You are on page 1of 182

i

EFFECT OF BINDER COMPOSITION ON DURABILITY AND MECHANICAL PROPERTIES OF HIGH PERFORMANCE SELF-COMPACTING CONCRETE

by Mehmet Emre Batopu B.S., Civil Engineering, Boazii University, 1997 M.S., Civil Engineering, Boazii University, 2000

Submitted to the Institute for Graduate Studies in Science and Engineering in partial fulfillment of the requirements for the degree of Doctor of Philosophy

Graduate Program in Civil Engineering Boazii University 2006

ii

EFFECT OF BINDER COMPOSITION ON DURABILITY AND MECHANICAL PROPERTIES OF HIGH PERFORMANCE SELF-COMPACTING CONCRETE

APPROVED BY:

Prof. Turan zturan (Thesis Supervisor) Prof. Cengiz Karako Prof. Hulusi zkul Prof. Mehmet Ali Tademir Assoc. Prof. Cem Yaln

DATE OF APPROVAL:

iii

To my dearest Wife and Daughters, Erva, Merve and Mina Begm BATOPU

iv

ACKNOWLEDGEMENTS

I would like to start by expressing my deep gratitude to Prof. Dr. Turan zturan, my thesis advisor, for his guidance, help and support throughout the course of my graduate studies, both masters and PhD. He always managed to save some of his valuable time for me in his busy schedule. I am also grateful to E. Gneyisi and M. Gesolu, assistants of the Boazii University Construction Materials Laboratory, for their help and support. I also would like to thank help-to-technician I. Gltekin and other lab staff for their assistance during the experiments. I would also like to thank Yapkim Yap Kimya Sanayi A.. for providing chemical admixtures (Glenium 51), ISFALT A.. for providing materials and technical support and NUH imento Sanayi A.. for providing cement. Finally, I would like to thank my family who has never deprived me from their support and encouragement during my academic life. Without their support I probably wouldnt be able to complete this work.

ABSTRACT

EFFECT OF BINDER COMPOSITION ON DURABILITY AND MECHANICAL PROPERTIES OF HIGH PERFORMANCE SELF-COMPACTING CONCRETE

This study was carried out in order to investigate the effect of binder composition on fresh and mechanical properties of self-compacting concrete (SCC). Total of eight different concrete mixtures have been produced. All mixes had a total binder powder content of 550 kg/m3. One of these was a control mix with only cement and no powder replacement. The other seven had varied percentages (5, 10, 27, and 36 per cent) of fly ash (FA), silica fume (SF) and limestone powder (LP) replacing Portland cement in the mix. In the fresh state, mixes were tested according to the self-compactability criteria tests; such as flow table, Utube, V-funnel, L-box and visual segregation rating. The mechanical tests consisted of compressive strength, splitting tensile strength, lollipop pullout. Another set of results was obtained from pullout, compression, ultrasound, and rebound hammer tests applied on specimens sawn from different sections of full size columns (10x20x200 cm). The cut surfaces of these columns also allowed the investigation of variations in concrete microstructure with changing depth along the column length. The concrete-steel interface was also examined using video-microscope techniques. The durability properties were assessed using water absorption, sorptivity, rapid chloride permeability, deicing salt scaling, carbonation, sulfate resistance and drying shrinkage tests performed on different number of specimens for each mix. In the fresh state, all the mixes exhibited satisfactory self-compactability and resistance to segregation which was also credited for the homogeneity of the full size columns properties. The results showed that the combined use of SF and FA replacement of cement improved the mechanical and durability properties of the SCC, where as the LP especially in large percentages reduces these properties. The only exception is the drying shrinkage, which is reduced by LP and increased by SF replacement of cement.

vi

ZET

BALAYICI ER N N YKSEK PERFORMANSLI KEND L NDEN YERLEEN BETONUN DAYANIM VE DAYANIKLILIK ZELL KLER ZER NDEK ETK LER

Bu alma balayc ieriinin yksek performansl kendiliinden yerleen betonun (KYB) dayanm ve dayankllk zellikleri zerindeki etkilerini aratrmak amac ile gerekletirilmitir. Toplam olarak sekiz farkl beton karm retilmitir. Karmlarn tamamnda balayc miktar 550 kg/m3 olarak belirlenmitir. Bunlardan biri, hibir mineral katk eklenmeyen ve sadece imento ihtiva eden kontrol karmdr. Dier yedi karmda farkl oranlarda (yzde 5, 10, 27, 36) uucu kl (UK), silis duman (SD) ve ta tozu (TT) imento yerine kullanlmtr. Taze betonda kendiliinden yerleebilme ltlerine uygunluk testleri olan, yaylma, U-tp, V-huni, L-kutu ve grsel ayrma testleri gerekletirilmitir. Mekanik zellikler basn dayanm, yarmada ekme dayanm ve aderans testleri ile incelenmitir. Bir dier sonu kmesi ise byk boyutta dklen kolonlardan (10x20x200 cm) farkl derinliklerde kesilip kartlan numunelerden elde edilmitir. Bu numunelere de basn, ekme dayanm, ultrasonik ses hz ve yzey sertlii testleri uygulanmtr. Farkl derinliklerde elde edilen kesit grntleri stnde detayl grnt analizleri yaplm ve kolon boyunca mikro-yapdaki deiimler aratrlmtr. Kolonlara yerletirilen yatay donatlarda beton-donat ara yz video-mikroskop ile incelenmitir. Dayankllk zellikleri ise su emme, klcal geirimlilik, hzl klorr geirgenlii, buz zc tuzlara kar diren, karbonatlama, slfat dayanm ve kuruma rtresi testleri ile belirlenmitir. Karmlarn tm kendiliinden yerleme zelliini salamakta ve ayrma gstermemektedir, kolon elemanlarnn yapsndaki homojenlik de buna balanabilir. Sonulara gre KYBde imentonun yerine katk olarak SD ve UKnn birlikte kullanm, ta tozuna gre daha iyi mekanik ve dayankllk zelliklerinin olumasn salamaktadr. Bunun tek istisnas TT kullanm ile azalan ve SD ile artan kuruma rtresi deerleridir.

vii

TABLE OF CONTENTS

DEDICATION.iii iii ACKNOWLEDGEMENTiv iv ABSTRACT. v v ZET vi vi LIST OF FIGURES .?? xi LIST OF TABLES ..?? xvi LIST OF SYMBOLS / ABREVIATIONS...?? xx 1. INTRODUCTION. .. 1 1 2. LITERATURE REVIEW. 4 4 2.1. 2.2. 2.3. Brief History of Self-Compacting Concrete4 4 Theory of Self-Compactibility.. 5 Types of Self-Compacting Concrete.. 77 2.3.1. 2.3.2. 2.4. 2.4.1. 2.4.2. 2.4.3. Powder Type SCC 7 7 Viscosity Type SCC. 8 8 Mix Design10 10 Binder Composition..13 13 Fresh Concrete Properties.16 16 2.4.3.1. 2.4.3.2. 2.4.3.3. 2.4.3.4. 2.4.3.5. 2.4.3.6. 2.4.4. 2.4.4.1. 2.4.4.2. 2.4.4.3. 2.4.5. 2.4.5.1. Flow Table Test.16 16 U-Tube Test...19 19 V-Funnel Test21 21 L-Box.21 21 Rheological Studies...22 22 Segregation25 25 Compressive Strength....26 26 Splitting Tensile Strength..29 29 Bond Strength31 31 Rapid Chloride Permeability.32 32

Previous Research on Self-Compacting Concrete..10 10

Mechanical Properties...26 26

Durability Properties.32 32

viii

2.4.5.2. 2.4.5.3. 2.4.5.4. 2.4.5.5. 2.4.5.6. 2.4.5.7. 3.1.

Sulphate Resistance...34 34 Freeze-Thaw Resistance36 36 Deicing Salt-Scaling Resistance38 37 Carbonation...39 39 Capillary Absorption.40 40 Drying Shrinkage..41 41

3. EXPERIMENTAL PROGRAM AND METHODOLOGY...44 43 Materials.44 43 3.1.1. Cement..44 43 3.1.2. 3.1.3. 3.1.4. 3.1.5. 3.1.6. 3.2. 3.3. 3.4. Fly Ash ... 44 45 Silica Fume..44 45 Limestone Powder ..44 45 Aggregate.46 47 Superplasticizer....47 48

Concrete Mixture... 49 48 Production of Test Specimens...49 50 Test Procedures.50 51 3.4.1. Fresh Concrete Properties50 51 3.4.1.1. 3.4.1.2. 3.4.1.3. 3.4.1.4. 3.4.1.5. 3.4.2.1. 3.4.2.2. 3.4.2.3. 3.4.2.4. 3.4.2.5. 3.4.3. 3.4.3.1. 3.4.3.2. 3.4.3.3. 3.4.2. Slump Flow Test..50 51 U-Tunnel Test..51 52 V- Funnel Test. 51 52 L-Box Test 52 51

Segregation...51 52 Hardened Concrete Properties.52 53 Compressive Strength..53 52 Splitting Tensile Strength 53 52 Pull-out Strength. 53 54 Test on Full Size Structural Elements..55 56 Ultrasonic Pulse Velocity Test.60 61 Water Absorption.60 61 Capillary Absorption60 61 Chloride Ion Permeability62 63

Permeability Tests 61 60

ix

3.4.3.4. 3.4.4. 3.4.4.1. 3.4.4.2. 3.4.5. 4.1. 4.2.

Carbonation 66 65 Deicing Salt Scaling Resistance. 66 65 Sulphate Resistance 68 67

Durability Tests. 66 65

Drying Shrinkage 69 68

4. TEST RESULTS AND EVALUATION 70 69 Fresh Concrete Properties 70 69 Hardened Concrete Properties.. 73 72 4.2.1. Mechanical Properties. 73 72 4.2.1.1. 4.2.1.2. 4.2.1.3. 4.2.1.4. 4.2.1.5. 4.2.2. 4.2.2.1. 4.2.2.2. 4.2.2.3. 4.2.2.4. 4.2.3. 4.2.3.1. 4.2.3.2. 4.2.4. 4.3. 4.4. Compressive Strength.. 72 73 Splitting Tensile Strength. 74 75 Pull-out Strength.. 75 76 Ultrasonic Pulse Velocity Test. 76 77 Test on Full Size Structural Elements.. 76 77 Water Absorption.. 109 108 Sorptivity 109 108 Chloride Ion Permeability.. 110 109 Carbonation 110 109 Deicing Salt Scaling Resistance. 111 110 Sulphate Resistance 112 111

Permeability Tests 109 108

Durability Tests. 111 110

Drying Shrinkage... 113 112

Cost Analysis of SCC. 114 113 Analytical Correlations between Concrete Properties 120 119 4.4.1. 4.4.2. 4.4.3. 4.4.4. 4.4.5. Relationship between Water Absorption and Sorptivity (S).. 121 120 Relationship between Total Charge Passed (Q) and Sorptivity (S). 120 Relationship between Total Charge Passed (Q) and Water Absorption 121 Relationship between Total Charge Passed (Q) and Initial Current (IC). 121 Relationship between Total Charge Passed (Q) and

Compressive Strength (fc).. 123 4.4.6. 4.4.7. 4.4.8. 4.4.9. 4.5. Relationships between Splitting Tensile Strength, Compressive Strength and Pull-out Strength. 123 Relationships between Compressive Strength, Pull-out Strength and UPV of Column and Lab Specimens...112 125 Relationship between Compressive Strength and UPV. 126 Relationship between In-situ Mechanical Properties and Distance from the Top of the Columns..112 127 Overall Evaluation of the SCC Mixes.. 137 5. CONCLUSIONS .213 141 6. RECOMMENDATIONS FOR FUTURE WORK ..217 144 APPENDIX A: PHOTOGRAPHS. 145 REFERENCES ..225 153

xi

LIST OF FIGURES

Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. Figure 2.6. Figure 2.7. Figure 2.8.

Excess paste theory .

Relation between spherical factor of powders and paste fluidity 15 Schematic representation of slump flow.. 17 Flow spread test mould and data representation 18 Schematic representation of U-type test apparatus 20 Schematic representation of the V-Funnel apparatus 21 Schematic representation of the L-Flow apparatus. 22 Relationship between torque and speed in Newtonian and Bingham Model. 23

Figure 2.9. Figure 2.10. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 3.5.

Two-point workability apparatus 24 Pullout specimen . 32 Aggregate grading curve and zones.. 47 Pull-out Test Set-up54 Schematic representation of the column specimen 59 Experimental setup for sorptivity test 62 RCPT experimental setup. 64

xii

Figure 3.6. Figure 4.1. Figure 4.2. Figure 4.3. Figure 4.4. Figure 4.5. Figure 4.6. Figure 4.7. Figure 4.8. Figure 4.9. Figure 4.10. Figure 4.11. Figure 4.12. Figure 4.13. Figure 4.14.

The deicing salt scaling specimen with galvanize dike.67 Compressive strength test results at 7 days.. 84 Compressive strength test results at 28 days.. 84 Compressive strength test results at 90 days .. 85 Compressive strength developments at 7, 28, 90, and 900 days.. 85 Splitting tensile strength test results at 7 days.. 86 Splitting tensile strength test results at 28 days.86 Splitting tensile strength test results at 90 days 87 Splitting tensile strength developments at 7, 28, and 90 days 87 Pull-out strength test results at 7 days 88 Pull-out strength test results at 28 days.. 88 Pull-out strength test results at 90 days 89 Pull-out strength developments at 7, 28, and 90 days 89 90 Ultrasonic pulse velocity results at 28 and 90 days Results of pullout strength of lab specimens compared with column specimens 91

Figure 4.15.

Results of compressive strength of lab specimens compared with 92 column specimens

xiii

Figure 4.16.

Results of Schmitt rebound hammer values changing with depth of column.. 93

Figure 4.17.

Results of ultrasonic pulse velocity test changing with depth of column..... 94

Figure 4.18. Figure 4.19. Figure 4.20. Figure 4.21. Figure 4.22.

Linear expansion versus time for concretes exposed to sulphate attack. 118 Drying shrinkage test results. 118 Relationship between sorptivity and water absorption at 28 - 90 days 128 Relationship between total charge passed and sorptivity at 28- 90 days. 129 Relationship between water absorption and total charge passed at 28 and 90 days. 129

Figure 4.23.

Relationship between initial current and total charge passed at 28 and 90 days 130

Figure 4.24. Figure 4.25.

Relationship between initial current and total charge passed for all data. 130 Relationship between comp. strength and total charge passed at 28 and 90 days 131

Figure 4.26.

Relationship between compressive and splitting tensile strength at 7, 28 and 90 days.. 131

Figure 4.27.

Relationship between splitting tensile and pull-out strength at 7, 28 and 90 days.. 132

Figure 4.28.

Relationship between compressive strength and pull-out strength at 7, 28 and 90 days.. 132

xiv

Figure 4.29.

Relationship between compressive strength of column and lab specimens... 133

Figure 4.30. Figure 4.31. Figure 4.32. Figure 4.33.

Relationship between pull-out strength of column and lab specimens.. 133 Relationship between UPV of column and lab specimens. 134 Relationship between compressive strength and UPV at 28 and 90 days.. 134 Relationship between compressive strength and distance from the top of the column.. 135

Figure 4.34.

Relationship between pull-out strength and distance from the top of the column. 135

Figure 4.35. Figure 4.36.

Relationship between UPV and distance from the top of the column... 136 Relationship between Schmitt hammer rebound value and distance from the top of the column. 136

Figure 4.37.

Relationship between cost of materials and compressive strength of SCC 137

Figure A.1. Figure A.2.

Slump flow test in a time sequence 145 Column specimen formwork (a), demolded column specimen before sawing operation(b) 146

Figure A.3.

Deicing salt scaling resistance specimens after increasing numbers of F&T cycles. 147

Figure A.4.

Deicing salt scaling specimens after 50th cycle, with severely deteriorated surfaces148

xv

Figure A.5. Figure A.6.

Test set-up for Rapid Chloride Permeability. 149 Carbonation test; spraying of phenolphthalein (a), close-up photo of M6 where slight carbonation was observed (b) 150

Figure A.7.

Video-microscope images of concrete-steel interface without any visible defects .. 151

Figure A.8.

Video-microscope images of concrete-steel interface without any visible defects. 152

xvi

LIST OF TABLES

Table 2.1. Table 3.1.

Mixture proportion guidelines for SCC1 11 Physical, chemical and strength properties of cement and powder materials.. 45

Table 3.2. Table 3.3. Table 3.4. Table 3.5. Table 3.6. Table 3.7. Table 3.8. Table 4.1. Table 4.2. Table 4.3. Table 4.4. Table 4.5. Table 4.6.

Sieve analysis and physical properties of the aggregates 46 Analysis report of the superplasticizer (Glenium 51).. 48 Mix proportioning of concretes 49 Classification of the column sections... 58 60 The criterion for the ultrasonic pulse velocity.. 64 Interpretation of results obtained using RCPT T 277-89.. 66 The rating scale for ASTM C 672 69 Properties of fresh concrete 70 General acceptance criteria for SCC and EN 206-1:2000 classes. 79 Compressive strength results at 7, 28 and 90 days 79 Splitting tensile strength results at 7, 28 and 90 days 80 Pull-out strength results at 7, 28 and 90 days 80 Ultrasonic pulse velocity test results.

xvii

Table 4.7. Table 4.8. Table 4.9.

Test results of column sections at 28 days.. 81 Correlation between the properties of column and lab specimens 83 Different stages of image analysis in calculating the area ratio for column M195

Table 4.10.

Different stages of image analysis in calculating the area ratio for column M2. 96

Table 4.11.

Different stages of image analysis in calculating the area ratio for column M3 . 97

Table 4.12.

Different stages of image analysis in calculating the area ratio for column M4 98

Table 4.13.

Different stages of image analysis in calculating the area ratio for column M5. 99

Table 4.14.

Different stages of image analysis in calculating the area ratio for column M6.. 100

Table 4.15.

Area ratios of concrete column cross-sections found by image analysis and used in segregation coefficient calculations157 101

Table 4.16. Table 4.17. Table 4.18. Table 4.19.

Segregation coefficients of the concrete mixtures casted in columns 101 Stages of image analysis at the steel-concrete interface for column M1...155 102 Stages of image analysis at the steel-concrete interface for column M2.. 103 Stages of image analysis at the steel-concrete interface for column M3.. 104

xviii

Table 4.20. Table 4.21. Table 4.22. Table 4.23. Table 4.24.

Stages of image analysis at the steel-concrete interface for column M4... 105 Stages of image analysis at the steel-concrete interface for column M5... 106 Stages of image analysis at the steel-concrete interface for column M6... 107 Water absorption test results at 28 and 90 days.. 114 Sorptivity test results of the concrete mixtures at the age of 28 and 90 days.. 114

Table 4.25. Table 4.26

Rapid chloride permeability test results at 28 and 90 days. 115 Visual rating results of the deicing salt scaling tests on concrete specimens 115

Table 4.27. Table 4.28.

116 Depth of carbonation of concretes exposed to outdoor environment... Compressive strength of concretes stored in water and in 116 sulfate solution.

Table 4.29. Table 4.30. Table 4.31.

117 Cost analysis of SCC mixes used in this study 122 Initial current readings of RCPT tests.. Results of the regression analyses for the correlation between the 127 mechanical properties and distance from the top of the columns.

Table 4.32. Table 4.33.

139 Relative values of mechanical properties of SCC mixes.. 140 Relative values of durability properties of SCC mixes

xix

Table 4.34.

Relative values of overall properties of SCC mixes including cost of materials 140

xx

LIST OF SYMBOLS / ABREVIATIONS

C C D db ESS fc n fp n ft
n

Yield value (mobility) Plastic viscosity Percentage damage during sulphate exposure Bar diameter Explained sum of squares Compressive strength at age n Pull-out strength at age n Splitting tensile strength at age n Concentration of coarse aggregate Mean of coarse aggregate concentration Cumulative water absorption per unit area of inflow surface Embedment length (bonded length) Angular speed Ultimate applied load Total charge passed during RCPT Coefficient of correlation Coefficient of determination Viscosity Flow diameter Residual sum of squares Sorptivity Torque Elapsed time during sorptivity test Total sum of squares Volume of total aggregate Volume of binder Volume of coarse aggregate Actual value of the cost of the mix Lowest value of the cost among the mixes

Gi
G i

ls N P Q R R
2

Rm rn, Dn RSS S T t TSS Vagg Vb Vc VC VCmin

xxi

VCr VD VDmin VDr Vf VM VMmax VMr Vp Vs Vw m R AEA FA F&T IC ITZ LP OPC RCPT PBFS PLS SC SCC SF UPV WA VMA

Relative value of the cost of the mix Actual value of a durability / permeability property of the mix Lowest value of a durability /permeability property among the mixes Relative value of a durability / permeability property of the mix Volume of fine aggregate Actual value of a mechanical property of the mix Highest value of a mechanical property among the mixes Relative value of a mechanical property of the mix Volume of paste Volume of total solids Volume of water Flowability Concrete strength without sulphate exposure Concrete strength after sulphate exposure Bond strength Air entrainment admixture Fly ash Freezing and thawing Initial current reading during RCPT Interfacial transition zone Limestone powder Ordinary Portland cement Rapid chloride permeability test Powdered blast furnace slag Powdered limestone Segregation coefficient Self-compacting concrete Silica fume Ultrasonic pulse velocity Water absorption Viscosity modifying admixture

1. INTRODUCTION

The concept of self-compacting concrete (SCC) was first introduced in Japan in 1988 as part of an effort to achieve durable concrete structures. In order to achieve the design properties of the concrete mixtures, it was realized that the greatest obstacle at site was poor compaction. The solution to these problems came along with other advantages by the use of SCC, which can be compacted into every corner of a formwork, purely by means of its own weight and without the need for vibrating compaction (Okamura and Ouchi, 1999). Although the SCCs greatest claim is given above, the experimental studies on SCC are mainly focused on lab specimens rather than larger scale representations of the real-size constructional elements.

It has been shown that having a powder content (materials finer than 0.1 mm) of 500 to 600 kg/m3 is critical for attaining self-compactability in concrete (Topu and Khurana, 2000). Since the excessive use of cement, as a fine material in concrete is uneconomical (along with other drawbacks), mix-proportions with replacement powders such as silica fume, fly ash, limestone powder is investigated. Here the aim was to distinguish between the effects of different mineral powders (used separately or combined with varying contents, properties and finenesses) in SCC with reasonable explanations.

The fundamental objective of this work was to provide information on the hardened properties of self-compacting concrete produced using different powders as cement replacements. To this end, the engineering properties of SCC in sampled specimens were to be compared with those of in full size structures. So this study aimed to combine and compare results of the tests conducted on the laboratory scale and the larger scale specimens to investigate the extent of the observed self-compactability on the lab floor. The concrete was produced in a laboratory pan-mixer, and the following mechanical properties of sampled specimens were tested: compressive and flexural strength, shrinkage, bond strength with reinforcement. In addition, in order to assess durability, longer-term comparative tests were made on freeze-thaw resistance, water absorption, sulphate

resistance, deicing salt scaling, drying shrinkage, rapid chloride permeability and carbonation. Testing of fresh concrete was a part of the work, also necessary for mix design verification and acceptance purposes. To supplement this, special tests were made on settlement and segregation. The full-size elements were tested in two ways. Firstly a number of in-situ tests were carried out in order to determine mechanical properties at changing heights of vertical elements, and secondly, the cut surfaces of the elements were examined through image analysis to determine coarse aggregate segregation at changing heights of the vertical elements. Thus the scope of this study might be summarized as; Comparison of basic mechanical properties of hardened SCC with varying binder composition. Comparison of some aspects of the durability of hardened SCC with varying binder composition. Evaluation of the effects of different mineral powders (used separately of combined). Verification of properties of hardened SCC in full-size columns using in situ test methods. Evaluation of the relation between mechanical and durability properties of SCC. Evaluation of the problems associated with mineral admixtures, such as thermal cracking or incompatibility with superplastizers. Comparison of the economical aspects of using different mineral admixtures for specific concrete properties. Since self-compacting concrete is a relatively new topic there are some areas of research where the existing data is limited making this study significant. (1) Further study is necessary for the comprehensive evaluation of durability in SCC (combined with mechanical properties), since most of the recent work is on fresh concrete properties and mix design. For instance, there is contradiction among researchers about the effect of mineral admixtures on carbonation and shrinkage of concrete.

(2) There is also limited data on the frost durability and scaling resistance of SCC which is susceptible to such damage with a high potential of being used in congested reinforced sections of beams and girders of bridges. (3) More data is also required to correlate the results of simple (relatively new) empirical tests used in determining fresh properties of SCC. (4) Last but not the least, in evaluating the effects of mineral admixtures, cost analysis should be done since economical considerations is one of the main drawbacks of the wider use of SCC in the construction industry. Commenting on the future of SCC, Okamura et al. (Okamura, 1999) concluded that the rational mix-design method and appropriated testing method at site have almost been established for SCC and these eliminate some of the obstacles for making SCC widely used. The next task is to distribute the manufacturing, qualification and construction techniques for SCC. In addition, new structural design and construction systems making full use of SCC should be introduced, such as the sandwich structures in the immersed tunnel in Kobe, Japan. As mentioned above the main obstacle against the wide use of SCC is surely its price. The superplastizers used are expensive to produce and some of the inorganic fillers such as silica fume are also expensive due to high demand. But it should not be forgotten that high w/c (0.6) present concrete is not durable, and consequently often requires repairs and causes inappropriate level of service. Therefore, the initial cost should not be evaluated alone, but together with life-cycle cost. It is also worth noting that SCC containing up to 50% fly ash replacement can attain 28-day target strengths of 35 MPa. It is also shown by cost analysis that such a SCC mix has no significant extra cost compared to ordinary concrete and would be flowable with a slump flow of 500 mm and more resistant to segregation and thermal cracking, but might exhibit higher bleeding water and long setting time (Bouzoubaa, et al. 2002). When self-compacting concrete becomes so widely used that it will be seen as the standard concrete rather than an expensive special concrete, it would be much more convenient for engineers to design durable and reliable concrete structures worldwide.

2. LITERATURE REVIEW

2.1.

Brief History of Self-Compacting Concrete

The concept of self-compacting concrete (SCC) was first introduced in Japan in 1988 as part of an effort to achieve durable concrete structures. In order to achieve the design properties of the concrete mixtures, it was realized that the greatest obstacle at site was poor compaction. The need for sufficient compaction increased the demand for skilled workmanship (recently declining in Japan) and time required for concrete placement. The solution to these problems came along with other advantages by the use of SCC, which can be compacted into every corner of a formwork, purely by means of its own weight and without the need for vibrating compaction (Okamura and Ouchi, 1999).

