured resistivity was the same as the resistivity measured for a saturated concrete
cylinder (the line on the right in Figure 4.78).

202
The second explanation is just an assumption. The affected pore structure was almost
saturated prior to the resistivity test because in some cases more water saturation did not
change the measured resistivity values.
It is important to mention that in neither of the cases, the current flow was not restricted
by the concrete edges since the probe spacings were less than specimen thickness.
As a practical conclusion, the optimum probe spacing must be calculated and applied to
the Wenner probe. In this case, the measured values represent the changes in the concrete
micro structure. In case of different probe spacings, it is recommended to fully saturate
the concrete depth which is affected by the applied current path. Since the Wenner
current penetration depth is approximately equal to the probe spacing, water sorptivity
coefficient which was calculated in Section 4.3, is useful for saturating the required
depth.

4.4.2.7 Surface electrical resistivity as an indicator for other properties of concrete


It was found that the surface electrical resistivity of concrete can be well correlated with
the other properties of concrete measured by the other tests in this research program.
Some destructive tests must be used for measurement of these physical properties of
concrete while measuring the surface electrical resistivity of as-constructed concrete by
the Wenner probe is non-destructive and time saving. Hence, the electrical resistivity of
concrete can be used to estimate the other properties of concrete provided that the
correlation between the resistivity and those properties are available.

4.4.2.7.1 Surface electrical resistivity and compressive strength


Both electrical resistivity and compressive strength are influenced by the concrete
porosity, as shown in Figures 4.80 and 4.81. Although higher resistivity concrete had
higher compressive strength, compressive strength is not the reason for this.

203

250

Surface Electrical Resistivity (K.cm)

0.08 x
y = 0.62 e
R2 = 0.91

200

150

100

50

0
20

30

40

50

60

70

80

Compressiv e Stre ngth (MPa)

Figure 4.80: Surface resistivity of concrete cylinders versus compressive strength

300

Surface Electrical Resistivity (K .cm)

0.1x

y = 0.23 e
R2 = 0.85

250

200

150

100

50

0
20

30

40

50

60

70

Compressive Strength (MPa)

Figure 4.81: Surface resistivity of concrete slabs versus compressive strength

80

204
It can be seen that the relations were not linear because the surface resistivity is more
influenced by the factors affecting the pore system than compressive strength. For
example, in one mix design, lowering the W/CM ratio (or adding SCMs) increases the
compressive strength by 16% where as the surface electrical resistivity was increased by
up to 60% at the same age.
Different moisture distribution in circular concrete slabs resulted in a lower R-square
coefficient in Figure 4.81 than Figure 4.80.

4.4.2.7.2 Surface electrical resistivity and total charge passed


The electrical current reported by the RCPT is not only the charge taken by the dissolved
ions in the pore system, but also the charge carried by the penetrating chloride ion. This
is one of the major disadvantages of the RCPT. On the other hand, during the surface
electrical test the induced current is carried by only chemical ion available in the pore
system. Now that there is a relation between the RCPT and surface resistivity values with
high R-square values (Figures 4.82, 4.83, and 4.84) the RCPT results can be replaced
with the surface electrical values to avoid chloride ion effects on the total charge passed.

200

Surface Electrical Resistivity (K.cm)

180
160
140

y = 15954 x

120

-0.94

R = 0.98

100
80
60
40
20
0
0

1000

2000

3000

4000

5000

6000

RCPT Passing Charge (Coulombs)

Figure 4.82: Surface resistivity of concrete cylinders versus the RCPT passing charge

7000

205
Since MTO level of concrete acceptance according to the RCPT is 1000 coulombs,
modified correlation between the Wenner probe resistivity and the RCPT charge passed
in useful as shown in Figure 4.83.

Surface Electrical Resistivity (K.cm)

180
160

y = 13632x

140

-0.91

R = 0.96
120
100
80
60
40
20
0
0

100

200

300

400

500

600

700

800

900

1000

RCPT Passing Charge (Coulombs)

Figure 4.83: Modified surface resistivity versus the RCPT results based on MTO specification

In addition, surface electrical resistivity of concrete slabs (tested in fully saturated


condition) can be correlated wit the RCPT results as shown in Figure 4.84.

206

260

Surface Electrical Resistivity (K .cm)

240

-1.1

y = 54692x

220

R = 0.95

200
180
160
140
120
100
80
60
40
20
0
0

1000

2000

3000

4000

5000

6000

7000

RCPT Passing Charge (coulombs)

Figure 4.84: Surface resistivity of concrete slabs versus the RCPT passing charge

4.4.2.7.3 Surface electrical resistivity and the other types of electrical resistivity
A well-defined correlation was observed between the surface electrical resistivity and the
first 5 min. RCPT resistivity as shown in Figures 4.85 and 4.86.

207

140

Surface Electrical Resistivity (K.cm)

120
y = 1.48 x - 0.60
2
R = 0.99
100

80

60

40

20

0
0

10

20

30

40

50

60

70

80

90

RCPT First 5 min. Electrical Resistivity (K.cm)

Figure 4.85: Surface resistivity of concrete cylinders versus the RCPT resistivity

160

Surface Electrical Resistivity (K .cm)

140

y = 1.82 x - 7.04
2
R = 0.97

120
100

80
60
40

20
0
0

10

20

30

40

50

60

70

80

RCPT First 5 min. Electrical Resistivity (K .cm)

Figure 4.86: Surface resistivity of concrete slabs versus the RCPT resistivity

90

208
In the application of the four-electrode method, a good correlation between surface
electrical resistivity and Monfore bulk resistivity was obtained as shown in Figures 4.87
and 4.88.

Surface Electrical Resistivity (K .cm)

150

y = 1.24 x + 0.79
R2 = 0.99

125

100

75

50

25

0
0

20

40

60

80

100

DC-Cyclic Bulk Resistivity (K .cm)

Figure 4.87: Surface resistivity of fully saturated cylinders versus Monfore resistivity

Surface Electrical Resistivity (K .cm)

175

y = 1.60 x - 6.63

150

R = 0.96
125
100
75
50
25
0
0

20

40

60

80

100

DC-Cyclic Bulk Resistivity (K .cm)

Figure 4.88: Surface resistivity of not fully saturated slabs versus Monfore resistivity

209
Surface electrical resistivity of concrete cylinders was better correlated to the Monfore
resistivity (higher R-square values) than surface electrical resistivity of concrete slabs
since surface resistivities of slabs cured at 50% RH room were higher than that in fully
saturated cylinders. These test results demonstrate that the curing conditions of concrete
are very important for measuring electrical resistivity of concrete. Therefore, multiplying
surface resistivity values of concrete by 1.24 will result in DC-cyclic bulk resistivity
value of same concrete provided (Figure 4.87).
In all cases, surface electrical resistivity was higher than any other type of electrical
resistivity because of the wall effects of cylindrical moulds and surface drying. Wall
effect causes finer aggregate and more cement paste distribution on specimens surface
(boundary effect). Also concrete surface dries before concrete core.
The little difference between the correlation coefficients in Figures 4.87 and 4.88 was
caused by the different moisture content of concrete slabs from cylinders. Figures 4.85
and 4.87 were plotted based on the values measured on fully saturated cylinders, so both
axes were based on saturated values.
The resistivity of surface concrete can be affected by wall effects, moisture content, water
bleeding, and aggregate segregation. Higher segregation, lower moisture content, and
presence of more cement paste on the surface layer of the concrete specimens resulted in
higher surface electrical resistivity.

4.4.2.7.4 Surface electrical resistivity and water sorptivity coefficient


Both pore structure connectivity and the quantity and mobility of the pore water are
affected the surface electrical resistivity. Also capillary pore tortuosity is affected by the
physical characteristics of the pore system and influences the water sorptivity results.
Therefore, surface electrical resistivity and water sorptivity results can be correlated as
shown in Figure 4.89.

210

Surface Electrical Resistivity (K.cm)

160
140
-1.55

y = 446.3 x
2
R = 0.95

120
100
80
60
40
20
0
0

10

12
-4

14

16

18

Water Sorptivity Coefficient (x10 mm/sec )

Figure 4.89: Surface resistivity of concrete cylinders and water sorptivity coefficient

In as much as the water sorptivity coefficients of concrete slabs were influenced by the
different moisture contents at the time of testing, the relation between the filed sorptivity
values and surface resistivity is not reliable.

4.4.2.7.5 Surface electrical resistivity and diffusion of chloride ion through concrete
Rebar corrosion initiation depends on chloride ion penetration. On the other hand,
concrete surface electrical resistivity can be related to the susceptibility for chloride ion
penetration (as shown in Figures 4.90 and 4.91).

211

50
45
40

30

-12

m /s)

35

(10

RCPT Chloride Migration Coefficient

-1.21

y = 279.8 x
R2 = 0.92

25
20
15
10
5
0
0

20

40

60

80

100

120

140

Surface Electrical Resistivity (K .cm)

Figure 4.90: Surface resistivity of concrete cylinders versus chloride migration coefficient

50
45

30

m /s)

35

(10

-12

RCPT Chloride Migration Coefficient

-0.98

y = 126.2 x
R2 = 0.88

40

25
20
15
10
5
0
0

50

100

150

200

Surface Ele ctrical Resistivity (K .cm)

Figure 4.91: Surface resistivity of concrete slabs versus chloride migration coefficient

250

212
Higher chloride migration coefficient is caused by the bigger pore size distribution and
more pores connectivity. Higher chloride migration coefficient, lower surface electrical
resistivity. Different moisture distribution in circular concrete slabs resulted in lower Rsquare coefficient in Figure 4.91 than that in Figure 4.90.

