AN EVALUATION OF THE USE OF SELF-CONSOLIDATING CONCRETE (SCC) FOR DRILLED SHAFT APPLICATIONS Except where reference is made to the work of others, the work described in this thesis is my own or was done in collaboration with my advisory committee. This thesis does not include proprietary or classified information. Joseph Donald Bailey Certificate of Approval: Dan A. Brown Anton K. Schindler Gottlieb Associate Professor Assistant Professor Civil Engineering Civil Engineering George E. Ramey Stephen L. McFarland Professor Dean Civil Engineering Graduate School i AN EVALUATION OF THE USE OF SELF-CONSOLIDATING CONCRETE (SCC) FOR DRILLED SHAFT APPLICATIONS Joseph Donald Bailey A Thesis Submitted to The Graduate Faculty of Auburn University In Partial Fulfillment of the Requirements for the Degree of Master of Science Auburn University August 8, 2005 ii ii AN EVALUATION OF THE USE OF SELF-CONSOLIDATING CONCRETE (SCC) FOR DRILLED SHAFT APPLICATIONS Joseph Donald Bailey Permission is granted to Auburn University to make copies of this thesis at its discretion, upon the request of individuals or institutions and at their expense. The author reserves all publication rights. Signature of Author Date Copy sent to: Name Date iii iii AN EVALUATION OF THE USE OF SELF-CONSOLIDATING CONCRETE (SCC) FOR DRILLED SHAFT APPLICATIONS Joseph Donald Bailey Master of Science, August 8, 2005 (B.S. Auburn University 2002) 234 Typed Pages Directed by Dr. Anton K. Schindler and Dr. Dan. A. Brown In drilled shaft construction, the recent development of sophisticated techniques for integrity and load testing has lead to the ability to asses the quality of drilled shaft foundations in terms of integrity and load carrying capabilities after they have been cast. Although this ability has lead to better assessment of drilled shaft foundations, it has also given insight to problems that are associated with materials and construction practices that has lead to defects or less than optimal performance for drilled shaft foundations. This study examines the more common problems associated with drilled shaft foundations to emphasize the importance of constructability in design and workability in the construction materials. The majority of these problems consist of the failure to adequately consider one or more of the following issues: iv iv ? Workability of concrete for the duration of the pour ? Compatibility of the highly congested rebar cages and concrete being placed ? Bleeding and segregation The main purpose of this study is to evaluate the use of self-consolidating concrete (SCC) as a viable material to overcome these issues due to its high flowability, passing ability, resistance to segregation, and reduced bleeding. A laboratory study will examine the difference between ordinary drilled shaft concrete and self-consolidating concrete (SCC) for both fresh and hardened properties. The fresh properties include filling ability, passing ability, segregation resistance, workability over time, bleeding characteristics, and controlled setting, while the hardened properties include the comparison of the compressive strength, elastic modulus, drying shrinkage, and permeability. The laboratory results for both the ordinary drilled shaft concrete (ODSC) and SCC mixtures were evaluated and compared. The results show that SCC can be used to address many of the problems associated with drilled shaft construction because of the inherent workability, passing ability, resistance to segregation, and reduced bleeding of this type of concrete. Furthermore, the data suggest that the use of SCC in drilled shaft applications can provide similar or improved hardened concrete properties, which includes the compressive strength, elastic modulus, drying shrinkage, and permeability. However, some potential concerns for this material in drilled shaft applications include the general and overall lack of experience and research, lack of standardized tests, lack of well-defined mixture proportioning, and larger changes in workability compared to ODSC mixtures. v Style manual or journal used The Chicago Manual of Style Computer software used Microsoft Word, Microsoft Excel, Microsoft Powerpoint vi TABLE OF CONTENTS LIST OF TABLES ?.??????????????????????????.x LIST OF FIGURES ???????????????..??????????....xii CHAPTER ONE: INTRODUCTION????????????????????..1 1.1 Statement of problem??????????????????????1 1.2 Research Objectives??????...?...???????.??????...3 1.3 Research Scope??.??.?..??????????????????.4 CHAPTER TWO: SELF-CONSOLIDATING CONCRETE???????????.5 2.1 Introduction?????????????????????????...5 2.2 SCC Testing Procedures???????????????????...?7 2.2.1 Slump Flow Test 2.2.2 L- Box Test 2.2.3 J-Ring Test 2.2.4 Segregation Column Test 2.3 Fresh Concrete Properties???????????...????????18 2.3.1 Rheology 2.3.2 Filling Ability 2.3.3 Passing Ability 2.3.4 Segregation Resistance 2.4 Hardened Concrete Properties????????????????...?.29 2.4.1 Compressive Strength 2.4.2 Modulus of Elasticity 2.4.3 Drying Shrinkage 2.4.4 Permeability 2.5 Summary and Conclusions.???..??????????...????..48 CHAPTER THREE: EXPERIENCES WITH DRILLED SHAFT CONCRETE???.52 3.1 Introduction????...??.??????????????????.52 3.2 Workability of Drilled Shaft Concrete???????????????53 3.3 Compatibility between Congested Rebar Cages and Concrete?????...74 vii vii 3.4 Segregation and Bleeding???????????????????...81 3.5 Summary and Conclusions??????????????????.....87 CHAPTER FOUR: LABORATORY TESTING PROGRAM AND MATERIALS ?...89 4.1 Introduction ????????????????????????....89 4.2 Requirements for ODSC and SCC Mixtures ???????????.....89 4.3 Laboratory Testing Program ?????????????????.?.91 4.3.1 Phase I ? Effect of Type and Dosage of HRWRA 4.3.2 Phase II ? Effect of Retarder Dosage 4.3.3 Phase III ? Appropriate SCC Mixing Procedure 4.3.4 Phase IV ? Selection of SCC Properties 4.3.5 Phase V ? Methods to Modify the Viscosity of SCC Mixtures 4.4 Raw Material Sources??????????????????......?.112 4.4.1 Cementitious Materials 4.4.2 Aggregates 4.4.3 Chemical Admixtures CHAPTER FIVE: LABORATORY EQUIPMENT, SPECIMENS, AND PROCEDURES ????????????????????.???????.126 5.1 Introduction ????????????????????...?....?....126 5.2 Batching and Mixing Procedure ???????????????...?127 5.3 Fresh Property Testing ????????????????????..129 5.3.1 Slump Test 5.3.2 Slump Flow Test 5.3.3 J-Ring Test 5.3.4 L-Box Test 5.3.5 Segregation Column Test 5.3.6 Bleeding 5.3.7 Unit Weight and Air Content 5.3.8 Making and Curing Specimens 5.3.9 Time of Set 5.4 Hardened Concrete Properties ?????????????????..145 5.4.1 Compressive Strength 5.4.2 Modulus of Elasticity 5.4.3 Drying Shrinkage 5.4.4 Permeability viii CHAPTER SIX: PRESENTATION AND ANALYSIS OF RESULTS ?????....151 6.1 Introduction ?????...???????????????????151 6.2 Phase I - Selection of Type and Dosage of HWRWA ????????..151 6.3 Phase II ? Effect of Retarder Dosage ???????????????153 6.4 Phase III - Appropriate SCC Mixing Procedure ???????...???156 6.5 Phase IV - Selection of SCC Properties ??????????.???...158 6.5.1 Fresh Concrete Properties and Workability 6.5.2 Segregation and Bleeding Results 6.5.3 Passing Ability: J-Ring and L-Box 6.5.4 Compressive Strength 6.5.5 Modulus of Elasticity 6.5.6 Drying Shrinkage 6.5.7 Permeability 6.5.8 Comparison between Laboratory and Field Conditions 6.6 Phase V - Methods to Modify the Viscosity of SCC Mixtures ???.?....190 6.6.1 Fresh Concrete Properties 6.6.2 Hardened Concrete Properties 6.6.3 Effect of VMA on Fresh Concrete Properties 6.7 Summary of Research Findings?????????????????.210 CHAPTER SEVEN: PROPOSED EXPERIMENTAL FIELD STUDY.?????...214 7.1 Introduction ???????????????????????......214 7.2 Test Shafts ????????????...???????????.....215 7.3 Fresh Concrete Property Testing ????????????????..215 7.4 Hardened Concrete Property Testing ????????????.??...216 7.5 Placement Monitoring ????????????????????...217 7.6 Testing of Non-exhumed Shafts ???????????????.?..217 7.7 Testing of Exhumed Shafts ??????????????????...218 7.8 Instrumentation ??????????????????????.....218 CHAPTER EIGHT: SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS.219 8.1 Introduction???????????????????????...?219 8.2 Summary and Conclusions ??????????????????...219 8.3 Recommendations??????????????????????.222 ix LIST OF TABLES Table 2.1 Visual Stability Index (VSI) Rating??????????..????.12 Table 2.2 J-Ring Passing Ability Rating?????????????????15 Table 2.3 Compressive Strength Results from Turcry et al.(2003)............................35 Table 2.4 Mixture Proportions (Kim et al. 1998)....................................................... 43 Table 4.1 Concrete Mixture Proportions for Phase I ???????????.....93 Table 4.2 Concrete Mixture Proportions for Phase II ????????????95 Table 4.3 Concrete Mixture Proportions for Phase III ???????????..98 Table 4.4 Concrete Mixture Proportions for Phase IV ???????????.103 Table 4.5 Concrete Mixture Proportions for Phase V ??????????......110 Table 4.6 Chemical Analysis Results for the Giant Type I Cement ??????113 Table 4.7 Chemical Analysis Results for the SEFA Group Class F Fly Ash ??..114 Table 4.8 Chemical Analysis Results for the Micron 3 Fly Ash ???????..115 Table 4.9 Specific Gravities for Cementitious Materials ??????????.116 Table 4.10 Specific Gravities and Absorption Capacities for Aggregate Sources ?.120 Table 5.1 Visual Stability Index (VSI) Rating ??????????????133 Table 5.2 J-Ring Passing Ability Rating ???????.?????????136 Table 6.1 Fresh Concrete Properties for Phase I ?????????????.152 Table 6.2 Fresh Concrete Properties for Phase II ?????????????155 Table 6.3 Fresh Concrete Properties for Phase III ????????????..157 Table 6.4 Fresh Concrete Properties for ODSC Mixtures ?????????..159 x x Table 6.5 Fresh Concrete Properties for SCC Mixtures ??????????.160 Table 6.6 Segregation and Bleeding Results ??????????????..167 Table 6.7 Passing Ability Results for SCC Mixtures ???????????.168 Table 6.8 Fresh Concrete Properties for both Laboratory and Field Conditions ?185 Table 6.9 Fresh Concrete Properties for Phase V ????????????...194 Table 6.10 Tabulated Results of the Effect of VMA on Fresh Concrete Properties..209 xi LIST OF FIGURES Figure 2.1 Upright Slump Cone Method ??????.???????????9 Figure 2.2 Inverted Slump Cone Method ????????????????....9 Figure 2.3 Visual Stability Index Rating ????????...??.?????...12 Figure 2.4 L-Box Testing Apparatus ??????????????????.14 Figure 2.5 J-Ring Testing Apparatus ?????????????????.....15 Figure 2.6 Segregation Column Testing Apparatus ?????????.???..17 Figure 2.7 Bingham Fluid Model ???????????????????...19 Figure 2.8 Mechanism of Blocking ???????????????????24 Figure 2.9 Blocking Ratios versus Reinforcement Spacing for Different Size Aggregate ???????.?...???????????????.25 Figure 2.10 Strength versus Cement-to-Water Ratio ????????????...30 Figure 2.11 Compressive Strength versus Porosity ?????????????.31 Figure 2.12 Strength versus Gel-to-Space Ratio Based on Tested Mortar Cubes??33 Figure 2.13 Average Compressive Strengths of SCC and Conventional Bridge Concretes with w/c of 0.40 ?????????????????...34 Figure 2.14 Elastic Modulus versus Sand-to-Aggregate Ratio ....................................38 Figure 2.15 Aggregate Content versus Water-to-Cement Ratio ????????..40 Figure 2.16 Shrinkage Modification Factor for Different Sand-to-Aggregate Ratios ??????????????????????????41 Figure 2.17 Shrinkage versus Cement Content ...........................................................42 Figure 2.18 Drying Shrinkage versus Age (Kim et al. 1998) ?????????..43 xii xii Figure 2.19 SCC versus Normal Concrete....................................................................44 Figure 2.20 Drying Shrinkage versus Sand-to-Aggregate Ratio????????...45 Figure 2.21 Coefficient of Permeability versus Capillary Porosity ???????.46 Figure 2.22 Coefficient of Permeability