In 1988, Okamura completed the first prototype of SCC using materials already on the market. The prototype performed satisfactorily with regard to drying and hardening shrinkage, heat of hydration, denseness after hardening, and other properties. This concrete was defined as follows at the three stages of concrete: (1) fresh: self-compacting (2) early age: avoidance of initial defects (3) hardened: protection against external factors (Okamura and Ouchi, 1999) After the development of the prototype, SCC caught the attention of research institutes of large construction companies in Japan and soon found many practical applications. The most important reasons for this quick popularity were (Bouzoubaa and Lachemi 2002); (1) reducing construction time and labor cost (important since 50% of total construction cost is related to manpower (Byfors, 1999); (2) assuring of proper compaction; especially in confined zones where vibrating compaction is difficult and therefore eliminating the need for compaction, forming better finished surfaces, reducing leveling work (il, 2000);

(3) eliminating noise due to vibration; effective especially at pre-cast concrete industry and urban construction sites, for better working environment. During the anchorage construction of Akashi-Kaikyo Bridge (1.991 meter, longest suspension bridge in the world with 500.000 m3 of concrete poured at the piers) in 1998, SCC proved to eliminate segregation although it fell as much as 3 meters and contained large size coarse aggregates (40 mm). Another positive effect was observed in the 20% reduction in the construction period. There are many other recent projects where the workmanship (up to 80% reduction), and construction time has been reduced similarity with the use of SCC (Nagataki, 2000). Self-compacting concrete is shown also to be proper repair material which has good filling capacity, and that the absence of segregation enhances the characteristics with the old concrete, even in an inverted position (a common practice in repairing bridge girders damaged by frost). Microstructure studies also indicated that SCC performed as well as conventional dry-mix shotcrete, making it a very interesting new repair material (Lacombe et al., 1999). SCC has a great potential for new application methods in construction since it allows the use of new techniques, which were not feasible before with conventional concrete. 2.2. Theory of Self-compactability According to Okamura self-compactability might be achieved by separating the concrete into two constitutes, coarse aggregates and mortar and then adjusting the rheology of the mortar to achieve self-compactability by incorporating a variety of mineral additives, plasticizers, and thickeners. Contrarily, A.W. Saak argues that optimizing the particle size distribution of the binder (for example, cement, silica fume, fly ash, etc. ) and of fine and coarse aggregates based on packing considerations provide a better understanding of the physical properties required to achieve SCC (Saak et al., 2001).

It has been argued that high packing density cant be used as criteria for producing highly flowable concrete, but according to recent experimental work, particle-packing density should be used in conjunction with interparticle separation (IPS) in order to explain the physical behavior of the paste (Saak et al., 2001). In concrete, the volume of the binder and particle size distribution of the aggragetes governs the IPS. For a given particle size distribution, the volume of the binder should be sufficient to fill the interstitial voids between the aggregates to produce a desired IPS. These arguments can also be explained by the Excess Paste Theory proposed by Kennedy (Kennedy, 1940), which states that to attain workability, it is necessary to have not only enough cement paste to cover the surface area of the aggregates, so as to minimize the friction between them, but also more of it to give better flowability. The application of this theory is illustrated in Figure 2.1 (Noguchi et al., 2001).

Aggregate

Excess paste

Add paste Void Thickness of Excess paste

( Compacted ) ( Dispersion )

Figure 2.1 Excess paste theory

2.3.

Types of Self Compacting Concrete

The method for achieving self-compactability involves not only high deformability of paste or mortar, but also resistance to segregation between coarse aggregate and mortar when the concrete flows through the confined zone of reinforcing bars. Yield stress of a paste is easily decreased with the addition of a superplastizer, but yield stress must be above a critical minimum to insure that segregation of aggregates does not occur when the mix is at rest (Saak et al., 2001). Since segregation prevention is critical in producing SSC, generally three methods were proposed to obtain adequate segregation resistance (Nagataki, 2000), (Topu and Khurana, 2000). (1) limiting coarse aggregate content and reducing the water to powder ratio by increasing the powder content using cement and/or mineral admixtures such as limestone powder, fly ash, or granulated blast furnace slag powder (Powder Type SCC), (2) adding viscosity agent with superplastizer (Viscosity Type SCC), (3) doing both of the above (Combination Type SCC).

2.3.1. Powder Type SCC

To make the concrete deform well, it is beneficial to reduce the friction between the solid particles, which includes coarse aggregate, fine aggregate and all types of powder. To reduce the aggregate-aggregate friction, it is necessary to reduce the possibility of aggregate inter-particle contact. One way to achieve this is to increase the aggregate inter-particle distance by reducing the aggregate content or, in other word, increasing the paste content. The friction among powder materials is not possible to reduce by increasing the inter-particle distance through increasing the water content of the paste. The paste phase itself needs to have excellent deformability and that require dispersion of fine particles which in turn require surface active agents like superplasticizers. The use of too high a water content leads to segregation and undesirable performance of the hardened concrete e.g. strength and durability. The shape of the powder particles has an effect on the demand of water and superplasticizers. The use of spherical pozzolans such as FA is considered effective for

this purpose. A reduction of the friction between aggregate and powder particles tends to reduce the resistance to segregation. It may then be effective not to increase the deformability of the paste and concrete as a whole, but rather to increase the viscosity of the paste (Skarendahl, 1999).

In mixing SCC increasing the paste volume by increasing powder content consequently leads to lower content of coarse aggregates. The theory behind can be best explained by changes in internal stress of the mix. The frequency of collisions and contact between aggregate particles can increase as the relative distance between the particles decreases and the internal stress can increase when concrete is deformed, particularly near obstacles. It has been revealed that the energy required for flowing is consumed by the increased internal stress, resulting in blockage of aggregate particles. Limiting the coarse aggregate content, whose energy consumption is particularly intense, to a level lower than normal proportions is effective in avoiding this kind of blockage (Okamura and Ouchi, 1999). Limiting coarse aggregate also serves to avoid segregation on superplastizer addition and a simple approach consists of increasing the sand content replacing 4 to 5 % coarse aggregate (Bouzoubaa and Lachemi, 2002). But this reduction of coarse aggregate leads to a use of high volume of cement, which in turn causes cost, temperature increase. Previous investigations showed that fly ash improves rheological properties and reduces thermal cracking due to excessive heat of hydration. Most of the studies contained concrete mixes with different percentages of fly ash replacing cement not exceeding 30% replacement (Bouzoubaa and Lachemi, 2002).

2.3.2. Viscosity Type SCC

Highly viscous paste is also required to avoid the blockage of coarse aggregate when concrete flows through obstacles. When concrete is deformed, paste with a high viscosity also prevents localized increases in the internal stress due to the approaching coarse aggregate particles. High deformability can be achieved only by the employment

of a superplasticizer, keeping the water-powder ratio to be very low value (Okamura and Ouchi, 1999).

The action of superplastizers in SCC takes place as these large molecules surround the surface of hydrating cement particles. Since the cement particles are normally positively charged, the dipolar water molecules are attracted on their surface and for a sheath of water with the help of SP. This sheath prevents the cement particles from flocculating, and its dispersive action also allows more surface of cement to be in contact with water. While increasing the water/cement ratio to enhance flowability leads to a decrease in viscosity, SP increase can make the paste more flowable with little reduction in viscosity, and avoids segregation. But it is also worth mentioning that there is a limit for the increase in flowability with increasing SP dosage (Venugopal et al., 2000).

Viscosity-modifying admixtures (viscosity agent) are used as an alternative approach to producing SCC with high powder content (Bilberg, 2000), but they also increase the materials cost. So the best alternative is using mineral additives like fly ash, blast furnace slag, or limestone powder to enhance stability.

A study by M. Collepardi et al. aimed to optimize the type and the dosage of superplasticizer in Class F flyash concrete, so that compressive strengths as high as those of superplastizer silica fume concretes could be obtained (Collepardi, 2001). In this study the superplastizer dosage and the pozzolan addition ranged from 2 to 4% and from 12 to 20% respectively by weight of cement. The cement factor varied from 255 to 400 kg/m3. The results of this work indicated that only in the presence of ASTM Type III Portland cement, superplastizer flyash concrete can be as strong as the corresponding silica fume concrete, particularly at relatively high cement factors ( 300 kg/m3). In their discussion the results are evaluated as follows. The ASTM Type III cement has high C3S content and higher fineness, the hydration of this cement produces more Ca(OH)2 , to be used in the pozzolanic reactions of both flyash and silica fume. But the higher amount of Ca(OH)2 should favor fly ash more than silica fume, since fly ash is potentially less reactive for the

10

lower fineness and lower content of amorphous silica. A less pronounced similiar advantage for fly ash is the increased cement content. Another study by zkul et al. points out that at least 500 kg/m3 of fine material (> 100 m), including the cement, is necessary to produce a self-compacting concrete (zkul et al., 2000). The concrete mixes used in this study contained fly ash and silica fume, and it is observed that 50 kg/m3 of fly ash addition to 450 kg/m3 of cement is not sufficient for self-compactability. The best results are observed for 100 and 150 kg/m3 of fly ash addition, but a value of 200 kg/m3 fly ash caused rapid flow loss as observed with an excessive increase in cement content. 2.4. Previous Research on Self-compacting Concrete

2.4.1. Mix Design Self-compactability can be largely affected by the characteristics of materials and the mix-proportion. A simple mix-proportioning system for mixing SCC has found general acceptance (Okamura and Ouchi, 1999). The coarse and fine aggregate contents are fixed so that the self-compactability can be achieved easily by adjusting the water-powder ratio and the superplasticizer dosage only. The catch-points are listed as; (1) Coarse aggregate content in concrete is fixed at 50% of the solid volume (2) Powder content (materials finer than 0.1 mm): 500-600 kg/m3 (3) Fine aggregate (sand 0.1 to 5mm) content is fixed at 40% of the mortar volume. (4) Water-powder ratio in volume is assumed as 0.9 to 1.2, depending on the properties of the powder. (5) Maximum size of aggregate: up to 20 mm (6) Superplastizer (polycarboxylate ethers), AEA, VMA dosage and the final water-powder ratio are determined so as to ensure self-compactability (using following tests in (7)-(9)). (7) Slump flow: 60 to 70 cm (8) V-funnel: 8 to 12 sec (9) U-box: h> 30 cm

11

There are other guidelines for SCC using the recommended volume ratios of each constitute and there is little variation in the suggested guidelines given by different authors as given in Table 2.1 (Saak et al., 2001).

Table 2.1 Mixture proportion guidelines for SCC.

Researchers Okamura Yurgui et al Ambrose et al

Vc / Vagg 0.64 0.54 0.44

Vf / Vagg 0.36 0.46 0.56

Vb / Vs 0.22 0.24 0.18

(Vb + Vf) / Vagg 0.64 0.78 0.78

where Vc: volume of coarse aggregate; Vf : volume of fine aggregate.; Vagg : volume of total aggregate.; Vb : volume of binder (solids); and Vs = v. of total solids (aggregates + binder).

The mortar or the paste of SCC needs to be deformable as well as being viscous. This can be achieved by proper dosage of superplastizer, which ensure a low waterpowder ratio for the high deformability. Okamura et al. suggested that, the best way to determine the proper water-powder ratio and superplastizer dosage is to make trial mixes (Okamura and Ouchi, 1999). So when the mix proportion is decided, selfcompactability has to be tested by U-type test, slump-flow and funnel tests. The relations between the properties of the mortar in SCC and the mix proportion have been investigated and then formulated by Okamura and Ouchi. These formulae can be used for establishing a rational method for adjusting the water-powder ratio and superplasticizer dosage for achieving appropriate deformability and viscosity. In these formula Ouchi et al. defines indexes for flowability and viscosity as m and Rm , respectively (Ouchi et al., 1996).

m = (r1r2 r02) /r02

(2.1)

12

where r1, r2 are the measured flow diameter and r0 is the flow cones diameter. Rm = 10/t (2.2)

where t is the measured time (in seconds) for mortar to flow through the V-funnel. These indexes are practical to use since they are obtained by simple tests. Larger m value indicates higher flowability and larger Rm indicates higher viscosity. Ozawa et al. investigated these indexes and concluded that mortar with a m of 5 and an Rm of 1 is the most appropriate mixture for achieving SCC (Ozawa et al., 1996). Another suggestion for m and Rm values comes from Edamatsu et al., saying 3 m 7 and 1 Rm 2 is sufficient for SCC (Edamatsu et al., 1999). Since a rational-mix design method which determines the necessary superplastizer amount, Sp / P, and water-powder ratio, Vw / Vp, the (since these values are specific for each different material) best way to attain self-compactability is making assumptions and that checking m and Rm values. Ouchi et al. also emphasizes the fact the chemical action of the superplastizer might be influenced by combination with the powder, which has to be assessed with use of trial mixes (Ouchi et al., 1996).

In mixing conventional concrete, water-cement ratio is fixed to ensure predetermined design strength. With SCC, however the water-powder ratio has to be selected taking self-compactability into account because self-compactability is very sensitive to this ratio. Despite this fact, usually the water-powder ratio is still low enough to satisfy the strength for ordinary structures unless the most of powder materials in use is not reactive (which is not the case for silica fume, fly ash). Mixing procedure of SCC is also different for conventional concrete and a typical SCC mixing procedure is presented below (Khayat et al., 2000). (1) All concrete mixtures are prepared in 60 L batches in rotating drum mixer (2) First, sand and coarse aggregate are mixed (homogenized) for 30 sec. (3) Then 75% mixing water and superplastizer is added and mixed for 30 sec.

13

(4) Cement and mineral additives are added and mixing is resumed for 1 min. (5) The remaining water is added and mixed for 3 min. After 2 min at rest the concrete is mixed for a final additional 2 min.

2.4.2. Binder Composition The pozzolanic additives such as fly ash, slag, silica fume, zeolite silica, etc. are often used to produce SCC in order to improve its strength, workability and durability and to reduce the cost as well (Jianxiong et al., 1999).

The roles of these pozzolanic additives are as follows; (1) increasing the hydration products due to additional pozzolanic reactions; reducing porosity can improve durability and resistance to chemical attack with increases strength. (2) reducing heat of hydration and reaction rate; reducing micro cracks and transition zones (3) increases filling effect of micro aggregate, adjusting grading of the components to achieve optimum compact, adjusting cohesiveness (4) improving workability due to increased fines content and shape effect (Fang et al., 1999). (5) reducing use of cement, achieving economic and environmental benefits (Jianxiong et al., 1999).

The most common knowledge about increasing the powder content in the concrete paste is that it generates an exponential increase in yield value and plastic viscosity. When a new generation superplastizer is added to this paste the yield value is significantly reduced with little reduction in plastic viscosity and this is the principle of forming SCC. A drawback of high powder content in SCC is the potential formation of thermal cracks, especially in mass concreting. For this reason, such concrete is mostly used with granulated blast furnace slag powder, fly ash and limestone powder. When water to powder ratio of the concrete is low, the resulting w/c ratio leads to high

14

compressive strength values of over 40 MPa. When limestone powder is used as the mineral admixture, its content might be adjusted to attain normal strength concrete (Nagataki, 2000).

The rheological studies by Nischer clearly shows that flowability of the paste increases as the Blaine fineness of the different additions of fly ash, lime stone powder increases, and the same increase also evident in replacement of cement with 50% blast furnace slag powder (Nischer, 2000). According to Saak et al., SCC paste with 30% silica fume (replacing cement) exhibits an increase in yield stress and equilibrium viscosity as the SP dosage increases (Saak et al., 2000). They further noted that at zero yield stress, replacement of cement by 30 % silica fume leads to a 36% reduction in equilibrium viscosity compared to the all-cement mix, and attributed this to the increase in the solid packing density and spherical morphology of the silica fume particles.

Khayat et al. produced SCC containing 3% silica fume and 20% fly ash replacement, with water-to-binder ratios and binder contents ranging from 0.37 to 0.50 and 360 to 600 kg/m3, respectievly (Khayat et al., 2000). Topu et al. produced SCC with 350 kg/m3 cement with 200 kg/m3 fly ash (%36) from ayrhan power plant, with water-to-binder ratio of 0.35 and 1.2% superplastizer (Glenium 27) (Topu et al., 2000) and compared homogeneity according to compressive strength of lab and site samples of SCC and ordinary concrete. Their slump flow, V-funnel, and U-box values were 4963 cm, 10 sec, 32 cm respectively and all were found to be within the appropriate range for SCC [4]. According to Khurana et al., a silica fume content of 50 kg/m3 makes a total fines content of 450 to 500 kg/m3 sufficient for producing SCC. They also reported that for other types of fines (fly ash, limestone or quartz filler) the total fines content should be between 500 to 600 kg/m3 (Khurana et al., 2000). Comparing the use of these powders, they noted that fly ash is very advantageous for ready mix concrete since it is readily available at a low price and also limestone filler is preferred among precast industry since it has an even lower price and it produces a lighter color concrete more appreciated by the customers.

15

In order to distinguish between the effects of different mineral powders in SCC, Nagataki (Nagataki, 2000) defined the term spherical factor (packing bulk density / specific gravity) and concluded that powders having higher spherical factors improve the fluidity of the paste better. In measurement of fluidity of cement paste, the replacement ratio of inorganic powders was 40% and the water to powder ratio is 90% by volume. The results revealed that silica fume performs better than fly ash due to higher spherical factor as seen in Figure 2.2. Nagataki also investigated the effect of particle size distribution on fluidity and made studies with different particle sizes of limestone powder. The use of combination of fine and coarse powder resulted in higher packing density and larger amount of free water and highest fluidity is also observed at that mix (Nagataki, 2000).

6,0 5,0 Fluidity (1/Pa.s) 4,0 3,0 2,0 1,0 0,0 0,50 OPC PFA PBFS PCS PLS PSS

0,55

0,60

0,65

0,70

Spherical factor of powders

Figure 2.2 Relation between spherical factor of powders and paste fluidity (Nagataki, 2000) Other experiments carried out by Ozawa focused on the influence of mineral admixtures, like fly ash and blast furnace slag, on the flowing ability and segregation resistance of self-compacting concrete. He found out that the flowing ability of the concrete improved remarkably when Portland cement was partially replaced with fly ash and blast furnace slag. After trying different proportions of admixtures, he concluded that

16

10-20% of fly ash and 25-45% of slag cement, by mass, showed the best flowing ability and strength characteristics (Ozawa and Okamura, 1996) Meanwhile Subrahamanian et al. were working on new mix designs in India aimed to investigate and improve the Okamura-Ozawa studies (Subrahmanian et al., 2002). They were trying to determine different coarse and fine aggregate contents from those developed by Okamura. The coarse aggregate content was varied, along with water-powder (cement, fly ash and slag) ratio, being 50%, 48% and 46% of the solid volume. The U-tube trials were repeated for different water-powder ratios ranging from 0.3 to 0.7 in steps of 0.10. On the basis of these trials, it was discovered that self-compactability could be achieved when the coarse aggregate content was restricted to 46 percent instead of 50 percent tried by Okamura. In the next series of experiments, the coarse aggregate content was fixed at 46 percent and the sand content in the mortar portion was varied from 36 percent to 44 percent on a solid volume basis in steps of 2 percent. Again, the water-powder ratio was varied from 0.3 to 0.7 and based on the U-tube trials a sand content of 42 percent was selected. In order to show the necessity of using a viscosity-modifying agent along with a superplasticizer, to reduce the segregation and bleeding, the mixture proportion developed by the two researchers was used to cast a few trial specimens. In these trials, viscositymodifying agent was not used. The cast specimens were heavily reinforced slabs having 2400x600x80 mm and no vibration or any other method of compaction was used. However, careful qualitative observations revealed that the proportions needed to be delicately adjusted within narrow limits to eliminate bleeding as well as settlement of coarse aggregate. It was difficult to obtain a mixture that was at the same time fluid but did not bleed. This led to the conclusion that slight changes in water content or granulometry of aggregate may result either in a mixture with inadequate flowing ability, or alternatively one with a tendency for coarse aggregate to segregate. Therefore, it became necessary to incorporate a viscosity-modifying agent in the concrete mixture. The fresh concrete experimental results for replacing part of the binder with ultrafine materials indicate a vast variation. So it can be concluded that the selection of a fine material for improved concrete workability is not a simple phenomena and it cannot be predicted from the physical or chemical characteristic of the admixture, and can only be determined by trial mix designs.

17

2.4.3. Fresh Concrete Properties 2.4.3.1. Flow Table Test. The slump flow test is one of the most popular methods in evaluating the consistency of concrete, both in lab and on site due to its ease of operation and the portability, where the slump is greater than 24 cm (Noor and Umoto, 1999). The slump flow test (Figure 2.3) specified by the Japan Society of Civil Engineers (JSCE) judges the capability of concrete to deform under its own weight against the friction of the surface with no other external restraint present. Although slump flow test has been designed for testing deformability it must be sometimes misleading since even concrete with the same slump flow can have different behavior when passing through such obstacles as reinforcing bars, depending on their mix proportion. This is why there are some other tests (U, V,L type tests) devised to make this differentiation. Khayat et al. used the slump flow for their measurements and took readings of the diameter (D1,D2) of spread along two perpendicular lines as shown in Figure 2.3. Their slump values ranged from 260 to 280 mm and their mean slump flow values (D1+D2 / 2) ranged between 500 and 610 mm, which are comparable to the minimum accepted value of 520 mm for SCC (Beupre et al., 1999). According to Nagataki, a slump flow ranging from 500 to 700 mm is considered as the slump required for a concrete to be self-compacted (Nagataki, 2000).

slump

slump cone

D1 D2

slump flow = (D1+D2)/2

Figure 2.3 Schematic representation of slump flow

18

Venugopal et al. devised another method for flow spread test and used a mould as shown in Figure 2.4a (Venugopal et al., 2000). The bottom diameter of the mould is 100 mm, so they defined a term for relative flow area, R, as follows;

(D R=

100 2 D = 1 2 100 100

(2.3)

Their values for R typically ranged from 0.2 to 15, and they also found that, for a paste made with particular powder the relative flow area, R, and the water content by volume (Vw/Vp) are linearly related as shown in Figure 2.4.b. The characteristics of the powder with water+SP are defined by two parameters, intercept and slope of the line. The intercept is retained water ratio, which can be thought of as comparing water absorbed on the powder surface together with that required to fill the voids in the powder system and to provide sufficient dispersal of the particles. The slope angle is the deformation coefficient, which is a measure of the sensitivity of the fluidity of the paste to increasing water content (Venugopal et al., 2000).

Figure 2.4 Flow spread test mould and data representation (Venugopal et al., 2000)

19

According to Fang et al. the slump of SCC increases with increased fly ash content in the range of 20-40% when water binder ratio was constant (Fang et al., 1999). This might be explained by the shape effect due to the spherical, smooth surface of the fly ash particles, which act as lubricating balls in the mixture. They also observed a loss in fluidity above 30% grounded fine fly ash replacement, related to increased absorption effect between FA and cement particles after grinding.

Fang et al. also investigated the effect of silica fume and concluded that although strength of concrete was increased to some extent, fluidity of the paste decreased. Some of the slumps did not meet the requirements for SCC, and more SP dosage and higher water binder ratios were required. The reasons for these were given as; (1) SF particles are fine and lightweight: they are easily absorbed on larger particles at low water binder ratios and form flocculation (2) Agitation time is insufficient (3) Silica fume is not compatible with all superplastizers.

Considering these and high price of silica fume, they concluded that it is best to use fly ash and blast furnace slag powder rather than SF (Fang et al., 1999).

According to Jianxiong et al. increased fly ash addition (above 45% cement) reduces the flowability of SCC due to increased water requirement due to increased specific surface. They also reported that the combined addition of fly ash and superfine slag reveals an improved flowability comparing the slump flow (Jianxiong et al., 1999). Gram et al. recorded that, slump flow for SCC increases with filler content (Gram et al., 2000). Studies by Hokkaido verified that fly ash is superior to other powders (blast furnace slag and limestone powder) in its effect to enhance flowability and control heat of hydration (Hokkaido, 1998). 2.4.3.2. U- Tube Test. A highly flowable concrete is not necessarily self-compacting, because SCC should not only flow under its own weight, but should also fill the entire form and achieve uniform consolidation without segregation. This characteristic of SCC is

20

called the filing capacity and several test methods are designed to measure filling capacity, U test being the most common of all.

Among the many testing methods proposed for evaluating self-compactability, Okamura and Ouchi (Okamura and Ouchi, 1999) claims that, U-type test proposed by the Taisei group seems, at this stage to be the most appropriate (Figure 2.5.). In this test, the degree of compactability can be indicated by the height that the concrete reaches after flowing through an obstacle. Concrete with the filling height of over 300 mm can be judged as self-compacting.

The results by Khayat et al. (Khayat et al., 2002), indicate that a SCC made with high binder content of 550 kg/m3, low sand/paste (S/Pt) volume 0.60 to 0.66, and low coarse aggregate volume of 300 to 330 l/m3 is more suitable to ensure high filling capacity of densely reinforced sections than a concrete of similar slump flow (650 mm) with moderate binder content of 425 kg/m3, 0.70 to 0.85 S/Pt volume, and 275 to 405 l/m3 of coarse aggregate.

Concrete fill

Opening of gate

400 mm Sliding gate 571 mm Filling height

Reinforcing bars 4 @50 mm = 200 mm 140 mm 280 mm

Figure 2.5 Schematic representation of U-type test apparatus.

21

2.4.3.3. V- Funnel Test. Funnel test has been proposed for testing viscosity (Figure 2.6) Slump flow and V-funnel flow time might be determined at different times (e.g. 6 and 60 min after first contact of cement with water). The flow time is determined using a simple procedure; the funnel is completely filled with fresh concrete, and the flow time is measured as the time between the opening of the orifice and the complete emptying of the funnel. According to Bouzoubaa, a funnel flow time of less than 6 sec was recommended for SCC (Bouzoubaa and Lachemi, 2002).

concrete fill 500 mm

450 mm 2

1 slide gate 150 mm

75 mm 75 mm Figure 2.6 Schematic representation of the V-Funnel apparatus 2.4.3.4. L Box. Khayat et al. used the L-box for determining flowability and recorded Lflow spread, L-flow slope and flow duration as fresh concrete properties. (Figure 2.7.) They found that in general the mixes with higher w/cm exhibited greater flow spread and flatter surface slopes, and concluded that such mixes might be expected to exhibit greater deformability and filling capacity despite their similar slump flow values (Beupre et al., 1999).