4.4.2.8 Wenner probe as a practical instrument


Although some critics have reported that the Wenner probe is not an accurate instrument
because of the different measured values at different locations of a concrete member,
statistical analyses summarized in Tables 4.16 and 4.18 have rejected this statement. As
shown before, the coefficient of variation (COV) of surface electrical resistivity
measurements were very low (fewer than 20% in all mixes) independent from the type of
the specimen and mix design. The values measured by the four-electrode resistivitymeter represent the surface resistivity of concrete with a low COV provided that enough
measurements have been done.
The Wenner probe values can be used for many applications, but among of them, one
application is the most practical. The surface electrical resistivity of concrete can be
measured non-destructively in less than three seconds. This value can be related to the
other properties of concrete which are time-consuming and destructive to be measured.
Figure 4.92 shows a practical relation between the surface electrical resistivity and other
properties of concrete (e.g. the RCPT coulombs and resistivity, water sorptivity, and etc.)
measured by the standard tests in this research program.

213

Figure 4.92: Practical relation between the surface resistivity values and other tests values

214
In addition, early age surface electrical resistivity (prior to formwork removal) can be
used for the prediction of 28 day compressive strength. The quality control of the
compressive strength of concrete is typically performed on standardized specimens such
as 100 x 200 mm cylinders used in Section 4.1 at the age of 28 days. However
construction process is continuous because of the cost and deadline issues after formwork
removal. Ferreira and Jalali (2010) have shown that the 7 day Wenner resistivity can be
used for the prediction for 28 day compressive strength since this argument has been
proved by both experimental and theoretical approaches (Ferreira and Jalali, 2010).
The prediction graphs based on 7 day Wenner resistivity values and 28 day compressive
strengths are shown in Figure 4.93.

80

28-day Compressive Strength (MPa)

70
(Slabs) y = 1.87x + 30.85
R2 = 0.98

60

(Cylinders) y = 0.91 x + 32.35

50

R2 = 0.97

40
30
20
10
0
0

10

15

20

25

30

35

7-day Surface Electrical Resistivity (K .cm)

Figure 4.93: Regression analysis of compressive strength and surface electrical resistivity

40

CHAPTER 5
CO/CLUSIO/S A/D RECOMME/DATIO/S
A durable concrete is not necessarily a high strength concrete, but rather it is an
impermeable concrete. The ability of aggressive fluids to penetrate into concrete is
influenced by concrete porosity and connectivity of the capillary pore structure. Concrete
penetrability can be measured by standard tests which are time-consuming and
destructive. However, electrical resistivity tests of concrete can also be used as indicator
of concrete penetrability. A series of experiments (compressive strength, RCP, first 5
minute RCPT resistivity, depth of chloride ion penetration by silver nitrate spraying (NT
492) using the modified ASTM C1202, Monfore DC-cyclic bulk resistivity, and surface
electrical resistivity) were performed in this research project. Based on data from this
research program, the following conclusions and recommendations have been made.

5.1 Conclusions
1- Reducing W/CM ratio, adding SCMs to mixtures, and increasing the length of
moist curing improved concrete durability. This finding was established from the results
of the RCP test, bulk resistivity, initial and secondary water sorptivity, and the surface
electrical resistivity test. Also the depth of chloride penetration (NT 492) increased as the
W/CM ratio increased, while adding SCMs and increasing the length of moist curing
decreased the depth of chloride ion penetrated.
2- Silica fume improved concrete properties at early ages and slag contributed at
later ages. Early age concrete resistivity was significantly higher in ternary mixes
contained silica fume and slag than binary mixes contained slag, while at later ages the
resistivity binary mixes contained slag was higher than that in similar W/CM ratio plain
cement concrete mixes.
3- The RCPT charge passed and the calculated first 5 min. RCPT resistivity were
influenced by pore solution conductivity, while depth of chloride ion penetrated into
concrete was only dependant on the physical characteristics of pore structure. For
example, at 91 days of age, electrical conductivity was low while chloride penetration

215

216
coefficient was zero (in concrete mixes containing 8% silica fume and 25% slag with
W/CM ratio 0.35, 0.38, and 0.40).
4- Although the total charge passed through concrete in the RCP test was influenced
by physical characteristics of the pore system, it was more sensitive to the change in pore
fluid conductivity resulting from the cementitious materials used in mix design than
W/CM ratio. Replacing 8% Portland cement by silica fume increased the 5 min. RCPT
resistivity by a factor of 3 at 28 days, while a reduction in W/CM ratio from 0.40 to 0.35
increased the 5 min. RCPT resistivity by 18% in concrete mixes contained silica fume
and slag and 28% in a binary mix containing 25% slag.
5- Concrete permeability is directly proportional to the quantity and mobility of the
pore water which are affected by the porosity and the pore structure of the hardened
cement paste. Using SCMs and reducing W/CM ratio decreased sorptivity coefficient and
RCPT charges.
6- Calculating the chloride penetration coefficient by silver nitrate spraying method
(NT 492) proved that RCPT technique is a reliable test for concretes containing SCMs
such as silica fume.
7- Extrapolating the 6 h. coulomb values from 30 minute readings is a reliable
technique to avoid the heat effects on the values, but this is only a concern with concrete
having a level of penetrability higher than very low, 1000 coulombs. By dividing
extrapolated charges by 0.82, 6 h RCPT coulombs can be estimated.
8- The internal moisture content of concrete specimens used for the laboratory
sorptivity test (ASTM C1585) was between 50% to 60%, while concrete slabs used for
the field sorptivity test had different moisture contents because of the lengthy period
required to condition the 406 x 75 mm circular slabs. Field sorptivity test results proved
that water sorptivity was increased by decreasing levels of concrete saturation through the
time period of the project. Therefore, one problem to be overcome by a field sorptivity
test is the influence of concrete moisture content on test results.
9- The rate of water absorption (initial and secondary) and the amount of water
absorbed by the concrete decreased with increased maturity. Also water sorptivity was
reduced as W/CM ratio decreased and where SCMs were used.

217
10- Three types of resistivity were measured in this research project: surface
resistivity by the Wenner probe, Monfore DC-cyclic bulk resistivity, and the first 5 min.
RCPT resistivity. Surface electrical resistivity was the fastest (less than 5 seconds),
followed by first 5 min. RCPT resistivity, and Monfore resistivity (15 minutes). These
resistivity values were well correlated (the difference between these electrical resistivities
remained constant in each mixture as concrete aged, so linear functions could relate
them) which represented that the influencing factors were similar.
11- Surface resistivity using a Wenner probe can be used as an indicator for in-place
quality of concrete and therefore long-term concrete durability. Electrical resistivity (both
bulk and surface resistivity) is a function of concrete moisture content, the ionic
conductivity of concrete pore water, and physical characteristics of the pore structure:
a. Concrete moisture content significantly affected the surface electrical
resistivity measured by the Wenner probe as concrete slabs had shown high
resistivity values before the field sorptivity test performed (RH was about
50%) while resistivity of fully saturated surfaces (after conducting the field
sorptivity test) was lower.
b. Decreasing W/CM ratio resulted in higher surface electrical resistivity values
because lower W/CM ratio resulted in lower porosity and a more
discontinuous pore structure.
c. A high electrical resistivity was not necessarily due to a fine or discontinuous
pore structure, because ionic concentration within the pore fluid influenced
the magnitude of passing current.
12- Specimen surface properties influenced surface electrical resistivity values
measured by the Wenner probe. Surface electrical resistivity values of concrete cylinders
were higher than Monfore bulk resistivity and RCPT resistivity due to the wall effects of
cylindrical molds and potentially surface drying.
13- Since the RCPT values were well correlated with the Wenner array results, the
surface electrical resistivity could be used as an indicator of the ASTM C1202 resistance
to chloride ion penetration currently used by MTO and CSA A23.1.
14- From these experiments there was an inverse correlation between concrete
resistivity and non-steady-state chloride diffusion coefficients obtained from NT 492.

218
15- The Monfore bulk resistivity test results were independent of specimen geometry
since bulk resistivity values measured by the DC-cyclic resistivity-meter for full length
cylinders and concrete discs were not significantly different.
16- A good correlation between surface electrical resistivity measured by the Wenner
4-probe array and other permeability indicator tests were found in this research.
Therefore, electrical resistivity measured by the Wenner probe can be used as an
indication of concrete permeability although electrical conductivity depends on both pore
structure and chemistry of pore solution.
17- The average surface electrical resistivity-to-bulk resistivity value ratio (

surface
)
bulk

was 1.24. This number is based on resistivity values of concrete specimens at same
moisture content. In addition, the surface electrical resistivity-to- first 5 min. RCPT
resistivity ratio (

surface
) was 1.48.
RCPT

18- Surface electrical resistivity values can be used as indicators for RCPT charge
passed although there was not a linear relation between the Wenner values and RCPT
Coulombs.
19- Measuring both the field water sorptivity and surface electrical resistivity by the
Wenner probe are relatively, quick, simple and non-destructive if used in the field. In
addition, the outer concrete surface is being tested, so covercrete quality and alternative
curing methods can be studied. However, both are influenced by surface moisture
conditions.
20- The 28-day compressive strength of concrete (MPa) can be estimated by
multiplying 7-day Wenner resistivity (K.cm) by 1.87.