versus Water-to-Cement Ratio ?????.46 Figure 2.23 RCPT Values of SCC and Normal Concrete ???????????48 Figure 3.1 Drilled Shaft Mixture with a Slump of Approximately 6.5 Inches ??...55 Figure 3.2 Drilled Shaft Mixture with a Slump of 8-9 Inches ?????????56 Figure 3.3 Illustration of Tremie Placement ???????????????...59 Figure 3.4 Illustration of Entrapped Debris due to the Eruption of Fresh Concrete through Stiff Concrete ??.?????????????????..60 Figure 3.5 Illustration of Entrapped Debris Seams due to Extraction of the Tremie..61 Figure 3.6 Example 1 of Shaft Defects due to the Loss of Workability ?????.62 Figure 3.7 Example 2 of Shaft Defects due to the Loss of Workability ?????.62 Figure 3.8 Example 3 of Shaft Defects due to the Loss of Workability ?????.63 Figure 3.9 Example 4 of Shaft Defects due to the Loss of Workability ?????.63 Figure 3.10 Example 5 of Shaft Defects due to the Loss of Workability?????..64 Figure 3.11 Illustration of a Cased Hole ?????????????????...65 Figure 3.12 Necking and Arching due to the Extraction of the Casing when Workablility is Lost ??????????????.??????..66 Figure 3.13 Shaft Defects due to the Extraction of the Casing ?????????.67 Figure 3.14 Slump Loss versus Time Relationship ?????????????..68 Figure 3.15 Slump Loss at 70 o F for Two Different Brands of Type I Cement ...??.69 Figure 3.16 Shaft Defects due to the Loss of Workability (left), After Removal of Surface Flaws for Repairs ??????????????..????70 Figure 3.17 Shaft Defects due to the Loss of Workability from Construction Delays..72 xiii xiii Figure 3.18 Screening of the Concrete due to Heavily Reinforced Rebar Cages ?.?76 Figure 3.19 Elevation Difference between the Inside and Outside of a Rebar Cage due to Screening of the Concrete Flow ?????????????..77 Figure 3.20 Shaft Defects due to the Screening of the Concrete Flow ??????.77 Figure 3.21 Oval Shaped Column with Figure 8 Rebar Cage ?????????..79 Figure 3.22 Shaft Defect due to Double Reinforcement Cages ?????????80 Figure 3.23 Illustration of Bleeding in Freshly Placed Concrete ????????..84 Figure 3.24 Example 1 of Bleed Channels and Surface Streaks due to Bleed Water Traveling along the Casing ??????????????????84 Figure 3.25 Example 2 of Bleed Channels and Surface Streaks due to Bleed Water Traveling along the Casing ??????????????????85 Figure 3.26 Shaft Defect due to Bleed Water Traveling along the Casing ????...86 Figure 4.1 Flow Chart for Phase I ???????????????????..94 Figure 4.2 Flow Chart for Phase II ???????????????????96 Figure 4.3 Flow Chart for Phase III ??????????????????.100 Figure 4.4 Example of Identification System for Ordinary Drilled Shaft Concrete 102 Figure 4.5 Example of Identification System for SCC Mixtures ???????.102 Figure 4.6 Flow Chart for Phase IV ??????????????????.106 Figure 4.7 Example of Identification System for Phase V ?????????..108 Figure 4.8 Flow Chart for Phase V ??????????????????..111 Figure 4.9 Particle Size Analyzer ???????????????????117 Figure 4.10 Laser Particle Size Analyzer Results ?????????????..118 Figure 4.11 Raw Material being Delivered from South Carolina ???????..119 Figure 4.12 No. 67 Coarse Aggregate Gradation for South Carolina Material ??.121 Figure 4.13 No. 789 Coarse Aggregate Gradation for South Carolina Material ?...121 xiv xiv Figure 4.14 Fine Aggregate Gradation for South Carolina Material ??????.122 Figure 4.15 No. 67 Coarse Aggregate Gradation for Alabama Material ????...122 Figure 4.16 No. 89 Coarse Aggregate Gradation for Alabama Material ????...123 Figure 4.17 Fine Aggregate Gradation for Alabama Material ????????...123 Figure 5.1 New Indoor Mixing Facility ????????????????...127 Figure 5.2 12 ft 3 Concrete Mixer used for this Research ??????????.129 Figure 5.3 Testing Equipment for Slump Test ??????????????.130 Figure 5.4 Testing Equipment for Slump Flow Test ???????????...132 Figure 5.5 Visual Stability Index Rating ????????????????..133 Figure 5.6 Testing Equipment for J-Ring Test ??????????????135 Figure 5.7 L-Box Dimensions ????????????????????..137 Figure 5.8 Testing Equipment for L-Box Test ??????????????.137 Figure 5.9 Testing Equipment for Segregation Column Test ?..???????..139 Figure 5.10 Testing Equipment for Bleeding Test ?????????????.142 Figure 5.11 Testing Equipment for Unit Weight and Air Content ???????143 Figure 5.12 Testing Equipment for Time of Setting ????????????..145 Figure 5.13 600 Kip Forney Compression Machine ????????.????.146 Figure 5.14 Concrete Specimen with Compressometer Attached ???????.148 Figure 5.15 Humboldt Length Comparator, Mold, and Concrete Specimen ???.149 Figure 5.16 Model 164 Test Set and Proove? It Cells ????????????150 Figure 6.1 Slump Flow vs. Concrete Age for Phase I ???????????..153 Figure 6.2 Slump Flow vs. Concrete Age for Phase II ???????????155 Figure 6.3 Penetration Resistance vs. Concrete Age for Phase II ???????156 xv xv Figure 6.4 Slump Flow vs. Concrete Age for Phase III ??????????...158 Figure 6.5 Workability of ODSC Mixture (approximately 8.25 inches) ????.161 Figure 6.6 Workability of SCC Mixture (approximately 20 inch slump flow) ?...161 Figure 6.7 Slump vs. Concrete Age ??????????????????.163 Figure 6.8 Slump Flow vs. Concrete Age ????????????????164 Figure 6.9 Slump vs. Concrete Age for all Concrete Mixtures ????????164 Figure 6.10 Penetration Resistance vs. Concrete Age ???????????...165 Figure 6.11 Compressive Strength vs. Concrete Age for Phase IV ???????171 Figure 6.12 Modulus of Elasticity vs. Concrete Age for Phase IV ???????.173 Figure 6.13 Predicted vs. Measured Elastic Modulus according to ACI 318 (2002) Equation for Phase IV ???????????????????174 Figure 6.14 Predicted vs. Measured Elastic Modulus According to ACI 363 (2002) Equation for Phase IV ????????????????.??...174 Figure 6.15 Drying Shrinkage vs. Concrete Age ?????????????...176 Figure 6.16 91-Day Permeability Results ????????????????..179 Figure 6.17 365-Day Permeability Results ????????????????.179 Figure 6.18 Attaining Raw Materials from Stock Piles ???????????..182 Figure 6.19 Unloading Raw Materials onto Conveyer Belt ??????????183 Figure 6.20 Raw Materials being Delivered to Hopper via Conveyer Belt ????183 Figure 6.21 Raw Materials being Mixed by Ready Mix Truck ????????..184 Figure 6.22 Slump Flow vs. Mixing Time under Laboratory and Field Conditions ..186 Figure 6.23 Temperature Profile Obtained from I-Button ???.???????.188 Figure 6.24 Penetration Resistance vs. Concrete Age under Laboratory and Field Conditions ???????????????????????...188 xvi xvi Figure 6.25 Compressive Strength vs. Concrete Age under Laboratory and Field Conditions ??????????????????????..?189 Figure 6.26 Modulus of Elasticity vs. Concrete Age under Laboratory and Field Conditions ?????...?????????????????....190 Figure 6.27 Compressive Strength vs. Concrete Age for VMA Mixtures????...201 Figure 6.28 Modulus of Elasticity vs. Concrete Age for VMA Mixtures ???.?.201 Figure 6.29 Compressive Strength vs. Concrete Age for Fly Ash Mixtures???...202 Figure 6.30 Modulus of Elasticity vs. Concrete Age for Fly Ash Mixtures??.?..202 Figure 6.31 Compressive Strength vs. Concrete Age for Silica Fume Mixtures ?....203 . Figure 6.32 Modulus of Elasticity vs. Concrete Age for Silica Fume Mixtures?.?203 Figure 6.33 Compressive Strength vs. Concrete Age for GGBFS Mixtures ???..204 Figure 6.34 Modulus of Elasticity vs. Concrete Age for GGBFS Mixtures ???..204 Figure 6.35 Compressive Strength vs. Concrete Age for Micron 3 Mixtures???205 Figure 6.36 Modulus of Elasticity vs. Concrete Age for Micron 3 Mixtures ???205 Figure 6.37 Compressive Strength vs. Concrete Age for Limestone Mixtures ??.206 Figure 6.38 Modulus of Elasticity vs. Concrete Age for Limestone Mixtures ??..206 Figure 6.39 Predicted vs. Measured Elastic Modulus According to ACI 318 (2002) Equation for Phase V ??????????????????......207 Figure 6.40 Predicted vs. Measured Elastic Modulus According to ACI 363 (2002) Equation for Phase V ??????????????????.?.207 Figure 6.41 Graphical Results of the Effect of VMA on Fresh Concrete Properties .210 Figure 7.1 Proposed Field Site ????????????????????.215 xvii xvii CHAPTER 1 INTRODUCTION 1.1 STATEMENT OF PROBLEM In recent years, drilled shaft concrete mixtures are facing increasing demands for fluidity. One of the primary reasons for this increased need for fluidity is the utilization of larger diameter shafts, deeper shafts, and congested rebar cages. The use of larger diameter shafts that are designed for large bending moments and seismic conditions require high amounts of reinforcement to be placed within the shaft. Consequently, the rebar cages have become progressively more congested and resistive to concrete flow. The addition of numerous access tubes for integrity testing has also lead to increased congestion in the rebar cages. These heavily congested rebar cages have lead to increased blockage due to the contact within the aggregates when the concrete is forced through the rebar cage. Even when a concrete mixture has sufficient workability, blockage could occur due to the bridging of the coarse aggregate at the vicinity of the reinforcement bars. In addition, tremie placement of drilled shaft concrete may require prolonged periods for concrete placement, which may result in a loss of concrete workability before the shaft is completed. It has been observed by experienced engineers that this loss of workability has lead to structural defects due to the entrapment of debris within the shaft. These large diameter and deep shafts also call for large amounts of concrete to be placed. Case studies have shown that with these massive concrete pours 1 large amounts of bleed water can be generated and begins to rise up in the column. The rising bleed water can result in larger interfacial transition zones, loss of bond between the reinforcement and concrete, surface streaking, and vertical bleed channels in the interior of the shaft. Many state DOT specifications have not kept appropriate workability considerations as a special aspect of drilled shaft concrete to meet these increasing demands for fluidity. For this reason, other viable materials such as self-consolidating concrete (SCC) could be possible solutions for this increased need for fluidity. Although the use of SCC has been largely implemented in the precast/prestressed industry, the potential benefits in drilled shaft construction are enormous. SCC has the potential to address many of the problems associated with drilled shaft construction because of the inherent workability, passing ability, resistance to segregation, and reduced bleeding of this type of concrete. The general requirements for SCC mixtures are as follows: ? Reduced volume of coarse aggregate: as the size and amount of coarse aggregate plays an important role on the passing ability of the concrete mixture. ? Increased volume of paste: the friction between the aggregate controls the spreading and the filling ability of the concrete. ? Increased volume of very fine material: this ensures sufficient workability while reducing the risk of segregation and bleeding. ? High dosage of high-range water reducing admixtures: these provide the necessary fluidity. ? Viscosity modifying admixture (VMA): these admixtures can potentially reduce bleeding and coarse aggregate segregation. 2 SCC has not currently been used for drilled shaft construction in North America. Some of the potential impediments include the general and overall lack of experience and research with these concrete mixtures for drilled shaft construction. As a result, extensive research must be conducted to generate interest in the construction community and to develop DOT acceptance of the use of SCC in drilled shaft construction, especially when inspection is difficult. 