22

Figure 2.7 Schematic representation of the L-Flow apparatus (Beupre et al., 1999) 2.4.3.5. Rheological Studies Although the repeatability, reliability and simplicity of the above empirical tests are advantageous, there are many claims that a better understanding of concrete behavior can be gained from rheological assessments. On the other hand, it has been shown that self-compactability of concrete evaluated by U-box test can be related to rheological parameters, such as yield stress and plastic viscosity of the mortar. Ferraris et al. (Ferraris et al., 2000) investigated the workability of SCC with different rheometers, Utype and V-tunnel tests and concluded that the plastic viscosity and yield stress values do not correlate with V-funnel or U-flow test. They also claimed that slump flow is not enough to determine whether a flowable concrete is SCC. A similar criticism to conventional workability test methods comes from Villareal et al., arguing that a rheological Bingham model should be used to evaluate workability of SCC (Tattersall, 1991). When a torque is applied to an impeller immersed in a Newtonian liquid it is observed that the there is a linear relationship between the angular speed (N) of the impeller and torque (T). This relation is expressed as; T= C.N (2.4)

23

where C is the slope of the linear plot of N versus T and is proportional to fluidity and viscosity of the liquid. It is obvious that this graph goes through the origin implying that the liquid gets into motion without any initial stress to be imposed, but in the case of concrete some minimum stress or force is necessary to get it moving at all. Which means that it posses a yield value, and consequently the flow curve cannot pass through the origin. So concrete does not behave like a Newtonian liquid, although the relationship between torque and speeds is a simple straight line see Figure 2.8, at it has an intercept on the torque axis. The equation of the line may be written; T= C + C. N (2.5)

where T is the torque at angular speed N, C is a measure of yield value (mobility) and C is a measure of plastic viscosity. This relation is defined by the Bingham model and therefore concrete is called a Bingham material.

Newtonian

Bingham

Speed, N

Slope= C''

C'

Torque, T

Figure 2.8 Relationship between torque and speed in Newtonian and Bingham Model Tattersall and Banfill classified the test methods for rheology of fresh concrete, and formed two groups of testing; empirical and rigorously defined (Tattersall, 1991). They defined the disadvantages of empirical testing as the incomparability of the results obtained by different methods, presence of a single-point test representing only a single operating condition. The approach of Tattersall and Banfill has been to use the Bingham model for fluid flow to represent the rheological behavior of fresh concrete. This is a relatively simple linear mathematical model that relates shear rate and stress applied to fresh concrete

24

with constants; the yield stress and plastic viscosity. Tattersall tried to simplify the measurements of concrete rheology using the two-point workability test. This would be done by obtaining a flow curve from measurements of the torque required to rotate a suitable impeller immersed in concrete at several different speeds using the setup shown in Figure 2.9. The application of this model to fresh concrete is an approximation, but it appears to work reasonably well at relatively low shear rates. And in addition to the twopoint testing method there are some recently developed and more complex rheometers, such as BML, BTRHEOM, CEMAGREF-IMG, and IBB Rheometer. (Tattersall, 1991). .

Figure 2.9 Two-point workability apparatus (Tattersall, 1991)

25

2.4.3.6. Segregation. Segregation is defined by Neville, as the separation of the constituent of a heterogeneous mixture so that their distribution is no longer uniform (Neville, 1986). Concrete is naturally susceptible to segregation due to the different particle sizes and specific gravities of its constituents. This natural tendency can be controlled by suitable grading, care in handling, or using admixtures to attain flowability and appropriate viscosity as in SCC.

There are two basic types of segregation. In the first coarser aggregate separates out as they travel further down a slope or settle more than finer particles. In the second form of segregation, which is mostly observed in wet mixes, grout is separated from the mix.

There are different ways to determine the segregation in fresh concrete. The segregation of the aggregates might be monitored during slump flow test. The concrete having proper self-compactability should not exhibit any segregation even at the periphery of slumped material (Beupre et al., 1999). Another common method in testing segregation in hardened concrete is to cut the specimens in two and to observe the cross-section for any apparent segregation of coarse aggregate (Saak et al., 2001).

Using a modified sum of squares approach (Khayat, 1998), mean of coarse aggregate concentration and the segregation coefficient were calculated using the following expressions Eq. 2.6.

G G = i i =1 N

(G G )
i

SC =

i =1

(2.6)

where G , Gi , N, and SC correspond to the mean of coarse aggregate concentration, the concentration of coarse aggregate (percentage of surface area of concrete section), the number of cross-sections (6 in this case) and the segregation coefficient (percent) respectively. It has been reported that typical values of segregation coefficients of a stable, self-compacting concrete can be lower than 7 per cent (RILEM Report on SCC, 2000).

26

zkul et al., investigated segregation using cores taken from different locations of a 1500x200x1000 mm specimen along with visual segregation monitoring during flow tests (zkul et al., 2000). In the study by Khayat et al. segregation was determined by gently pouring a fresh concrete sample from a 2 L container over a 5 mm mesh to observe the quantity of mortar passing through the screen after 5 min (Khayat et al., 2000). The segregation index, SI, is taken as the ratio of the mortar passing through the screen to that contained in the 2 L concrete sample. A stable concrete should exhibit an SI value lower than 5%. Bouzoubaa investigated segregation index of SCC using the same method (Bouzoubaa and Lachemi, 2002) and reported that segregation index of SCC with similar water-to-cementitious materials ratio (0.45) decreases with an increase of the percentage of the fly ash used.

Considering the powder type SCC that contains 40% greater volume of powder compared to ordinary concrete and experiments have shown that, this increase in powder content enhances stability and segregation resistance of the mix (Sonebi et al., 2000).

2.4.4. Mechanical Properties

2.4.4.1. Compressive Strength. When testing the compressive strength of SCC, generally, 150x200 cylinders are tested according to ASTM C 39 or BS 1881: Part 116 B. There are also elastic modulus and core compressive strength evaluations to determine the homogeneity in larger SCC samples.

Working with SCC, Beupre et al., reported that higher strengths (64 MPa) at 28 days (ASTM C 39) were obtained in non-air entrained concrete made with low w/cm (0.35) and silica fume (%3 replacement) (Beupre et al. 1999). According to test core studies by Khayat (Khayat et al. 2001), Zhu (Zhu et al., 2001) and also zkul (zkul et al., 2000) there are no significant difference in the uniformity of in situ properties between the SCC mixes and the corresponding well compacted conventional mixes. However, compressive strength and modulus of elasticity were greater for SCC samples than those obtained from the medium fluidity conventional concrete.

27

In a different approach, Paultre et al. reported that SCC, despite its high deformability, could develop lower in-place (in a reinforced column) compressive strength (10% lower) than strength determined on control cylinders of same concrete (Khayat et al., 2001).

It has been reported that SCC containing up to 50% fly ash (Class F) replacement can attain 28-day target strengths of 35 MPa (Bouzoubaa and Lachemi, 2002). It is also shown by cost analysis that SCC made with 50% of fly ash and with a water-to-cementitious materials ratio of 0.45 can reach a 28-day compressive strength of 35 MPa with no significant extra cost compared to ordinary concrete. Such a SCC would be flowable with a slump flow of 500 mm and flow time of 3 sec, more resistant to segregation and thermal cracking, but might exhibit higher bleeding water and long setting time. It should be noted that, this cost comparison is in contradiction with some research reporting average cost of SCC as 25 to 50 % more than conventional concrete (Khayat et al., 2002).

According to compressive strength studies by Topu et al on SCC, the cementitious factor for fly ash is approximately 0.25 (4 kg of fly ash is equivalent to 1 kg of cement) and that for silica fume is about 2.5 (Topu and Khurana, 2000). Generally, studies with fly ash replacement have shown that it is possible to produce high volume fly ash concrete (5060% cement by weight) with compressive strength values similar or even higher than conventional concrete with the use of proper high range water reducers and air entrainment admixtures

Fang et al. concluded that 20, 30 and 40 % fly ash replacement of cement reduces the early strength of SCC, but the 28 day compressive strength values reaches (40% FA) or even exceeds (2030% FA) the values for corresponding concretes without fly ash (Fang et al., 1999). Jianxiaong et al. found that SCC can be produced with superfine sand (0.63-1.2 FM) and 30-60% slag and fly ash, having 28 days compressive strengths of 50-86 MPa (Jianxiaong, 1999). And Takada et al. (Takada et al., 1998) also found that silica fume addition increased the strength of SCC with an increased demand for SP. On the other hand, fly ash increases workability without additional dosage of SP, but its contribution to strength is limited although increasing with time (Jacobs and Hunkeler, 1999).

28

Perssons study (Persson, 2001), reveals that elastic modulus, creep, and shrinkage of SCC did not differ significantly from the same properties in normal concrete. Other researchers (Gram et al., 2000), (Sonebi and Bartos, 2000) support the finding for modulus of elasticity adding that compressive strength of SCC is better than ordinary concrete both at 1 and 28 days. They related this strength increase to the homogeneity of the SCC with much less voids due to water separation and smaller cement grains forming better packed smaller crystals, rather than big ones. At a given strength the modulus of elasticity of SCC is found to be lower than that of a common concrete due to the smaller maximum grain size and the higher amount of cement paste in SCC (Jacobs and Hunkeler, 1999). Ray and Chattopadhyay carried out studies on the effects of 4, 8, 12, 16% of silica fume by weight of cement on compressive strength. Concretes with a content of 8% silica fume showed the highest compressive strength values after 28 days (45 MPa), followed by concretes having 4, 12, and 16%. Addition of silica fume at all percentages improved the flexural strength, with a significant rise for a 4% SF to 8.5 MPa (Ray and Chattopadhyay 1999). The influence of silica fume on workability and compressive strength of concretes were the major research objectives for Duval and Kadri. Concretes that have been investigated had low water-cement ratios (0.25 to 0.40). The type I Portland cement was replaced by 10-30% by mass silica fume and superplasticizer was added. It was found that silica fume increased best the compressive strength (25%) and the workability of concretes when its content was between 4 and 8 %. Duval and Kadri also found out that if silica fume exceeds 15% of the cementitious material, both compressive and tensile strengths are reduced (Duval and Kadri 1998). Yan et al., studied the hydration process of cement-fly ash pastes with superplasticizer. At low water-binder ratio, the presence of fly ash improved the fluidity of fresh paste. Nagataki et al. measured the strength of fly ash high-performance concrete containing variable amounts of silica fume. Flowability increased with fly ash content whereas strength increased with silica fume content. Pore size distribution was determined to show the positive effect of adding silica fume to fly ash (Hawkins et al., 1998).

29

An extensive survey of the literature conducted by the Portland Cement Association concluded that "in general, the use of up to 5 % limestone does not affect the performance of Portland cement." Even higher contents of ground limestone could potentially be utilized in lower water-to-cement ratio (< 0.45) systems, where a substantial fraction of the cement clinker particles remains unhydrated, effectively acting as a rather expensive filler material (Bentz, 2005) . Compressive strength decreases as limestone is added because the amount of cement in the mixture decreases as the limestone powder increases (Selih et al., 2003). It has been also observed by other researchers that limestone powder reduces compressive strength (Lert, et al. 2000). Limestone accelerates the hydration of C3S to form a calcium silicate hydrate that incorporates CO2 in its structure. When mixed with limestone, the hydration of C3A led to the formation of calcium monocarboaluminate. Study of the mixture C3SC3A-limestone-anhydrous calcium sulfate showed that initially ettringite and a calcium silicate aluminate hydrate gel incorporating CO2 were formed. contribute to the final strength of concrete to a very limited extent. Sprung and Siebel found that the use of inert material as a very fine filler can lead to an increase in strength due to improved packing of the particles (i.e., filling of voids between the cement grains). This effect is seen at early ages, but unlike the case with fly ash or other pozzolanic materials, does not produce additional increases in strength with continued curing. When limestone is included in large quantities (15% to 25%) it acts as a diluent so that strengths are lower than for comparable Portland cement concretes (Sprung and Siebel, 1991). Later, calcium monocarboaluminate was formed (Hawkins et al., 1998). All of these reactions seem to

2.4.4.2. Splitting Tensile Strength. For high performance silica fume concretes it was found that the fracture resistance (toughness and energy) decreased and the brittleness increased with the age of the concrete. This was attributed to the increase in the strength of the cement paste and the interface leading to fewer bond cracking and more aggregate rupture. Hence it appears as the tendency of a higher splitting tensile strength of SCC. Likely as not, the reason for this fact is given by the better microstructure, especially the

30

smaller total porosity and the more even pore size distribution within the interfacial transition zone of SCC. Further a denser cement matrix is present due to the higher content of ultra-fines. SCC can be obtained in such a way, by adding chemical and mineral admixtures, so that its splitting tensile and compressive strengths are higher than those of normal vibrated concrete. An average increase in compressive strength of 60% has been obtained for SCC, whereas 30% was the increase in splitting tensile strength. Also, due to the use of chemical and mineral admixtures, SCC has shown smaller interface micro-cracks than normal concrete, fact which led to a better bonding between aggregate and cement paste and to an increase in splitting tensile and compressive strengths. A measure of the better bonding was the greater percentage of the fractured aggregate in SCC (20-25%) compared to the 10% for normal concrete. Pfeifer (Pfeifer, 1967) studied the effect of natural fine sand, in replacement of the lightweight fines, on seven structural lightweight concrete splitting tensile strengths. The test results showed equal splitting strengths for all continuously moist cured lightweight and normal weight concretes when the compressive strengths were equal. However, the splitting strengths of the lightweight concretes were generally reduced when the cylinders were allowed to dry before testing. The use of fine sand minimized this strength reduction and with the exception of a few lightweight aggregates, the natural sand provided improvements in the workability and finishability of the plastic concrete. By using 6 x 12 inches cylinders for splitting tensile strength tests, Pfeifer discovered that splitting tensile strength of air-dried lightweight concrete generally increased with increasing natural sand content. Studies regarding the effects of curing and drying environments on splitting tensile strength of lightweight and normal weight concretes were carried out by Hanson in two test series (Hanson, 1968). The first series showed that the duration of the initial moist curing period prior to drying at 50 percent relative humidity had little effect on the splitting tensile strength. While there was a loss of splitting strength for the lightweight concrete early in the drying periods, continued storage in the drying atmosphere led to considerable gain in the splitting strengths. In the second series, concretes were subjected to drying for 21 days at different levels of relative humidity after initial moist curing for 7 days. Only minor changes of splitting strength were found as the relative humidity was varied.

31

Investigations involving the splitting tensile strength of very high-strength concrete (fc > 50 MPa) for penetration-resistant structures were carried out by ONeil et al. Due to the addition of silica fume, high-range water reducing admixtures and special curing conditions provided, the tensile strength of the concrete was higher than that of conventional concrete (ONeil et al., 2002). 2.4.4.3. Bond Strength. The resistance of a reinforcing bar against pulling out forces is mainly provided by three factors (elik, 1999). These factors are: - Adhesion between steel and concrete - Friction between steel and concrete - Bearing of the deformation on the steel surface against surrounding concrete. SCC is composed to be flowable enough to enhance capacity of cast members in the absence of internal or external vibration. The high flowability has to be combined with cohesiveness to minimize bleeding, segregation, and settlement of the cast concrete before hardening. Otherwise there is a risk of accumulation of porous cement paste (especially worsened by bleeding) under the lower part of the horizontally embedded reinforcement. This would reduce the bond to the reinforcement. A similar effect will be obtained if the concrete does not have enough deformation capacity to fully encapsulate the reinforcing bars. From studies of bond of reinforcement in experimental walls it is concluded that the bond of horizontal bars is comparable between the tested SSC and a normally vibrated control concrete. It is concluded that the bond in the top of the wall as compared to the bottom of the wall is less reduced for SCC compared with the control. The top-bar effect is thus less marked for the tested SCC. The bond on SCC is equal to or even better than that of normal vibrated concrete (Skarendahl, 1999). Due to the fact that the load bearing capacity of a reinforced concrete structure is considerably influenced by the bond behavior between the reinforcing bars and the concrete investigations on the bond behavior between the re-bars and the SCC were necessary, especially considering the time development of the bond strength. These investigations showed, that the main parameters which influence the bond behavior are the

32

surface of the re-bars, the number of load cycles, the mix design, the direction of concreting, as well as the geometry of the (pull-out) test specimens (Figure 2.10) (Dehn et al., 2002). Experimental results showed higher compressive strengths (36%) and splitting tensile strengths (28%) of the SCC specimens compared to normal concrete specimens. Also, the bond behavior measured at 1, 3, 7 and 28 days after concreting was better for SCC than that of normally vibrated concrete. (Druta, 2003)

14 rebar to be embeded in concrete

Direction of concrete placement

Empty concrete formwork

100 mm 100 mm 150 mm

Figure 2.10 Pullout specimen


2.4.5. Durability Properties

2.4.5.1. Rapid Chloride Permeability (RCP). Chloride permeability is one of the most important durability criteria of concrete. The penetration of chlorides into reinforced concrete accelerates the corrosion of reinforcement and the consequent deterioration of concrete structures located in high chloride concentration environments such as marine structures.

In the study by Tang et al., the chloride diffusivity of SCC was determined using the draft Nordtest method (Tang et al., 1998). In this method an external potential of 10 to 60 volts is applied across a 50 mm thick specimen for 24 hours. After testing the specimen is split in two and the exposed surfaces are sprayed with 0.1 N silver nitrate solution. The chloride penetration depths are then measured with a ruler. Their findings indicate that chloride diffusivity of SCC increases with decreasing w/c ratio and increasing

33

strength, contradicting with common previous data. This is attributed to the unexpected behavior of the unevenly dispersed mineral filler (limestone).

In the test by Jacobs et al., the chloride diffusion coefficients were determined by forcing the chlorides to penetrate into the SCC samples by means of an electrical field and measuring the chloride penetration depths at the end of the test (Jacobs et al., 1999). The results were satisfactory considering the durability criteria, similar to the findings by Trgardh (Trgardh, 1998).

Khayat evaluated the rapid chloride permeability (ASTM C 1202) at 28 to 35 days using samples measuring 50 mm in length and 95 mm in diameter obtained from the centers of 100x200 mm cylinders (Khayat, 2000). His findings indicated that SCC (w/cm: 0.45) incorporating 3% SF and 20% FA performed worse than corresponding mixes with 40% slag powder instead of fly ash. The total charge passed was measured as 2000 coulombs and 880 coulombs respectively.

In another study by Caijun Shi et al., it was also demonstrated that SCC incorporating glass powder has higher chloride migration resistance compared to concrete having fly ash which has relatively lower pozzalan reactivity (Caijun et al., 2005).

Having said all the above it should also be noted that fly ash plays a critical role in transforming larger pores of concrete into small pores and reducing micro cracking in the transition zone. R Pattel et al. supported this fact by showing SCC incorporating 30-60 % FA rated low to very low in RCP tests (Pattel et al., 2004).

Shi et al., pointed out that the test method to rank the chloride penetration resistance by using the rapid chloride permeability test may be not appropriate for concrete made with some mineral admixtures. Test results may be linked to the chemical composition of the pore solution. Electrical conductivity of the pore solution may be modified by the presence of the admixture and therefore is not entirely characteristic of chloride transport.

34

Ramezanianpour et al., compared compressive strength, volume stability, carbonation, corrosion, and chloride diffusivity of concrete. Concrete mixtures containing silica fume, blast-furnace slag, di-atomeous earth or trass, were compared to plain concrete. Samples were subjected to simulated marine environment with wetting and drying cycles and to several other curing modes. Generally, the admixtures improved concrete properties. The best results were obtained for the silica fume concrete (Hawkins et al., 1998).

2.4.5.2. Sulphate Resistance. Sulphate resistance is an important durability property since concrete can be severely damaged by the reactions of water-soluble sulfates. This occurs as the following; first sulfate reacts with Ca(OH)2, commonly known as lime, present in the paste to form calcium sulfate (gypsum), then gypsum reacts with C3A, one of the main constituents of cement, and forms calcium sulphoaluminate. This reaction yields products of greater volume than those of the original reactants, resulting in expansion and finally, this expansion deteriorates the concrete paste (Edamatsu et al., 1999). Yeinobali et al. used both of these methods to evaluate sulphate resistance of mortars containing SF and they also devised their own method and stored specimens (50 mm cubes) in 10% Na2SO4 and 8.4 per cent MgSO4, making both solutions having 67.6 g/l SO4 ion concentration. They evaluated the results by using changes in appearance, mass, and compressive strength at the end of 60 weeks (Yeinobal and Dilek, 1996). The results showed that SF improved durability in Na2SO4 solution, where as it had detrimental effects in MgSO4 solution. This difference between the effects of Na2SO4 and MgSO4 solutions is also observed in other research on mineral admixtures (Cohen et al., 1988), (Hooton et al, 1990). A study on Class F and Class C fly ashes using USBR 4908 have shown that 25-35% Class F FA replacement had positive effects on sulphate resistance, where as Class C FA addition had a highly variable effect (Freeman and Carrasquillo, 1992). There is also evidence that expansion due to sulphate attack increases as the high lime FA content increases. There are some other studies done with new methods, such as using 150x150 cubes in 1-5 per cent Na2SO4 for 360 days and determining expansion and increase in weight (Piasta and Hebda, 1991), or using 100x100x250 prisms in 3- 7 per cent MgSO4 and Na2SO4 and measuring expansion at 28-660 days (Mangat and Khatib, 1995).

35

Ray and Chattopadhyay carried out studies on the effects of 4, 8, 12, and 16 per cent of silica fume by weight of cement on resistance against chemicals (acids and sulfates) of concretes. For testing of resistance against acids and sulfates 50 mm cubic samples were oven dried at 105C and immersed in 2% HNO3, 2% H2SO4, and 5% Na2SO4 solutions for 45 days (Ray and Chattopadhyay, 1999). The weight and strength losses were noted with reference to a set of undisturbed samples cured in water. Conventional concrete exhibited slight bleeding, but this phenomenon was completely eliminated when silica fume has been added in the mixture. Also, the values of air content decreased with the increase in silica fume content. They dropped from 5.5% for normal concrete to 3.5% for 16% of silica fume replacement. As regarding the resistance against acids and sulfates, test results showed that immersion in H2SO4 has caused maximum loss in weight and strength, followed by HNO3 and Na2SO4. The maximum strength loss of 24% has occurred for mixtures without any silica fume, whereas the minimum loss of 12% occurred for mixtures containing 4% of silica fume (Druta, 2003). Some tests have shown improved sulfate resistance for cements containing limestone. This effect appears to be related to the rate of potential deterioration. If the test period is extended, the performance of limestone cements is equivalent to that of the controls. Sulfate resistance is primarily a function of C3A content and water: cement ratio. Locher indicated that the addition of fly ash may increase the sulfate resistance of highly sulfateresistant cements. Gonzlez and Irassar measured the sulfate resistance of type II and type V cements, with 0-20% of limestone filler. Expansion, flexural, and compressive strengths, in addition to XRD and solution consumption measurements were carried out. After one year of exposure, a replacement level of 10% did improve sulfate resistance, while the 20% level was detrimental from hydration characteristics standpoint (Hawkins et al., 1998). SCC with limestone powder exhibited larger increase of mass due to water uptake after curing in distilled water, seawater or sodium sulphate, than NC did. No difference was found between resistance to sulphate attack of concrete with 5% silica fume and concrete without silica fume (Hooton et al, 1990). Concrete with sulphate-resisting cement showed considerably worse durability than concrete with normal Portland cement. The self-compacting concrete with high amount of fly ash showed the best resistance to sulphate attack at pH = 2 (Persson, 2001).

36

Other research has also shown that incorporating silica fume into a concrete mix increases the sulfate attack. It is suggested that this could be due to refined the pore structure of silica fume-incorporated mixes or to the reduction in calcium hydroxide content in the presence of silica fume, which reduces the extent of gypsum formation and hence, increase sulfate resistance. (Mangat et al., 1999) 2.4.5.3. Freeze-Thaw Resistance. Concrete has long been known to be vulnerable to frost attack. As the temperatures of saturated hardened concrete is lowered, the water held in the capillary pores in the cement paste freezes in a manner similar to the freezing of capillaries in rock, and expansion of the concrete takes place. If subsequent thawing is followed by refreezing, further expansion takes place, so that repeated cycles of freezing and thawing have a cumulative effect, and one can envisage an analogy between this and fatigue failure (Batopu, 2000). A comparative study have shown that freezing and thawing resistance of concrete can also be evaluated through freezing and thawing test in the presence of deicer salt (4 percent NaCl solution) (Batopu, 2000). Actually, deicer salt scaling and freezing and thawing tests were conducted simultaneously on the same specimens. In the deicing salt scaling test 100x100x500 mm prismatic specimens were placed in a freezing environment (at -18 3 C) for 16 to 18 hours at selected ages. At the end of this period the specimens were removed from the freezer and placed in the curing room at 23 1.7 C and a relative humidity of 45 to 55% for 6 to 8 hours. These freezing and thawing cycles were evaluated separately as a freezing and thawing test with the above cycle conditions, by comparing the flexural and equivalent cube strengths of specimens after F&T cycle application (max 50 cycles in this case). Many studies on the durability of high volume Class F fly ash concrete have shown that with proper superplastizer addition, air entrainment and/or curing (Batopu, 2000), it is possible to produce concrete with satisfactory durability against freezing and thawing, carbonation, chloride ion penetration, shrinkage and alkali silica reaction (Joshi, 1998). It should be noted that similar durability studies on SCC containing fly ash or other mineral admixtures are limited.

37

Gram et al., claims that frost resistance of SCC is better than ordinary concrete with same water to cement ratio (Gram et al., 2000). The tests were performed according to Swedish standard SS 13 72 44. The salt frost resistance was also found excellent since SCC has more air voids than ordinary concrete even though there is no air entraining additives, the superplastizers cause proper entrainment leading to salt frost-scaling resistance.

In the study by Khayat et al., the freezing thawing resistance of SCC was determined using 75x100x400 mm specimens which were wet cured for 14 days and exposed to 300 cycles of freezing and thawing (ASTM C 666A) and length changes were recorded (Beupre, 1999). The results basically show air entrainment and increased SF dosage (8 to 25 %) increases freezing thawing resistance of SCC, but silica fume inclusion also increases the resistance at low w/cm, even without air entrainment. Another study by Khayat reported that SCC made with 3% SF and 20% FA replacement performed satisfactorily regarding frost resistance (ASTM C 666) (Khayat, 2000).