219
5.3 Recommendations
The Wenner probe surface resistivity as a technique for in-situ evaluation of durability of
covercrete was studied. It can be used as a regular quality assurance procedure. The
following recommendations are made for use of the Wenner probe technique and for
modifying ASTM C1202:

1- In the lab, the Wenner probe provided meaningful results quickly. However, for
use in the field, these results (as well as field sorptivity test results) must be calibrated
with in-situ moisture contents of concrete surfaces and with laboratory testing
procedures. In the field, powder samples (10 mm deep for the field sorptivity test and up
to the probe spacing for the Wenner probe test) should be collected in order to obtain a
representative measure of the in-situ moisture condition of structure. Finally calibration
curves must be plotted for every mixture.
2- Although the low coefficient of variation (COV) of results indicated that the
different Wenner resistivity values measured on the surface of the 406 x 75 mm circular
slabs were not statistically variable, measurement locations should be close together (not
far than 100 mm). For using this non-destructive instrument, the number of
measurements performed on each slab, four diagonal and four orthogonal, can be reduced
to four (diagonal) for a 1300 cm2 area.
3- With the Wenner probe, calculating the optimum probe spacing and the cell
constant conversion factor are necessary for measuring the true surface electrical
resistivity of concrete, which is independent from probe spacing.
4- Moreover effects of changes in physical properties of pore structure, such as
continuity of porous or pore size distribution by the water sorptivity test for different
moisture histories of concrete (in pre-conditioning concrete specimens, fully dried in a
vacuum over silica gel), should be studied.
5- In concrete mixtures containing materials that affect pore solution chemistry (e.g.
silica fume), calculating the migration coefficient by the silver nitrate spraying method
(NT 492) was an indicator of the chloride penetrability of concrete since it eliminated
conductivity biases caused by the pore solution chemistry. It is recommended that ASTM
C1202-07 be modified to incorporate these additional measurements.

220

CHAPTER 6
REFERE/CES

ACI Guide 302R-04, Guide for Concrete Floor and Slab Construction, American
Concrete Institute Technical Committee Document 302, 2004, 16 pages

ACI Guide 304R-00, Guide for Measuring, Mixing, Transporting and Placing Concrete,
American Concrete Institute Technical Committee Document 304, 2000, 41 pages

ACI Guide 309R-05, Guide for Consolidation of Concrete, American Concrete Institute
Technical Committee Document 309, 2005, 35 pages

Ahmed M. S., Kayali O., and Anderson W., Evaluation of Binary and Ternary Blends of
Pozzolanic Materials Using the Rapid Chloride Permeability Test, Journal of Materials
in Civil Engineering, Vol. 21, No. 9, 2009, pp. 446-453

Al-Khaiat, H., Fatuuhi, N., Carbonation of Concrete Exposed to Hot and Arid Climate,
Journal of Materials in Civil Engineering, Vol. 12, No. 2, 2002, pp. 97-107

Andrade C., Types of Models of Service Life of Reinforcement: The Case of the
Resistivity, Concrete Research Letters, Vol. 1, No. 2, June 2010, pp. 73- 80

ASTM C33-07, Standard Specification for Concrete Aggregates, American Society of


Testing and Materials, Pennsylvania, 2007

ASTM C39-05, Standard Test Method for Compressive Strength of Cylindrical Concrete
Specimens, American Society of Testing and Materials, Philadelphia, PA, 2005

220

221
ASTM C39-05a, Standard Test Method for Compressive Strength of Cylindrical
Concrete Specimens, American Society of Testing and Materials, Pennsylvania, 2005

ASTM C70-01, Standard Test Method for Surface Moisture in Fine Aggregate, American
Society of Testing and Materials, Pennsylvania, 2001

ASTM C127-04, Standard Test Method for Density, Relative Density (Specific Gravity),
and Absorption of Coarse Aggregate, American Society of Testing and Materials,
Pennsylvania, 2004

ASTM C128-04a, Standard Test Method for Density, Relative Density (Specific Gravity),
and Absorption of Fine Aggregate, American Society of Testing and Materials,
Pennsylvania, 2004

ASTM C136-06, Standard Test Method Sieve Analysis of Fine and Coarse Aggregates,
American Society of Testing and Materials, Pennsylvania, 2006

ASTM C143-08, Standard Test Method for Slump of Hydraulic-Cement Concrete,


American Society of Testing and Materials, Pennsylvania, 2008

ASTM C157-08, Standard Test Method for Length Change of Hardened HydraulicCement Mortar and Concrete, American Society of Testing and Materials, Pennsylvania,
2008

ASTM C173-10, Standard Test Method for Air Content of Freshly Mixed Concrete by the
Volumetric Method, American Society of Testing and Materials, Pennsylvania, 2010

ASTM C192-07, Standard Practice for Making and Curing Concrete Test Specimens in
the Laboratory, American Society of Testing and Materials, Pennsylvania, 2007

222
ASTM C231-08b, Standard Test Method Air Content of Freshly Mixed Concrete by the
Pressure Method, American Society of Testing and Materials, Pennsylvania, 2008

ASTM C233-07, Standard Test Method for Air-entraining Admixtures for Concrete,
American Society of Testing and Materials, Pennsylvania, 2007

ASTM C260-06, Standard Specification for Air-Entraining Admixtures for Concrete,


American Society of Testing and Materials, Pennsylvania, 2006

ASTM C470-02, Standard specification for Molds for Forming Concrete Test Cylinders
Vertically, American Society of Testing and Materials, Pennsylvania, 2002

ASTM C494-08, Standard Specification for Chemical Admixtures for Concrete,


American Society of Testing and Materials, Pennsylvania, 2008

ASTM C617-03, Standard Practice for Capping Cylindrical Concrete Specimens,


American Society of Testing and Materials, Pennsylvania, 2003

ASTM C873-04, Standard Test Method for Compressive Strength of Concrete Cylinders
Cast in Place in Cylindrical Molds, American Society of Testing and Materials,
Pennsylvania, 2004

ASTM C1077-07a, Standard Practice for Laboratories Testing Concrete and Concrete
Aggregates for Use in Construction and Criteria for Laboratory Evaluation, American
Society of Testing and Materials, Pennsylvania, 2007

ASTM C1202-07, Standard Test Method for Electrical Indication of Concretes Ability to
Resist Chloride Ion Penetration, American Society of Testing and Materials,
Pennsylvania, 2007

223
ASTM C1240-05, Standard Specification for Silica Fume Used in Cementitious
Mixtures, American Society of Testing and Materials, Pennsylvania, 2005

ASTM C1585-04, Standard Test method for Measurement of Rate of Absorption of Water
by Hydraulic-Cement Concretes, American Society of Testing and Materials,
Pennsylvania, 2004

ASTM C1602-06, Standard Specification for Mixing Water Used in the Production of
Hydraulic Cement Concrete, American Society of Testing and Materials, Pennsylvania,
2006

Bassuoni M. T., Nehdi M. L., and Greenough T. R., Enhancing the Reliability of
Evaluating Chloride Ingress in Concrete Using the ASTM C1202 Rapid Chloride
Penetrability Test, Journal of ASTM International, Vol. 3, No. 3 , 2006, pp. 1-13

Bensted J., Thaumasite - Background and Nature in Deterioration of Cements, Mortars


and Concretes, Cement and Concrete Composites , Vol. 21, No. 2, 1999, pp. 117-121

Bertolini L., Carsana M., Redaelli M., Conservation of Historical Reinforced Concrete
Structures Damaged by Carbonation Induced Corrosion by Means of Electrochemical
Realkalisation, Journal of Cultural Heritage, Vol. 9, No. 4, 2008, pp. 376-385

Bertolini L. and Polder R. B., Concrete Resistivity and Reinforcement Corrosion Rate as
a Function of Temperature and Humidity of the Environment, TNO report 97-BT-R0574,
Netherland, 1997

Bryant J. W., Weyers R. E., and Garza J. M., In-Place Resistivity of Bridge Deck
Concrete Mixtures, ACI Materials Journal, Vol. 106, No. 2, 2009, pp. 114-122

Buehlef M. G. and Thurber W. R., A Planar Four-Probe Structure for Measuring Bulk
Resistivity, IEEE Transactions on Electron Devices, Vol. 23, No. 8, 1976, pp. 968-974

224
Butefuhr M., Fischer C., Gehlen C., Menzel K., and Nurnberger U., On-Site Investigation
on Concrete Resistivity a Parameter of Durability Calculation of Reinforced Concrete
Structures, Materials and Corrosion, Vol. 57, No. 12, 2006, pp. 932-939