1.2 RESEARCH OBJECTIVES The primary objectives of this research are to evaluate the use of SCC in drilled shaft construction, identify appropriate testing techniques and characteristics for this specific application, and potential problems or concerns with the use of SCC in drilled shaft construction. A laboratory testing program as well as a later full-scale field study will examine the difference between ordinary drilled shaft concrete and SCC for both fresh and hardened properties. The fresh properties include filling ability, passing ability, segregation resistance, workability over time, bleeding characteristics, and controlled setting, while the hardened properties include the comparison of the compressive strength, elastic modulus, drying shrinkage, and permeability. It is anticipated that this research will lead to additional interest in this topic from state and national transportation agencies. While there is considerable research being directed toward SCC as a material, this research will be primarily focused upon the application of this technology to the drilled shaft industry. 3 1.3 RESEARCH SCOPE This research project was subject to a literature review of published material related to SCC. Chapter 2 of this thesis contains information regarding the factors that influence both the fresh and hardened concrete properties of SCC. These fresh properties primarily consist of the filling ability, passing ability, and segregation resistance, while the hardened concrete properties include compressive strength, modulus of elasticity, drying shrinkage, and permeability. Chapter 3 will address several aspects of design and construction that are essential for high-quality drilled shaft concrete and problems that are encountered in drilled shaft construction that leads to poor quality drilled shaft foundations. Selected examples of more common problems associated with drilled shaft concrete are cited in this chapter so that these problems can be understood. Chapters 4 and 5 present the laboratory testing program, raw materials, and testing procedures for both fresh and hardened concrete properties utilized for this research project. Chapter 6 provides an in-depth discussion and analysis of the laboratory testing program as well as providing for a comparison between ordinary drilled shaft concrete and SCC for fresh and hardened concrete properties. Chapter 7 presents a proposed field study to be conducted in South Carolina. This proposed field study will examine the difference between ordinary drilled shaft concrete and SCC for both fresh and hardened properties under simulated field conditions. Finally, Chapter 8 offers conclusions and recommendations based on the results and analysis provided in Chapter 6. 4 CHAPTER 2 LITERATURE REVIEW OF SELF-CONSOLIDATING CONCRETE A review of literature pertaining to relevant research topics associated with self- consolidating concrete (SCC) is presented in this chapter. The topics covered include a background of SCC, SCC testing procedures, fresh concrete properties of SCC, and hardened concrete properties of SCC. 2.1 INTRODUCTION In the early 1980s, durability issues related to concrete structures were a major concern and topic of interest in Japan. The gradual reduction in high-quality construction practice and concrete placement by unskilled labor resulted in deficient concrete structures (Okamura and Ouchi 1999). In an effort to produce durable concrete structures independent of the quality of construction practice, Professor Hajime Okamura at the University of Tokyo developed self-consolidating concrete that would be able to consolidate under its own weight without the need for external vibration. Self-consolidating concrete is able to fill formwork, encapsulate reinforcement bars, and consolidate under its own weight. At the same time it is cohesive enough to maintain its homogeneity without segregation or bleeding. This makes SCC useful in applications where placing concrete is difficult, such as heavily-reinforced concrete structures or where formwork is complex. Rilem Report 23 from Technical Committee 174-SCC (Skarendahl 2000) suggests that three main functional requirements of SCC are as follows: 5 1. Filling Ability: The ability of the concrete to completely fill formwork and encapsulate reinforcement without the use of external vibration. 2. Passing Ability: The ability of the concrete to pass through restrictive sections of formwork and tightly spaced reinforcement bars without blockage due to interlocking of aggregate. 3. Segregation Resistance: The ability of the concrete to keep particles in a homogenous suspension throughout mixing, transportation, and placement. The objective with the development of SCC is to overcome problems associated with conventional concrete that include improper consolidation, inability to pass and encapsulate reinforcement bars, and the incapability to adequately fill formwork. However, the acceptance and application of SCC in North America requires an orderly and conscious approach because suppliers, contractors, engineers, and architects are concerned with different aspects of the concrete?s performance (Khayat and Daczko 2003). In addition, this approach is imperative because SCC is being developed by multiple agencies with different approaches, and there is a lack of standardized tests to assess the quality of SCC. This has led well-known organizations, such as ASTM and RILEM, to address this issue and develop standardized testing procedures. As standardized tests and well-defined mixture proportioning become available, the familiarity and use among suppliers, contractors, engineers, and architects will increase with confidence allowing the user to quantify the benefits of SCC. Some of these benefits include reduced labor cost, superior finish, reduced need for surface patching, no vibration, and reduction in noise pollution. 6 2.2 SCC TESTING PROCEDURES The Filling Ability, Passing Ability, and Segregation Resistance of a SCC mixture must be evaluated through appropriate test methods to determine its quality. These properties are not independent from each other, but interrelated in some aspect. For example, the deformation capacity or filling ability is in part related to the viscosity, where the viscosity is strongly related to the segregation resistance. For that reason, some of the tests developed for SCC evaluate one or more of the fresh properties. Numerous tests have been developed for the evaluation of the fresh properties of SCC; however, a review of test methods that are relevant to this research and drilled shaft applications is presented in this section. The tests covered include the slump flow test, L- Box, J-Ring, and segregation column. The information presented in this section is offered only as a guide until standardized tests are developed for SCC testing. 2.2.1 Slump Flow Test The assessment of SCC to flow involves the evaluation of the filling ability or deformation capacity. The slump flow test is one of the most common and popular test to evaluate the deformation capacity of SCC because the procedure and apparatus are relatively simple (Takada and Tangtermsirikul 2000). The slump flow test is used to assess the filling ability of SCC in the absence of obstructions. The rate of deformability can be assessed by determining the T 50 time while the segregation resistance can to a certain degree be visually inspected. The T 50 time corresponds to the time that it takes for the concrete to flow 20 inches during the slump flow test. The slump flow test consists of filling an ordinary slump cone with SCC without any rodding. The cone is then lifted and the mean diameter is measured after the 7 concrete has ceased flowing. The apparatus for the slump flow test can be seen in Figure 2.1. The PCI (2003) states that it can be argued that because the slump flow test only assesses the free flow, unrestricted by boundaries, it is not representative of what happens in concrete construction. However, the slump flow test is useful in evaluating the consistency of SCC as delivered to the job site, and it gives some indication of segregation resistance by visual observation (PCI 2003). The slump flow test can be performed in the upright or inverted position, and the method utilized should be used consistently without switching from one to the other. The upright and inverted methods can be seen in Figures 2.1 and 2.2, respectively. The inverted method allows the slump cone to be easily filled by pouring the sample into a larger opening, which reduces the amount of spillage. Furthermore, the inverted method does not require a person to stand on the slump flow table because the weight of the concrete holds the cone downward onto the table. Ramsburg (2003) reports that it is possible that only one technician can conduct the slump flow by either method, though most users state that two technicians are needed for the upright position. However, one may argue that the two methods will produce different slump flow values. A recent study conducted by Oldcastle Inc. evaluated the difference in slump flow values between the two methods using three different mix designs with three different levels of performance (Ramsburg 2003). The three different levels of performance were slump flows less than 25 inches, slump flows over 25 inches, and slump flows over 25 inches with noted segregation and bleeding. The results indicate that there is only a minimal difference in the slump flow between the two methods. 8 Figure 2.1 ? Upright Slump Cone Method (PCI 2003) Figure 2.2 ? Inverted Slump Cone Method (PCI 2003) 9 As previously reported, the T 50 time corresponds to the time that it takes for the concrete to flow 20 inches during the slump flow test. The T 50 time is often used to evaluate the viscosity or rate of deformability of SCC. However, Takada and Tangtermsirikul (2000) reported that the T 50 time is not a direct measure of the viscosity of the mix independent of the slump flow value. For instance, a larger slump flow value will produce a lower T 50 time when the viscosity of the mix is constant. The T 50 time can be used to evaluate the difference in viscosity of mixtures only when the slump flow value is constant (Takada and Tangtermsirikul 2000). Additionally, the T 50 time can be used as an indication of production uniformity of a given SCC mixture. The PCI (2003) states that the T 50 time should not be used as a factor to reject a batch of SCC, but rather as a quality control test. However, recent provisions provided by the FDOT state that the T 50 time should be between 2 and 7 seconds for acceptance purposes (Mujtaba 2004). Ramsburg (2003) reports that the T 50 times are somewhat arbitrary due to the difficulty of starting and stopping a clock while conducting the slump flow test. This issue is more of a concern as the T 50 times become lower, where the possible intervals are less than 1.5 seconds. Furthermore, the tests conducted by Oldcastle Inc. show an increase in T 50 times when performing the inverted slump flow test (Ramsburg 2003). It was further determined that as the slump values increased, the difference in T 50 times were found to be less obvious between the two methods. Moreover, the inverted method could possibly improve the accuracy of the T 50 times when the intervals are only a few seconds; thus, a small difference in viscosity could be more noticeable with the inverted method (Ramsburg 2003). 