In the freezing and thawing test applied by Jacobs et al. concrete specimens were saturated with water and then subjected to 10 frost/thaw cycles (cooling down to 25 C in ethylene bath and thawing in water) within 3 days (Jacobs et al.,1999). They determined the deicing salt resistance using the same method, but freezing in 35% CaCl2 solution. The results show that fly ash addition and air-entrainment are necessary for adequate freezethaw resistance.

2.4.5.4. Deicing Salt-Scaling Resistance. In the case of road slabs, frost not only affects the durability of concrete but leads also to the use of deicing salts, which exert an adverse effect on concrete by increasing the severity of the frost action. The salts commonly used are NaCl and CaCl2 and their repeated application with intervening periods of freezing or drying results (Neville, 1986) in surface scaling of the concrete. The salts produce osmotic pressure and cause movement of water toward the top layer of the slab where freezing takes place.

38

A layer of pure ice does not crack, even when held at -18 C for several hours. This may explain why scaling is not serious when pure water is used in the surface layer. However when any amount of salt is present then the ice contains pockets or channels of brine that forms mechanical flaws, and encourage fracture of the layer. Pure ice does not crack in the tests temperature range (down to -20 C) and in high concentration of salt solutions the ice is too weak to exert appreciable stress. But at low, moderate concentration (3% NaCl) cracking of the ice creates tensile stress where the ice joins the surface of the paste and damage may therefore result from the fatigue after many cycles. On the other hand, salt-induced swelling causes cycles of stress that create some micro-cracking, thereby weakening the surface, and may cause buckling of thin layers of paste over aggregate particles (Valenza, et al., 2005) One action of the deicers is to enhance the corrosion of steel (Neville, 1986). The deicer melts the snow or ice, which is often ponded by adjacent ice the resulting liquid is absorbed and, because of its lowered freezing point, remains liquid. As more ice melts, the melt water becomes diluted until its freezing point rises to near the freezing point of water. Freezing then occurs. Thus, freezing and thawing occur as often as without the use of deicers, or even more often since a possibly insulating layer of ice has been destroyed. In consequence, deicers can be said to increase saturation, possibly to increase the number of cycles of freezing and thawing, and to promote corrosion of steel. Although freezing and thawing tests done in accordance with ASTM Standard C 666A, do give indications of scaling resistance, the real test for scaling is that in the presence of deicer salts as suggested by ASTM Standard C 672. Scaling tests can be considered more practical because this problem is much more widespread and more common than the internal micro-cracking generated by rapid freeze-thaw tests (Batopu, 2000). For this reason, tests were conducted in accordance with ASTM C 672 (Standard Test Method for Scaling Resistance of Concrete Surfaces Exposed to Deicing Chemicals). Khayat et al. (Khayat, 2000), (Beupre, 1999) used two 280x230x75 mm specimens to evaluate the deicer salt scaling resistance of SCC. The slabs received 14 days of wet curing, before 50 cycles of freezing and thawing (ASTM C 672). The top surfaces of the slabs were ponded with a solution of 3% NaCl that was changed weekly. In addition to

39

visual rating of the degree of scaling, the mass of scaled particles was determined, and presented as cumulative mass loss. The results revealed that non air-entrained, high w/cm mixtures performed worse than air-entrained specimens, as might be expected. Parallel research (Khayat, 2000) on SCC containing 3%SF and 20%FA or 40%SG have shown that scaling loss of concrete was limited to 0.20 kg/m2, considerable lower than 0.50 kg/m2 maximum limit recommended by Quebec Standards.

Studies on salt scaling and freezing and thawing indicate that it is possible to produce air entrained high fluidity concrete with proper resistance against damage despite the highly fluid nature of the concrete, which can lead to some bleeding, thus making the upper most layer of concrete more susceptible to scaling (Beupre, 1999). Studies on bleeding have shown that bleeding increases with increasing W/ C+FA (fly ash) ratio, but an increase of fly ash from 40 to 60 % replacement did not significantly influence bleeding water of the SCC (Bouzoubaa, 2002).

2.4.5.5. Carbonation. Carbonation in concrete occurs when CO2 present in the atmosphere reacts, in the presence of moisture, with hydrated cement minerals. Ca(OH)2 carbonates to CaCO3, but other cement compounds are also decomposed, hydrated silica, alumina, and ferric oxide being produced (Neville, 1986). One of the most important durability problems in reinforced concrete related to carbonation is the corrosion of the reinforcement. The corrosion of the reinforcing steel is accelerated when concrete carbonates. This is attributed to the reduction in alkalinity (pH dropping from over 12 to 8) of the concrete cover, which protects the reinforcement against corrosion if moisture and oxygen could ingress. Sirivivatnanon and Khatri studied the dosage of fly ash in concrete for different applications. It was found that for above-ground structures, carbonation rate was increased for dosages greater than 40%. Hwang et al. 73 investigated the carbonation of mortars in which part of the fine aggregate is replaced by fly ash. It was found that the carbonation depth decreased with increasing fly ash content (Hawkins et al., 1998).

40

Khayat et al. evaluated the carbonation of concrete with SF and FA with w/c 0.450.80 and cement content 256-456 kg/m3. To measure the depth of carbonation, a slice with a minimum thickness of 20 mm was broken off a 100x100x500 mm prism and phenolphthalein indicator solution was applied according to above procedure up to 3200 days (Khayat et al., 2002). Their results indicated that concretes containing SF and FA are likely to carbonate more when exposed to prolonged hot exposure conditions.

Malami et al. used 0.1% phenolphthalein solution in 95% ethyl alcohol (modified RILEM method), which gave a more intense color of the alkaline area, and the carbonation depth was measured on the next day. They performed the tests at 0, 2, 4, 6, up to 36 months and their results indicated that 4 and 15% FA addition didnt influence the carbonation depth, but higher percentages (30-50%) increased the depth of carbonation (Malami et al., 1994).

Studies by Vennesland and Carette have shown that SF has little effect or might even increase the rate of carbonation especially at high w/cm values. Contrarily, Gneysi reported an increase in carbonation of concrete (10%SF, 15%FA) at lower w/b values. He also added that the lowest degrees of carbonation were observed when SF and FA were used in combination (Gneysi, 1999).

2.4.5.6. Capillary Absorption. Capillary absorption is a criteria commonly used to assess the water permeability of a concrete mix which is a crucial property with regard to overall durability of concrete. The water permeability gives clues about the porosity of the concrete, the degree of connection of the pore structure and concretes potential performance against penetration of aggressive chemicals. 2.4.5.7. Drying Shrinkage. A potentially negative aspect of SCC is shrinkage. Since generally a large amount of fine material is used in the mixtures (particularly those without VMA) and the NMS is limited, the concrete typically has higher shrinkage. Increased shrinkage may result in more cracks in restrained concrete elements, which can accelerate

41

the deterioration of both the concrete and the reinforcement. Shrinkage, after hardening of concrete, is the decrease of concrete volume with time. This decrease is due to changes in the moisture content of the concrete and physicochemical reactions (hydration of cement). Shrinkage is commonly expressed as a dimensionless strain (in./in. or m/m), under steady conditions of relative humidity and temperature. Some studies report a higher drying shrinkage for SCC than for normal concrete while others report the opposite. The same contradictions are found in literature regarding the auto-generous shrinkage, i.e. sometimes it is quoted to be higher for SCC, and sometimes it is quoted to be lower. The choice of constituent materials will to a lesser or bigger extent influences the properties. For instance it has been reported that the use of limestone with suitable fineness can reduce shrinkage of SCC.(Bui et al, 1999)

Since shrinkage in concrete takes place through mechanisms in the concrete paste factors affecting the behavior of the paste also affect shrinkage. It is further pointed out by experimental work (Neville, 1986) that pastes made with both Portland and high-alumina cements, and pure ground calcium monoaluminate exhibit essentially similar shrinkage, the fundamental cause of shrinkage must be sought in the physical structure of the gel rather than in its chemical and mineralogical character. For instance, the specimen size affects the shrinkage rate and larger specimens were observed to exhibit larger shrinkage values due to larger surface for loss of water by evaporation.

Investigations on drying shrinkage of SCC containing 3% SF and 20% FA have shown that the mixes had moderate shrinkage values after 180 days (380-600 m/m ), despite the high water content and low content of coarse aggregate (Khayat, 2000). zkul et al., also investigated shrinkage on 100x100x500 mm prismatic specimens and initial measurements were taken 2 days after casting (zkul et al., 2000). The specimens were kept in a room with a R.H. of 655 and at 2020 C temperature. The results showed that ordinary Portland cement specimens had higher shrinkage than the ones with Portland composite and blended cement. B. Perssons study, reveals that elastic modulus, creep, and shrinkage of SCC did not differ significantly from the same properties

42

in normal concrete (Persson, 2001). Villain et al. studied drying shrinkage of SC mortar and found that shrinkage in the flowing mortar is less than the ordinary one. They attributed this to no segregation and good homogeneity of self-compacting mortar (Villian et al., 1998). On the contrary, Gram et al. found that plastic shrinkage in SC is higher than for ordinary concrete (Gram et al., 2000).

Bouzoubaa et al. studied SCC incorporating fly ash and used 76x102x390 mm prisms to determine drying shrinkage. They stored the specimens in lime saturated water for 7 days prior to transfer to a conditioned chamber at 20 2 C and 50% R.H. The drying shrinkage of the prisms was measured at 7, 14, 28, 56, 112, and 224 days after initial curing according to ASTM C 157. Shrinkage measurements indicated that at 224 days there was no significant difference between SCC and ordinary concrete specimens. Fly ash content didnt cause a difference either, but the increase in W/C+FA increases shrinkage (Bouzoubaa et al., 2002).

In the case of limestone powder, drying shrinkage and carbonation depths show mixed results, in some cases being increased and in some decreased in limestone Portland cements when compared to control Portland cements. The differences are of limited practical significance in the context of overall concrete mix variations (Hawkins et al, 1998). Jianxiong et al. also found that the shrinkage of SCC is similar (4.01x10-4) to or even lower than that of ordinary concrete probably due to lower water-binder ratio and less cement content [18]. Similarly, Sonebi et al. reported that drying shrinkage of ordinary concrete at 28 days was about 25-30% higher than that of SCC (Jianxiong et al., 1999). The length changes due to shrinkage of SCC mixes were found to be higher, compared to common concrete with higher w/c ratios but similar water contents and the difference in shrinkage became less evident with age. Jacobs et al. attributed this initial difference to the more fluid consistency and to the lower modulus of elasticity of SCC (Jacobs and Hunkeler, 1999).

43

3. EXPERIMENTAL PROGRAM AND METHODOLOGY

This study was carried out in order to investigate the effect of binder composition on the fresh, mechanical and durability properties of self-compacting concrete. As mentioned earlier, having a powder content (materials finer than 0.1 mm): 500-600 kg/m3 is critical for attaining self-compactability in concrete. Since the excessive use of cement, as a fine material in concrete is uneconomical (along with other drawbacks as mentioned in above sections), mix-proportions with replacement powders such as silica fume, fly ash, limestone powder is investigated. In this study, such concrete mixes made with changing proportions of powders are tested for their mechanical and durability performances as well as fresh rheological behavior. In the fresh state, mixes were tested according to the self-compactability criteria tests; such as flow Table, U-tube, V-funnel, L-box and visual segregation rating. The mechanical tests consisted of compressive strength, splitting tensile strength, lollipop pullout. Another set of results was obtained from pullout, compression, ultrasound, rebound hammer tests applied on specimens sawn from different sections of 2m high columns (100x200x2000 mm). The cut surfaces of these columns also allowed the investigation of any variation in concrete microstructure with changing depth along the column length. In addition, sorptivity, water absorption, rapid chloride permeability, deicing-salt scaling, sulfate resistance, carbonation, drying shrinkage tests were implemented as durability tests.
3.1.Materials 3.1.1. Cement

The cement used in this study is classified as ASTM Type I Portland cement (PC 42.5). The data for the physical and chemical properties and rate of strength gain for this particular cement are given in Table 3.1. as taken from Nuh imento Sanayi A.. The C3A percentage is known to play a critical role in sulphate resistance of the cement. Here ASTM Standard C150-78a specifies a certain type of cement, Type V, as sulphate resistant cement where the C3A content is limited to 5% and also the total content of C4AF plus

44

twice the C3A content is restricted to 20 percent. For the cement used in this study both of these values are slightly exceeded therefore the cement may not be identified as sulphate resistant. This makes the evaluation of the effects of mineral additives on sulphate resistance more meaningful.
3.1.2. Fly Ash

The fly ash used in this study is classified as Class C according to ASTM C 618. The ash has a lignite coal source used in the Soma Thermal Power Plant in the Aegean region of Turkey and it is this coal source which makes it Class C fly ash with relatively high calcium content and cementations properties. Since the ASTM C618 describes the Class C fly ash as having a CaO content typically more than 10 %, whereas Class F fly ash has less than 10 %. The properties of this fly ash are summarized in Table 3.1.
3.1.3. Silica Fume

The chemical analysis and physically properties of the silica fume used is summarized in Table 3.1. The specific gravity of silica fume is 2.25 g/cm3. The analysis report was taken from Yapkim Yap Kimya Sanayi A..

3.1.4. Limestone Powder

The limestone powder used, had a specific gravity of 2.74 gr/cm3. The analysis report taken from SFALT A.., is presented in Table 3.1.

45

Table 3.1 Physical, chemical and strength properties of cement and powder materials. Portland Cement (PC 42.5) 21.1 4.7 63.71 4.01 0.98 2.57 . . . 0.73 2.08 3.17 3517 26.30 45.60 59.10 54.29 19.62 5.68 12.19

Analysis Report SiO2 (%) Al2O3 (%) CaO (%) Fe2O3 (%) MgO (%) SO3 (%) Cl- (%) Na2O (%) K2O (%) Insoluble Residue (%) Loss on Ignition (%) Specific Gravity (g/cm3) Blaine Fineness (cm2/g) fcc (2 days) (MPa) fcc (7 days) (MPa) fcc (28 days) (MPa) C3S (%) C2S (%) C3A (%) C4AF (%)

Class C Fly Ash 49.10 22.98 13.61 5.41 1.44 1.54 0.02 0.30 1.13 . 2.10 2.40 3928 . . . . . . .

Silica Fume 92.26 0.89 0.49 1.97 0.96 0.33 0.09 0.42 1.31 . . 2.25 221300 . . . . . . .

Limestone Powder 1.97 0.86 47.89 0.84 1.78 . . 0.02 0.01 . 46.13 2.74 3420 . . . . . . .

46

3.1.5. Aggregate

Fine aggregate used in this study is river sand (0-8 mm) and crushed limestone sand. Crushed limestone was chosen as coarse aggregate and used in two separate batches; fine (4-8 mm) and coarse (8-20 mm). Both of these aggregates were sieved, washed and surface dried before use. The maximum particle of aggregates was 20 mm. Sieve analysis and physical properties of the aggregates used in the concrete mix are presented in Table 3.2. The sieve analysis results of the final mix design are given in Figure 3.1. Table 3.2 Sieve analysis and physical properties of the aggregates Sieve size (mm) 31.5 16.0 8.0 4.0 2.0 1.0 0.5 0.25 Passing (%) Crushed Sand Stone Sand 100 100 100 100 100 100 99 19 100 100 100 100 57 40 27 17

Crushed Stone I 100 100 81 23 3 2 2 2

Crushed Stone II 100,0 59.5 1.0 0.5 0.4 0.4 0.4 0.4

47

Aggregate Grading Curve


Percentage Passing
120,00 100,00 80,00 60,00 40,00 20,00 0,00 A32 B32 C32 Agg. Mix

0,25

0,5

16

31,5

Sieve Size, mm

Figure 3.1 Aggregate grading curve and zones (according to TS 706)


3.1.6. Superplasticizer

Modified polycarboxylic ether type; Glenium 51 was used as a liquid superplasticizer. Glenium 51 had a specific gravity of 1.07 g/cm3 and its analysis report, given by Yapkim Yap Kimya Sanayi A.. is shown in Table 3.3.

48

Table 3.3 Analysis report of the superplasticizer (Glenium 51)

Type Color State Specific Gravity (g/cm3) Total Alkali Content (Na2O eqv.) Chloride Content Nitrate Content

Modified Polycarboxylic Ether Dark Brown Homogeneous Liquid 1.07 Max. 1 < 0.1 % None

3.2. Concrete Mixture

During the experiments, a total of eight different concrete mixtures have been produced. All mixes had a total binder powder content of 550 kg/m3. One of these was a control concrete with only Portland cement and no powder replacement. The other seven had varied percentages of fly ash (FA), silica fume (SF) and limestone powder (LP) replacing Portland cement in the mix as presented in Table 3.4. Numerous trial mixes were made in order to determine both the optimum water content and superplasticizer dosage used in the final mixes. During the trials; slump flow spread, segregation, excessive bleeding, and consistency were observed for the mixes. Finally, a constant water-binder ratio of 0.33 is chosen for all of the mixes. Total cementitious material (cement + FA + SF + LP) is kept constant at 550 kg/m3 for all concretes. Glenium 51 was incorporated into the mixture with a dosage of 1.30% of total binder material, to attain a slump flow of 600-700 mm without segregation, obtaining selfcompactability.

49

Table 3.4 Mix proportioning of concretes

Material Cement (kg/m3) Fly Ash (kg/m3) Silica Fume (kg/m3) Limestone (kg/m3) w/b Water (kg/m3) River Sand (kg/m3) Crushed. Sand (kg/m3) Coarse Aggregate 3 (kg/m ) Fine Aggregate (kg/m3) Superplasticizer (% weight of binder) Powder

M1 550 0 0 0 0.33 182 440.6 289.2 374.2 576.2 1.30

M2 350 200 0 0 0.33 182 430.1 282.3 365.3 562.5 1.30

M3 350 150 50 0 0.33 182 428.6 281.3 364.0 560.5 1.30

M4 350 150 25 25 0.33 182 429.9 282.1 365.1 562.1 1.30

M5 350 150 0 50 0.33 182 431.1 282.9 366.1 563.7 1.30

M6 350 0 0 200 0.33 182 434.0 284.8 368.6 567.5 1.30

M7 350 0 50 150 0.33 182 431.5 283.2 366.5 564.3 1.30

M8 350 50 0 150 0.33 182 433.0 284.2 367.8 566.2 1.30

3.3. Production of Test Specimens

Concrete mixtures were produced in batches of approximately 0.04 m3 and powerdriven revolving pan mixer was used for mixing, in accordance with ASTM C 192. Different sizes of specimens were casted from each batch to ensure that the specimens for a specified test are produced from separate batches in order to observe any variations. Casting was done continuously without any vibration being applied. The specimens were kept in laboratory conditions and demolded after 24 hours. After demolding the specimens were subjected to water curing until their test dates.

50

It was observed during trial mixing, that the fresh concrete behavior was clearly affected by the mixing sequence and timing. Therefore a specific mixing procedure (Khayat, 2000) as described earlier in literature review was adopted;

(1) (2)

All concrete mixtures are prepared in 40 L batches in a pan-mixer First, sand and coarse aggregate are mixed (homogenized) for 30 sec. (3) Then 75% mixing water and superplastizer is added and mixed for 30 sec.

(4) (5)

Cement and mineral additives are added and mixing is resumed for 1 min. The remaining water is added and mixed for 3 min. After 2 min at rest the concrete is mixed for a final additional 2 min.

3.4. Test Procedures

In this section, the mechanical and durability tests which were conducted on concrete specimens are described.
3.4.1. Fresh Concrete Properties

There are some special workability tests designed for SCC as mentioned earlier. Such workability tests applied in this study were slump flow, U-tube, V-funnel and L-box. Another fresh concrete property, unit weight was also determined immediately after mixing in accordance with ASTM C 138. For each fresh mix characteristics such as workability, surface finishing, bleeding, segregation and setting time were also noted for making comparisons between mixes. 3.4.1.1.Slump Flow Test. In order to evaluate the consistency of concrete, the method described by the Japan Society of Civil Engineers (JSCE) (see Figure 2.3) is applied. This method judges the capability of concrete to deform under its own weight against the friction of the surface with no other external restraint present. In the slump flow test, measurements were taken at the diameter (D1,D2) of spread along two perpendicular lines.

51

The measurements are evaluated according to the SCC criteria stating that a slump flow ranging from 500 to 700 mm is necessary to define a concrete mix to be self-compacting. 3.4.1.2. U- Tunnel Test. Among several methods designed to measure filling capacity fresh concrete mixtures, the U test being the most common of all, shown in Figure 2.4 was selected. In this test, the height the concrete reaches after flowing through an obstacle was measured as an indication of the degree of compactability. Concrete with the filling height of over 300 mm can be judged as self-compacting.

3.4.1.3. V- Funnel Test. In evaluating the viscosity of the mixtures the V-Funnel test was conducted as shown in Figure 2.5. The flow time is determined using a simple procedure; the funnel is completely filled with fresh concrete, and the flow time is measured as the time between the opening of the orifice and the complete emptying of the funnel. According to SCC criteria, a funnel flow time of 3-12 sec was recommended for SCC. The observations during these tests might also be used for the assessment of the segregation resistance of the mixes. 3.4.1.4. L Box Test. L-box was used for determining flowability and recorded L-flow spread, L-flow slope and flow duration as fresh concrete properties. (Figure 2.6.) It was also observed which mixes exhibit greater flow spread and flatter surface slopes, to conclude that such mixes might be expected to exhibit greater deformability and passing ability despite their similar slump flow values. 3.4.1.5. Segregation. There are different ways to determine the segregation in fresh concrete. In this study, segregation of the aggregates was monitored during slump flow test. The concrete having proper self-compactability should not exhibit any segregation even at the periphery of slumped material. Segregation was also evaluated on cut surfaces of hardened concrete specimens (2m columns). The evaluation of the concrete crosssections at different depths of the columns provided detailed information on the segregation of coarse aggregate.

52

3.4.2. Hardened Concrete Properties

3.4.2.1.Compressive Strength. When testing the compressive strength of SCC, generally, 150x300 cylinders are tested according to ASTM C 39 or BS 1881: Part 116 B. Three 150x300 cylinders for each mix were tested at the ages of 7, 28, and 90 days. There are also elastic modulus and core compressive strength evaluations to determine the homogeneity in larger SCC samples.

3.4.2.2.Splitting Tensile Strength. The splitting tensile strength determination was generally carried out in accordance with the ASTM C 496. Here again, three 150x300 cylinders for each mix were tested at the ages of 7, 28, and 90 days. The tensile strength was calculated by the equation;

ft =

2P DL

(3.1)

where ft is the tensile strength, in MPa; P is the maximum load applied, in N; and D and L the diameter and the length of the cylindrical specimen, in mm. If flexural strength is to be evaluated, the method given in ASTM C 293 is the most common and the flexural strength is calculated using following simple relation;

f =

Mc I

(3.2)

where f is the flexural stress, in MPa; M is the maximum moment ( (PL / 4) in this case), in Nmm; c is the largest distance from the neutral surface ( (h / 2) in this case), in mm; and I is the moment of inertia of the element ( (bh3 / 12) for beams), in mm4. Here b is the width of the specimen and h is the depth of the specimen. If the compressive strength determination is to be conducted on the portions of beams broken in flexure, this can be done in accordance with ASTM C 116 Standard Test Method for Compressive Strength of Concrete Using Portions of Beams Broken in Flexure. So, the

53

equivalent cube strengths of beam portions broken in the above flexural strength test can be determined. 3.4.2.3. Pull-out Test. The pull-out studies were carried out on three specimens for each mix. The specimens are 100x100x150 prisms and are casted with 14-deformed rebar placed horizontally at the centre of the specimen. This horizontal arrangement is aimed to investigate the degree of adhesion between horizontal rebars and SCC.

54

Clamps Pulled out rebar A A Steel plate & plaster Specimen Rigid steel frame

Clamps Section A -A Rebar

Pull-out machine Hole

Steel plate 100.100.8

Figure 3.3. Pull-out Test Set-up The 100x100x150 prisms will be tested according to RILEM standards until failure by using friction grips and a 50 tons capacity load cell using the setup in Figure 3.3. There will be three specimens for each mix tested at 7, 28, and 90 days. The bond behavior was determined under uniform static loading using pullout specimens having a uniform concrete cover around the reinforcing bar. The bar diameter for the whole test series was 10 mm and the concrete cover around it had a diameter of 10 cm and a length of also 10 cm. to avoid an unwanted force transfer between the reinforcing bar and the concrete in the

55

unbonded area, the re-bars were encased with a plastic tube and sealed with a highly elastic silicone material. The re-bars were placed concentrically and the concrete was cast parallel to the loading direction. The tests were carried out in an electro mechanic testing machine where the specimens were loaded with a loading rate of 0,0008 mm/sec. The applied force of the machine was measured corresponding to the slip displacement of the reinforcing bar on the non-loaded side. The increase of the slip path was constantly monitored during the whole testing period. The test was terminated when pull-out failure occurred, the reinforced steel began to yield, or the surrounding concrete cover failed in split. The bond strength was calculated as follows:

P .d b .l s

(3.3)

where P, db, and ls correspond to the ultimate applied load, bar diameter and embedment length (bonded length), respectively.

3.4.2.4. Test on Full Size Structural Elements. These tests were aimed to measure the variation of concrete properties with height. In order to permit the range of comparisons required, two full-size (0.1 x 0,2 x 2.0 m) columns were cast for all mixes. The quality of the in-situ concrete in the 2-metre high columns was assessed in four ways: Core test (10x10 cm cubes cut out at different depths)- assesses in-situ compressive strength Rebound (Schmidt) hammer test - assesses near-surface quality and uniformity Pull-out test (10x10x20 cm regions with horizontally embedded reinforcement cut out to form a pullout specimens) assesses in-situ pull-out strength Image analysis on cross-section- assesses the segregation indices The full-scale SCC columns were cast by pouring the concrete directly from the top of the formwork with a free fall of up to 2 meters through the reinforcement as shown in Figure A.3.

56

Changes in compressive strength, rebound hammer value, ultrasonic pulse velocity, and bond to horizontally embedded reinforcing bars at different depths were also evaluated. Rebound hammer tests were carried out according to the procedures described in Draft European standard prEN12398: Testing concrete Non-destructive testing Determination of rebound number. Testing was carried out at the age of 28 days, and a minimum of 5 readings was taken at each location along the height of columns.