Chatterji S. , A Discussion of the Papers, ''A Novel Method for Describing Chloride Ion
Transport due to an Electrical Gradient in Concrete: Part 1 and Part 2'' by K. Stanish,
R.D. Hooton, M.D.A. Thomas, Cement and Concrete Research, Vol. 35, No. 9, 2005, pp.
1865-1867

Chini A. R., Muszynski L. C., and Hicks J., Determination of Acceptance Permeability
Characteristics for Performance-Related Specifications for Portland Cement Concrete,
Final report submitted to FDOT (MASc. Thesis), University of Florida, Department of
Civil Engineering, 2003

Clark L. A., The Management of the Thaumasite Form of Sulphate Attack in the UK,
Magazine of Concrete Research, Vol. 59, No. 7, 2007, pp. 465-467

Collepardi M., Ordinary and Long-Term Durability of Reinforced Concrete Structures, In


V. M. Malhotra (Ed.), Durability of Concrete, ACI Committee, Vol. 192, 2000, pp. 1-18
CSA A3001-08, Cementitious Materials for Use in Concrete, Canadian Standard
Association, Toronto, 2008
DeSouza S. J. and Hooton R. D., Test Methods for the Evaluation of the Durability of
Covercrete (MASc. Thesis), University of Toronto, Department of Civil Engineering,
1996

DeSouza S. J., Hooton R. D., and Bickley J. A., Evaluation of Laboratory Drying
Procedures Relevant to Field Conditions for Concrete Sorptivity Measurements,
American Society of Testing and Materials, Vol. 19, No. 2, 1997, pp. 59-63

225
DeSouza S. J., Hooton R. D., and Bickley J. A., A Field Test for Evaluation High
Performance Concrete Covercrete Quality, Canadian Journal of Civil Engineering, <RC
Research Press, Vol. 25, No. 3, 1998, pp. 551-556

Dhir R. K., Shaaban I. G., Claisee P. A., and Bryars E. A., Preconditioning in Situ
Concrete for Permeation Testing, Part

1: Initial Surface Absorption, Magazine of

Concrete Research, Vol. 45, No. 163, 1993, pp. 113-118

Edvardsen C., Chloride Migration Coefficients from Non-Steady-State Migration


Experiments at Environment-Friendly Green Concrete, www.gronbeton.dk/artikler/
Chloride%20migration%20coefficients.pdf , 2002

El-Dieb A. S., Hooton R. D., and Thomas M. D. A., Electrical Resistivity of Concrete
Measured by Different Methods (Unpublished paper), University of Toronto, Department
of Civil Engineering

Elkey, W. and Sellevold E. J., Electrical Resistivity of Concrete, Published Report, No.
80, <orwegian Road Research Laboratory, Oslo, Norway, 1995 , 36 pp.

Ewins A. J., Resistivity Measurements in Concrete, British Journal of <DT, Vol. 32, No.
3, 1990, pp. 120-126

Feliu S., Andrade C., Gonzalez J. A., and Alonso C., A New Method for In-situ
Measurement of Electrical Resistivity of Reinforced Concrete, Materials and Structures,
Vol. 29, No.6, 1996, pp. 362-365

Ferreira R. M., Jalali S., NDT Measurements for the Prediction of 28-day Compressive
Strength, <DT & E International, Vol. 43, No. 2, 2010, pp. 55-61

Florida DOT FM 5-578, Method of Test for Concrete Resistivity as an Electrical


Indicator of Its Permeability, January 2004

226
Forster S.W., Concrete Durability-Influencing Factors and Testing, Farmington Hills, MI,
Durability of Concrete, ACI Committee, Vol. 191, 2000, pp. 1-10

Gowers K. R. and Millard S. G., Measurement of Concrete Resistivity for Assessment of


Corrosion Severity of Steel Using Wenner Technique, ACI Material Journal, Vol. 96,
No. 5, 1999, pp. 536-541

Hansson I. L. H. and Hansson C. M., Electrical Resistivity Measurements of Portland


Cement Based Materials, Cement and Concrete Research, Vol. 13, No. 5, 1983, pp. 675683

Hooton R.D., Evolution of North American Standard for Sulfate Resistance: an Historic
Perspective and Recent Developments, Proceedings, RIELM Workshop on Performance
of Cement-Based Materials in Aggressive Aqueous Environments, Ghent, Belgium,
September 2007, pp. 75-77

Hooton R.D., Thomas M.D.A., Stanish K., Prediction of Chloride Penetration in


Concrete, Federal Highway Administration, Report No. FHWA-RD-00-142, October
2001

Ishida T. and Li C. H., Modeling of Carbonation Based on Thermo-Hygro Physics with


Strong Coupling of Mass Transport and Equilibrium in Micro-pore Structure of Concrete,
http://www.jsce.or.jp/committee/concrete/e/newsletter/newsletter14/isida.pdf, 2008

Jianyong L. and Pei T., Effect of Slag and Silica Fume on Mechanical Properties of High
Strength Concrete, Cement and Concrete Research, Vol. 27, No. 6, 1997, pp. 833-837

Kosmatka S. H., Kerkhoff B., Panarese W. C., MacLeod N. F., and McGrath R. J.,
Design and Control of Concrete Mixtures, Seventh Canadian Eition, Cement Association
of Canada, 2002

227
Kessler R. J., Power R. G., and Paredes M. A., Resistivity Measurements of Water
Saturated Concrete as an Indicator of Permeability, Corrosion 2005, April 3-7, 2005,
Houston, TX, pp. 1-10

Kessler R. J., Power R. G., Vivas E., Paredes M. A., and Virmani Y.P., Surface
Resistivity

as

an

Indicator

of

Concrete

Chloride

Penetration

Resistance,

http://concreteresistivity.com/Surface%20Resistivity.pdf, 2008

Kosmatka S. H., Kerkhoff B., and Panarese W. C., Design and Control of Concrete
Mixtures, Portland Cement Association, Skokie, Illinois, 2002

Lataste J. F., Sirieix C., Breysse D., and Frappa M., Electrical resistivity measurement
applied to cracking assessment on reinforced concrete structures in civil engineering,
<DT & E International, Vol. 36, No. 6, 2003, pp. 383-394

Lopez W. and Gonzalez J. A., Influence of the Degree of Pore Saturation on the
Resistivity of Concrete and the Corrosion Rate of Steel Reinforcement, Cement and
Concrete Research, Vol. 23, No. 2, 1993, pp. 368-376

McCarter W. J., Starrs G., Kandasami S., Jones R., and Chrisp M., Electrode
Configuration for Resistivity Measurements on Concrete, ACI Materials Journal, Vol.
106, No. 3, 2009, pp. 258-264

Millard S. G., Harrison J. A., and Edwards A. J., Measurements of the Electrical
Resistivity of Reinforced Concrete Structures for the Assessment of Corrosion Risk,
British Journal of <DT, Vol. 13, No. 11, 1989, pp. 617-621

Millard S. G. and Gowers K. R., The Influence of Surface Layers upon the Measurement
of Concrete Resistivity, Durability of Concrete, Second International Conference, ACI
SP-126, Montreal, Canada, 1991, pp. 1197-1220

228
Monfore G. E., The Electrical Resistivity of Concrete, Journal of the PCA Research
Development Laboratories, Vol. 10, No. 2, 1968, pp. 35-48

Monkman S. and Shao Y., Assessing the Carbonation Behaviour of Cementitious


Materials, Journal of Materials in Civil Engineering, Vol. 18, No. 6, 2006, pp. 768-776

Morris W., Moreno E. I., and Sagues A. A., Practical Evaluation of Resistivity of
Concrete in Test Cylinders Using A Wenner Array Probe, Cement and Concrete
Research, Vol. 26, No. 12, 1996, pp. 1779-1787

Neville A. M., Properties of Concrete, Fourth Edition, Pearson Education Ltd., 1995

Newlands M. D., Jones M. R., Kandasami S., and Harrison T. A., Sensitivity of
Electrodes Contact Solutions and Contact Pressure in Assessing Electrical Resistivity of
Concrete, Materials and Structures, Vol. 41, No.4, 2008, pp. 621-632

Nokken M. R. and Hooton R. D., Dependence of Rate of Absorption on Degree of


Saturation of Concrete, Journal of Cement, Concrete, and Aggregates (CCA), Vol. 24,
No. 1, 2002, pp. 20-24

Nokken M. R. and Hooton R. D., Electrical Conductivity as a Prequalification and


Quality Control, Concrete International, Vol. 28, No. 10, 2006, pp. 61-66

Parrott L. J., Moisture Conditioning and Transport Properties of Concrete Test


Specimens, Materials and Structures, Vol. 27, No. 8, 1994, pp. 460-468

Polder R. B., Test Methods for on Site Measurement of Resistivity of Concrete - a


RILEM TC-154 Technical Recommendation, Construction and Building Materials, Vol.
15, No. 2-3, 2001, pp. 125-131

229
Pun P., Kojuncdic T., Hooton R.D., Kojundic T., and Fidjestol P., Influence of Silica
Fume on Chloride Resistance of Concrete, Proceedings of PCI/FHWA International
Symposium on High Performance Concrete, New Orleans, Louisiana, October 1997, pp.
245256.