10 One of the most critical requirements for SCC is that it must not segregate during or after placement. The slump flow test provides an indication of the segregation resistance by visual observation. Therefore, the visual stability index (VSI) was developed to determine the ability of a SCC mixture to resist segregation. The VSI procedure is to assign a numerical rating from 0 to 3, in increments of 0.5, based on the homogeneity of the mixture after the slump flow test has been conducted. To differentiate the textural properties of SCC, it should be ranked according to Table 2.1 and Figure 2.3. The basis for the VSI is that when the segregation resistance is not sufficient, the coarse aggregate will tend to stay in the center of the slump flow patty and mortar at the SCC border (Takada and Tangtermsirikul 2000, and PCI 2003). In the case of minor segregation, a border of mortar without coarse aggregate can occur at the edge of the slump flow patty (PCI 2003). Since the slump flow patty has no significant depth through which settlement of aggregate can occur, the visual inspection of SCC in the wheelbarrow or mixer should be part of the process in determining the VSI rating (PCI 2003). In fact, Bonen and Shah (2004) state that visual evaluation of segregation from the slump flow patty is an inadequate measure for predicting segregation resistance in the static state. Khayat et al. (2004) reports that the VSI rating of the slump flow patty is considered part of the dynamic stability given the fact the concrete can exhibit some non- uniform texture following some mixing and transport; whereas, the VSI can be considered as a static stability index when it is observed in the wheelbarrow or mixer following some period of rest time. 11 Table 2.1 ? Visual Stability Index (VSI) Rating (Khayat et al. 2004) Rating Criteria 0 No evidence of segregation in slump flow patty, mixer drum, or wheelbarrow 1 No mortar halo in slump flow patty, but some slight bleeding on surface of concrete in mixer drum and/or wheelbarrow 2 Slight mortar halo (<10mm) in slump flow patty and noticeable layer of mortar on surface of testing concrete in mixer drum and wheelbarrow 3 Clearly segregating by evidence of large mortar halo (>10mm) and thick layer of mortar and/or bleed water on surface of testing concrete in mixer drum or wheelbarrow Figure 2.3 - Visual Stability Index Rating (Degussa Construction Chemicals 2004) 12 2.2.2 L- Box Test The L-Box is used to assess the passing ability of a SCC mixture. This test is suitable for laboratory and perhaps for site purposes (PCI 2003). The apparatus consist of a rectangular-section box in shape of an ?L? with a vertical and horizontal section separated by a removable gate as shown in Figure 2.4. The L-Box is equipped with reinforcement bars separated by narrow openings that are designed to evaluate the passing ability of a SCC mixture. The reinforcement bars can be different diameters and spaced at different intervals (PCI 2003). The PCI (2003) suggests that three times the maximum aggregate size may be appropriate for the reinforcement spacing. The L-Box test is conducted by filling the vertical section of the L-Box with SCC, and then the removable gate is lifted to allow the SCC to flow into the horizontal section (PCI 2003). After the flow has ceased, the height of the SCC at the end of the horizontal section (H 2 ) and the remaining SCC in the vertical section (H 1 ) is measured and expressed as a blocking ratio. This blocking ratio (H 2 /H 1 ) is an indication of the passing ability of a SCC mixture. The closer the blocking ratio is to 1, the better the passing ability of the SCC mixture. Petersson (2000) reports that according to Swedish experience, a blocking ratio of 0.80- 0.85 is an acceptable range of values. 13 Figure 2.4 ? L-Box Testing Apparatus (PCI 2003) 2.2.3 J-Ring Test The J-Ring test is used to determine the passing ability of a SCC mixture. This test is suitable for laboratory and perhaps for site purposes (PCI 2003). The equipment consists of an open steel circular ring, drilled vertically to accept sections of reinforcement bars as shown in Figure 2.5. The rods can be different diameters and spaced at different intervals (PCI 2003). The PCI (2003) suggests that three times the maximum size aggregate may be appropriate for the reinforcement spacing. The spacing of the rods at different intervals will impose a more or less severe test of the passing ability depending on the application. The J-Ring test is used in conjunction with the slump flow test. The combination of the two tests will allow the assessment of the passing ability of the SCC mix design. 14 This test is performed in the same manner as the slump flow test with the addition of the J-Ring. The difference between the slump flow and the J-Ring flow is compared, and then a passing ability rating is assigned according to Table 2.2. The larger the difference between the slump flow and the J-Ring flow indicates less passing ability. Recent provisions provided by the FDOT state that the difference between the slump flow and the J-Ring flow should be no more than 2 inches (Mujtaba 2004). Table 2.2 - J-Ring Passing Ability Rating (ASTM J-Ring Draft 2004) Difference between Slump Flow and J-Ring Flow Passing Ability Rating Remarks 0 ? X ? 1 inch 0 No visible blocking 1 < X ? 2 inches 1 Minimal to noticeable blocking X > 2 inches 2 Noticeable to extreme blocking Figure 2.5 - J-Ring Testing Apparatus (ASTM J-Ring Draft 2004) 15 2.2.4 Segregation Column Test The segregation column test is used to determine the stability and segregation resistance of a SCC mixture. This test can be used for both laboratory and perhaps site purposes (ASTM Segregation Column Draft 2004). The equipment consists of an 8-inch diameter by 26-inch tall schedule 40 PVC pipe. The PVC pipe is separated into 4 equal sections each measuring 6.5-inches in height. A collector plate measuring 20-in. x 20-in. is used to collect the concrete from the different sections of the column. The segregation column and the collector plate can be seen in Figure 2.6. The segregation column test is conducted by placing a sample of concrete in the cylinder mold in one lift without any means of mechanical vibration. After the concrete is allowed to sit for 15 minutes, the concrete column is separated into four equal sections using the collector plate. The concrete from the top and bottom section is wet-washed through a No. 4 sieve leaving the coarse aggregate on the sieve. The mass of aggregate from these sections of the column is obtained, and a segregation index is determined using Equation 2.1. MB TB CA CACA SI )( ? = Eq. 2.1 In Equation 2.1, SI is the segregation index, CA T is the mass of coarse aggregate in the top section, CA B is the mass of coarse aggregate in the bottom section, and CA BM is the mass coarse aggregate per section of the column according to Equation 2.2. [ ] MMBM CACACA ??= 0052.0007.0 Eq. 2.2 In Equation 2.2, CA M is the mass of coarse aggregate in 1 yd 3 of concrete. 16 Figure 2.6 - Segregation Column Testing Apparatus (ASTM Segregation Column Draft 2004) 17 2.3 FRESH CONCRETE PROPERTIES SCC is characterized by its filling ability, passing ability, and segregation resistance. SCC has to have a low yield stress value to ensure high flowability, small aggregate particles to prevent blockage, and adequate viscosity to prevent segregation. Thus, these characteristics must be discussed in further detail to allow the user to develop a well-designed SCC mixture. 2.3.1 Rheology Before discussing the various ways to modify SCC characteristics, it is helpful to discuss a few basic principles of rheology. Rheology can be described as the study of deformation and flow of matter under stress (Mindess et al. 2003). Understanding and knowledge of rheology behavior has been essential in the development of self- consolidating concrete and influences the performance of the fresh concrete properties. The rheology of concrete, paste, or mortar may be characterized by its yield stress and plastic viscosity. The rheology of fresh concrete, including self-consolidating concrete, is most often defined by the Bingham Fluid Model using Equation 2.3 (Mindess et al. 2003). ???? *+= o Eq. 2.3 In Equation 2.3, ? is the shear stress in psi, ? o is the yield stress in psi, ? is the plastic viscosity in psi . seconds, and ? is the shear strain in 1/seconds. Figure 2.7 shows that the Bingham Fluid Model requires a yield stress to obtain a strain that is followed by increasing shear stress with increasing shear strain (Khayat and Tangtermsirikul 2000). Khayat and Tangtermsirikul (2000) report that the target rheological properties for SCC are a low yield stress value together with an adequate plastic viscosity. 18 Figure 2.7 - Bingham Fluid Model A number of rheometers, e.g. BML viscometer and BTRHEOM rheometer, have been developed to measure the true rheological properties of fresh concrete. Rheometers provide two parameters, namely the initial yield stress value and the plastic viscosity, to characterize the fresh properties. These rheometers are useful in evaluating what the effects of different materials, such as cements, fillers, aggregates, mineral admixtures, and chemical admixtures have on the yield stress and plastic viscosity. According to Emborg (1999), rheometers are considered to be the most accurate way to describe the real behavior of fresh concrete. However, rheometers are based on different principles, and the results from different rheometers can not be easily compared. While it is likely that the use of rheology tests and rheometers will increase in the future, rheometers are expensive to purchase and existing tests are primarily used for initial mixture proportioning, testing, and research efforts. 19 2.3.2 Filling Ability Self-consolidating concrete must be able to fill formwork and encapsulate reinforcement without the use of external vibration. The high deformation capacity of SCC is related to the yield stress; thus, the yield stress must be reduced in order to ensure that SCC can flow around obstacles and achieve good filling ability. The deformation capacity can be increased by the reduction of interparticle friction between the solid particles, which include the paste, coarse aggregate, and fine aggregate (Khayat and Tangtermsirikul 2000). The interparticle friction between the paste particles requires the dispersion of fine material by superplasticizers. Khayat and Tangtermsirikul (2000) state that unlike water that reduces both the yield stress and viscosity, superplasticizers reduce the yield stress and cause a limited decrease in viscosity. As a result, the addition of superplasticizers can provide highly flowable concrete without a significant reduction in cohesiveness. In order to reduce interparticle friction due to aggregate-aggregate contact, Khayat and Tangtermsirikul (2000) recommend that the interparticle distance between the aggregate be increased. This is achieved by reducing the total aggregate content and increasing the paste content. Thus, the following actions should be taken to achieve adequate filling ability (Khayat and Tangtermsirikul 2000): 1. Increase the deformability of the paste ? Balanced water-to-cementitious materials ratio (Balanced so that adequate deformability and deformation velocity can be achieved) ? Superplasticizers 2. Reduced interparticle friction ? Low coarse aggregate volume ? Higher paste contents 20 Although different superplasticizers are available in the market today, almost all new and innovative superplasticizers are polycarboxylate based mixtures (Bonen and Shah 2004). It must be noted that the following discussion is based on the work of Bury and Christensen (2003). These polycarboxylate based superplasticizers are characterized by their strong dispersing action, controlled slump retention, enhanced concrete stability, enhanced pumping ability, and enhanced finishing ability. These superplasticizers function by imparting a negative charge on the cement particles that cause them to repel from one another. Traditional superplasticizers also function in this manner, but the new polycarboxylate based superplasticizers have side chains with varying lengths that aid in keeping the cement particles apart allowing water to surround more surface area of the cement particle (steric hindrance). The