There were 5 embedded bars (14 deformed) in each column and the distances from the bottom of the column to the centre of the bars were varying from 200 to 1800 mm. The centre to centre spacing between adjacent bars was 400 mm as shown in Figure 4. The plywood forms had oiled, smooth interior surfaces and were rigidly supported with embracing wooden bars to prevent any deformation. The casted columns were cured with wet burlap and removed from the formwork after 7 days and cured in lab conditions, for 28 days before sawing operation. Through sawing, 5 pullout specimens (200x100x100 mm) at different depths were obtained for each column as shown in Figure 3.4. Then, a 50 mm slice was removed from the bottom of the pullout specimens and the remaining 150x100x100 mm pullout specimens were tested according to RILEM TC9-RC-1983 standards, same as the lab pullout specimens. The total number of specimens obtained from a column is summarized in Table 3.5. Here, the specimens were labeled as M1.1, the first part (M1) represents the type of the mix and the second figure (1) denotes the distance from the top of column in the direction of concrete placement. So, M1.1 is the upper most specimen of the column casted with Mix 1, where M1.6 is the lower most specimen as shown in Figure 3.4.

The interfacial transition zone (ITZ) between cement paste and aggregate and around steel reinforcement has been an area of particular interest associated with engineering and durability properties of cementitious composites and structural reinforced concrete. As a result of its more porous and heterogeneous/anisotropic nature, the ITZ was recognized to be the weak link in the cement and concrete composites and to have a considerable influence on the engineering and durability properties. However, it has been very difficult to assess the micro-mechanical properties of the ITZ in real concrete specimens, as it required specialized techniques and apparatus. Therefore, the ITZ investigations in this

57

study were performed using a video-microscope. The 50 mm slices removed from the unloaded end of the pullout specimens was used for video-microscope investigation of the steel concrete interface with a magnification of x75 and x150. The formation of defected areas in the interfacial transition zone was investigated. 100x100 mm-cubic specimens were sawn just below each pullout specimen and tested for compressive strength parallel to casting direction. The remaining 200x200x100 mm parts were used for rebound hammer and ultrasonic pulse velocity tests. The 200x100 upper surfaces of these samples were also used in image analysis to evaluate the segregation index. Image analysis was done at cross-sections at heights of 5, 25, 65, 105, 145, and 195 cm from the bottom of the columns. First the photographs of the cross-sections were taken and digitalized. Then, using an image analysis software, ImageJ 1.30v, first the image was converted to grey scale 8-bit then black and white threshold with upper and lower threshold values set to enable particle analysis. The particle analysis gave the results for the percentage of aggregates on the cross-section of the total area. Using a modified sum of squares approach (Khayat, 1998), mean of coarse aggregate concentration and the segregation coefficient were calculated using the expressions (2.4) as mentioned above. It has been reported that typical values of segregation coefficients of a stable, self-compacting concrete can be lower than 7%.(RILEM Report on SCC, 2000)

58

Table 3.5 Classification of the column sections

Distance from the Notation top of the Dimension (cm) column (cm) M 1.1 M 1.1p M 1.2 M 1.2p M 1.3 M 1.3p M 1.4 M 1.4p M 1.5 M 1.5p M 1.6 5-15 15-25 25-35 55-65 65-75 95-105 105-115 135-145 145-155 175-185 185-195 Two 10x10x10 cubes One 10x10x15 pull-out specimen Two 10x10x10 cubes One 10x10x15 pull-out specimen Two 10x10x10 cubes One 10x10x15 pull-out specimen Two 10x10x10 cubes One 10x10x15 pull-out specimen Two 10x10x10 cubes One 10x10x15 pull-out specimen Two 10x10x10 cubes Compression Pull-out Strength Compression Pull-out Strength Compression Pull-out Strength Compression Pull-out Strength Compression Pull-out Strength Compression Tests Applied

59

Figure 3.4 Schematic representation of the column specimen

60

3.4.2.5. Ultrasonic Pulse Velocity Test. The ultra-sonic pulse velocity test was conducted according to ASTM C 597-71. The specimens tested were 100x100x500 prisms at the age of 28 days. These results would also be used as reference results for freezing and thawing, sulfate resistance tests. The ASTM Standards classifies the quality of concrete based on pulse velocity using the following criterion given in Table 3.6. Table 3.6. The criterion for the ultrasonic pulse velocity

Pulse Velocity ( km / s ) > 4.5 3.5-4.5 3.0-3.5 2.0-3.0 < 2.0

Quality of Concrete Excellent Good Doubtful Poor Very Poor

3.4.3. Permeability Tests

3.4.3.1.Water Absorption. Water absorption determination was done on the bottom 75mm sections of 100x200 mm cylinder specimens which had their 50-mm thick middle section removed for chloride ion permeability tests. So, for each concrete mix two cylinder sections cut from two cylinder specimens were tested at the age of 28 and 90 days. Before testing the specimens were oven dried for 48 hrs and weighed, and then they were immersed in water for 48 hrs and weighed. The total water absorbed was expressed as percentage of dry weight. 3.4.3.2. Capillary Absorption. The sorptivity of concrete is a quantity that measures the unsaturated flow of fluid into concrete. Sorptivity is a measure of the capillary forces exerted by the pore structure causing fluids to be drawn in to the body of the material. For one dimensional flow, it can be stated that;

61

i = S .t 1 / 2

(3.5)

where i, S, and t correspond to the cumulative water absorption per unit area of inflow surface, the sorptivity and elapsed time respectively. Capillary absorption (sorptivity) test was carried out using the set up illustrated in Figure 3.5. Water sorptivity determination was conducted on the top 75mm sections of 100x200 mm cylinder specimens at the age of 28 and 90 days. Before testing, the lateral faces of the specimens are coated with paraffin in order to assure that the water was absorbed only through the bottom surface, but not the lateral surface. Then the specimens are placed on rods in a glass container filled with enough water as to cover the 5 mm level above the base of specimens. The initial mass of the sample is taken and at time 0 is immersed to depth of 5 mm in the water. At selected times (typically 1, 2, 4, 9, 16, 25 minutes) the sample is removed from the water, the stopwatch stopped, excess water removed with a damp paper towel and the sample weighed. It is then replaced in the water and the stop watch started again. The gain in mass per unit area over the density of water is plotted versus the square root of the elapsed time. The slope of the line of best fit of these points (ignoring the origin) is reported as the sorptivity.

62

Concrete Specimen

Glass Container Water Level (5 mm above the bottom of the specimen) Water

Pieces of plastic keeping specimens above the bottom of the container

Figure 3.5. - Experimental setup for sorptivity test. 3.4.3.3. Chloride Ion Permeability. In this stage of the study, the chloride permeability of the concrete specimens was determined as a measure of concretes durability performance. For this purpose the rapid chloride permeability method described by AASHTO T227 or ASTM C 1202 was conducted. The experimental setup for the rapid chloride permeability test is given in Figure 3.6. The chloride ion permeability of the specimens was measured at the age of 28 and 90 days. For each batch of concrete, two specimens were tested at the same time for their rapid chloride permeability. There were a total of four cylindrical 100x200 mm specimens cast for each mixture and two of these were tested at 28 and the other two was tested at 90 days. Two 50 mm thick specimens were cut from the midsection of each cylinder. The specimens were then allowed to surface dry in air. In order to prevent evaporation of water from the saturated specimen a rapid setting coating was applied onto the lateral surface of the specimens. Then the specimens are subjected to a vacuum-saturation procedure for 1 hr. As a final stage of conditioning before rapid chloride permeability test, the specimens were kept immersed in water in the curing room at 20 C and 50 % R.H for 18 2 hr. Following these conditioning procedures, the specimens were transferred to a test cell, as shown in Figure 3.3. After the specimens were fixed in position the positive

63

electrode side of the cell (+) was filled with 0.30 N (1.2 percent) NaOH solution, while the negative electrode side (-) was filled with 0.50 N (3 percent) NaCl solution. Then the electrical connections of voltage application and data readout system were made and lead wires were attached to the cell banana posts. Then, a direct voltage of 60.0 0.1 V was applied across the two electrodes. Due to this applied voltage the chloride ions in the NaCl solution, being negatively charged, were attracted by the opposite positive electrode (+) and they penetrate through the pores of saturated concrete. The data acquisition system was adjusted to record the current passing through the specimens with 10 sec intervals over a 6 hours period (21600 sec). This current passing is interpreted as the result of penetration of chloride ions through the concrete. After 6 hours the test was terminated and the result was plotted in a current (amperes) versus time (sec) graph. The area under this curve was integrated to give the ampere-seconds, or coulombs, of charge passed through each concrete specimen during the 6 hours of testing. The total charge passed is evaluated as a measure of electrical conductance of the concrete during the test. A basic classification rating given by AASHTO T 277 (or ASTM 1202) was used in interpreting the chloride permeability of different concretes based on rapid chloride permeability test results. The rating consists of five classes from High to Negligible on the basis of total coulomb value as given in Table 3.7.

64

Table 3.7. Interpretation of results obtained using RCPT T 277-89 Charge Passed (coulombs) >4,000 2,000 4,000 1,000 2,000 100 1,000 < 100 Chloride Permeability High Moderate Low Very Low Negligible Typical of High w/c ratio (< 0.6) conventional portland cement concrete Moderate w/c ratio ( 0.4 0.5 ) conventional portland cement concrete Low w/c ratio (< 0.4) conventional portland cement concrete Latex-modified concrete, Internally sealed concrete Polymer-impregnated concrete, Polymer concrete
HOLE FOR FILLING RECEPTACLE BRASS MESH ELECTRODE BRASS O RING BOLT HOLE

A. ACRYLIC RECEPTACLE
POWER SUPPLY DATA LOGGER

60 V

11

Resistance

0.3 M NaOH SOLUTION NUT

3 % NaCl SOLUTION CONCRETE SPECIMEN BOLT

B. TEST SETUP

Figure 3.6. Experimental set-up for Rapid Chloride Permeability Test

65

3.4.3.4. Carbonation. Carbonation studies were performed on the 100x100x500 prisms. In this study, the measurement of carbonation was determined using the common phenolphthalein test. The test procedure is described in RILEM recommendations CPC-18, Measurement of hardened concrete carbonation depth, 1988. The extent of carbonation was determined by spraying phenolphthalein on a freshly broken concrete surface. Then the free Ca (OH)2 is coloured pink while the carbonated portion is uncoloured. The carbonation depth along the surface is then measured with a steel ruler of 0.5-mm accuracy. There were three specimens for each mix which were kept in lab conditions to enhance carbonation.
3.4.4. Durability Tests

3.4.4.1. Deicing Salt Scaling Resistance. The ASTM C 672 test was conducted on 100x100x500 mm prismatic specimens. A total of four specimens were casted for each concrete mixture. Two of these were exposed to the deicing salt scaling test at the ages of 28 and 90 days, while the remaining two were kept as reference specimens, one for each age. The standard suggests the placement of a dike about 25 mm wide and 20 mm high along the perimeter of the top surface of the specimen. This dike serves to maintain the brine pond (salt solution) on top of the specimen through out the period of the tests. The standard doesnt dictate the type of material to be used in forming the dike. It suggests mortar and epoxy mortar for this purpose, but due to practical reason which makes it necessary for the specimens to be stacked on top of each other in the limited freezing space, a more durable dike material has to be selected. So, fabricated galvanized frames of 50 mm depth were chosen as an appropriate dike material due to their durability and noncorrosiveness. After surface finishing operation, the frames were inserted into the fresh concrete of the top surface of the specimens and the concrete was left to harden. After the relevant curing period (28 and 90 days), the flat surface of the specimen, surrounded by the galvanize frame was covered with approximately 6 mm of a solution of calcium chloride (CaCl) and water, having a concentration such that each 100 ml of solution contains 4 g of anhydrous calcium chloride (0.66 N). Before the placement of the salt solution, the inside perimeter of galvanized frame was coated with a thin line of silicone based paste to prevent

66

the unwanted ingress of water through the interface between the frame and concrete. The drawing of the specimen with galvanized dike is shown in Figure 3.7. Then the specimens were placed in a freezing environment (at -18 3 C) for 16 to 18 hours. At the end of this time the specimens were removed from the freezer and placed in curing room at 23 1.7 C and a relative humidity of 45 to 55% for 6 to 8 hours. Brine was added between each cycle as necessary to maintain the proper depth of solution. This cycle was repeated daily, flushing off the surface thoroughly at the end of each 5 cycles. The scaled off particles were collected and weighed after drying. After making a visual examination by taking photographs and notes, the solution was replaced and the test was continued. The specimens were kept frozen during the inevitable interruptions in the daily cycling. The standard points out that generally 50 cycles are sufficient to evaluate a surface or surface treatment. However, where comparative tests are being made, it is recommended that the test be continued beyond the recommended minimum number of cycles if differences have not developed. In this study, 50 cycles have been found to be more than enough for making comparisons between normal and fly ash concrete, since all specimens apparently reached the final deterioration rating before the 50th cycle. The evaluation of the test results were made with the help of basic rating scale given by the ASTM C 672, shown in Table 3.8. Table 3.8. The rating scale for ASTM C 672

Rating 0 1 2 3 4 5

Condition of Surface No scaling Very slight scaling ( 3.2 mm depth, max, no coarse aggregate visible ) Slight to moderate scaling Moderate scaling ( some coarse aggregate visible ) Moderate to severe scaling Severe scaling ( coarse aggregate visible over entire surface )

67

Salt Solution Pond Silicone Lining Galvanized Dike Concrete Specimen

Figure 3.7. The deicing salt scaling specimen with galvanize dike 3.4.4.2. Sulphate Resistance. In the USBR 4908 Sulphate Exposure Test 76x152 mm cylinders are used with about one-third of a 1.9 cm stainless steel gage stud protruding from each end. The gage studs allow the monitoring of specimens for linear expansion. The specimens are continuously soaked in 10% Na2SO4 solution and changes in length and mass are monitored for 360 days. There are similar methods devised in literature and the specimens used in this study were 100x100x250 prisms in 10 % Na2SO4 and measuring expansion, compressive strength loss at 3, 15, and 30 months. The solutions are renewed at specific time intervals when expansion measurements were made. There were two 100x100x250 prisms for each mix. Since, the loss in strength is a quantity of primary interest and is one important measure of degree of attack, compressive strength values of reference specimens stored in water are compared to the strength values of specimens which were subjected to long-term sulfate attack. Then the percentage damage, D is calculated using the following relation:

68

D = 1 R .100

(3.6)

where R and correspond to strength after exposure to sulphate environment for period of T and strength in the absence of sulfate environment after time T respectively.
3.4.5. Drying Shrinkage

In BS 812 standards the specimens are 100 x 100 x 500 mm prisms and kept at 20 2
0

C and 50 5 % relative humidity during measurements of length change. The Swedish

standards SS 13 72 15 includes the use of 100 x 100 x 400 mm prisms stored under the same ambient conditions. According to ASTM C 157 drying shrinkage test is performed on 100x100x375 mm specimens. The specimens are demolded after one day and stored in lime-saturated water until 28 days. They are then transferred to a curing room with 50% R.H. and a 2020 C temperature and length changes are monitored at specific intervals. In this study, the drying shrinkage was tested as the length change of 100x100x250 mm beams. The beams were cured in water for six days from the time of demolding (24 hours after casting) and then exposed to drying at 50 % RH and 202 0C. The measurements were done with an extensometer and with measuring points of steel studs fixed at the two ends of the beams. The scale corresponds to 0.005 mm (i.e. 0.01 ). The tests were carried out at 7, 14, 28, 56, 112, 224 days to monitor the length changes of two prisms for each mix.

69

4. TEST RESULTS AND EVALUATION

4.1.

Fresh Concrete Properties

The fresh properties of the designed SCC mixes are summarized in Table 4.1. And some of typical SCC acceptance criteria used in Japan and Europe and SSC classes defined by EN 206-1:2000 (EFNARC, 2002) are given in Table 4.2.

Table 4.1. Properties of fresh concrete

Mixture Type Slump Flow (mm) T500 (sec) Tf (sec) U-tube (mm) V-funnel (sec) L-box, H2/H1 (%) Segregation resistance Bleeding (visual) Theor. (kg/m3) Unit Weight

M1 737 5 47 340 12 95 good none 2419 2400

M2 750 8 42 350 10 97 good none 2379 2335

M3 665 5 39 340 5 96 good none 2373 2310

M4 700 6 50 330 9 94 good none 2377 2330

M5 700 6 42 330 8 95 good none 2382 2340

M6 600 8 55 350 20 93 good slight 2393 2350

M7 525 good slight 2384 2329

M8 550 good slight 2390 2353

Exp. Unit Weight (kg/m3)

70

Table 4.2. General acceptance criteria for SCC and EN 206-1:2000 classes

Typical range of values and SCC classes Test Method Slump Flow (mm) L-box, H2/H1 V-funnel (sec) T500 (sec) Unit min. mm ratio sec sec 500-650 (SF 1) 0.8
8 (VF1) 2 (VS1)

max. 650-750 (SF2) 1.0 9 27 (VF2) > 2 (VS2) 760-850 (SF 3)

Comparing the results with the SCC criteria, it can be seen that all designed mixes exhibited satisfactory properties as fresh concrete. However, the properties from four measurements such as slump flow, U-tube, V-funnel and L-box are not very consistent. M2 showed highest slump flowability (750 mm), but had relatively higher V-funnel flow time (10 s). The results for the U-tube and L-box tests were the most comparable probably due to the similar nature of the test setups but the V-funnel results were generally inconsistent. In fact, the V-funnel test can evaluate the viscosity in relative terms only under the condition of the constant slump flow value. In such a case, a longer flow time represents a higher viscosity of the mixture and it directly relates to a better resistance to segregation. (Skarendahl et al., 1999). Therefore, there are still differences in the results of these tests, even though they are all used for testing the flowability or passing ability of SCC concrete mixtures. The mixes, are classified under consistency classes SF (by slump flow values) and viscosity classes as VS (by T500 value) and VF (by V-funnel time) the results are given in Table 4.3. The slump flow test photographs are presented in Figure A.1.

71

Table 4.3 Consistency and viscosity classes of the mixes

Mix M1 M2 M3 M4 M5 M6 M7 M8

SF Class SF2 SF2 SF2 SF2 SF2 SF1 SF1 SF1

VS Class VS2 VS2 VS2 VS2 VS2 VS2 . .

VF Class VF2 VF2 VF1 VF2 VF1 VF2 . .

In the slump results, there is an evident decrease in flowability where SF is added in M3, M4 and M7. This might be explained by the increased surface area of the SF particles increasing the water demand. Since, the water demand and workability are controlled by particle shape, particle size distribution, particle packing effect and the smoothness of surface texture. It is also known that at low water-binder ratio, the presence of fly ash improved the fluidity of fresh paste. The spherical shape of FA also minimizes the particles surface to volume ratio, resulting in low fluid demands. It is also reported that a higher packing density was obtained with spherical shapes as compared to crushed particles like, limestone powder, in a wet state. The reduction in workability for the limestone powder replacement might be due to the nodular and angular shape of the limestone particles which hinder the workability of the paste. This explains the relatively negative effect of limestone powder on the fresh concrete properties. The mix with only FA replacement performed better probably because of the common explanation of fly ash having a special shape effect due to its spherical particles which easily roll over one another, reducing inter-particle friction and therefore increasing workability.. The combined effect of the different mineral admixtures should be explained

72

along with resultant packing density. It was explained that, at an optimal particle size, the packing density was maximum, which help to achieve maximum flowabilty. When handling the mixes it was observed that they had consistency that resembled putty and all were very cohesive, but were remarkably flowable. Visual observations indicated all these concrete mixtures showed good segregation resistance and none of the SCC mixes exhibited any bleed water on the surface except M6, and to some extend M7 and M8. The slump flow photographs are given in Figure A.1.
4.2. Hardened Concrete Properties 4.2.1. Mechanical Properties

4.2.1.1.Compressive Strength. The compressive strength results of the concrete mixtures with different cement replacements at different ages are given in Table 4.3. and Figures 4.1.-4.4. Test were performed at 7, 28 and 90 days, but the compressive strength results of the control specimens for the sulphate resistance tests were also included in the compressive strength development graph to demonstrate the ultimate strength gain at a later age (30 months). The SCC mixtures had different amounts for cement replacement. The M1 mixture was a control mix with 100% Portland cement, where as M2 had fly ash (FA) replacing 36% of cement by mass. M3 has 27% FA and 10% silica fume (SF) replacement and M4 had 27% FA, 5% SF and 5% limestone powder (LP). M5 had 27% FA, 10% LP and M6 had 36% LP replacement. M7 had 27% LP, 10% SF and M8 had 27% LP, 10% FA replacement of cement. If a general evaluation is made, the results at 7 days shows that the compressive strength for the 100% cement mixture is higher than the mineral admixture replacement mixtures indicating that the pozzolanic reactions of FA and SF are not sufficient to enhance compressive strength at this early stage. This agrees with the conclusion; fly ash usually reduces the strength of concrete at early ages. But at 28 days the slower pozzolanic reactions come into play and the mix with %27 FA and 10% SF (M3) has the highest compressive strength results as expected. The compressive strength results of M3 are 20% higher when compared to M1. FA replacement alone in M2 causes a 7% increase in 28 days compressive strength. Here the role of limestone powder is also better understood as it only acts like inert filler reducing the compressive strength of the

73

paste in M6, M7 and M8 where M6 has the lowest strength and highest limestone powder replacement of 36%. The strength reduction relative to the control mixture M1 was 17%, 7%, and 6% for M6, M7 and M8 respectively The SF and FA additions should be accounted for the relative strength increases for M7 and M8. At the final age of 90 days, M3 still has the highest compressive strength value with a 10% increase in strength compared to M1. Once again it was observed that, SF with its high SiO2 content and extreme fineness enhances the compressive strength through the pozzolanic reactions and also by the physical effect of filling the voids in the microstructure of concrete paste. FA alone (M2) is slightly better than M1, having a 2% relative increase in compressive strength. Here again the limestone powder replacement has negative effects on compressive strength causing 18%, 13% and 16% relative decreases in M6, M7, and M8, respectively. The addition of limestone powder generally reduces the compressive strength of the hardened concrete. This might be explained due to the fact that the increase of the amount of limestone powder dilutes the pozzolanic reaction. And consequently compressive strength decreases as limestone powder is added because the amount of cement in the mixture decreases as the inert limestone powder increases. The strength development depends primarily on the formation of hydrated calcium silicates as the main hydration product, which is precipitated into the water filled spaces to form a more compact body. The crystallization of the initially formed hydrates, having strong binding forces, and their transformation into other hydration products having weaker binding forces affect the strength (Heikal et al., 2000). Although limestone powder addition is assumed to have accelerating effects on C3S hydration and its slow conversion to a monocarboaluminate phase was included in the reactions in cement paste. (Gibbs et al, 1999) the overall effect of limestone powder replacement of cement on compressive strength is negative. In the case of SF and FA, filling of the voids between the larger cement particles, and increasing production of secondary hydrates by pozzolanic reactions with the lime resulting from the primary hydration enhances compressive strength. This effect is more pronounced for silica fume which has much larger specific surface and pozzolanic reactivity compared to fly ash. It is also worth nothing that the fly ash used in this study

74

was a Class C fly ash with a CaO content of just over 10% to meet the Class C criteria than a Class F fly ash with dominating pozzolanic character. 4.2.1.2.Splitting Tensile Strength. The splitting tensile strength results are presented in Table 4.4. and Figures 4.5.-4.8. . For easy comparison, the tensile/compressive strength ratios are also included. The 7 day splitting tensile strength results show that M1 with Portland cement only has the highest value where as M5, M6, M7, M8 are in the lower strength group. This might be due to rather high limestone powder content of these mixes and insufficient pozzolanic reactions in the case of FA and SF being present. The strength gain at 28 days for the mixes with mineral admixtures results in higher strengths compared to M1 for all mixes except M5, M6, M7, and M8. Comparing all the concrete mixes, the splitting tensile strength at 28 days was in the order of M3 > M2 > M4 > M1 > M5 > M8 > M7 > M6. Here again the M6 mix performs worst since it has 36% limestone powder replacing cement. In the case of 90 days, it is observed that the mixes containing FA and SF exhibited a steady increase in splitting tensile strength due to their prolonged pozzolanic reactions. Again, comparing the mixes, splitting tensile strength at 90 days was in the order of M3 > M2 > M4 > M1 > M5 > M7 > M8 > M6. But this time the relative loss of splitting tensile strength due to limestone powder was more pronounced. When the splitting tensile strength values are compared with the respective compressive strength values for all mixes, it is observed that the fs/fc ratio, as presented in Table 4.4., ranged between 0.07-0.09 showing a good correlation between the corresponding strengths. Generally it was observed that SF had positive effect on splitting tensile strength. This was attributed to the increase in the strength of the cement paste and the interface leading to fewer bond cracking and more aggregate rupture and eventually to a higher splitting tensile strength of concrete. The reason for this fact might be explained by the better microstructure, especially the smaller total porosity and the more even pore size distribution within the interfacial transition zone of concrete which has a denser cement matrix due to the higher fineness of SF. The mixes with SF and FA probably showed smaller interface micro-cracks leading to a better bonding between aggregate and cement paste and to an increase in splitting tensile strengths. A measure of the better bonding is the greater percentage of the fractured aggregate in SCC (20-25%) compared to the 10% for normal concrete (Druta, 2003). The same visual observations of the splitting specimens

75

supported this finding and the broken surfaces had more aggregate failure for the higher splitting tensile strength values. Unfortunately, there are no presentable photographs that clearly show this effect of aggregate failure that was noted at the time of the experiments. The results showed that, the replacement of cement with limestone powder in large amounts (much more that 5%, used in practice) reduces the splitting tensile strength. This is attributed to the formation of a weaker bond in the interaction zone of the aggregate and cement matrix. The very fine LSP has the function of filler and creates a denser matrix compared with a matrix of normal strength concrete and a less dense matrix in comparison to a SF matrix. The weaker bond in the transition zone allows cracks around the aggregates and results in lower splitting tensile strength. 4.2.3. Pull-out Test. The pull-out studies revealed the results as shown in Table 4.4. and presented in Figures 4.9.-4.12. The results of the pull-out strength tests indicate two groups of similar strength concretes for all ages. The first group is the concretes with higher pullout strength consisting of M1, M2, M3, M4 and the second group with considerably lower strength at 7 days but slightly lower strength at 90 days is the M5, M6, M7, M8 concretes characterized by the incorporation of limestone powder. Even though the results show a similar trend as the compressive strength and splitting tensile strength tests, the differences due to use of different binder compositions are much less distinguishable. At 90 days M2 (36% FA) and M4 (27% FA, 5% SF, 5% LSP) performed better than any of the other mixes and the pullout strength at 90 days was in the order of M2 > M4 > M3 > M1 > M5 > M6 > M7 > M8. This was not expected because M3 had the best mechanical performance in the compression and splitting tests because of the higher SF content in 27% FA concrete. It is known that SF reduces the wall effect around the aggregate and reinforcement, thus allowing better packing of cement grains at the interface of the cement paste with aggregate or reinforcement. The limestone powder has lowering effect on pullout strength, due to same explanations for the reduced splitting strength, basically because of the weaker interfacial zone between the cement paste and the reinforcing steel bar. 4.2.1.4. Ultrasonic Pulse Velocity Test. The ultra-sonic pulse velocity tests conducted on 100x100x500 prisms at the age of 28 and 90 days and exhibited the results in Table 4.6. and Figure 4.13. All the specimens performed well in the UPV test and were classified as