RILEM Technical Committee, Update of the Recommendation of RILEM TC 189-NEC


Non-destructive Evaluation of the Concrete Cover (Comparative Test Part I, Comparative
Test of Penetrability Methods), Materials & Structures, Vol. 38, No. 284, 2005, pp. 895
906

Saha T. K. and Purkait P., Investigation of Polarization and Depolarization Current


Measurements for the Assessment of Oil-paper Insulation of Aged Transformers,
Dielectrics and Electrical Insulation, IEEE Transactions , Vol. 11, No. 1, 2004, pp. 144154

Savas B. Z., Effect of Microstructure on Durability of Concrete (PhD Thesis), North


Carolina State University, Department of Civil Engineering, Raleigh NC, 1999

Sengul O. and Gjorv O. E., Electrical Resistivity Measurements for Quality Control
During Concrete Construction, ACI Materials Journal, Vol. 105, No. 6, 2008, pp. 541547

Sengul O. and Gjorv O. E., Effect of Embedded steel on Electrical Resistivity


Measurements on Concrete Structures, ACI Materials Journal, Vol. 106, No. 1, 2009,
pp. 11-18

Scrivener K. L., Crumbie A. K., and Laugesen P., The Interfacial Transition Zone (ITZ)
Between Cement Paste and Aggregate in Concrete, Interface Science, Vol. 12, No. 4,
2004, pp. 411- 421

230
Shehata M.H., Thomas M.D.A., and Bleszynski R.F., The Effects of Fly ash Composition
on the Chemistry of Pore Solution in Hydrated Cement Pastes, Cement and Concrete
Research, Vol.29, No. 12, 1999, pp. 1915- 1920

Shi C., Effect of Mixing Proportions of Concrete on its Electrical Conductivity and the
Rapid Chloride Permeability Test (ASTM C1202 or ASSHTO T277) Results, Cement
and Concrete Research, Vol. 34, No. 3, 2004, pp. 537-545

Smith K. M., Schokker A. J., and Tikalsky P. J., Performance of Supplementary


Cementitious Materials in Concrete Resistivity and Corrosion Monitoring Evaluations,
ACI Materials Journal, Vol. 101, No. 5, 2004, pp. 385-390

Stanish K., Hooton R. D., and Thomas M. D. A., Testing the Chloride Penetration
Resistance of Concrete: A Literature Review, Department of Civil Engineering
University of Toronto, Toronto, Ontario, Canada, FHWA Contract DTFH61-97-R 00022
Prediction of Chloride Penetration in Concrete, 1997

Stanish K., Hooton R. D., and Thomas M. D. A., A Novel Method for Describing
Chloride Ion Transport due to an Electrical Gradient in Concrete: Part 1. Theoretical
description, Cement and Concrete Research, Vol. 34, No. 1, 2004, pp. 43-49

Stanish K., Hooton R. D., and Thomas M. D. A., A Novel Method for Describing
Chloride Ion Transport due to an Electrical Gradient in Concrete: Part 2. Experimental
study, Cement and Concrete Research, Vol. 34, No.1, 2004, pp. 51-57

Thaulow N. and Sahu S., Mechanism of Concrete Deterioration due to Salt


Crystallization, Materials Characterization, Vol. 53, No. 2-4, 2004, pp. 123-127

Tsui N., Flatt R. J., and Scherer G. W., Crystallization Damage by Sodium Sulfate,
Journal of Cultural Heritage, Vol. 4, No. 2, 2003, pp. 109-115

231
Viera M., de Almeida I. R., and Gonclaves A. F., Influence of Moisture Curing on
Durability of Fly Ash Concrete for Road Pavements. In Marlhotra V.M. (Ed.), Durability
of Concrete, American Concrete Institute, Farmington Hills, MI, 2000, pp. 91-102

APPE/DIX A
CO/CRETES MIX DESIG/

HPC (SFSL-0.35)
Silica fume cement Slag 25 %

410
25%
0%
0%
0.35
80
2.41%
2.72%

W/C=0.35

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Silica Fume Cement (GUb-8SF)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3080
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
307.5
102.5
0.0
0.0
714.3
1075.8
143.5
2343.6

232

Volume
3
m
0.100
0.036
0.000
0.000
0.263
0.398
0.144
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
307.5
102.5
0.0
0.0
723.3
1090.4
119.8
2343.6

Batch
Mass (kg)
24.6
8.2
0.0
0.0
57.9
87.2
9.6

233

HPC+ (SFSL-0.35+)
Silica fume cement Slag 25 %

410
25%
0%
0%
0.35
82
2.44%
2.40%

W/C=0.35

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Silica Fume Cement (GUb-8SF)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3080
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
307.5
102.5
0.0
0.0
714.3
1075.8
143.5
2343.6

Volume
3
m
0.100
0.036
0.000
0.000
0.263
0.398
0.144
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
307.5
102.5
0.0
0.0
723.5
1086.9
123.2
2343.6

Batch
Mass (kg)
25.2
8.4
0.0
0.0
59.3
89.1
10.1

234

SFSL-0.40
Silica fume cement Slag 25 %

375
25%
0%
0%
0.4
80
2.71%
3.01%

W/C=0.40

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Silica Fume Cement (GUb-8SF)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3080
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
281.3
93.8
0.0
0.0
728.1
1075.8
150.0
2328.9

Volume
3
m
0.091
0.033
0.000
0.000
0.268
0.398
0.150
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
281.3
93.8
0.0
0.0
739.6
1093.6
120.7
2328.9

Batch
Mass (kg)
22.5
7.5
0.0
0.0
59.2
87.5
9.7

235

PCSL-0.40
PC10 Slag 25 %

375
25%
0%
0%
0.4
82
2.13%
2.54%

W/C=0.40

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Portland Cement (GU)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density kg/m
3134
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
281.3
93.8
0.0
0.0
732.4
1075.8
150.0
2333.2

Volume
3
m
0.090
0.033
0.000
0.000
0.269
0.398
0.150
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
281.3
93.8
0.0
0.0
739.6
1088.4
130.2
2333.2

Batch
Mass (kg)
23.0
7.7
0.0
0.0
60.6
89.3
10.7

236

PCSL-0.40+
PC10 Slag 25 %

375
25%
0%
0%
0.4
82
2.98%
3.01%

W/C=0.40

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Portland Cement (GU)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3134
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
281.3
93.8
0.0
0.0
732.4
1075.8
150.0
2333.2

Volume
3
m
0.090
0.033
0.000
0.000
0.269
0.398
0.150
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
281.3
93.8
0.0
0.0
746.0
1093.6
118.6
2333.2

Batch
Mass (kg)
23.1
7.7
0.0
0.0
61.2
89.7
9.7

237

PCSL-0.45
PC10 Slag 25 %

355
25%
0%
0%
0.45
82
2.23%
2.07%

W/C=0.45

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Portland Cement (GU)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3134
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
266.3
88.8
0.0
0.0
723.7
1075.8
159.8
2314.2

Volume
3
m
0.085
0.031
0.000
0.000
0.266
0.398
0.160
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
266.3
88.8
0.0
0.0
731.5
1083.3
144.4
2314.2

Batch
Mass (kg)
21.8
7.3
0.0
0.0
60
88.8
11.8

238

PCSL-0.45+
PC10 Slag 25 %

355
25%
0%
0%
0.45
82
2.80%
2.02%

W/C=0.45

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Portland Cement (GU)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3134
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
266.3
88.8
0.0
0.0
723.7
1075.8
159.8
2314.2

Volume
3
m
0.085
0.031
0.000
0.000
0.266
0.398
0.160
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
266.3
88.8
0.0
0.0
735.7
1082.7
140.8
2314.2

Batch
Mass (kg)
21.8
7.3
0.0
0.0
60.3
88.8
11.5

239

PC-0.45
PC 10 Slag 0 %

355
0%
0%
0%
0.45
82
2.25%
1.90%

W/C=0.45

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Portland Cement (GU)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3134
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
355.0
0.0
0.0
0.0
731.2
1075.8
159.8
2321.8

Volume
3
m
0.113
0.000
0.000
0.000
0.269
0.398
0.160
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
355.0
0.0
0.0
0.0
739.3
1081.4
146.1
2321.8

Batch
Mass (kg)
29.1
0.0
0.0
0.0
60.6
88.7
12.0

240

PC-0.45+
PC 10 Slag 0 %

355
0%
0%
0%
0.45
82
1.81%
3.24%

W/C=0.45

Cementitious content, kg/m


% Slag
% Silica Fume
% FA
Water-cement ratio
Batch Volume, Litres
Sand moisture content
Coarse Agg. Moisture content

Portland Cement (GU)


Slag
Silica Fume
FA
Sand
Coarse Agg.
Water
Air

1.16%
1.38%

Density
3
kg/m
3134
2854
2300
2600
2720
2702
1000

Design Mass
3
per m (kg)
355.0
0.0
0.0
0.0
731.2
1075.8
159.8
2321.8

Volume
3
m
0.113
0.000
0.000
0.000
0.269
0.398
0.160
0.060
1.000

= Abs. Of sand
= Abs. Of Coarse Agg.