effectiveness of superplasticizers last only as long as there is sufficient molecules available to cover the surface area of the cement particles. Therefore, it is likely that with time and prolonged mixing the effectiveness of the superplasticizers will become inadequate and the workability of the mix will be lost. Repeated addition of superplasticizers may be beneficial from the standpoint of workability; however, it may increase bleeding, segregation, and change the amount of entrained air (Neville 1996). Neville (1996) goes on to suggest that the workability regained from the re-dosage may decrease at a faster rate. Therefore, the re-dosage should be applied immediately prior to placement according to the recommendations of the admixture supplier. Bonen and Shah (2004) state that the use of fine material, such as silica fume, will increase the yield stress because of greater water absorbance. Thus, it would be expected that the superplasticizer dosage should be increased to obtain the same slump flow. 21 Ferraris et al. (2001) reported that with a constant amount of water and cementitious material, the addition of 8% and 12% silica fume increased the superplasticizer dosage by 30% and 50% over the control mix. In fact, it is usually reported that if the volume concentration of solids is held constant, the addition of a fine material will decrease the workability (Ferraris et al. 2001). The most common reason for this reduction in workability is due to increased surface area of the fine material, which will increase the superplasticizer demand for the same water content. However, it is reported in some cases that the addition of fine material can decrease the water demand. Ferraris et al. (2001) states that the reduction of water demand for fine material, especially fly ash, is due to spherical particles that easily move past each other reducing the interparticle friction. The aggregate shape also influences the filling ability to a certain degree. It is reported that flat and elongated particles will lead to a decrease in the workability (Hodgson 2003). This is due to the fact that angular aggregate will have more mechanical interlocking and will need more work to overcome interparticle friction. For example, Petersson (1999) states that when crushed sand is used, the fluidity is decreased for the same amount of superplasticizer. Conversely, rounded aggregates will act like ?ball bearings? allowing the particles to easily move past each other, which will increase the workability for a constant paste and water content. It is generally considered that rounded and smaller aggregate particles will increase the filling ability of concrete. 22 2.3.3 Passing Ability The level of passing ability of SCC is a function of the stability, coarse aggregate content, coarse aggregate size, and reinforcement spacing. SCC with excellent deformability but with insufficient cohesiveness will lead to local aggregate segregation between the paste and coarse aggregate at the vicinity of the reinforcement that could lead to severe blockage (Khayat et al. 2004). This will not only lead to decreased passing ability, but it will also lead to an increase in the concentration of coarse aggregate at the reinforcement. The passing ability is also affected when the coarse aggregate size is large and/or the coarse aggregate content is high. This mechanism of blocking can be explained by the two dimensional model shown in Figure 2.8 (Khayat and Tangtermsirikul 2000). Figure 2.8 illustrates that the aggregate particles around an opening must change their path of travel to properly pass through the reinforcement. As a result, aggregate particles may collide at the reinforcement opening. This aggregate interaction may cause the aggregate to form a stable arch at the vicinity of the reinforcement opening (Khayat and Tangtermsirikul 2000). Therefore, to achieve suitable passing ability the following steps should be considered (Khayat and Tangtermsirikul 2000): 1. Enhance the cohesiveness to reduce segregation of aggregate ? Low water-to-cementitious materials ratio ? Viscosity modifying admixture 2. Compatible clear spacing and aggregate characteristics ? Low coarse aggregate contents ? Smaller maximum aggregate size 23 Figure 2.8 - Mechanism of Blocking (Khayat and Tangtermsirikul 2000) Khayat et al. (2004) conducted studies on the passing ability by evaluating the dynamic stability of SCC using the L-Box and J-Ring apparatus. It was reported that SCC mixtures prepared with 843 lb/yd 3 of cement with relatively low viscosity exhibited greater passing ability than SCC mixtures prepared with 650 lb/yd 3 of cement. Khayat et al. (2004) reported that the concrete mixtures prepared with 650 lb/yd 3 contained more coarse aggregate that increased the risk of collision and interaction among solid particles, which lead to a reduced ability to flow between the reinforcement bars. The greater tendency of aggregate blockage resulted in lower passing ability in both the J-Ring and L-Box tests. Studies conducted by Kim et al. (1998) also indicate that the passing ability was increased with decreasing coarse aggregate content. The results further indicate that SCC mixtures prepared with volume ratios of coarse aggregate-to-concrete of 0.27 and 0.31 demonstrated a higher passing ability than volume ratios of 0.35 and 0.39 at a constant water-to-binder ratio. Studies conducted at the Swedish Cement and Concrete Institute investigated the blocking in the L-Box apparatus using different maximum size aggregates and reinforcement spacing (Petersson 1999). The paste content in the mixes remained constant with slump flow values ranging from 25.5 to 28.5 inches. The results in Figure 24 2.9 indicate that as the reinforcement spacing increases relative to the maximum aggregate size the passing ability is also increased. Furthermore, closer examination of Figure 2.9 reveals that for the same reinforcement spacing the use of smaller aggregates produced higher blocking ratios. Thus, it can be concluded that the maximum size aggregate and reinforcement spacing has an effect on the passing ability of SCC. Figure 2.9 - Blocking Ratios versus Reinforcement Spacing for Different Size Aggregate (Petersson 1999) 2.3.4 Segregation Resistance One of the most important requirements of SCC is that it must not segregate during or after placement. SCC is much more prone to segregation than normal concrete due to the sharp reduction in viscosity caused by the high dosages of superplasticizers. Thus, it is essential that SCC have a high resistance to segregation so that there is 25 homogenous distribution of materials in the hardened concrete. Khayat and Tangtermsirikul (2000) report that SCC should not show signs of segregation in static or dynamic conditions, which include bleeding of water, paste and aggregate segregation, and non-uniformity in air pore size distribution. Thus, the following steps should be considered to produce sufficient segregation resistance (Khayat and Tangtermsirikul 2000): 1. Reduce the segregation of solid particles ? Reduced maximum size aggregate ? Low water-to-cementitious materials ratio ? Viscosity modifying admixture 2. Minimize bleeding due to free water ? Low water content ? Low water-to-cementitious materials ratio ? Powders with high surface area ? Viscosity modifying admixture Segregation of aggregate is related to a number of variables that consist of the particle size, particle specific gravity, and the proportions of the mixture. Bonen and Shah (2004) report that the use of larger coarse aggregates will settle much faster than smaller coarse aggregates when the density and viscosity of the suspension is held constant. Additionally, gradation is also an important factor in determining the proper coarse aggregate, especially where reinforcement is highly congested and/or the formwork has small dimensions. Hodgson (2003) reports that a gap-graded coarse aggregate promotes a greater degree of segregation than well-graded coarse aggregate. Thus, the coarse aggregate chosen for SCC is typically round in shape, well-graded, and smaller in maximum size than that used for conventional concrete. 26 Another method to increase the segregation resistance besides reduction in aggregate size is to increase the cohesiveness of the mixture. This is typically done by one of the three methods listed below. All three methods use superplasticizers to increase the fluidity of the mixture, but the difference between them is the method used to prevent the segregation (Hodgson 2003). 1. Powder method 2. Viscosity modifying admixture method, ?VMA? method 3. Combination method The powder method utilizes an increase in the volume of fines and low water content to reduce the amount of free water. Free water is defined as water that is not adhered to the solid particles and move independently from the solids in a mixture. Furthermore, Khayat and Tangtermsirikul (2000) state that it is essential to reduce the amount of free water in the mixture because an increase in free water content will decrease the viscosity of a SCC mixture. The most common methods for reducing the amount of free water is to use powder materials with a high surface area, use a low water- to-binder ratio, or both. For example, for a constant water content, the addition of high surface area material can absorb a greater amount of free water compared to cement particles. Thus, the plastic viscosity of the mix is increased due to greater water absorbance. Furthermore, the reduction of the water-to-binder ratio will increase the phase-to-phase cohesion that will increase the segregation resistance. Khayat et al. (1999) states that fine powder includes cement and supplementary cementitious material that is combined to enhance grain-size distribution, packing density, and reduced interparticle friction to lower the water demand for a necessary viscosity. 27 The VMA method utilizes lower cement contents, a superplasticizer, and a viscosity modifying admixture (VMA). The addition of a VMA may increase the viscosity of a mix to the extent that the water-to-cementitious ratio need not be increased (Bonen and Shah 2004). VMAs can provide adequate stability, reduce bleeding, and segregation resistance over a wider range of fluidity. There are two basic types of VMAs that are available in the market, and each VMA is based on the mechanism in which they function (Degussa Construction Chemicals 2004): 1. Thickening Type VMA- This VMA functions by thickening the concrete, making it cohesive without significantly affecting the fluidity. By increasing the viscosity of the mixture, the VMA makes the mixture more stable and less prone to segregation (Degussa Construction Chemicals 2004). These are typically polyethylene glycol (PEG) based VMAs. 2. Binding Type VMA- This VMA functions by binding the water within the concrete mixture. This VMA not only increases the viscosity of the mixture, but it also reduces bleeding. The binding type VMA is more potent in modifying the viscosity of the mixture than the thickening type (Degussa Construction Chemicals 2004). Welan Gum is an example of this type of VMA. Bury and Christensen (2003) state that the use of a VMA also increases the number of applications for SCC because more mixtures can be proportioned for a wider range of applications. For example, a VMA can be used with mixtures made with gap-graded materials. In fact, Berke et al. (2003) suggests that when poorly graded material and low powder contents are used, the use of a VMA can prove invaluable. In addition, because moisture contents in fine aggregate can change throughout daily operation, the use of a 28 VMA has been proven to be very valuable in overcoming deficiencies due to uncontrolled moisture (Bury and Christensen 2003). The combination approach utilizes a VMA with limited water content. The VMA in this method primarily is used to reduce the variability of the SCC that can arise from uncontrolled moisture and placement conditions. The VMA also controls bleeding and renders the concrete more robust, while the low water content provides the necessary viscosity (Khayat et al. 1999). 