76

good quality concrete. At 28 days the ultrasonic pulse velocities are in the order of M1>M3> M2>M5> M4>M6. This is reasonable when compared to the compressive strength values of the same mixes at 28 days. At 90 days the order becomes M3>M1> M2>M5> M4>M6 showing an increase in UPV of the M3 mix which is attributed to the prolonged reaction of SF and FA. The results show good correlation with other mechanical properties, especially the compressive strength and it is assumed that same reasons for better or worse performance of the mixes apply for ultrasonic pulse velocity. Simply, the denser the cement paste the higher the ultrasonic pulse velocity, hence higher the compressive strength. This property of UPV makes it a valuable tool in assessment of hardened concrete compressive strength in a non-destructive and simple manner. 4.2.1.5. Test on Full Size Structural Elements In order to determine the variation of mechanical properties and homogeneity, there were numerous tests applied on the columns sections and the results are summarized in Table 4.7 and presented in Figures 4.14.-4.17. The photographic views of the formwork and columns specimen are presented in Figure A. 2. The compressive strength results are given in Table 4.7. There is no significant variance among the results of the same mix at different depths of the column. The obvious trend is that, the bottom specimens have higher strength values than the top specimens, probably due to higher degree of compaction due to the weight of the concrete above. This also shows that the segregation has not occurred while concrete is being poured down a 2 m high formwork, at any extent to lower the compressive strength. This indicates that the mixes had adequate stability. Same arguments apply to pullout strength data shown in Table 4.7. The pullout strength shows a steady increase as the depth of the concrete above the specimen increases. Rebound hammer and ultrasonic pulse velocity tests have the same trends as presented in Table 4.8. Generally the near surface quality of the concrete, expressed as Rebound hammer values, increased along the height of the columns towards the bottom. The relations between the properties of the columns specimens and the corresponding lab specimens at the age of 28 days are summarized in Table 4.7.. In-situ strength achieved in the columns for all SCC mixes was in the range of 75-85 % of the 28 day strength for standard cylinder specimens. The ultrasonic pulse velocity readings were even

77

closer, being 80-100% of the values for lab specimens. The pull-out strength achieved in the columns for all SCC mixes was in the range of 70-97%. The differences among the various mixes were again parallel with the observations made in the laboratory specimens. The columns specimens have lowered compressive and pullout strength values than the laboratory specimens and this might be explained by the relatively favorable curing conditions of the laboratory specimens. The use of limestone filler especially alone is observed to cause an overall decrease in mechanical properties, whereas fly ash and silica fume have positive effects. In general the bond strength is dependent on the position of the reinforcing bar in the formwork (top bar effect), and the height of fresh concrete above and beyond the bar during casting. But the SCC mixes evaluated in this study shows a perfect filling ability around in the formwork and around the reinforcing bars and no sign of segregation. Therefore, the difference between good and poor bond conditions are lower, the top bar effect is less distinctive. The excellent filling effect of the SCC mixes were also observed at the finished surfaces of the columns when the formwork was removed, the surface were perfectly smooth, without any honeycombing or air voids. The image analysis of the columns cross-sections revealed the coarse aggregate distribution at different heights of the columns as shown in Table 4.9.-4.15. Using the equation (3.4) the area ratios which are equal to the area of the coarse aggregates divided by the total area of the cross-section are converted into segregation coefficients and given in Table 4.16. During the image analysis of the pullout strength specimens, the steelconcrete interface was investigated on the 50mm thick slices sawn from the back of the specimens. The steel concrete interface was analyzed by a video-microscope with a magnification of x75 and x150. Figure A.7 and A.8 shows the video-microscope images of the interface between the concrete and steel reinforcement. These investigations showed that there were no significant defects on the interfacial transition zone. The defects were expected to form due to the rising bleeding water or air voids getting trapped under the embedded reinforcement bars. But it seems the SCC mixes had enough deformation capacity to fully encapsulate the

78

reinforcing bars without any segregation or bleeding. The image analysis of the pullout specimens also pointed out the coarse aggregate distribution throughout the height of the columns and the relation between the embedded bar. The coarse aggregates were found to be evenly distributed along the height and surrounding the rebars without any uneven accumulation. The Tables 4.17-4.22 shows the results of image analysis done on the pullout specimens. These results show that all the mixes had high homogeneity, stability and low segregation. Again the use of limestone filler lowers stability values whereas silica fume and fly ash increases the segregation resistance of the mix.

79

Table 4.3. Compressive strength results at 7, 28 and 90 days Comp. Strength, 7 days (MPa) 44.50 38.50 41.40 35.40 38.40 30.50 31.48 35.92 Comp. Strength, 28 days (MPa) 50.70 54.00 60.40 49.10 48.60 42.20 47.29 47.53 Comp. Strength, 90 days (MPa) 57.20 58.10 63.10 54.20 54.20 46.70 49.58 47.72

Type of Concrete M1 M2 M3 M4 M5 M6 M7 M8

Table 4.4. Splitting tensile strength results at 7, 28 and 90 days Splitting Tensile Type of fs / fc Strength, ratio, Concrete 7 days 7 days (MPa) M1 M2 M3 M4 M5 M6 M7 M8 3.30 3.29 3.13 2.89 2.60 2.63 2.81 2.31 0.07 0.09 0.08 0.08 0.07 0.09 0.09 0.06 Splitting Tensile fs / fc Strength, ratio, 28 days 28 days (MPa) 3.32 3.62 3.86 3.49 3.27 3.13 3.20 3.27 0.07 0.07 0.06 0.07 0.07 0.07 0.07 0.07 Splitting Tensile fs / fc Strength, ratio, 90 days 90 days (MPa) 4.77 4.95 5.24 4.88 4.59 3.88 3.93 3.55 0.08 0.09 0.08 0.09 0.08 0.08 0.08 0.07

80

Table 4.5. Pull-out strength results at 7, 28 and 90 days Pull-out Type of Strength, fp / fc Concrete 7 days ratio, (MPa) 7 days M1 M2 M3 M4 M5 M6 M7 M8 8.88 9.56 9.50 8.60 5.92 6.40 6.14 6.44 0.20 0.25 0.23 0.24 0.15 0.21 0.20 0.18 Pull-out Strength, fp / fc 28 days ratio, (MPa) 28 days 10.00 11.04 10.76 10.74 9.68 9.08 8.45 8.92 0.20 0.20 0.18 0.22 0.20 0.22 0.18 0.19 Pull-out Strength, fp / fc 90 days ratio, (MPa) 90 days 10.40 11.12 10.68 11.04 10.08 9.64 9.22 9.14 0.18 0.19 0.17 0.20 0.19 0.21 0.19 0.19

Table 4.6. Ultrasonic pulse velocity test results

Type Concrete M1 M2 M3 M4 M5 M6

of UPV, 28 days (km/sec) 4.39 4.30 4.35 4.26 4.28 4.17

UPV, 90 days (km/sec) 4.42 4.35 4.45 4.28 4.31 4.19

Concrete quality Good Good Good Good Good Good

81

Table 4.7. Test results of column sections at 28 days

Specimen (increasing depth) M1.1 M1.2 M1.3 M1.4 M1.5 M1.6

Comp. Equivalent Strength of Comp. Strength (100x100) (150x300) cube cylinder (MPa) (MPa) 47.40 59.37 60.99 57.94 58.03 58.70 35.27 47.10 48.38 45.59 45.66 46.57

Pullout Strength (MPa) 8.64 9.48 9.20 10.48 10.92

Schmitt Rebound Hammer

Ultrasonic Pulse Velocity (km/sec)

36 39 40 40

3.37 3.30 3.62 3.73

M2.1 M2.2 M2.3 M2.4 M2.5 M2.6

51.81 57.15 55.12 60.82 56.78 59.82

39.79 44.97 43.02 48.25 44.68 47.46

8.88 9.56 10.36 10.52 11.56 33 34 34 35 3.82 4.46 4.42 4.47

M3.1 M3.2 M3.3 M3.4 M3.5 M3.6

54.54 60.41 56.68 58.19 58.66 59.09

42.56 47.92 44.60 45.78 46.54 46.88

7.72 8.20 9.24 8.04 10.52 38 38 40 42 4.37 4.37 4.38 4.47

82

Table 4.7. (cont.) Test results of column sections at 28 days

Specimen

Comp. Equivalent Strength of Comp. Strength (100x100) (150x300) cylinder cube (MPa) (MPa) 42.44 50.24 49.98 49.99 49.94 52.23 32.59 38.27 47.58 37.78 37.75 40.11

Pullout Strength (MPa) 4.12 7.16 4.52 8.48 9.80

Schmitt Rebound Hammer

Ultrasonic Pulse Velocity (km/sec)

M4.1 M4.2 M4.3 M4.4 M4.5 M4.6

32 35 35 37

3.64 3.53 3.53 3.57

M5.1 M5.2 M5.3 M5.4 M5.5 M5.6

48.82 49.49 48.94 55.34 57.45 54.33

36.62 37.41 36.70 43.19 45.21 42.40

5.64 6.32 9.92 11.24 11.32 32 32 32 33 3.48 3.50 3.51 3.51

M6.1 M6.2 M6.3 M6.4 M6.5 M6.6

41.17 37.58 43.75 43.23 43.69 43.40

31.62 29.57 33.33 32.94 33.28 33.07

6.44 7.80 7.56 6.76 6.36 31 32 32 29 3.43 3.41 3.44 3.44

83

Figure 4.8 Correlation between the properties of column and lab specimens

Mix Average Compressive Strength columns specimens (MPa) (C) of

M1

M2

M3

M4

M5

M6

44.76 44.70 45.71 39.01 40.26 32.30

Compressive Strength of lab specimens 50.70 54.00 60.40 49.10 48.60 42.20 (MPa) (L) Percentage (%) (C/L) 88 83 76 79 7.49 83 8.89 77 7.01 9.08 77 3.43 4.17 82

Average Pullout Strength of column 9.74 specimens (MPa) (C) Pullout Strength of lab specimens (MPa) (L) Percentage (%) (C/L)

10.18 8.74

10.00 11.04 10.76 10.74 9.68 97 92 4.29 4.30 100 81 4.40 4.35 100 70 3.57 4.26 84 92 3.50 4.28 82

Average Ultrasonic Pulse Velocity of 3.51 columns specimens (km/sec) (C) Ultrasonic Pulse Velocity specimens (km/sec) (L) Percentage (%) (C/L) of lab 4.39 80

84

Compressive Strength, fc (MPa)

70 60 50 40 30 20 10 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.1. Compressive strength test results at 28 days

Compressive Strength, fc (MPa)

70 60 50 40 30 20 10 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.2. Compressive strength test results at 90 days

85

Compressive Strength, fc (MPa)

70 60 50 40 30 20 10 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.3. Compressive strength test results at 90 days

Compressive Strenght, fc (MPa)

70 60 50 40 30 20 0 200 400 600 800 1000 Time (days)


Figure 4.4. Compressive strength developments at 7, 28, 90, and 900 days

m1 m2 m3 m4 m5 m6

86

Splitting Tensile Strength, fs (MPa)

6 5 4 3 2 1 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.5. Splitting tensile strength test results at 7 days


Splitting Tensile Strength, fs (MPa) 6 5 4 3 2 1 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.6. Splitting tensile strength test results at 28 days

87

6 Splitting Tensile Strength, fs (MPa) 5 4 3 2 1 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.7. Splitting tensile strength test results at 90 days

Splitting Tensile Strenght, fc (MPa)

6 5 5 4 4 3 3 2 0 20 40 60 80 100 Time (days) m1 m2 m3 m4 m5 m6 m7 m8

Figure 4.8. Splitting tensile strength developments at 7, 28, and 90 days

88

14 Pull-out Strength (MPa) 12 10 8 6 4 2 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.9. Pull-out strength test results at 7 days


14 Pull-out Strength (MPa) 12 10 8 6 4 2 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.10. Pull-out strength test results at 28 days

89

14 Pull-out Strength (MPa) 12 10 8 6 4 2 0 m1 m2 m3 m4 m5 m6 m7 m8 Concrete Mixes

Figure 4.11. Pull-out strength test results at 90 days

14 Pull-out Strenght (MPa) m1 12 10 8 6 4 0 20 40 60 80 100 Time (days)


Figure 4.12. Pull-out strength developments at 7, 28, and 90 days

m2 m3 m4 m5 m6 m7 m8

90

4,5 Ultrasonic pulse velocity (km/s) 4,4 4,3 4,2 4,1 4 m1 m2 m3 m4 m5 m6 Concrete Mixes 28 days 90 days

Figure 4.13. Ultrasonic pulse velocity results at 28 and 90 days

Figure 4.14. Results of pullout strength of lab specimens compared with column specimens ( in the column series 1 is the upper most specimens and 5 is the specimens at the bottom of the column)

91

Figure 4.15. Results of compressive strength of lab specimens compared with column specimens (in the column series 1 is the upper most specimens and 5 is the specimens at the bottom of the column)

92

Figure 4.16. Results of Schmitt rebound hammer values changing with depth of column. ( in the column series 1 is the upper most specimens and 4 is the specimens at the bottom of the column)

93

Figure 4.17. Results of ultrasonic pulse velocity test changing with depth of column. ( in the column series 1 is the upper most specimens and 4 is the specimens at the bottom of the column)

94

95

Table 4.9. Different stages of image analysis in calculating the area ratio for column M1

96

Table 4.10. Different stages of image analysis in calculating the area ratio for column M2

97

Table 4.11. Different stages of image analysis in calculating the area ratio for column M3

98

Table 4.12. Different stages of image analysis in calculating the area ratio for column M4

99

Table 4.13. Different stages of image analysis in calculating the area ratio for column M5

100

Table 4.14. Different stages of image analysis in calculating the area ratio for column M6

101

Table 4.15. Area ratios of concrete column cross-sections found by image analysis and used in segregation coefficient calculations Area Ratio= Areas of Coarse Aggregate / Total Area of Cross-section Specimen location SC1 top SC2 SC3 SC4 SC5 SC6 bottom M1 0.32 0.36 0.32 0.30 0.35 0.36 M2 0.28 0.36 0.34 0.29 0.30 0.33 M3 0.37 0.32 0.34 0.33 0.32 0.33 M4 0.30 0.33 0.36 0.32 0.30 0.35 M5 0.26 0.34 0.26 0.25 0.33 0.35 M6 0.27 0.33 0.30 0.33 0.29 0.37

Table 4.16. Segregation coefficients of the concrete mixtures casted in columns

Type of Concrete M1 M2 M3 M4 M5 M6

Segregation coefficients (%) 2.29 2.71 1.61 2.31 3.25 4.22

102

Table 4.17. Stages of image analysis at the steel-concrete interface for column M1

103

Table 4.18. Stages of image analysis at the steel-concrete interface for column M2

104

Table 4.19. Stages of image analysis at the steel-concrete interface for column M3

105

Table 4.20. Stages of image analysis at the steel-concrete interface for column M4

106

Table 4.21. Stages of image analysis at the steel-concrete interface for column M5

107

Table 4.22. Stages of image analysis at the steel-concrete interface for column M6

108

4.2.2. Permeability Tests

4.2.2.1.Water Absorption. Water absorption determination was done on the bottom 75mm sections of 100x200 mm cylindrical specimens and the results are presented in Table 4.23. Comparing all the concrete mixes, the rate of water absorption was in the order of M6 > M8 > M7 > M5 > M2 > M4 > M1 > M3 at the age of 28 days and M6 > M8 > M7 > M5 > M4 > M1> M2 > M3 at the age of 90 days. The results suggest that M6, M7 and M8 with high volume LSP has higher water absorption and M3 (27% FA, 10% SF) exhibits the lowest water absorption at all ages. Such results suggest that the near-surface concrete of M3 was denser and more resistant to fluid ingress than the corresponding limestone powder or FA alone mixes. Although limestone powder was assumed to have a filling effect, those mixes incorporating high volumes of LSP, exhibits higher water absorption than M1, containing cement alone. This is likely to be due to the high volume of limestone powder hindering the overall cementitious reactions, hence creating a more porous microstructure. On the contrary, the increase of cement hydration products owing to SF and FA and other factors, such as the better dispersion of cement and powder particles and formation of denser interfacial transition zone also contribute to the lower near-surface absorption of M3, M2.and M4. In general, any factor tending to improve the compressive strength of the concrete will have a beneficial effect upon the water-tightness. Therefore a similar trend in the compressive strength is observed for water absorption results for different binder compositions. 4.2.2.2. Sorptivity. Sorptivity (capillary absorption) test was conducted on the top 75mm sections of 100x200 mm cylinder specimens at the age of 28 and 90 days. The results are shown in Table 4.24. When the results are compared for all the concrete mixes, the capillary absorption expressed as sorptivity was in the order of M6 > M7 > M8 > M5 > M2 > M4 > M1 > M3 at the age of 28 days and M6 > M8 > M7 > M2 > M5 > M1> M4 > M3 at the end of 90 days. The results suggest that M6 has highest sorptivity values in both ages. These findings are in good agreement with the water absorption results and suggest that SF and FA decreases sorptivity whereas LSP increases it. It is noted that, permeability is related to pore structure. Pore size is less important than the connectivity of the pore system. So even though limestone powder doesnt help to create smaller volume pores it might have reduced the permeability by reducing the connectivity of the pores. But looking

109

at the results of this study, the use of high volume limestone powder seems not to reduce the connectivity of the pores. It is worth noting that the effect of the pozzolanic reactions would have been more pronounced if the concretes were subjected to better curing conditions. It was shown especially for FA that, the high temperature curing decreased the porosity of fly ash concrete as compared to low and standard temperature curing as applied in this case. (Batopu, 2000) 4.2.2.3. Chloride Ion Permeability. The chloride ion permeability test was conducted according to RCPT set-up and the results are found to be as shown in Table 4.25 and Figure A.5. The results show that M3 has the lowest chloride permeability, rated low at 28 days and very low at 90 days. The total charge passed is 668 coulombs for M3 whereas the value reached 3251 coulombs for M6, 2603 coulombs for M7, and 2542 coulombs for M8. The second best result regarding the resistance to chloride ion permeability was for M4. The 28 day value was 1868 coulombs which decreased down to 1004 coulombs at 90 days. Generally, there is a decrease in total charge passed with age which is quite expectable since the chloride permeability is governed by the pore system of the concrete, which improves with age. So, here again the formation of a less porous, denser microstructure and a discontinuous pore system becomes critical for reduced chloride ion permeability. The pozzolanic reactions seem to be able to develop a discontinuous pore system more readily. The pronounced reduction in pore size in pastes containing silica fume is attributed to both its high pozzolanic reactivity, and also to its very small particle size which allows it to pack efficiently between the cement grains, thereby subdividing the space. The combined effect of FA and SF (M3) seems to be improving the total chloride permeability resistance of the concrete mainly due to the better packing of the fines in the cement paste. In the case of limestone powder, the result is a porous microstructure increasing the chloride permeability same as water permeability. 4.2.2.4. Carbonation. The depths of carbonation of concretes exposed to outdoor environment are presented in Table 4.27.and Figure A.6. The results show that the carbonation was very limited for all concrete mixtures. In the first stage of the test the specimens were broken at the end of 3 months and phenolphthalein was sprayed on the

110

freshly broken surface showing no carbonation for all of the mixes. After 15 months the procedure was repeated and M6 was found to have very slight (less than 0.5 mm) carbonation on the periphery of the specimen. It was found that, in general, concretes made with Portland cement with limestone powder replacement (6 to 20% limestone) showed increased rates of carbonation as compared with those made with Portland cement only. The final results after 30 months indicated that limestone powder increased the carbonation of the concrete even though at a very minimum extent. So the M6 has 0.5 mm of carbonation depth, followed by M5 (27% FA, 10% LSP) and M2 (36% FA) with less than 0,5 mm of carbonation depth. The SF an FA addition reduced the outdoor environment carbonation to zero for the mixes, M1, M3, and M4.
4.2.3. Durability Tests

4.2.3.1. Deicing Salt Scaling Resistance. The visual rating results of the deicing salt scaling tests are presented in Table 4.26 and in Figure A.3 and A.4. Deicing salt scaling resistance test applied on the specimens at 28 days showed that generally the specimens had variable deicing salt scaling resistance. Some of them were extremely resistant (M3 > M1 > M2 > M5 >M4) even without air entrainment. This is important since the testing method devised by ASTM is known to be very severe, and the fact that no air entrainment was applied increases the severity of the testing conditions. Even though the tests conditions were harsh, the final stages of the M3, M1, M2, M5, M4 specimens were classified as no scaling or moderate to low scaling at the end of 50 cycles of freezing and thawing. The explanation of this observation might be the low water-binder ratio (0.33) or the superior surface finish of SCC, without bleeding or segregation. The use of SF and FA creates a denser paste and SF especially reduces bleeding, leading to greater strength in the superficial layer of paste and thereby reduces damage. It has been reported that, at about w/b = 0.38 the salt frost scaling of concrete with 40% fly ash and 5% silica fume (5.2% air content) was 20 times that of Portland cement concrete. (Bertil, 2002)

In the case of M6, M7 and M8 it seems that the excessive use limestone powder had detrimental effects on deicing salt scaling resistance. At the 5th rating of the visual rating scale of the test, all the coarse aggregates become visible over the entire surface, indicating

111

severe scaling. The M6, M7 specimens reached the 5th rating in the test at 40th cycle whereas M8 reached that rating at the 30th cycle making it more vulnerable to deicing salt scaling. It is concluded that the excessive use of limestone powder was insufficient to form a durable superficial layer on the top surface of the specimen. 4.2.3.2. Sulphate Resistance.. The effects of sulfate attack were evaluated as loss in compressive strength and linear expansion as shown in Table 4.28 and Figure 4.18 and Figure A.6.. Generally the specimens exhibited expectable sulphate resistance, with limited loss in compressive strength (1-7 % loss) and linear expansion (0.01-0.22 %). SF again exhibited positive effects on sulphate resistance. The positive influence of silica fume is also evident from the visual observations (see photo) indicating less surface damage for the specimen containing SF (M3, M4) where M6 specimen containing 36% LSP had some surface damage. In the Na2SO4 solution, no mass losses were observed in any of the specimens, and deterioration could only be characterized by reduction in strength and increase in length. This suggests that the mechanisms of deterioration are associated with expansion and weakening of the paste structure, but not with the hydrated material splitting off the surface of the concrete. FA and SF refine the pore structure and reduce the calcium hydroxide content which reduces the extent of gypsum formation and hence, increase sulfate resistance. So, concrete incorporating SF and FA has a compressive strength loss less than concrete without SF or FA when subjected to 5% Na2SO4 solution. When the linear expansions are compared M1 specimens have relatively the highest linear expansion although small in magnitude. M6 and M5 have similar linear expansions slightly lower than M1. The lowest expansion is observed for M3 followed by M2 and M4. The linear expansion results are in good agreement with the loss of compressive strength results. The small amount of linear expansion implies a less damaged concrete paste, hence a higher compressive strength. It is also important to note that, the C3A content of the parent cement determined the sulfate resistance. When pozzolans are used as partial cement replacement, they decrease the amount of C3A and react with calcium hydroxide to produce new cementing compounds filling the pores in the paste. As the amount of cement replacement with mineral additives increase strain values of the specimens decrease for all mineral additive type except limestone powder. (Dilek et al. , 2003)

112

4.2.4. Drying Shrinkage

In this study, 100 x 100 x 250 mm prisms were be kept at 50% R.H. and a 2020 C temperature and length changes are monitored at 7, 14, 28, 56, 112, 224 days. The results are shown in Figure 4.19. The results suggest that lowest shrinkage is observed for M6, the high volume LSP concrete and highest shrinkage is observed for FA- SF concrete M3. Shrinkage values for other mixes is in the order of M4 > M2 >M1 > M5 >M8 > M7 but they are very much similar. Here again the inclusion of LSP reduces shrinkage due to its lower fineness and less denser paste. This agrees with the results of previous research concluding that the use of limestone powder decreases shrinkage because of the fact that the inert limestone powder reduces the total amount of cement paste to be formed and consequently reduces drying shrinkage. For instance it has been reported that the use of limestone with suitable fineness can reduce shrinkage of SCC. (Bui et al., 1999) For the evaluation of the effect of SF and FA, the results indicate that the addition of silica fume increased drying shrinkage within the first 2 to 3 weeks of drying. Since shrinkage is the result of water consumption during the hydration process, when SF replaces part of cement, its high specific surface creates high water demand and the available water is consumed at an early age of hydration. More over the additional products of hydration by SF, reduces the space between the particles. Therefore the total porosity is reduced and the very dense paste matrix has a discontinuous pore structure of a very small pore size. On the contrary FA consumed water more slowly and results in lower self-desiccations and consequently lower shrinkage.