Adjusted Mass
3
per m (kg)
355.0
0.0
0.0
0.0
736.0
1096.2
134.6
2321.8

Batch
Mass (kg)
29.1
0.0
0.0
0.0
60.4
89.9
11.0

APPE/DIX B
COMPRESSIVE STRE/GTH TEST
Compressive strength of concrete cylinders is the most common test of hardened concrete
used as a measure of concrete quality, the most important engineering property.
According to ASTM C39-05 a concrete cylinder is subjected to a compressive axial load
until it becomes crashed. When the limit of compressive strength is reached, the concrete
cylinder is crushed and the strength (ultimate taken load divided by the section area) is
calculated.

Figure B.1: Crushed cylinder in compressive strength test

The results of this test are used as a basis for quality control of concrete proportioning,
mixing, curing, and placing operation (ASTM C39, 2005).

241

242

B.1 Influencing factors on concrete strength


The compressive strength is inversely proportional to concrete porosity, so any factor
reducing concrete porosity increases compressive strength (Neville, 1995). Three major
factors influence the mechanical strength as concrete is considered fully-compacted:
W/CM ratio, SCMs effects, and cement hydration.

B.1.1 W/CM ratio and concrete strength


When concrete is fully compacted, its strength is taken to be inversely proportional to the
W/CM ratio as shown in the following relation (Abrams, 1919 in Neville, 1995):
fc =

1
W / CM
2

where K1 and K2 are empirical constants.


Since more water results in more porosity, higher W/CM ratio results in lower
compressive strength and higher permeability (MacDonald and Northwood, 2000 in
Chini et al., 2003) as shown in Figure B.2.

Figure B.2: Relation between logarithm of strength and W/CM ratio (/eville, 1995)

243
It can be concluded from previous researches that in practice, the W/CM ratio is the
largest single factor in the strength of fully compacted concretes (Neville, 1995). In
addition, concretes with lower W/CM ratio express a long-term strength more rapidly
than higher W/CM ratio concretes because of the rapid establishment of the
discontinuous system of gel (Neville, 1995).

B.1.2 SCMs and concrete strength


SCMs reduce concrete porosity and discontinue capillary pore structure during their
secondary hydration (Chini, 2003). In other words, SCMs improve concrete properties in
two ways: changing the weak CH into the strong C-S-H during secondary hydration and
also reducing concrete porosity by its micro-filler effect (e.g. silica fume). Since silica
fume, one of the cementitious materials used in this research program, is extremely fine
(approximately 1/100th the size of an average cement particle), it fills the microscopic
voids between cement particles results in better paste-to-aggregate bond (Kosmatka et
al.,2002). More importantly silica fume reduces ITZ porosity (Neville, 1995).
Therefore, mixes containing SCMs have higher compressive strength (also split tensile
strength and rupture strength) and durability than plain cement concrete mixes (Jianyong
and Pei, 1997).

B.1.3 Cement hydration and concrete strength


Concrete mechanical strength increases as concrete hydrates, due to less total porosity
which is filled by cement hydration product as shown in Figure B.3.

244

Figure B.3: Compressive strength and age of concrete


(http://www.theconstructioncivil.com/2009/09/concrete-curing.html)

In addition, it can be concluded that curing scenario, affecting cement hydration,


influences the rate of strength gain in concrete. Any type of curing accelerating cement
hydration procedure, improves concrete compressive strength (e.g. compressive strength
of moist cured concrete is significantly higher than air cured concrete).

B.2 Methodology
The following steps are needed to be taken according to ASTM C39-05 during the
measurement of compressive strength. It is worth noting that a compression cylinder has
to be kept moist until testing (ASTM C39, 2005).

Step I) Grinding concrete cylinders: Since cylinders ends are in contact with the testing
machine platens, both ends must be flat by either grinding or capping (up to 70 MPa).
Any inclined end results in a compressive load which is not axial. The ends shall not be
out of plane by more than 0.05 mm. Neither end of the specimen shall be depart from
perpendicularity to the axis by more than 0.5 which is 1 mm in 100 mm diameter.

245

Step II) Cross-section area: Cross-section area of a cylinder is

D2
, where, D is the
4

average section diameter of the specimen (mm). The diameter of a cylinder is an average
of four measurements at right angles on both ends of the cylinder (two measurements on
each end). The diameter of cylinders is to be determined to the nearest 0.25 mm. The
length shall be measured to the nearest millimetre and shall have length-to-diameter
(L/D) ratio between 1.8 to 2.2. As ASTM C39-05 mentions, an individual diameter shall
not be different from any other diameter of the same cylinder by more than 2%.

Step III) Concrete cylinder is placed between platen plates of the testing machine.
Platen faces have a minimum dimension at least 3% greater than the diameter of the
specimen to be tested (ASTM C39, 2005). The rate of loading shall be applied at a
constant rate within the range of 0.15 to 0.35 MPa/S (Neville, 1995). Therefore, for the
research program the load is applied at a rate of movement corresponding to a loading
rate on the specimen of 2.0 0.1 KN/S. The load is applied until the specimen fails and
the peak load is recorded. Compressive strength is calculated by dividing the maximum
load by section area to the nearest 0.1 MPa and corrected for L/D ratio as shown in Table
B.1.
Table B.1: L/D correction factors for compressive strength test (ASTM C 39, 2005)

L/D

2.0

1.75

1.50

1.25

1.0

Factor

1.0

0.98

0.96

0.93

0.87

The type of failure and appearance of the fracture surface have to be one of the standard
types of fracture. Standard types of fracture for concrete cylinders and cubes are shown in
Figure B.4.

246

Figure B.4: Types of concrete specimen fracture (ASTM C39, 2003)

The type of failure and appearance of the fracture surface is recorded if different from
type (a). The cone failure results when the friction between the specimen and the platens
restrains the lateral force applied to the specimen by its horizontal expansion during
compressive strength test. Compressive strength of unusual types of fracture is not
acceptable.

247

B.3 Compressive strength test results

Day 3 (MPa)

Day 7 (MPa)

Mixture
CYL.1

CYL.2

CYL.3

HPC

44.92

46.99

45.11

HPC+

43.86

44.87

SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45

39.73

fc 3day

fc 7day

CYL.1

CYL.2

CYL.3

45.67

57.64

57.68

56.9

57.41

44.97

44.57

54.76

53.18

54.13

54.02

43.55

45.23

42.84

53.49

52.11

53.3

52.97

37.26
32.11

35.02
33.33

35.1
32.01

35.79
32.48

39.11
34.51

40.29
35.2

38.13
34.9

39.18
34.87

29.81

29.29

31.69

30.26

32.01

31.64

30.96

31.54

PCSL-0.45+

27.18

28.02

25.81

27.00

29.73

29.26

30.52

29.84

PC-0.45

32.44

32.19

30

31.54

35

33.8

32.8

33.87

PC-0.45+

28.74

32.5

30.1

30.45

32.1

30.19

33.52

31.94

Day 28 (MPa)

Day 56 (MPa)

Mixture
CYL.1

CYL.2

CYL.3

HPC

67.28

68.24

70.23

HPC+

63.88

59.2

SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45

60.23

fc 28day

fc 56day

CYL.1

CYL.2

CYL.3

68.58

73.11

75.54

72.54

73.73

62.54

61.87

65.81

65.25

66.1

65.72

59.64

59.65

59.84

62.92

62.24

60.52

61.89

48.14
44.25

50.31
43.12

49.57
45.91

49.34
44.43

51.98
48.11

52.81
47.45

51.28
46.39

52.02
47.32

39.74

41.56

39.55

40.28

43.28

44.9

44.48

44.22

PCSL-0.45+

36.68

37.95

39.6

38.08

41.71

42.36

41.11

41.73

PC-0.45

38.02

38.5

35.32

37.28

37.01

41.84

38.69

39.18

PC-0.45+

33.45

33.35

35.53

34.11

36.08

38.68

37.52

37.43

248

Day 91 (MPa)
Mixture

fc 91day

CYL.1

CYL.2

CYL.3

HPC

78.58

74.55

76.25

76.46

HPC+

70.02

67.53

67.12

68.22

SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45

62.58

64.58

62.68

63.28

48.48
47.92

58
50.83

59.99
49.38

55.49
49.38

46.85

45.81

44.8

45.82

PCSL-0.45+

45.8

43.43

44

44.41

PC-0.45

41.1

46.48

41.81

43.13

PC-0.45+

40.03

41.33

39.04

40.13

APPE/DIX C
LABORATORY SORPTIVITY TEST RESULTS
Sorptivity (x10-4 mm/sec)
Mixture
Initial