2.4 HARDENED CONCRETE PROPERTIES The hardened properties of concrete are often the most valued by design and quality control engineers. It has become evident that SCC can have a large variation in mechanical properties due to the different formulations used. Despite these variations, literature has shown that the mechanical properties of well-designed SCC are comparable or better than the corresponding properties of conventional concrete (Bonen and Shah 2004). Although many mechanical properties can be evaluated and compared, only the mechanical properties that are relevant to this research will be discussed in this section. 2.4.1 Compressive Strength The property that is most often specified for concrete design and quality control is the compressive strength. The testing of the compressive strength is relatively easy to perform in the laboratory, and the compressive strength is universally accepted as a general index of concrete strength. There are many determining factors that influence the 29 compressive strength of concrete; however, the compressive strength is best described by the water-to-cementitious ratio and the porosity relationship. When fully compacted, the concrete strength is taken to be inversely proportional to the water-to-cementitious ratio. In 1919, Duff Abrams established the following relationship between the water-to-cementitious ratio and the compressive strength (Neville 1996). cw c K K f / 2 1 = Eq. 2.4 In Equation 2.4, f c is the compressive strength, w/c is the water-to-cementitious ratio, and K 1 and K 2 are empirical constants. However, at water-to-cementitious ratios less than 0.38 the maximum possible hydration of the cement is less than 100%. Therefore, the slope of strength gain rate is slowed as the water-to-cementitious ratio is reduced as shown in Figure 2.10. Figure 2.10 - Strength versus Cement-to-Water Ratio (Neville 1996) 30 The influence of the water-to-cementitious ratio on compressive strength is not truly a constitutive law because the water-to-cementitious ratio rule does not include all factors that influence the compressive strength (Neville 1996). Neville (1996) suggests that it may be more appropriate to relate the compressive strength of concrete to the concentration of solid products of hydration of cement in the space available for these products. The porosity relationship can be considered one of the most important factors in cement based material because it affects both the cement paste matrix and the interfacial transition zone (ITZ). In general, there exist a relationship between the porosity and the strength of solids that can be described by Equation 2.5 (Neville 1996). ( ) n cc pff ?= 1 0, Eq. 2.5 In Equation 2.5, f c is the strength of the material, f c,o is the strength at zero porosity, n is a coefficient, and p is the porosity. The coefficient n depends on factors that include the cement composition, paste age, and temperature. This relationship between porosity and compressive strength can be seen in Figure 2.11. Figure 2.11 ? Compressive Strength versus Porosity (Neville 1996) 31 Powers and Brownyard determined that the compressive strength is related to the gel-to-space ratio (Mindess et al. 2003). The gel-to-space ratio is defined as the ratio of the gel volume over the summation of the volume of the gel, capillary pores, and air voids. Powers and Brownyard concluded that the compressive strength of the hydrated cement is closely related to the following equation: Eq. 2.6 3 Xaf c ?= In equation 2.6, f c is the strength of the material at a given porosity (p), a is the intrinsic strength of the material at zero porosity, and x is the gel-to-space ratio. Where this ratio is defined as the summation of the hydrated cement paste to the sum of the volume of the hydrated cement and capillary pores. The relationship between the gel-to-space ratio and the compressive strength can be seen in Figure 2.12. Bonen and Shah (2004) report that the compressive strength of SCC is also best approximated by the porosity content. They further suggest that similar to the Powers and Brownyard formulation, the compressive strength of SCC is defined in terms of the binder-to-space ratio, b rather than the gel-to- space ratio, x. The binder-to-space ratio is defined as the binder volume over the summation of the volume of the binder, capillary pores, and air voids. The binder volume is the sum of the gel volume, filler volume, and the solid volume of the superplasticizers and VMA. 32 Figure 2.12 - Strength versus Gel-to-Space Ratio Based on Tested Mortar Cubes (Neville 1996) In most cases, high contents of fine material are placed within SCC to increase segregation resistance. This addition of fine material can be capable of producing a denser microstructure that will decrease both the porosity of the cement paste matrix and interfacial transition zone of SCC. A study conducted at the Swedish Cement and Concrete Research Center investigated the microstructure in SCC and conventional bridge concretes using image analyzing and light microscopy techniques (Tragardh 1999). It was concluded that the porosity of the bulk paste and the ITZ was significantly higher in the conventional concrete compared to SCC with the same water-to- cementitious ratio (Tragardh 1999). It must be noted that although the SCC and conventional concrete had the same water-to-cementitious ratio, the SCC incorporated high amounts of fine inert limestone filler that produced a low water-to-binder ratio. The incorporation of fine material in the SCC allowed particles to pack more efficiently 33 around the aggregates; therefore, leading to a decrease in porosity around the ITZ. Tragardh (1999) further concluded that pores were more evenly distributed between the ITZ and bulk paste, and the effect of microbleeding, which leads to an increase of the local water-to-cementitious ratio at the interfacial zone, was found to be much less in SCC. It was further determined that the compressive strength for SCC was higher than the conventional concrete at the same water-to-cementitious ratio due to this improvement in the microstructure (Tragardh 1999). The results from the compressive strength test can be seen in Figure 2.13. This investigation determined that the microstructure properties not only improved in SCC, but shows a strong correlation between these microstructure properties and the measured properties (Tragardh 1999). Figure 2.13 ? Average Compressive Strengths of SCC and Conventional Bridge Concretes with w/c of 0.40 (Tragardh 1999) A study conducted at the Master Builders Research and Development Center also compared the engineering properties of SCC and conventional concrete (Attiogbe et 34 al. 2003). The study used both conventional concrete and SCC with a cement content of 640 lb/yd 3 and 160 lb/yd 3 of fly ash with a water-to-cementitious ratio of 0.37. The specimens were either steamed or air cured in molds for 24 hours. Compressive tests were then conducted at the ages of 1 and 28 days. The research indicated that the early- age porosity was lower in the SCC than the conventional concrete (Attiogbe et al. 2003). Attiogbe et al. (2003) reported that the compressive strengths for 1 day and 28 day of the steamed-cured SCC were comparable to the steamed-cured conventional concrete. It was further reported that the compressive strength for the air-cured SCC exceeded the strengths of the air-cured conventional concrete. Furthermore, Turcry et al. (2003) reported on values that indicate that at similar water-to-cementitious ratios, the compressive strength of SCC is comparable or higher than conventional concrete. The results from this study can be seen by looking at Table 2.3. Table 2.3 - Compressive Strength Results from Turcry et al. (2003) SCC 1 OC 1 SCC 2 OC 2 Gravel (lb/yd 3 ) 1331 1803 1424 1912 Sand (lb/yd 3 ) 1449 1314 1364 1348 Cement (lb/yd 3 ) 590 607 590 665 Limestone Filler (lb/yd 3 ) 253 0 421 0 Water (lb/yd 3 ) 315 286 269 295 Water/Cement Ratio 0.53 0.47 0.46 0.44 Water/Powder Ratio 0.37 0.47 0.27 0.44 Sand/Aggregate Ratio (by mass) 0.52 0.41 0.48 0.41 Compressive Strength (psi) 6960 7400 8700 7251 Modulus of Elasticity (1 x 10 6 psi) 4.35 5.50 5.20 5.20 OC = Ordinary Concrete 35 2.4.2 Modulus of Elasticity For structural design, the modulus of elasticity of concrete increases approximately with the square root of its strength. According to the ACI 318 (2002), the modulus of elasticity of concrete can be best approximated by Equation 2.7. '33 51 c . cc fwE = Eq 2.7 In Equation 2.7, E c is the modulus of elasticity in psi, w c is the unit weight in lb/ft 3 , and f c ? is the compressive strength in psi. It is suggested that this formulation is valid for density ranging from 90 to 155 pcf and compressive strengths up to 6000 psi. The ACI Committee 363 (2002) ?State-of-the-Art- Report on High Strength Concrete? also found that the equation used by ACI 318 (2002) was valid only up to compressive strengths up to 6000 psi. According to the ACI Committee 363 (2002) ?State-of-the-Art- Report on High Strength Concrete?, high strength concrete has compressive strengths ranging from 6,000 to 12,000 psi. By comparing stress-strain curves for normal and high strength concrete, high strength concrete has a more linear and steeper slope in the ascending portion of the curve as well as a higher strain at the maximum stress (Jones 2004). Therefore, the ACI Committee 363 (2002) recommends the following equation for estimating the modulus of elasticity for high strength concrete, which is valid for compressive strengths ranging from 3,000 to 12,000 psi. In Equation 2.8, E c is the modulus of elasticity in psi and f c ? is the compressive strength in psi. 000,000,1'000,40 += cc fE Eq. 2.8 36 In contrast to the compressive strength, the modulus of elasticity of concrete is significantly affected by the modulus and volume fraction of the aggregate (Bonen and Shah 2004). Bonen and Shah (2004) report that as the volume fraction of aggregate is increased, the modulus of elasticity is increased. This is an important factor for SCC since the aggregate volume fraction is typically lower compared to conventional concrete. It is then expected that the elastic modulus of SCC would be lower than conventional concrete with the same strength (Bonen and Shah 2004). In addition to the limited aggregate volume fraction, SCC typically incorporates a higher sand-to-aggregate ratio to increase its segregation resistance and passing ability. Therefore, a closer examination must be presented to determine how the sand-to-aggregate ratio and aggregate volume fraction influence the modulus of elasticity. Su et al. (2002) investigated the effect of the sand-to-aggregate ratio on the elastic modulus of SCC. Their study consisted of varying the sand-to-aggregate ratio of SCC from 0.30 to 0.55 while maintaining a constant aggregate volume fraction of 0.6 and a water-to-cementitious ratio of 0.39. By maintaining a constant water-to-cementitious ratio and aggregate volume fraction, only the varying sand-to-aggregate ratio should influence the elastic modulus. Su et al. (2002) also invested how the elastic modulus of the fine and coarse aggregate influence the modulus of elasticity. The results from this study can be seen by looking at Figure 2.14. Su et al. (2002) reported that when the elastic modulus of the fine aggregate was 2 times that of the coarse aggregate, the elastic modulus of concrete increased from 4.05 x 10 6 psi to 4.3 x 10 6 psi when the sand-to- aggregate ratio increased from 30% to 47.5%. It is further reported