113

4.3. Cost Analysis of SCC

A primary concern for the wide use of SCC in the construction industry, especially in Turkey, is the cost. This is worth mentioning, since currently the commercially available SCC supplied by a few ready mix concrete manufactures costs twice as much as ordinary concrete. For instance, the price of C25 ready mix concrete is approximately 80 YTL/m3 where as the price goes up to 92 or 94 YTL for C45 and C50. The official price (MSB Poz 324) for self-compacting concrete given by the Ministry of Defense is 172 YTL/m3. To study the relative costs of the materials used, the prices (as of December 2005) of the materials in Marmara region were considered and presented in Table 4.29. The cost analysis show that the prices of the SCC mixes are comparable to commercial ready mix concretes C45 and C50 of similar strength. And there is a significant decrease in cost as the mineral admixtures especially limestone powder replaces some of the cement. This reduction in cost is down to 16% when comparing M1 to M6 which also has a 16% relative reduction in 28 days compressive strength. On the other hand, M2 and M3 have lower costs and higher strength values than M1 with Portland cement alone. It is also observed that the costs of SCC for all mixes are much less than the actual market price. Therefore it is concluded that SCC prices might compete with ordinary concrete prices with the proper material selection. Eventually, the actual case in the construction industry also implies the same finding. In a study focusing on the analysis of 11 years of case studies concerning SCC usage in large quantities, it has been pointed out that among 51 case studies (dating from 1993 to 2003), in most cases the powder contained more than 30% of mineral addition. And it is not surprising that limestone was the most often used, in 28 cases (Donome, 2005). It is also evaluated that raw material costs alone should not be considered the criteria for acceptance or rejection of the system. All costs should be included to form an overall cost for 1 m3 of concrete such as mould and assembly, casting and compaction, stripping, finishing, vibration. Even the lack of vibration and finishing workmanship might greatly reduce the overall SCC cost compared to ordinary concrete. There are obviously other aspects affecting the overall cost such as reducing construction time and labor, filling the forms with heavy reinforcement, lowering heat of hydration, enhancing final mechanical

114

and durability properties, etc. to be considered while using SCC with mineral admixtures in a specific civil engineering project. As a conclusion; from holistic considerations, SCC is surely more cost-effective than ordinary concrete. Table 4.23. Water absorption test results at 28 and 90 days Water Absorption (%) at 28 days 2.68 3.47 2.58 3.26 4.28 5.50 5.28 5.34 Water Absorption (%) at 90 days 2.55 2.48 1.23 3.07 3.91 5.03 4.85 4.96

Type of Concrete M1 M2 M3 M4 M5 M6 M7 M8

Table 4.24. Sorptivity test results of the concrete mixtures at the age of 28 and 90 days Sorptivity value at 28 days (mm/min1/2 ) 0.066 0.091 0.046 0.079 0.104 0.131 0.121 0.117 Sorptivity value at 90 days (mm/min1/2 ) 0.046 0.078 0.023 0.026 0.056 0.095 0.085 0.088

Mixture Type M1 M2 M3 M4 M5 M6 M7 M8

115

Table 4.25. Rapid chloride permeability test results at 28 and 90 days Total Charge Chloride of Passed Permeability at 28 days Rating (coulombs) 3283 3244 1086 1868 2759 3689 2654 2760 moderate moderate Low Low moderate moderate moderate moderate Total Charge Chloride Passed Permeability at 90 days Rating (coulombs) 2929 2859 668 1004 2280 3251 2603 2542 moderate moderate very low low moderate moderate moderate moderate

Type Concrete M1 M2 M3 M4 M5 M6 M7 M8

Table 4.26 Visual rating results of the deicing salt scaling tests on concrete specimens

Visual rating results after every 5 cycles until 50 cycles of F-T.

Type of 5 Concrete M1 M2 M3 M4 M5 M6 M7 M8 0 0 0 0 0 0-1 0-1 0-1

10 0 0 0 0 0 1-2 1-2 1-2

15 0 0-1 0 0-1 0 2-3 1-2 2-3

20 0 0-1 0 0-1 0-1 3 2-3 3-4

25 0 1 0 1 0-1 3 3 4-5

30 0 1 0 1-2 1 4 3-4 5

35 0 1 0 1-2 1-2 4 4 5

40 0 1 0 1-2 1-2 5 5 5

45 0 1 0 2 1-2 5 5 5

50 0 1 0 2 2 5 5 5

116

Table 4.27. Depth of carbonation of concretes exposed to outdoor environment

Type Concrete M1 M2 M3 M4 M5 M6

of

Carbonation depth. mm Exposure period, 3 months 0 0 0 0 0 0 Exposure period, 15 months 0 0 0 0 0 less than 0.5 Exposure period, 30 months 0 less than 0.5 0 0 less than 0.5 0.5

Table 4.28. Compressive strength of concretes stored in water and in sulfate solution

Type Concrete M1 M2 M3 M4 M5 M6

of

Compressive Strength, Compressive months (MPa) Strength, 28 days in 10 (MPa) in water Na2SO4 50.70 54.00 60.40 49.10 48.60 42.20 54.41 55.06 63.33 54.15 54.65 47.87 50.96 54.49 62.21 53.55 51.15 54.49

30 Percentage damage, % D 6.30 1.00 1.80 1.40 6.40 4.60

Linear expansion, %

0,00

0,05

0,10

0,15

0,20

0,25

0,30

Figure 4.18. Linear expansion versus time for concretes exposed to sulphate attack.

Table 4.29. Cost analysis of SCC mixes used in this study

Material

Unit Price (YTL/ ton)

200
Immersion time in sulfate solution, days

M1

M2

M3 31.50 2.25 4.50 0.00 0.27 7.70 4.50 2.91 4.48 38.90 97.01 60.40

M4 31.50 2.25 2.25 0.18 0.27 7.74 4.51 2.92 4.50 38.90 95.02 49.10

M5 31.50 2.25 0.00 0.35 0.27 7.76 4.53 2.93 4.51 38.90 92.99 48.60

M6 31.50 0.00 0.00 1.40 0.27 7.81 4.54 2.94 4.54 38.90 91.91 42.20

M7 31.50 0.00 4.50 1.05 0.27 7.76 4.53 2.93 4.51 38.90 95.94 47.29

M8 31.50 0.75 0.00 1.05 0.27 7.79 4.54 2.94 4.53 38.90 92.27 47.53

Cement 90.00 Fly Ash 15.00 Silica Fume 90.00 Limestone Powder 7.00 Water 1.50 River Sand 18.00 Crushed Sand 16.00 Fine Aggregate 8.00 Coarse Aggregate 8.00 Superplastizer 5440.00 Total Cost (YTL/m3) fc 28 days (MPa)

49.50 0.00 0.00 0.00 0.27 7.92 4.62 2.99 4.61 38.90 108.81 50.70

31.50 3.00 0.00 0.00 0.27 7.74 4.51 2.92 4.50 38.90 93.34 54.00

400 600 800 1000

m6

m5

m4

m3

m2

m1

117

118

0,30 0,25
Linear expansion, % m1 m2 m3 m4 m5 m6

0,20 0,15 0,10 0,05 0,00 0 200 400 600 800 1000
Immersion time in sulfate solution, days

Figure 4.18. Linear expansion versus time for concretes exposed to sulphate attack

1000 800 Shrinkage (10E-6)

M1
600 400 200 0 0 20 40 60 80 100 120 140 Age (days) 160 180 200 220 240

M2 M3 M4 M5 M6 M7 M8

Figure 4.19. Drying shrinkage test results

119

4.4. Analytical Correlations between Concrete Properties

In order to evaluate the relationships between various concrete properties, the scatter plots of their values were produced. Then trend lines were drawn showing where the data points cluster around a line. These lines all appeared linear indicating a linear association between the relevant properties. These lines also formed a basis for the evaluation of a regression model and quantitative relations between designated properties. Here the strength of these relationships was quantified by coefficient of determination, R2, a value which is probably the most popular measure of fit in statistical modeling. R2 is a measure that can be computed for a fitted model, takes values between 0 and 1, becomes larger as the model fits better, and provides a simple and clear interpretation. In statistics, the R2 is the proportion of a sample variance response variable that is explained by the predictor (explanatory) variables when a linear regression is done. The formula for R2 is

R2 =

E SS R = 1 SS TSS TSS

(4.1.)

where ESS is the explained sum of squares, RSS is the residual sum of squares, and TSS is the total sum of squares. It is obvious that, as R2 approaches unity as a regression approaches a perfect fit. (i.e., R2= 1- sum ((data regression)2)) / sum ((data datamean)2)) The correlation coefficient R, which is the square root of R2, can assume values between -1 and 1. To interpret the direction of the relationship between variables, one looks at the signs (plus or minus) of the regression. The sign of R (+ or -) is the same as the sign of the slope of the trend line. When R has a positive value the proposed relation is a direct relation and when R has a negative value the relation is an inverse relation.

120

4.4.1. Relationship between Water Absorption and Sorptivity (S)

Sorptivity coefficient is an important measure of durability and is frequently related to other properties of concrete (Gneysi, 2004). In this study sorptivity coefficients of SCC with different amounts of cement replacement admixtures were evaluated along with water absorption values in percentage. The plots of them are illustrated in Figure 4.20. For the different SCC mixes, the following relationships between the sorptivity and water absorption values at 28 and 90 days were presented in Equation 4.1 and 4.2, respectively. The analysis also showed that the water absorption and sorptivity results are linearly correlated with R2 values of 0.93 and 0.63 for 28 and 90 days, respectively. So the higher the water absorption the higher is the water sorptivity, which is reasonable since both of these values are directly related to and affected by the pore structure and density of the resulting concrete. S28 = 0.023 WA28 - 0.002 S90 = 0.016 WA90+ 0.006 (4.1) (4.2)

The plots also showed that the sorptivity value is more age dependent that water absorption. This is attributed to the fact that water absorption is a mechanism governed by the larger pores in the concrete where as sorptivity is effected by the continuity of the smaller capillary pore structure. So as time goes on from 28 to 90 days the prolonged cementitious reactions induced by mineral admixtures (SF, FA) still have positive effects on the sorptivity of the concrete as they block more of the capillary pores.
4.4.2. Relationship between Total Charge Passed (Q) and Sorptivity (S)

The total charge passed (Q) is determined by the rapid chloride permeability test and is an indication of the resistance of the concrete to aggressive chemical attack. Water sorptivity can also be regarded as a measure of concrete durability since it is also a measure of the concrete permeability. Since both of these values are set based on similar mechanisms, their interdependence is worth studying. The plots of their relation at 28 and 90 days are given in Figure 4. 21. The linear relation between these values is given as the Equations 4.3 and 4.4. The graphs also showed that the R2 values range between 0.40 and

121

0.69 and the data obtained at 28 days was more scattered compared to 90 days maybe because of the fact that some of the admixtures had prolonged pozzolanic reactions making up for the initial deficiencies in time. S28 = 1.8x10E4 C28 + 960 S90 = 2.7x10E4 C90 + 577 (4.3) (4.4)

The plots indicate once more, that total charge passed and sorptivity decrease with age of concrete with similar age dependencies. This is due to the fact that both of these properties are closely related to the capillary pore structure of the concrete.
4.4.3. Relationship between Total Charge Passed (Q) and Water Absorption

In order to asses the interdependence of the total charge passed (Q) and water absorption similar methods were used as sorptivity. And it was found that total charge passed as measured in rapid chloride permeability test is not as strongly correlated with water absorption when compared to sorptivity. This supports the fact that chloride permeability is affected by the capillary porosity rather than the overall porosity of concrete. The relation between total charge passed (Q) and water absorption is depicted in Figure 4.22. There is again a directly linear correlation but there is a large scatter and the R2 values are as low as 0.22 and 0.34.
4.4.4. Relationship between Total Charge Passed (Q) and Initial Current (IC)

The initial current readings at the start of the 6 hour RCPT test are given in Table 4.30 and the plots of initial current versus total charge passed are shown in Figures 4.234.24. The linear regression analysis for the results of 28 days, 90 days and combined, showed that there exists a strong linear correlation between initial current readings and total charge passed. The R2 values for these relations are 0.90, 0.89 and 0.88. The following expressions, Equation 4.5- 4.7 represent the relationships; Q = 21902 IC + 421 Q = 29322 IC 158 for 28 days for 90 days (4.5) (4.6)

122

Q = 25043 IC + 156

for all data

(4.7)

These results are in agreement with the results of other research studies given in literature (Gneysi, 2004). In past studies the correlation between the initial current and the total charge has been studied with a variety of mixtures with or without mineral admixtures. Nevertheless, the number of studies on this topic is still limited to be able to establish an empirical relationship between initial current (IC) and total charge passed (Q). It is obvious that the assessment of the chloride permeability of concrete would be much easier if the initial current readings were to substitute for the results of the 6 hours of RCPT. This method would be preferable not only for being faster and simpler, but also due to the fact that it does not change the nature of the specimens or reduce the repeatability of the test. Since, during the 6 hour RCPT, the changes in the concrete temperature, ambient temperature, and pore structure as charge flows and accidental leakages of the electrolyte cells significantly affects the total charge passed at the end of the test. Table 4.30. Initial current readings of RCPT tests Type of Mix M1 M2 M3 M4 M5 M6 M7 M8 Initial Current, IC (mA) 28 days 139.0 127.2 40.9 63.9 121.3 139.7 103.6 85.0 90 days 105.6 119.6 26.1 38.3 107.3 123.5 93.7 98.0

123

4.4.5.

Relationship between Total Charge Passed (Q) and Compressive Strength (fc)

In order to quantify the interdependence between the two properties of concrete, the RCPT chloride penetration resistance versus compressive strength relation is plotted as given in Figure 4.25. The analysis of the data clearly shows that there is an inverse relation between the two data sets. When the compressive strength for a specific concrete is low, then the total charge passed is high. This indicates that the factors making the concrete more permeable to chloride diffusion also hinders the adequate strength development in the concrete. Although there is an obvious relation, the R2 values varied between 0.46 and 0.32 implying that there is a poor linear relationship between the charge passed and the compressive strength due to the wide scatter of the data. This also agrees with the fact that not many efforts have been made in literature to relate coulombs from RCPT test to the corresponding compressive strength (Gneysi, 2004).
4.4.6. Relationships between Splitting Tensile Strength, Compressive Strength and Pull-out Strength

The plots of the relation between splitting tensile strength and compressive strength are given Figure 4.26. The analysis clearly shows a strong correlation between these results at 7, 28 and 90 days. Here again the early results of 7 days show relatively more scattering. The relations are given in equation form as in Equation 4.8- 4.11. f s7 = 0.045 f c7 + 1.19 f s28 = 0.043 f c 28 + 1.24 f s 90 = 0.101 f c 90 0.98
f s.all = 1.293.e 0.02 f c.. all

(4.8) (4.9) (4.10)


(4.11)

The analysis also showed that the splitting tensile strength and compressive strength results are linearly correlated with R2 values of 0.37, 0.89 and 0.88 for 7, 28 and 90 days respectively. Here the increase for splitting and compressive strength have similar trends for 7 and 28 days, but for 90 days it seems that splitting tensile strength values are more

124

effected by the increasing age. Thus, an exponential type of relation as given in Eq (4.11) may be given to fit the whole data with R2 = 0.76. The relation between splitting tensile strength and pull-out strength was also studied. This was particularly interesting since all the pull-out failures occurred as the samples split into three of two pieces. The relation between pull-out strength and splitting tensile strength is given in Figure 4 27. The R2 values are 0.73, 0.69 and 0.85 for 7, 28 and 90 days. The coefficients indicate a stronger correlation even at the early stages when compared to the relation between compressive strength and splitting tensile strength, but surprisingly the 28 and 90 days correlation is higher for compressive strength splitting tensile strength relation. The linear regression expressions are given below in Eq. 4.12, 4.13, and 4.14 where as a logarithmic type of relation as given in Eq. 4.15 is observed to fit all data best with R2 = 0.67. f p7 = 3.84 f s7 3.34 f p 28 = 3.28 f s 28 1.29 f p 90 = 1.18 f s 90 + 4.91 f p all = 5.97 Ln (f s all ) + 1.75 (4.12) (4.13) (4.14) (4.15)

In order to evaluate the same kind of relation between pull-out strength and compressive strength, the following plots were generated in Figure 4.28. Here again there is a directly linear relation between the two properties but the lower coefficients of correlation indicate a weaker relation than the ones among the above mentioned properties. The magnitudes of R2 for the relation between pull-out strength and compressive strength at 7, 28 and 90 days are 0.42, 0.51, and 0.61. Again a logarithmic relation seems to represent the relation between pull-out strength and compressive strength for the whole data set with R2 = 0.73. Similar behavior as for the splitting tensile and compressive strength relation is observed for the relation between pull-out and splitting tensile strength, the values increase with a similar trend from 7 to 28 days, but for 90 days the splitting tensile values have a slightly better increase with increasing age. This might be explained by fact that prolonged

125

cementitious reactions improve the strength of the cement-aggregate interface, causing a more pronounced increase in splitting tensile strength.
4.4.7. Relationships between Compressive Strength, Pull-out Strength and

Ultrasonic Pulse Velocity of Column and Lab Specimens

One of the aims of this study was the verification of properties of hardened SCC in full-size columns using in situ test methods. This allowed the comparison of the lab specimens casted and cured in relatively favorable conditions with the simulated real life experience in 2 m columns. The relationship between the in-situ properties of specimens cut from full size structural elements and laboratory specimens were to be evaluated. Therefore the following graphs were drawn as given in Figure 4.29 4.31. The relation between the compressive strength of column specimens and lab specimens showed a general trend of linear correlation. The R2 value of this relation was 0.78 and the relation is equation form was; f c column = 0.74 f c lab + 3.44 (4 .16)

This implies that the concrete compressive strength properties show a significant degree of variance among lab and in-situ concrete samples. The column samples have strength values of nearly 75% of the strength values of corresponding lab specimens. This shows that even though SCC has relatively stable properties; the strength values might show some variance according to different conditions of dimension, formwork, concrete placement and curing. Analyzing the pull-out strength relationship between column and lab specimens, it is easily seen that the data has greater scattering and a lower coefficient of correlation, R2 being equal to 0.22. This might be due to the fact that pull-out strength is more sensitive to changes in the compaction of the cement paste around the reinforcing bars in column specimens. When there is even a slight random defect in the interface between the concrete and steel, this greatly reduces the pull-out strength of the specimen. The effects of these random defects might be reduced by increasing the number of test specimens in each group.

126

Ultrasonic pulse velocity values of column specimens were lower than the velocities obtained in lab specimens. Moreover, the in-situ ultrasonic pulse velocities changed within a larger range compared to the lab specimens. The R2 value is 0.74 and the regression equation is given in Eq. 4.15. UPV column = 9.04 UPV lab 34.70 (4.17)

4.4.8. Relationship between Compressive Strength and Ultrasonic Pulse Velocity

The plots of the relation between compressive strength and ultrasonic pulse velocity are given Figure 4.32. The analysis clearly shows a linear correlation between these results at 28 and 90 days. The R2 values are 0.54 and 0.89, respectively. Here again the earlier results of 28 days show relatively more scattering. The relations might be given in equation form as in Equation 4.16- 4.17. f c 28 = 58.5 UPV 28 - 200.4 f c 90 = 53.8 UPV 90 - 177.5 (4.18) (4.19)

The results show that ultrasonic pulse velocity is a reasonable measure of compressive strength and it is assumed that same reasons for better or worse performance of the mixes apply for ultrasonic pulse velocity. Simply, the denser the cement paste the higher the ultrasonic pulse velocity, hence higher the compressive strength. This property of UPV makes it a valuable tool in assessment of hardened concrete compressive strength in a non-destructive and simple manner.

127

4.4.9. Relationship between In-situ Mechanical Properties and Distance from the Top of the Columns

It has been observed that some mechanical properties such as compressive strength or pull-out strength (e.g. top-bar effect) increases as we go down from the top to the bottom of a vertical concrete element. This might be best explained by the better compaction conditions at the bottom of the element, assuming that the concrete has good resistance to segregation. In this study the purpose of the in-situ test on full size columns was the verification of this phenomenon. The relation between distances of the specimens from the top of the columns and their compressive strength, pull-out strength, UPV and Schmitt rebound hammer values were plotted as given in Figures 4.45-4.48. The analysis have shown that the plots for all the mixes have trend-lines with positive slopes indicating that compressive and pull-out strength increases as the distance from the top increases. These results are in good agreement with the results of other previous research (zkul, 2000). The ultrasonic pulse velocity and rebound hammer values have a similar overall trend except for the limestone powder containing M6, M7 and M8 mixes for which the values are not very much affected by the depth of casting. The regression coefficients for all properties are summarized in Table 4.31. Table 4.31. Results of the regression analyses for the correlation between the mechanical properties and distance from the top of the columns. Coefficients of Linear Regression Mix Type Compressive Strength Pull-out Strength Ultrasonic Velocity Schmitt hammer Pulse Rebound M1 0.22 0.74 0.79 0.79 M2 0.48 0.96 0.61 0.90 M3 0.19 0.56 0.68 0.89 M4 0.49 0.66 0.27 0.88 M5 0.68 0.89 0.83 0.60 M6 0.44 0.08 0.30 0.30

The results of the regression analyses show that the there is a strong linear relation between the location of the specimens and the measured mechanical properties. M6 mix

128

might be an exception with more scattered results, implying that the overall homogeneity of the mix was not maintained as it was filling the 2 m high formwork. This was attributed to the relatively high limestone powder content of the mix, where the high limestone powder content, being an inert filler, resulted in a reduction in the overall cementations and pozzolanic reactions of the mix.

0,15 Sorptivity value (mm/min1/2) 0,125 R = 0.93 0,1 28 days 0,075 R = 0.63 0,05 0,025 0 0,00
2 2

90 days

1,00

2,00

3,00

4,00

5,00

6,00

Water Absorption (% )

Figure 4.20. Relationship between sorptivity and water absorption at 28 and 90 days

129

4000 Total charge passed (coulombs) 3500 3000 2500 2000 1500 1000 500 0 0 0,02 0,04 0,06 0,08 0,1 0,12 0,14 Sorptivity value (mm/min1/2 ) R = 0.69 R = 0.40
2 2

28 days 90 days

Figure 4.21. Relationship between total charge passed and sorptivity at 28 and 90 days
4000 Total charge passed (coulombs)

3000

R = 0.23 R = 0.34
2

2000

28 days 90 days

1000

0 0,00

1,00

2,00

3,00

4,00

5,00

6,00

Water Absorption (% )

Figure 4.22. Relationship between water absorption and total charge passed at 28 and 90 days

130

5000
Total Charge Passed, Q (coulombs)

4000
R = 0.90
2

3000 2000 1000 0 0,00


R = 0.90
2

28 days 90 days

0,05

0,10

0,15

Initial Current, IC (A)

Figure 4.23. Relationship between initial current and total charge passed at 28 and 90 days
5000 Total Charge Passed, Q (Coulombs)

4000 Q = 25043 IC + 156.19 R = 0.88 3000


2

2000

1000

0 0 0,05 Intial Current, IC (A) 0,1 0,15

Figure 4.24. Relationship between initial current and total charge passed for all data

131

4000 Total Charge Passed (coulombs) 3500 3000 2500 R = 0.46 2000 1500 1000 500 40 45 50 55 60 65 Compressive Strength (MPa)
2

28 days R = 0.32
2

90 days

Figure 4.25. Relationship between comp. strength and total charge passed at 28-90 days

6,00 Splitting Tensile Strength (MPa) R = 0.88


2

5,00

all data y = 1.3e


2 0.02x

R = 0.76 4,00 R = 0.37 3,00


2

7 days R = 0.89
2

28 days 90 days

2,00 20 30 40 50 60 70 Compressive Strength (MPa)

Figure 4.26. Relationship between compressive and splitting tensile strength at 7, 28 and 90 days

132

12,00 11,00 Pull-out Strength (MPa) 10,00 9,00 8,00 R = 0.73 7,00 6,00 5,00 2,00
2 2

R = 0.69

R = 0.85

all data y = 5.97 Ln(x) + 1.75 R = 0.67


2

7 days 28 days 90 days

2,50

3,00

3,50

4,00

4,50

5,00

5,50

6,00

Splitting Tensile Strength (MPa)

Figure 4.27. Relationship between splitting tensile and pull-out strength at 7, 28 and 90 days

12,00 11,00 Pull-out Strength (MPa) 10,00 9,00 8,00 7,00 6,00 5,00 25 30 35 40 45 50

R = 0.51

R = 0.61 all data y = 6.82 Ln(x) - 16.90 R = 0.42


2

7 days 28 days 90 days

R = 0.73

55

60

65

70

Compressive Strength (MPa)

Figure 4.28. Relationship between compressive strength and pull-out strength at 7, 28 and 90 days

133

Compressive Strength of Column Specimens at 28 days (MPa)

65 60 55 50 R = 0.78 45 40 35 30 30 35 40 45 50 55 60 65 Compressive Strength of Lab Specimens at 28 days (MPa)


2

Figure 4.29. Relationship between compressive strength of column and lab specimens

Pull-out Strength of Column Specimens at 28 days (Mpa)

11,5 10,5 R = 0.22 9,5 8,5 7,5 6,5 6,5 7,5 8,5 9,5 10,5 11,5 Pull-out Strength of Lab Specimens at 28 days (MPa)
2

Figure 4.30. Relationship between pull-out strength of column and lab specimens

134

6 Ultrasonic Pulse Velocity of Column Specimens at 28 days (km/sec) 5,5 5 4,5 4 3,5 3 2,5 2,50 R = 0.74
2

3,00

3,50

4,00

4,50

5,00

5,50

6,00

Ultrasonic Pulse Velocity of Lab Specimens at 28 days (km/sec)

Figure 4.31. Relationship between UPV of column and lab specimens


70 65 Compressive Strength (MPa) 60 55 50 45 40 35 4,00 R = 0.54
2

R = 0.89

28 days 90 days

4,25 Ultrasonic Pulse Velocity (km/sec)

4,50

Figure 4.32. Relationship between compressive strength and UPV at 28 and 90 days

135

65

60

M1 R = 0,22 M2 R = 0,48 M3 R2 = 0,19 M4 R2 = 0,49 M5 R =0,68 M6 R2 = 0,44

Compressive strenght (MPa)

55

50

45

40

35

30 0 50 100 150 200 Distance from the top of the column (cm)

Figure 4.33. Relationship between compressive strength and distance from the top of the column
14 12
Pull-out strength (MPa)
M1 M2 M3 M4 M5 M6

R = 0,74 R = 0,96 R = 0,56 R = 0,66 R = 0,89 R = 0,08


2 2 2 2 2

10 8 6 4 2 0 0 50 100 150 200


Distance from top of the column (cm)

Figure 4.34. Relationship between pull-out strength and distance from the top of the column

136

4,80 Ultrasonic pulse velocity (km/sec) 4,60 4,40 4,20 4,00 3,80 3,60 3,40 3,20 3,00 0 20 40 60 80 100 120 140 160 180 Distance from the top of the column (cm)
R2 = 0,79 R = 0,61 R2 = 0,68 R = 0,27 R = 0,83 R = 0,30
2 2 2 2

M1 M2 M3 M4 M5 M6

Figure 4.35. Relationship between UPV and distance from the top of the column
45

40 Rebound hammer number

M1 R2 = 0,79
35

M2 R2 = 0,90 M3 R2 = 0,89

30

M4 R2 = 0,88 M5 R2 = 0,60

25

M6 R2 = 0,30

20 0 50 100 150 200 Distance from the top of the column (cm)

Figure 4.36. Relationship between Schmitt hammer rebound value and distance from the top of the column

137

100 Cost of Materials of the Mix (YTL/m3)

97,5 R = 0.79 95
2

92,5

90 40 45 50 55 60 65 Compressive Strength at 28 days (MPa)

Figure 4.37. Relationship between cost of materials and compressive strength of SCC
4.5 Overall Evaluation of the SCC Mixes

Till this point of the study, the SCC mixes were tested in two main categories; mechanical tests and durability tests. Then the mixes were also evaluated according to their material costs. The relation between material costs and compressive strength of the SCC mixes is given in Figure 4.49 which shows that there is a linear relation between cost and compressive strength with a high correlation of R2 being equal to 0.79. But to be able to take into account every other property of the SCC mixes that have been measured, a different kind of approach has to be devised. In order to do that a scoring procedure was applied. For every group of test results, a relative value of mechanical, durability and permeability property was calculated for each mix as follows; For the mechanical properties; VMr = VM / VMmax * 100 where

138

VMr : relative value of a mechanical property of the mix VM: actual value of a mechanical property of the mix VMmax : highest value of a mechanical property among the mixes (best performance) For the durability and permeability properties; VDr = VDmin / VD * 100 where VDr : relative value of a durability or permeability property of the mix VD: actual value of a durability or permeability property of the mix VDmin : lowest value of a durability or permeability property among the mixes (best performance) And for cost of materials of the mixes; VCr = VCmin / VC * 100 where VCr : relative value of the cost of the mix VC: actual value of the cost of the mix VCmin : lowest value of the cost among the mixes (best price) So the relative values of mechanical and durability properties of each mix are given in Tables 4.32 4.33. The final table, Table 4.34 includes the cost of materials into the evaluation and the final overall average points of the mixes were determined showing the ranking of each mix. The results show that, in mechanical properties the relative values are closer together. M1, M2, M3, M4 and M5 range between 85 and 98 points, whereas the ones with excessive amounts of inert limestone powder, M6, M7 and M8 are in a lower group with average points of 76-79. However, the results of the durability tests show greater variance in relative points. Here again M3 has the highest points (93), but there is a significant difference between M3 and the second group of relative points of M1, M2, M4 (with 56-66 points) and M5 (46 points), just on the limits of the group. M6, M7, and M8 are again very close together with 36-37 relative points for overall durability performance.