Day 28
Secondary

Initial

Day 56
Secondary

Initial

Day 91
Secondary

BOTTOM
HPC
HPC+
SFSL - 0.40
PCSL - 0.40
PCSL - 0.40+
PCSL - 0.45
PCSL - 0.45+
PC - 0.45
PC - 0.45+

16.3
17.4
19.8
22.7
25.0
28.5
31.0
36.0
39.4

3.9
4.3
4.8
5.3
5.8
7.2
8.4
11.7
16.5

HPC
HPC+
SFSL - 0.40
PCSL - 0.40
PCSL - 0.40+
PCSL - 0.45
PCSL - 0.45+
PC - 0.45
PC - 0.45+

16.0
19.5
21.7
25.5
27.3
32.9
36.2
40.3
44.4

3.1
3.9
4.5
6.8
7.3
8.6
9.6
14.0
16.9

10.7
12.6
13.5
16.0
21.5
24.7
25.9
27.0
32.1

2.8
3.2
3.2
4.5
4.7
5.0
5.4
6.8
11.1

6.9
7.8
9.5
12.2
13.8
16.9
21.2
26.3
30.5

1.1
1.8
2.3
2.8
3.5
4.1
5.0
5.7
8.2

2.4
3.2
3.5
5.8
6.3
5.4
7.7
9.5
13.3

12.1
14.0
16.0
18.5
20.9
22.1
26.3
34.1
38.0

1.0
2.2
2.8
3.8
4.1
4.8
5.8
8.2
11.8

2.5
3.2
4.1
5.0
5.8
6.4
6.8
7.7
11.6

14.6
16.0
18.8
17.9
22.7
23.7
28.2
34.3
38.9

1.5
2.1
3.3
4.4
4.9
5.2
5.8
6.6
10.3

2.0
3.0
3.9
5.3
5.6
5.9
6.6
9.0
11.3

22.1
25.9
25.9
32.7
37.8
39.7
44.9
53.8
59.1

1.6
2.3
2.7
4.0
4.4
5.0
5.7
8.0
10.5

MIDDLE
13.4
16.0
16.7
21.0
24.2
25.8
33.8
35.1
39.6

TOP
HPC
HPC+
SFSL - 0.40
PCSL - 0.40
PCSL - 0.40+
PCSL - 0.45
PCSL - 0.45+
PC - 0.45
PC - 0.45+

18.6
21.5
24.0
27.8
30.3
33.3
37.6
42.2
46.0

3.2
3.6
5.4
6.1
7.4
8.2
8.5
11.9
15.0

HPC
HPC+
SFSL - 0.40
PCSL - 0.40
PCSL - 0.40+
PCSL - 0.45
PCSL - 0.45+
PC - 0.45
PC - 0.45+

30.1
34.1
36.3
41.6
44.6
49.7
52.8
61.8
76.8

3.6
4.4
5.5
6.0
6.8
8.0
8.9
13.0
14.6

15.2
19.5
21.0
23.3
26.5
28.6
32.1
35.9
40.2

CORED
25.3
28.6
29.9
35.8
40.9
42.3
50.8
56.3
63.1

249

APPE/DIX D
FIELD SORPTIVITY TEST RESULTS

Sorptivity

MIXTURE

CS 1

CS 2

(mm/min0.5)

AGE

HPC
HPC +
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

0.009
0.010
0.010
0.012
0.015
0.028
0.030
0.006
0.025

0.006
0.008
0.009
0.013
0.016
0.029
0.029
0.016
0.030

0.008
0.009
0.009
0.013
0.016
0.028
0.029
0.011
0.028

14
14
14
14
14
14
14
14
14

HPC
HPC +
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

0.009
0.015
0.015
0.017
0.032
0.042
0.054
0.017
0.025

0.012
0.008
0.012
0.019
0.031
0.053
0.060
0.018
0.037

0.010
0.011
0.013
0.018
0.031
0.047
0.057
0.018
0.031

28
28
28
28
28
28
28
28
28

HPC
HPC +
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

0.020
0.022
0.027
0.031
0.047
0.060
0.074
0.031
0.038

0.013
0.019
0.023
0.032
0.052
0.057
0.063
0.030
0.046

0.017
0.021
0.025
0.031
0.050
0.058
0.068
0.030
0.042

56
56
56
56
56
56
56
56
56

HPC
HPC +
SFSL 0.40
PCSL 0.40
PCSL 0.40+
PCSL 0.45
PCSL 0.45+
PC 0.45
PC 0.45+

0.022
0.028
0.030
0.049
0.071
0.089
0.115
0.044
0.057

0.019
0.024
0.026
0.050
0.072
0.100
0.116
0.048
0.078

0.020
0.026
0.028
0.049
0.072
0.094
0.115
0.046
0.067

91
91
91
91
91
91
91
91
91

250

APPE/DIX E
DC-CYCLIC BULK ELECTRICAL RESISTIVITY

E.1 Full length cylinders (100 x 200 mm)


Day 3
Mixture

Day 7

CYL.1

CYL.2

CYL.3

HPC

29.47

9.52

13.63

HPC+

28.20

9.29

SFSL-0.40

8.15

PCSL-0.40

3day (K.cm)

7day (K.cm)

CYL.1

CYL.2

CYL.3

17.54

27.97

29.58

21.29

26.28

13.28

16.92

27.00

32.01

15.40

24.80

4.69

14.38

9.07

28.07

9.26

24.23

20.52

8.36

6.32

4.29

6.32

5.82

9.36

14.34

9.84

PCSL-0.40+

3.37

3.65

4.26

3.76

8.29

4.49

4.23

5.67

PCSL-0.45

3.42

3.60

3.50

3.51

3.39

3.57

4.37

3.78

PCSL-0.45+

3.36

3.53

4.27

3.72

3.40

3.54

4.30

3.75

PC-0.45

3.35

3.45

3.30

3.37

5.00

4.85

4.94

4.93

PC-0.45+

3.36

3.68

3.56

3.53

4.81

4.65

4.25

4.57

Day 28
Mixture

Day 56

CYL.1

CYL.2

CYL.3

HPC

83.08

81.12

82.88

HPC+

82.97

76.78

SFSL-0.40

71.12

PCSL-0.40

28day (K.cm)

56day (K.cm)

CYL.1

CYL.2

CYL.3

82.36

92.43

94.52

96.66

94.54

77.74

79.16

85.89

86.34

91.92

88.05

72.80

73.63

72.52

77.30

78.11

79.29

78.23

23.48

23.81

22.66

23.32

29.00

28.91

30.38

29.43

PCSL-0.40+

17.51

17.36

19.36

18.08

24.81

25.91

24.18

24.97

PCSL-0.45

15.51

16.26

15.39

15.72

16.91

30.16

19.23

22.10

PCSL-0.45+

14.86

15.18

15.27

15.10

19.32

19.72

19.98

19.67

PC-0.45

7.15

8.15

5.12

6.81

9.12

10.68

9.50

9.77

PC-0.45+

5.90

4.50

14.07

8.16

8.34

9.42

8.15

8.64

251

252

Day 91
Mixture

91day (K.cm)

CYL.1

CYL.2

CYL.3

HPC

105.21

102.00

100.00

102.40

HPC+

90.56

97.51

90.00

92.69

SFSL-0.40

85.32

80.21

81.25

82.26

PCSL-0.40

40.88

43.80

41.16

41.95

PCSL-0.40+

36.36

30.08

31.49

32.64

PCSL-0.45

28.72

31.08

29.54

29.78

PCSL-0.45+

26.44

28.90

30.90

28.75

PC-0.45

11.77

13.55

12.61

12.64

PC-0.45+

10.69

9.90

10.59

10.39

253

E.2 Concrete cores (100 x 50 mm)

Day 3
Mixture

Middle Bottom

Day 7

3day (K.cm)

Middle Bottom

7day (K.cm)

HPC

4.39

5.42

4.91

20.13

25.10

22.62

HPC+

4.62

4.43

4.53

23.73

25.84

24.79

SFSL-0.40

4.93

3.97

4.45

24.33

24.23

24.28

PCSL-0.40

4.95

3.11

4.03

5.90

7.32

6.61

PCSL-0.40+

4.09

4.56

4.33

4.66

6.00

5.33

PCSL-0.45

4.57

3.16

3.87

4.12

5.13

4.63

PCSL-0.45+

4.00

3.10

3.55

4.16

5.08

4.62

PC-0.45

5.18

3.20

4.19

5.10

3.20

4.15

PC-0.45+

3.12

5.02

4.07

3.59

4.00

3.79

Day 28
Mixture

Middle Bottom

Day 56

28day (K.cm)

Middle Bottom

56day (K.cm)

HPC

80.35

74.34

77.35

91.38

91.26

91.32

HPC+

73.31

76.40

74.86

90.73

88.52

89.63

SFSL-0.40

64.97

63.67

64.32

82.71

73.93

78.32

PCSL-0.40

15.18

19.36

17.27

29.85

30.28

30.07

PCSL-0.40+

13.69

16.90

15.30

23.58

25.17

24.38

PCSL-0.45

12.98

13.64

13.31

20.92

22.34

21.63

PCSL-0.45+

11.53

11.62

11.57

22.34

22.42

22.38

PC-0.45

6.97

6.13

6.55

10.02

11.81

10.92

PC-0.45+

5.90

6.92

6.41

9.84

9.64

9.74

254

Day 91
Mixture

Middle Bottom

91day (K.cm)

HPC

110.00

89.20

99.60

HPC+

99.00

94.58

96.79

SFSL-0.40

87.00

81.99

84.50

PCSL-0.40

41.83

43.28

42.56

PCSL-0.40+

30.84

33.31

32.08

PCSL-0.45

28.89

29.34

29.12

PCSL-0.45+

26.88

27.86

27.37

PC-0.45

12.45

13.49

12.97

PC-0.45+

11.58

11.10

11.34

255

APPE/DIX F
SURFACE ELECTRICAL RESISTIVITY

F.1 Cylinders (100 x 200 mm)