that when the elastic modulus of the fine aggregate was half of the coarse aggregate, the elastic modulus 37 decreased from 3.26 x 10 6 psi to 3.05 x 10 6 psi when the sand-to-aggregate ratio increased from 30% to 47.5%. Su et al. (2002) concluded that when the elastic modulus of the fine and coarse aggregate are not much different and the total volume fraction of aggregate is constant, the elastic modulus of SCC is not significantly affected by the sand-to- aggregate ratio (Su et al. 2002). Figure 2.14 - Elastic Modulus versus Sand-to-Aggregate Ratio (Su et al. 2002) Turcry et al. (2003) further compared the elastic modulus of SCC to conventional concrete using two SCC mixtures and two conventional concrete mixtures. Turcry et al. (2003) maintained a sand-to-aggregate ratio of 0.41 for the conventional concrete and sand-to-aggregate ratios of 0.52 and 0.48 for the SCC mixtures. Table 2.3 indicates that the conventional concrete had a higher aggregate fraction as well as a lower sand-to- aggregate ratio. Turcry et al. (2003) concluded that due to the higher paste volume for SCC, the elastic modulus for SCC was lower than conventional concrete at the same 38 compressive strengths. On the other hand, Attiogbe et al. (2003) reported that the modulus of elasticity for the SCC mixtures corresponded well with the conventional concrete when the strength was held constant. The study conducted by Attiogbe et al. (2003) used the same water-to-cementitious ratio of 0.37 with sand-to-aggregate ratios by mass of 0.44 for conventional concrete and 0.53 for the SCC mixtures. The concrete mixtures contained cement and fly ash of 650 lb/ft 3 and 160 lb/yd 3 , respectively. Therefore, both the conventional concrete and SCC has comparable aggregate volume fractions with varying sand-to-aggregate ratios (Table 2.3). 2.4.3 Drying Shrinkage Shrinkage is a term that represents the strain caused by the loss of water from hardened concrete (Mindess et al. 2003). Shrinkage of concrete is a function of the paste properties, and the response of the paste to moisture loss is modified by several parameters. However, the most important parameter is exerted by the aggregate, which will restrain the shrinkage (Neville 1996). Equation 2.9 states that the ratio of shrinkage of concrete, S c , to the shrinkage of neat cement, S p , depends on the aggregate content, a, and the experimental values of n range from 1.2 to 1.7 (Neville 1996). Eq. 2.9 n pc aSS )1( ?= Since SCC incorporates high paste volumes and reduced aggregate content, Equation 2.9 reveals that SCC is prone to an increase in drying shrinkage compared to conventional concrete. Neville (1996) suggest that for a constant water-to-cementitious ratio, the reduction in aggregate content will cause an increase in drying shrinkage. This relationship between water-to-cementitious ratio and aggregate content can be seen in Figure 2.15. 39 Figure 2.15 - Aggregate Content versus Water-to-Cement Ratio (Neville 1996) Neville (1996) reports that the size and gradation of aggregate per se does not influence the magnitude of shrinkage, but the use of larger aggregate permits a leaner mix that results in lower shrinkage. On the other hand, the ACI Committee 209 (2002) suggests that the gradation of the aggregate may have an effect on the drying shrinkage. According to ACI Committee 209 (2002), for different sand-to-aggregate ratios the shrinkage should be modified according to Equation 2.10 and 2.11. It can be seen from Equation 2.10 and 2.11 and Figure 2.16 that as the sand-to-aggregate ratio is increased, which indicates more fine aggregate, the shrinkage will also increase. In Equation 2.10 and 2.11, ? is the sand-to-aggregate ratio. ?014.030.0 +=Shrinkage %5010mm) and thick layer of mortar and/or bleed water on surface of testing concrete in mixer drum or wheelbarrow Figure 5.5 - Visual Stability Index Rating (Degussa Construction Chemicals 2004) 133 The slump flow test was frequently performed in the inverted position in conjunction with the J-Ring test. Under these circumstances, the mold was placed on the base plate with its 4 inch diameter facing downward. Furthermore, the slump flow tests were conducted by two technicians. The second technician timed the test while the other lifted the mold. The use of such practice helped reduce the operator error and improve the accuracy of the T 50 times. 5.3.3 J-Ring Test The J-Ring test was performed on the SCC mixtures via the procedure recommended by the PCI (2003). The purpose of this test was to determine the passing ability of a SCC mixture by comparing the slump flow diameter to the J-Ring diameter. Since the slump flow test was used in conjunction with the J-Ring test, the equipment required to perform the J-Ring test is similar to the slump flow test with the addition of the J-Ring as shown in Figure 5.6. The J-Ring apparatus consisted of a rectangular section 1 1/8 by 1 inch open steel circular ring that measured approximately 12 inches in diameter. The open steel circular ring was drilled vertically to accept 16-5/8 inch diameter reinforcement bars measuring 4 inches in height with a center to center spacing of approximately 2 5/16 inches. 134 Figure 5.6 ? Testing Equipment for J-Ring Test Procedure (PCI 2003): 1. Moisten all equipment and place on a flat and rigid surface. 2. After the mixture procedure is complete, normally sample approximately 0.2 ft 3 of SCC in accordance with ASTM C 172 (1998). 3. Position the mold, with its 4 inch diameter facing downwards, in the center of the base plate. 4. Place the SCC sample into the mold by means of a bucket. 5. Strike off any excess SCC by means of rolling the tamping rod over the surface. 6. Remove any excess SCC around the base of the mold by means of an absorbent sponge. 7. Lift the mold vertically a distance of 9 +/- 3 inches in 3 seconds without lateral movement allowing the concrete to flow out freely. 8. Measure the final diameter of the SCC in two perpendicular directions. 9. Calculate the average of the two diameters and report the result to the nearest ? inch. 135 10. Conduct the slump test using the inverted method with the provided procedure in Section 5.2.2. 11. Calculate the difference between the J-Ring diameter and the slump flow diameter of the companion test. 12. Assign a passing ability rating according to Table 5.2. Table 5.2 - J-Ring Passing Ability Rating (ASTM J-Ring Draft 2004) Difference between Slump Flow and J-Ring Flow Passing Ability Rating Remarks 0 ? X ? 1 inch 0 No visible blocking 1 < X ? 2 inches 1 Minimal to noticeable blocking X > 2 inches 2 Noticeable to extreme blocking 5.3.4 L-Box Test The L-Box test was performed on the SCC mixtures via the procedure recommended by the PCI (2003). The purpose of this test was to determine the passing ability of a SCC mixture by determining the extent of blocking by reinforcement. The equipment required to perform the L-Box test is listed below and can be seen in Figure 5.7 and 5.8. Equipment: 1. L-Box made of non-absorbing material with the dimensions shown in Figure 5.7 2. 5/8 inch tamping rod 3. Scoop 4. Tape measure 136 Figure 5.7 ? L-Box Dimensions (PCI 2003) Figure 5.8 ? Testing Equipment for L-Box Test 137 Procedure (PCI 2003): 1. Moisten all equipment and place on flat and rigid surface. 2. After the mixture procedure is complete, sample approximately 0.5 ft 3 of SCC in accordance with ASTM C 172 (1998). 3. Ensure that the sliding gate can open freely and then close the gate. 4. Fill the vertical section of the L-Box with the SCC mixture. 5. Let the SCC mixture stand for approximately 1 minute. 6. Strike off any excess SCC by means of rolling the tamping rod over the surface. 7. Lift the gate and allow the SCC mixture to flow into the horizontal section. 8. When the flow has ceased, measure and record the ?H1? and ?H2? dimensions, shown in Figure 5.7. 9. Calculate the blocking ratio, H2/H1. 5.3.5 Segregation Column Test The segregation column test was performed on the SCC mixtures via the procedure recommended by an ASTM Segregation Column draft (2004). The purpose of this test was to determine the stability and segregation resistance of a SCC mixture. The equipment required to perform the segregation column test is listed below and can be seen in Figure 5.9. Equipment: 1. Balance or scale accurate to 0.1 lb. or to within 0.3% of the test load 2. Column mold made of schedule 40 PVC pipe measuring 8 inches in diameter and 26 inches in height and separated into 4 equal sections each measuring 6.5 inches in height 138 3. Collection plate made of a non-absorbent material measuring 20 inches square containing a semi-circular cut out section in the center measuring 8.5 inches across 4. 5/8 inch tamping rod 5. No. 4 Sieve Figure 5.9 ? Testing Equipment for Segregation Column Test Procedure (ASTM Segregation Column Draft 2004): 1. Assemble and moisten all equipment, then place on flat and rigid surface. 2. After the mixture procedure is complete, sample approximately 0.8 ft 3 of SCC in accordance with ASTM C 172 (1998). 139 3. Immediately fill the column mold in one continuous lift allowing the concrete to over fill the top. 4. Strike off any excess SCC by means of rolling the tamping rod over the surface. 5. Allow the SCC to set for 1 hour. This time was extended to 1 hour from the 20 minute recommendation provided by the ASTM Segregation Column draft (2004). It was expected that due to the extended set times for drilled shaft concrete that the segregation would be more pronounced after 1 hour. 6. Place the collection plate around the mold and firmly hold the top section of mold while removing the spring clamps. 7. Screed the SCC onto the collection plate by using a horizontal twisting motion of the PVC pipe. 8. Place the obtained SCC from the collection plate into a No. 4 Sieve. Wet wash the sample leaving only the coarse aggregate on the No. 4 sieve. 9. Repeat steps 7 and 8 for the bottom section of the column. 10. Calculate the segregation index using the Equation 5.1 and 5.2. MB TB CA CACA SI )( ? = Eq. 5.1 Where: SI is the segregation index, CA T is the mass of coarse aggregate in the top section (lbs), CA B is the mass of coarse aggregate in the bottom section (lbs), and CA BM is the mass coarse aggregate (lbs) per section of the column according to Equation 5.2. [ ] MMBM CACACA ??= 0052.0007.0 Eq. 5.2 Where: CA M is the mass of coarse aggregate (lbs) in 1 yd 3 of concrete. 140 5.3.6 Bleeding Test Bleeding tests were performed on all SCDOT concrete mixtures via the procedure in ASTM C 232 (1998), Standard Test Method for Bleeding of Concrete using Method A. The purpose of this test was to determine the relative quantity of mixing water that will bleed from a freshly mixed concrete sample. All equipment used met the requirements of this specification. The ? cubic foot container, scale, pipet, mallet, and glass graduate can be seen in Figure 5.10. After placement of the concrete mixture into the ? cubic foot container, the accumulated bleed water was drawn off the surface until cessation of bleeding. The accumulated bleed water was placed into a 100-mL graduate and recorded after each transfer. The total amount of bleed water was recorded and expressed as a percentage of the net mixing water contained in the test specimen. This procedure was followed as closely as possible for all SCC mixtures with the following modifications. These modifications were necessary to account for SCC characteristics. 1. The SCC mixtures were placed in the container in one continuous lift without any rodding. 2. The tapping of the sides 10 to15 times with the approximate mallet was not conducted for the SCC mixtures. 