139

As shown in Table 4.34, the cost of materials relative values are within a close range between 84-100 points. And when we average the relative points for mechanical, durability and price we get an overall average point of the mix. Here again, M3 has the highest point verifying that it is the mixture with best performance in all aspects. The poor durability performance of the M6, M7 and M8 mixes was compensated by their relatively lower costs of materials, due to the fact limestone powder is the least expensive admixture. The overall points range between 70 and 95, indicating that the differences among concrete performance diminish when all of the different aspects of concrete properties are taken into consideration. Table 4.32. Relative values of mechanical properties of SCC mixes
M1 M2 M3 M4 M5 M6 M7 M8

Compressive Strength, 7 days Compressive Strength, 28 days Compressive Strength, 90 days Splitting Tensile Strength, 7 days Splitting Tensile Strength, 28 days Splitting Tensile Strength, 90 days Pull-out Strength, 7 days Pull-out Strength, 28 days Pull-out Strength, 90 days UPV, 28 days UPV, 90 days Standard Deviation Average Mechanical Point

100 84 91 100 86 91 93 91 94 100 99 5.75


93

87 89 92 100 94 94 100 100 100 98 98 4.72


96

93 100 100 95 100 100 99 97 96 99 100 2.47


98

80 81 86 88 90 93 90 97 99 97 96 6.61
91

86 80 86 79 85 88 62 88 91 97 97 9.67
85

69 70 74 80 81 74 67 82 87 95 94 9.72
79

71 78 79 85 83 75 64 77 83 ----6.58
77

81 79 76 70 85 68 67 81 82 ----6.55
76

140

Table 4.33. Relative values of durability properties of SCC mixes


M1 M2 M3 M4 M5 M6 M7 M8

Water Absorption at 28 days Water Absorption at 90 days Sorptivity value at 28 days Sorptivity value at 90 days Total Charge Passed at 28 days Total Charge Passed at 90 days

96 48 70 50 33 23

74 50 51 29 33 23 80 72 95 100 50

100 100 100 100 100 100 100 66 100 56 100

79 41 58 88 58 67 60 67 100 71 33

60 31 44 41 39 29 60 79 95 16 13

47 24 35 24 29 21 0 100 90 22 13

49 25 38 27 41 26 0 84 -------

48 25 39 26 39 26 0 82 -------

Deicing Salt Scaling after 50 cycles 100 Drying shrinkage 75

Carbonation depth after 30 months 100 % damage in sulfate solution 16

Linear expansion in sulfate solution 9 Standard Deviation Average Durability Point

34.00 26.17 16.11 19.29 25.45 31.19 24.26 23.60


56 60 93 66 46 37 36 36

Table 4.34. Relative values of overall properties of SCC mixes including cost of materials
M1 M2 M3 M4 M5 M6 M7 M8

Average Durability Point Average Mechanical Point Price Point Standard Deviation

56 93 84

60 96 98

93 98 95

66 91 97

46 85 99

37 79 100

36 77 98

36 76 99

19.36 21.26 2.71

16.42 27.24 32.21 31.28 32.16

Overall Average Point

78

84

95

84

77

72

70

70

141

5. CONCLUSIONS

Following conclusions may be drawn from the results of the tests conducted in this study. 1. All the mixes had satisfactory self-compacting properties in the fresh state. The addition of fly ash and silica fume had positive effects on fresh workability and flowability of the pastes whereas limestone powder caused a decrease in these properties. This might be explained by the differences in the particle size, shape and texture among the fine materials used. 2. The compressive, splitting tensile, and pullout strengths and other mechanical properties were increased with the use of silica fume and fly ash, especially in the 90 days specimens. On the other hand the addition of limestone powder caused a decrease in these mechanical properties. 3. The hardened concrete tests conducted on the specimens sawn from the columns revealed results similar to the laboratory tests with minor differences that might be attributed to more favorable curing and compaction conditions of the lab specimens. The results also showed that there were insignificant differences among the properties of the column specimens taken at different depths. This showed that mixes were stable enough to resist segregation when casted in a 2000 mm congested column as the fresh concrete tests claimed. Here again limestone powder caused a relative decrease in the above properties compared to silica fume and then fly ash. 4. The image analysis and segregation coefficient studies also showed that the mixes were stable and self-compacting, showing minor variances in relevant properties with changing depth. 5. The microscopic studies of the interface between the concrete and steel reinforcing bars, showed that the SCC in all mixes perfectly encapsulated the

142

reinforcing bar, showing no signs of defects or irregular coarse aggregate accumulation. 6. The cumulative results of the water absorption, sorptivity and chloride ion permeability test showed that the mixes with SF and FA had lower permeability in all cases where as limestone powder in excess amounts increases the permeability of the concrete due to the formation of a porous hardened concrete. 7. In the deicing salt scaling test, all concretes performed satisfactorily except the mixes with high LSP contents. This was attributed to the dilution of the cementitious reactions, hence formation of a weakened superficial layer of concrete vulnerable to stresses induced during the deicing salt scaling test. It might be conclude that in concrete applications where the final concrete would be exposed to deicing salt scaling, the use of high volumes of limestone powder should not be recommended. 8. The carbonation of all the mixes was very limited possibly due to the low w/b ratio and low permeability of the concretes tested. Even the mixes with limestone powder incorporation exhibited very slight carbonation at the age of 30 months. 9. The sulfate resistances of the concrete mixes were also satisfactory even for a long immersion period of 30 months. High linear expansions and compressive strength loss was mostly observed for the concrete with excess limestone powder. And the percentage damage was lowest for SF FA concrete. 10. The drying shrinkage results showed that silica fume increased the shrinkage where as limestone powder decreased it. The high shrinkage of silica fume mix was explained by the higher surface area, higher water consumptions and reduction of space between the particles of SF, replacing cement. 11. Considering the cost analysis discussed in this study, the use of mineral admixtures and especially limestone powder becomes a powerful tool in decreasing the material costs without significant tradeoff in mechanical or durability properties.

143

12. Results of regression analyses on durability properties confirmed that there is a good correlation between water absorption and sorptivity for a given age, and initial current reading and total charge passed in RCPT. The relationships between some of the mechanical properties such as compressive strength and pull-out strength or UPV were also found to be directly linear relations with significantly high degrees of correlation. 13. The in-situ compressive strengths of column specimens were found to be interdependent with corresponding lab specimens of the same mixes. Generally all of the mechanical properties were found to be strongly correlated for column and lab specimens, verifying that the performances of the SCC mixes tested were stable in both cases. 14. The in-situ mechanical properties of the columns specimens were found to be increasing to a certain extent with increasing height. This effect was more pronounced for pull-out strength and Schmitt rebound values. 15. The overall evaluation of the mixes revealed that the mechanical properties are within a closer range than the durability properties. The mixes (M6, M7 and M8) with higher limestone powder additions have significantly lower durability points. But when price, as defined by the cost of materials, is taken into consideration the gap between the highest and the lowest overall points diminishes. It is also worth noting that, even though the choice of specific components generally seemed to be governed by local practice and availability, the exact composition of the SCC binder should be decided upon a holistic evaluation of the desired mechanical and durability properties, structural design and esthetics aspects and overall economy.

144

6. RECOMMENDATIONS FOR FUTURE WORK

For a better understanding of the effects of different cement replacements, the number of combinations of different powder materials might be increased. To be able to distinguish between the exact effects of individual powder materials (FA, SF, and LP) more combinations with 0 %, 10%, 20%... 50% cement replacements might be necessary. Using different kinds of powders that have not yet been studied also seems to be appropriate study for future work. Here the aim should be to lower the cement content and improve or at least not worsen the mechanical and durability properties by using readily available alternative powder materials. Another issue to be considered is the fresh concrete tests; there seems to be a gap between the common fresh concrete test results (slump flow, U-tube, L-box, V-funnel) and the actual rheology of the self-compacting concrete. It is therefore suggested that another set of results should be evaluated using a proper rheometer test set-up. Or a special custommade set-up might be devised after careful calculations and calibrations. The two-point test set-up seems to be a proper starting point for such a study since it is the simplest method. The number of specimens tested would greatly influence the reliability and statistical significance of the results.

145

APPENDIX A: PHOTOGRAPHIC VIEW OF TEST SET-UP AND SPECIMENS BEFORE AND AFTER TESTING

(c) (a) Figure A.1 Slump flow test in a time sequence (b)

(a)

(b)

Figure A.2 Column specimen formwork (a), demolded column specimen just before sawing operation (b)

146

147

(b) Figure A.3 Deicing salt scaling ressitance specimens after increasing numbers of F&T cycles (in the order of M1-M6 from left to right)

(a) (b) th Figure A.4. Deicing salt scaling specimens after 50 cycle, with severely deteriorated surfaces; (a) M7 and (b) M8

148

149

Figure A.5 Test set-up for Rapid Chloride Permeability

150

Figure A.6 Carbonation test; spraying of phenolphthalein (a), close-up photo of M6 where slight carbonation is observed (b)

(a) (b) Figure A.7 Video-microscope images of concrete-steel interface without any visible defects ( at x150 magnification) (a) cross-section at M 1.1 and (b) cross-section at M 1.5.

151

(a) (b) Figure A.8. Video-microscope images of concrete-steel interface without any visible defects ( at x150 magnification) (a) cross-section at M 5.1 and (b) cross-section at M 5.5.

152

153

REFERENCES

Batopu, M. E., 2000, Effect of Curing on Resistance to Deicing Salt Scaling, Freezing and Thawing and Chloride-ion Penetration of Fly Ash Concrete, M.S. Thesis, Boazii University. Bentz, D. P., 2005, Modeling the Influence of Limestone Filler on Cement Hydration Using CEMHYD3D, submitted to the Journal of Cement and Concrete Composites. Beupre, D., P. Lacombe and K. H. Khayat, 1999, Laboratory Investigation of Rheological Properties and Scaling Resistance of Air Entrained SSC, Materials and Structures, V.32, pp. 235-240, April. Bouzoubaa, N. and M. Lachemi, 2002, Self-compacting Concrete Incorporating High Volumes of Class F Fly Ash, Cement and Concrete Research, V.31, pp. 413-420. Bui K., and D. Montgomery, 1999, Drying Shrinkage of Self-Compacting Concrete Containing Milled Limestone, Proceedings 1st Int. RILEM Symposium on SCC, Stockholm, Sweden, pp 227-238, September. Byfors, J., 1999, SCC is an Important Step Towards Industrialization of the Building Industry, Proc. International RILEM Symposium (Self Compacting Concrete), Technical Research and Consulting on Cement Concrete, Switzerland, September. Caijun, S. and Y. Wu, 2005, Mixture Proportioning and Properties of Self-Consolidating Lightweight Concrete Containing Glass Powder, ACI Materials Journal, pp. 355-363, September. Collepardi, M., S. Monosi and M. Valente, 2001, Optimization of SP Type and Dosage in FA and SF Concretes, ACI SP 119-22. TFB

154

Cohen, M. and A. Bentur, 1988, Durability of Portland Cement-Silica Fume Pastes in Magnesium Sulfate and Sodium Sulfate Solutions, ACI Materials Journal, pp. 148-157, May-June. elik, B., 1997, Effect of Binding Material Composition on Mechanical Properties of Concrete Subjected to Different Exposure Conditions, M.S. Thesis, Boazii University. il, ., 2000, Kendiliinden Yerleen Beton, Vizyon, YKS Teknik Dergi, Say:2. Dehn, F., 2002, High Performance Self-Compacting Concretes for Bridge Construction, Conference Proceedings: 1st North American Conference on Design and Use of SCC, pp. 433-438, November. Dilek, F. T. and M. Tokyay, 2003, Sulfate Resistance of Laboratory-made Cements with Different Mineral Admixtures. Sixth CANMET/ACI Int. Conference on Durability of Concrete, Thessaloniki, Greece. Domone, P. L., 2005, Self-compacitng Concrete: An Anaysis of 11 Years of Case Studies, Cement & Concrete Composites, article in press, October. Druta, C., 2003, Tensile Strength and Bonding Characteristics of Self Compacting Concrete M.S. Thesis, Polytechnic University of Bucharest. Duval, R. and E. H. Kadri, 1998, Influence of Silica Fume on the Workability and the Compressive Strength of High-Performance Concretes, Cement and Concrete Research Journal, Vol. 28, Issue 4, pp.533-547. EFNARC; Specification and Guidelines for Self-Compacting Concrete, European Association fro Producers and Applicators of Specialist Building Products (EFNARC), 2002, pp. 32. Edamatsu, Y. and N. Nishida, 1999, A Rational Mix-Design Method for SCC Considering Interaction between Coarse Aggregate and Fine Mortar Particles, Proceedings of the 1st RILEM Symposium on Self-Compacting Concrete, pp. 309-320.

155

Fang, W., C. Jianxiong and Y. Changhui, 1999, Studies on SCC with High Volume Mineral Additives, Proc. International RILEM Symposium (Self Compacting Concrete), Por. 7, TFB Technical Research and Consulting on Cement Concrete, pp. 569-578, September. Ferraris, C. F., L. Brower, J. Daczko, and C. Ozyldirim, 1999, Workability of SelfCompacting Concrete, Journal of Research of NIST, Vol. 104, No. 5, pp.461-478. Freeman, R. and R. Carrasquillo, 1992, Effect of Intergrinding FA on the Sulfate Resistance of FA Concrete, ACI SP 132-17, Istanbul Conference. Gibbs, J., and W. Zhu, 1999, Strength of Hardened Self-Compacting Concrete, Proc. Int. RILEM Symposium (Self Compacting Concrete), Pro 7, TFB Technical Research and Consulting on Cement Concrete, pp. 199-209, September. Gram, H. E., P. Piiparinen, 1999, Properties of SCC- Especially Early Age and Long Term Shrinkage and Salt Frost Resistance, Proc. International RILEM Symposium (Self Compacting Concrete), Por. 7, TFB Technical Research and Consulting on Cement Concrete, Switzerland. Gneysi, E., 1999, Effect of Binding Material Composition on Strength and Chloride Permeability of High Strength Concrete, M.S. Thesis, Boazii University. Hanson, J. A., 1968, Effects of Curing and Drying Environments on Splitting Tensile Strength of Concrete, ACI Journal, pp.535-543. Hawkins, P., P. Tennis and R. Detwiler, 1998, The Use of Limestone in Portland Cement: A State-of-the-Art Review, Portland Cement Association Cements Research Progress, American Ceramic Society. Heikal, M., H. El-Diamony and M.S. Morsy, 2000, Limestone-filled Pozzolanic Cement, Cement and Conceret Research, Vol. 30, pp 1827-1834.

156

Hooton, R. and J. Emery, 1990, Sulfate Resistance of a Canadian Slag Cement, ACI Materials Journal, pp. 547-555, Nov-Dec. Jacobs, F. and F. Hunkeler, 1999, Design of SCC for Durable Concrete Structures, Proceedings of the 1st RILEM Symposium on Self-Compacting Concrete, pp. 397-407. Jianxiong, C., P. Xincheng, H. Yubin, 1999, A Study of SCC with Superfine Sand and Pozzolanic Additives, Proc. International RILEM Symposium (Self Compacting Concrete), Por. 7, TFB Technical Research and Consulting on Cement Concrete, pp. 549560, Switzerland, September. Kennedy, C.T., 1940, The Design of Concrete Mixes, Proceedings of the American Concrete Institute, Vol.36, pp. 373-400. Khayat, K. H., A. Ghezal, M. S. Hadriche, 2000, Development of Factorial Design Models for Proportioning SCC, Shigeyoshi Nagataki Symposium, Vision of Concrete: 21st Century. Khayat, K., H., 2000, Optimization and Performance of Air-Entrained SCC, ACI Materials Journal, pp. 526-535, Sep-Oct. Khayat, H., M. N. Haque and I. Fattuhi, 2002, Concrete Carbonation in Arid Climate, Materials and Structures, pp. 421-426, August. Khayat, K. H., K. Manai and A. Trudel, 1997, In situ Mechanical Properties of Wall Elements Cast Using SCC, ACI Materials Journal, pp. 491-500, Nov-Dec. Khayat, K. and P. Paultre, 2001, Structural Performance and In-Place Properties of SCC Used for Casting Highly Reinforced Columns, ACI Materials Journal, pp. 371-378, SepOct.

157

Khayat, K., C. Hu and M. J. Laye, 2002, Importance of Aggregate Packing Density on Workability of SCC, Conference Proceedings: 1st North American Conference on Design and Use of SCC, pp. 53-64, November. Lacombe, P., D. Beaupre and N. Pouliot, 1999, Rheology and Bonding Characteristics of Self-leveling Concrete as a Repair Material, Materials and Structures, V. 32, pp. 593-600 October. Lim, S. N. and T. H. Wee, 2000, Autogenous Shrinkage of GGBFS Concrete., ACI Materials Journal, V. 97, No.5, pp. 587-593, Sept.-Oct.

Ling, C.X. and K. T.

Lin, 1992, Shrinkage of Statically Compacted cement-

Phosphogypsum Mixtures., ACI Committee 209, SP 135, pp. 77-93, April. Mangat, P. and J. M. Khatib, 1995, Influence of FA, SF and Slag on Sulfate Resistance of Concrete, ACI Materials Journal, pp. 542-552, Sep-Oct. Malami, C., and V. Kaloidas, 1994, Carbonation and Porosity of Mortar Specimens with Pozzolanic and Hydraulic Cement Admixtures, Cement and Concrete Research Vol. 28, pp. 1444-1454. Nagataki, S., 2000, Advanced Technology for High Fluidity Concrete, Cement and Concrete Technology in the 2000s, 2nd Int. Symp. , Istanbul, Turkey, pp. 6-10, September. Neville, A. M., 1986, Properties of Concrete, Longman House, Burnt Mill, Harlow, Essex. Neville, A. and P. Aitcin, 1998, High Performance Concrete- An Overview, Materials and Structures, v. 31, pp. 111-117. Noor, M. A. and T. Uomoto, 1999, Three-Dimensional Discrete Element Simulation of Rheology Tests of SCC, Proceedings 1st Int. RILEM Symposium on SCC, Stockholm, Sweden, pp. 35-45, September.

158

Okamura, H. and M. Ouchi, 1999, Self-Compacting Concrete. Development, Present Use and Future, Proceedings of the 1st RILEM Symposium on Self-Compacting Concrete, A. Skarendahl and O. Petersson, Eds., pp. 3-14. ONeil, E. F., B. D. Neeley and J. D. Cargile, 2002, Tensile Properties of Very-HighStrength Concrete for Penetration-Resistant Concrete, US Army Engineer R & D Center. Ouchi, M., M. Hibino and H. Okamura, 1996, Effect of Superplasticizer on Selfcompactability of Fresh Concrete, Transportation Research Record, no 1574, pp. 37-40. Ozawa, K., N. Sakata and H. Okamura, 1996, Evaluation of Self-compactability of Fresh Concrete Using the Funnel Test, Collected Papers (University of Tokyo Department of Civil Engineering), v.34, pp. 59-75. Ozawa, K. and H. Okamura, 1996, Mix Design for Self-compacting High Performance Concrete, Collected Papers (University of Tokyo Department of Civil Engineering), v.34, pp. 151-175. zkul, M. H., . Doan, Z. avdar, A. R. Salam and N. Parlak, 2000, Effects of SCC Admixtures on Fresh and Hardened Concrete Properties, Cement and Concrete Technology in the 2000s, 2nd Int. Symp. , Istanbul, Turkey, September. Patel, R., K. Hossain, M. Shehata, M. Bouzoubaa, and M. Lachemi, 2004, Development of Statistical Models for Mixture Design of High-Volume Fly Ash SCC, ACI Materials Journal, v. 101, no. 4, pp. 1841-1845, July. Persson, B., 2001, Assessment of the Chloride Migration Coefficient, Internal Frost Resistance, Salt Frost Scaling and Sulphate Resistance of Self-Compacting Concrete, Report no. TVBM-3100, Lund University. Persson, B., 2001, A Comparison between Mechanical Properties of SCC and the Corresponding Properties of Normal Concrete, Cement and Concrete Research, v. 31, pp. 193-198.

159

Pfeifer, D. W., 1967, Sand Replacement in Structural Lightweight Concrete Splitting Tensile Strength, ACI Journal, pp.384-392. Piasta, W. G. and L. Hebda, 1991, Sulfate Expansion and Permeability of Concrete with Limestone Aggregate, Magazine of Concrete Research, V. 43, pp. 81-85. Ray, I. and R. Chattopadhyay, 1999, Effect of Silica Fume on Superplasticized Concrete, International Conference on Concretes, Dundee, Scotland. Saak, W. A., H. M. Jennings and S. P. Shah, 2001, New Methodology for Designing Self-Compacting Concrete, ACI Materials Journal, pp. 429-439, Nov.-Dec. Saak, W. A., 2000, Characterization and Modeling of the Rheology of Cement Paste, PhD Thesis, Northwestern University, Evanston, III. Selih, J. and J. Tritthart, 2003, Durability of Portland Limestone Powder- Cement Concrete, Sixth CANMET/ACI Int. Conference on Durability of Concrete, Thessaloniki, Greece. Skarendahl, A. and . Petersson, 1999 , Self-Compacting Concrete- State-of-the-art Report, RILEM Report 23, RILEM Publications, Cedex, France. Sonebi, M. and P. Bartos, 2002 Hardened SCC and Its Bond with Reinforcement, Materials and Structures, v.35, no. 252, pp. 462-469, September. Soroushian, P. and S. Ravanbakhsh, 1998, Control of Plastic Shrinkage Cracking, ACI Materials Journal, V. 95, No.4, pp. 429-435, July-August. Subramanian, S., and D. Chattopadhyay, 2002, Experiments for Mix Proportioning of Self-Compacting Concrete, Indian Concrete Journal, pp. 13-20, January.

160

Takada, K. and G. Pelova, 1999, Influence of Microfillers on Proportioning of Mortar in SCC, Proc. International RILEM Symposium (Self Compacting Concrete), Por. 7, TFB Technical Research and Consulting on Cement Concrete, pp. 537-547, September. Tang, L., J. Andalen and S. Hjelm, S. 1999, Chloride Diffusivity of Self Compacting Concrete, Proc. International RILEM Symposium (Self Compacting Concrete), Por. 7, TFB Technical Research and Consulting on Cement Concrete, pp. 187-198, September. Tattersall, G. H., 1991, Workability and Quality Control of Concrete, Chapman & Hall, London. Topu, O. and R. Khurana, 2000, Role of SP in the Development of SCC, Cement and Concrete Technology in the 2000s, 2nd Int. Symp , Istanbul, Turkey, 6-10 Sept. Topu, O., A. Salk, . il, M. nal, S. Polat and A. ste, 2000 , Application of SCC in Common Building Structure-A Case Study, Cement and Concrete Technology in the 2000s, 2nd Int. Symp. , Istanbul, Turkey, 6-10 Sept. Valenza J. J. and G. W. Schere, 2005, Mechanism of Salt Scaling, Materials and Structure no.38 pp. 479-488, May. Venugopal, C. N., S. K. Prasad and T. S. Nagaraj, 2000, SP- Its Optimal Dosage in Cement Based Composites, Cement and Concrete Technology in the 2000s, 2nd Int. Symp. , stanbul, Turkey, 6-10 September Villian, G., L. V. Baroghe and C. Kounkou, 1999, Comparative Study on the Induced Hydration Drying and Deformations of SC and Ordinary Mortars, Proc. International RILEM Symposium (Self Compacting Concrete), Por. 7, TFB Technical Research and Consulting on Cement Concrete, Switzerland, September. Yeinobal, A. and F. T. Dilek, 1996, Sulfate Resistance of Mortars Containing SF as Evaluated by Different Methods, ACI SP 153-42, Milwaukee Conference.

161

Zhu, W., J. Gibbs and P. J. Bartos, 2001, Uniformity of in situ Properties of SelfCompacting Concrete in Full-Scale Structural Elements, Cement and Concrete Composites, V.23, pp. 57-64.

You might also like