Day 3

Day 7

Wet Cylinders

Wet Cylinders

Mixture
a =25 mm

a =30 mm

a =40 mm

a =50 mm

a =25 mm

a =30 mm

a =40 mm

a =50 mm

HPC

15.57

17.13

24.54

28.17

36.13

42.88

54.18

71.53

HPC+

12.29

13.50

17.65

21.98

36.98

40.70

51.81

66.69

SFSL-0.40

17.81

19.75

24.62

32.50

26.00

37.29

46.63

60.62

PCSL-0.40

8.37

8.98

12.35

16.11

17.41

17.53

22.58

29.57

PCSL-0.40+

8.20

9.23

11.58

15.33

10.56

11.53

15.18

20.13

PCSL-0.45

6.00

6.63

8.27

11.77

8.37

10.03

12.73

16.83

PCSL-0.45+

4.70

5.00

6.63

8.97

8.22

9.00

11.28

15.23

PC-0.45

4.00

4.35

5.60

7.20

5.00

5.77

6.87

9.40

PC-0.45+

4.13

4.50

5.93

7.67

4.80

5.33

6.67

8.63

Mixture

Day 28

Day 56

Wet Cylinders

Wet Cylinders

a =25 mm

a =30 mm

a =40 mm

a =50 mm

a =25 mm

a =30 mm

a =40 mm

a =50 mm

HPC

106.73

125.72

163.67

207.28

112.75

133.83

167.75

211.91

HPC+

105.71

121.73

153.13

198.75

109.33

125.67

156.00

200.67

SFSL-0.40

88.33

108.51

137.82

180.41

97.59

113.29

146.65

189.27

PCSL-0.40

30.96

37.44

46.98

61.37

33.90

41.27

51.72

69.22

PCSL-0.40+

21.15

23.68

29.53

39.29

27.80

35.01

43.92

57.28

PCSL-0.45

21.03

23.43

28.79

39.25

21.60

24.53

32.45

47.43

PCSL-0.45+

19.30

21.12

27.44

36.60

27.09

29.30

35.93

46.32

PC-0.45

9.90

10.46

14.27

19.15

12.33

13.41

16.96

22.78

PC-0.45+

7.78

8.72

10.48

13.20

10.21

11.46

14.59

19.68

255

256

Day 91
Mixture
Wet Cylinders
a =25 mm

a =30 mm

a =40 mm

a =50 mm

HPC

121.69

144.38

174.69

220.58

HPC+

114.65

132.32

160.25

208.26

SFSL-0.40

109.56

122.02

156.35

196.25

PCSL-0.40

48.28

57.25

65.25

88.58

PCSL-0.40+

40.10

45.68

58.11

79.70

PCSL-0.45

40.10

45.03

56.70

74.17

PCSL-0.45+

30.33

35.41

45.77

60.88

PC-0.45

15.30

17.40

21.90

28.37

PC-0.45+

11.69

13.34

17.08

23.22

257

F.2 Circular slabs (406 x 75 mm)

Mixture

Day 3

Day 7

Slab #1

a =20 mm

HPC
HPC+
SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45
PCSL-0.45+
PC-0.45
PC-0.45+

7.11

9.03

9.93

5.80

7.18

10.11

a =30 mm a =40 mm a =50 mm

a =20 mm

a =30 mm

a =40 mm

a =50 mm

11.45

22.25

24.21

26.41

29.63

7.88

8.79

17.65

21.61

23.24

27.03

12.65

14.28

15.38

16.63

19.54

22.43

26.06

3.44

4.55

5.25

6.13

8.49

10.58

11.33

13.18

4.25

5.24

5.48

6.46

6.19

8.06

9.14

10.76

2.98

3.96

4.48

5.13

4.33

5.94

6.93

7.95

2.21

2.86

3.25

3.76

4.49

5.50

6.18

7.13

2.78

3.28

3.66

4.19

3.54

3.95

4.66

5.15

2.34

2.79

2.99

3.84

3.13

3.54

4.08

4.70

a =20mm

a =30 mm

a =40 mm

a =50 mm

Slab #2

a =20 mm

HPC
HPC+
SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45
PCSL-0.45+
PC-0.45
PC-0.45+

6.80

8.78

9.70

11.11

21.30

23.84

26.94

29.59

5.29

6.60

6.73

8.20

16.36

20.53

22.49

25.58

9.83

12.79

13.93

14.94

15.93

19.44

21.16

24.33

3.36

4.40

5.01

5.83

8.01

10.01

10.70

11.89

3.90

4.68

4.93

5.69

5.63

7.69

8.95

10.36

3.06

3.84

4.36

5.00

4.63

5.99

6.88

7.91

2.25

3.06

3.33

3.91

4.45

5.46

6.23

7.38

2.85

3.13

3.69

4.28

3.53

4.05

4.29

4.83

2.31

2.75

2.76

3.46

3.35

3.44

3.85

4.41

a =30 mm a =40 mm a =50 mm

258

Mixture

Slab #1
HPC
HPC+
SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45
PCSL-0.45+
PC-0.45
PC-0.45+

Slab #2
HPC
HPC+
SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45
PCSL-0.45+
PC-0.45
PC-0.45+

Day 14

Day 28

WET

WET

a =20 mm a =30 mm a =40 mm a =50 mm a =20 mm a =30 mm a =40 mm a =50 mm


71.15

84.91

87.75

97.86

120.05

154.10

156.80

170.05

66.88

78.75

81.06

90.09

108.40

141.60

134.93

153.23

50.21

58.36

61.14

70.18

99.50

126.88

119.70

138.55

16.73

20.61

22.54

26.41

28.03

31.13

32.15

34.23

16.26

18.85

19.59

23.25

23.88

29.30

31.40

33.35

13.56

18.50

17.93

20.84

18.85

23.48

25.73

30.43

13.03

15.01

16.74

17.89

14.40

20.33

21.25

23.45

6.86

7.64

8.23

8.74

9.60

12.00

12.15

13.03

6.23

7.79

7.81

8.54

8.98

10.78

9.43

10.15

a =20 mm a =30 mm a =40 mm a =50 mm a =20 mm a =30 mm a =40 mm a =50 mm


71.18

81.65

88.20

98.28

99.53

152.08

145.53

163.00

65.85

76.16

79.88

89.61

90.75

129.38

136.80

162.20

51.53

58.55

61.94

71.08

80.73

114.55

121.15

129.85

17.64

20.71

22.38

25.51

28.53

32.28

35.70

37.45

15.93

17.44

18.81

22.03

21.60

26.98

29.15

32.58

13.64

16.41

17.46

19.36

16.53

21.25

23.80

26.78

10.09

14.01

14.74

17.36

14.60

20.00

22.78

25.73

6.53

7.79

8.16

8.59

9.40

10.38

10.33

11.70

6.50

8.20

7.80

8.33

8.30

9.60

10.28

11.15

259

Mixture

Slab #1
HPC
HPC+
SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45
PCSL-0.45+
PC-0.45
PC-0.45+
Slab #2
HPC
HPC+
SFSL-0.40
PCSL-0.40
PCSL-0.40+
PCSL-0.45
PCSL-0.45+
PC-0.45
PC-0.45+

Day 56

Day 91

WET

WET

a =20 mm a =30 mm a =40 mm a =50 mm a =20 mm a =30 mm

a =40 mm a =50 mm

235.88

251.50

267.50

269.88

260.85

280.35

280.21

300.52

153.50

178.75

197.38

222.63

159.00

190.25

220.25

259.01

135.21

167.30

178.59

181.39

140.20

180.11

188.85

219.68

47.15

49.23

49.20

52.08

67.99

87.56

90.28

109.22

33.41

38.60

41.53

49.64

58.57

63.54

83.49

98.90

25.25

34.40

37.60

42.70

35.88

53.40

63.99

82.73

22.28

32.18

36.10

44.23

29.76

45.45

55.04

67.00

13.28

16.53

16.23

17.40

24.56

30.73

31.10

32.48

12.15

15.48

14.28

15.65

21.16

25.29

27.83

27.48

a =20 mm a =30 mm a =40 mm a =50 mm a =20 mm a =30 mm

a =40 mm a =50 mm

159.60

249.50

254.88

259.88

200.00

310.21

320.12

322.00

131.75

186.75

201.50

230.13

149.64

198.25

241.20

250.21

109.77

166.29

183.65

185.75

129.65

180.58

200.69

210.25

39.88

44.70

49.39

52.71

41.69

62.38

72.45

88.85

28.53

32.60

41.80

49.89

34.08

50.89

69.25

82.16

18.68

26.83

32.55

37.53

30.53

45.83

56.25

73.01

18.45

30.05

33.75

42.65

27.01

40.64

54.24

64.71

13.43

16.98

17.58

19.30

20.33

32.34

35.69

36.65

10.88

14.53

14.80

16.03

18.29

25.35

27.90

26.00

260

END
SUMMER 2010

FIN

You might also like