141 Figure 5.10 ? Testing Equipment for Bleeding Test 5.3.7 Unit Weight and Air Content The unit weight and air content was determined for all concrete mixtures via the procedure in ASTM C 138 (1998), Standard Test Method for Unit Weight, Yield, and Air Content. The purpose of this test was to determine the weight per cubic foot and air content of freshly mixed concrete. All equipment used met the requirements of this specification. The balance, 5/8 inch tamping rod, measure, strike off plate, and mallet can be seen in Figure 5.11. This procedure was followed as closely as possible for all SCC mixtures with the following modifications. These modifications were necessary to account for SCC characteristics. 1. The SCC mixtures were placed in the measure in one continuous lift without any rodding. 2. The tapping of the sides 10 to15 times with the approximate mallet was reduced to no more than 5 soft taps by hand or mallet for all SCC mixtures. The purpose of this exercise was to help alleviate any large 142 entrapped air pockets that remained along the cylinder walls while providing little or no consolidation effort. Figure 5.11 ? Testing Equipment for Unit Weight and Air Content 5.3.8 Making and Curing Specimens The making and curing of concrete specimens were performed according to ASTM C 192 (1998), Standard Practice for Making and Curing Concrete Test Specimens in the Laboratory. After strike off, all specimens were capped with a tightly sealed lid. The cylinders remained in the mixing room for a period of 24 hours or until 2 times the initial set was achieved due to extended setting times of the drilled shaft concrete mixture. Cylinders were then relocated from the mixing room, stripped of their molds, and placed in a moist curing room. The conditions of the moist cure room were held constant at a temperature of 73 o F and relative humidity of 100%. The cylinders remained in the curing room until testing. This procedure was followed as closely as 143 possible for all SCC mixtures with the following modifications. These modifications were necessary to account for SCC characteristics. 1. The SCC mixtures were placed into the cylinder molds in one continuous lift without any rodding. 2. The tapping of the sides 10 to 15 times with the approximate mallet was reduced to no more than 5 soft taps by hand or mallet for all SCC mixtures. The purpose of this exercise was to help alleviate any large entrapped air pockets that remained along the cylinder walls while providing little or no consolidation effort. 5.3.9 Time of Set Setting tests were performed on all SCDOT concrete mixtures via the procedure in ASTM C 403 (1998), Standard Test Method for Time of Setting of Concrete Mixtures by Penetration Resistance. The purpose of this test was to determine the time of setting for freshly mixed concrete mixture by means of penetration resistance. All equipment used met the requirements of this specification. The mortar container, penetration needles, pipet, and loading apparatus can be seen in Figure 5.12. Each mortar sample was obtained by vibrating a portion of the concrete mixture over a No. 4 sieve, and then placing the mortar into the metal container shown in Figure 5.12. The specimens were kept sealed by a tightly placed lid to prevent the occurrence of evaporation. Prior to the removal of the bleed water, a wedge was inserted under the container to facilitate the collection of bleed water. The lid was then removed for bleed water draw off and testing. The testing consisted of making no less than six penetrations until at least one penetration 144 resistance reading equaled or exceeded 4,000 psi. The specimens were maintained at mixing room temperature for the entire period until final set was achieved. There were no necessary modifications to this test to account for SCC characteristics. Figure 5.12 - Testing Equipment for Time of Setting 5.4 HARDENED CONCRETE PROPERTIES 5.4.1 Compressive Strength The compressive strength of the 6 x 12 inch cylindrical concrete specimens was tested in accordance with ASTM C 39 (1998), Standard Test Method for Compressive Strength of Cylindrical Concrete Specimens. The equipment used in the laboratory to determine the compressive strength of the specimens met all requirements set forth by this standard. The specimens were tested using unbonded caps that consisted of a steel 145 retaining ring and neoprene pads. The unbonded caps meet all requirements described by ASTM C 1231 (1998), Standard Practice for use of Unbonded Caps in Determination of Compressive Strength of Hardened Concrete Specimens. The load rate utilized for the 6 x 12 inch specimens was 35 psi/sec., which corresponds to a value of 60,000 lbs/min. Each specimen was loaded in a 600 kip Forney compression machine as shown in Figure 5.13 until failure occurred. Figure 5.13 ? 600 Kip Forney Compression Machine 5.4.2 Modulus of Elasticity The modulus of elasticity of the 6 x 12 inch cylindrical concrete specimens was tested in accordance with ASTM C 469 (1998), Standard Test Method for Static Modulus of Elasticity and Poisson?s Ratio Strength of Concrete in Compression. The purpose of this test was to determine the chord modulus of 6 x 12 inch specimens. The equipment used in the laboratory to determine the elastic modulus of the concrete specimens met all 146 requirements set forth by this standard. The specimens were tested using unbonded caps that consisted of a steel retaining ring and neoprene pads. The unbonded caps meet all requirements described by ASTM C 1231 (1998), Standard Practice for use of Unbonded Caps in Determination of Compressive Strength of Hardened Concrete Specimens. A Humboldt compressometer equipped with a digital dial gauge was used to determine the elastic modulus of the concrete specimen. The researcher ensured that the compressometer was positioned evenly from the top, bottom, and sides. The concrete specimen was subsequently placed in the 600 kip Forney compression machine and tested as shown in Figure 5.14. The load rate utilized for the 6 x 12 inch concrete specimens was 35 psi/sec., which corresponds to a value of 60,000 lbs/min. Each specimen was first loaded to 40% of the ultimate strength without recording any data. The purpose of this exercise was to ensure that all equipment was properly seated and working correctly. The load was re-applied while recording the appropriate data. After the data was recorded, the modulus of elasticity was determined according to Equation 5.3. ( ) 00005.0 2 12 ? ? = ? SS E Eq. 5.3 Where, E = Chord modulus of elasticity, psi, S 2 = Stress corresponding to 40% of the ultimate load, psi, S 1 = Stress corresponding to a longitudinal strain of 50 millionths, psi, ? 2 = longitudinal strain produce by S 2 147 Figure 5.14 ? Concrete Specimen with Compressometer Attached 5.4.3 Drying Shrinkage The length change of 3 x 3 x 12 inch concrete specimens was tested in accordance with ASTM C 157 (1998), Standard Test Method for Length Change of Hardened Hydraulic-Cement Mortar and Concrete. The purpose of this test was to determine the length change of the concrete specimens due to drying shrinkage. The equipment used in the laboratory to determine the length change of the concrete specimens met all requirements of this specification. The concrete specimens were allowed to cure for 28 days in a lime saturated bath. Afterwards, the concrete specimens were removed and placed into air storage. The air storage room met all requirements stated in this specification. A Humboldt length comparator equipped with a digital dial gauge was used to determine the length change of the concrete specimens. The use of a digital 148 gauge helped reduce operator error and allowed more accurate measurements to be taken. The length comparator can be seen by looking at Figure 5.15. Figure 5.15 ? Humboldt Length Comparator, Mold, and Concrete Specimen 5.4.4 Permeability The permeability of 4 x 8 inch concrete specimens was tested in accordance with ASTM C 1202 (1998), Standard Test Method for Electrical Indication of Concrete?s Ability to Resist Chloride Ion Penetration. The purpose of this test was to give an indication of the concrete?s permeability by determining the resistance to chloride ion penetration by electrical conductance. The equipment used in the laboratory to determine the permeability of the concrete specimens met all requirements set forth by this standard. Proove? It cells and a Model 164 Test Set with LED read outs, automatic shut off, and automatic processing equipment was used to determine the resistance of ion penetration 149 by electrical conductance. The Model 164 Test Set and Proove? It cells can be seen by looking at Figure 5.16. The concrete specimens were allowed to cure in moist curing room until time of testing. Afterwards, the concrete specimens were removed from the moist cure room and a 2 +/- 1/8 inch slice was cut from the 4 x 8 inch specimen using a water-cooled diamond saw. The cut specimens were subsequently conditioned for testing according to requirements set forth by this standard. After conditioning was completed, the specimens were removed from the vacuum desiccator and placed into the Proove? It cells. The use of Proove? It cells allowed the rapid preparation of the concrete specimens after the removal from the desiccator. Furthermore, the Proove? It cells did not require the use of cell sealant, which further decreased the preparation time after removal from the desiccator. Each specimen was tested for a period of 6 hours as required by this standard. Figure 5.16 ? Model 164 Test Set and Proove? It Cells 150 CHAPTER 6 PRESENTATION AND ANALYSIS OF RESULTS 6.1 INTRODUCTION The presentation and analysis of results is presented in this chapter in the following order: ? Phase I ? Selection of Type and Dosage of HRWRA ? Phase II ? Effect of Retarder Dosage ? Phase III ? Appropriate SCC Mixing Procedure ? Phase IV ? Selection of SCC Properties ? Phase V ? Methods to Modify the Viscosity of SCC Mixtures 6.2 PHASE I ? SELECTION OF TYPE AND DOSAGE OF HRWRA The appropriate type and dosage of HRWRA was selected to provide a concrete mixture that was within the proposed quality control limits and that showed adequate fresh concrete properties after 30 minutes of continuous mixing. Table 6.1 and Figure 6.1 presents the results obtained from Phase I. Figure 6.1 shows that the slump flow increased as the dosage of Glenium 3030 NS increases; this is expected since the fluidity of a concrete mixture will increase as the HRWRA dosage increases. It can also be concluded from this figure that the Glenium 3000 NS appeared to be more effective at the same dosage amount of Glenium 3030 NS. Although the Glenium 3000 NS was more effective than the Glenium 3030 NS at the same dosage amount, it produced an undesirable thick-flakey film on top of cured specimens. Table 6.1 and Figure 6.1 demonstrate that the slump flow decreased with continuous mixing. Thus, transportation time must be taken into consideration when 151 determining the appropriate HRWRA dosage. The results further indicate that dosages of 8 oz/cwt of Glenium 3000 NS and 10 oz/cwt Glenium 3030 NS were within the specified range of slump flow values of 21 ? 3 inches upon completion of mixing. The dosage of 8 oz/cwt of Glenium 3000 NS produced a concrete mixture with a VSI rating of 1.5 at a slump flow value of 23 ? inches, a VSI rating that is not acceptable for this research project. On the other hand, a dosage amount of 10 oz/cwt of Glenium 3030 NS produced a concrete mixture that had appropriate fresh concrete properties after 30 minutes of continuous mixing. Based upon these results, it was determined that the appropriate type of HRWRA to be used throughout this research project should be Glenium 3030 NS, with a dosage amount of 10 oz/cwt to produce desirable fresh concrete properties after 30 minutes of continuous mixing under controlled laboratory conditions. Table 6.1 ? Fresh Concrete Properties for Phase I 8 oz/cwt Glenium 3030 NS 8 oz/cwt Glenium 3000 NS 10 oz/cwt Glenium 3030 NS 12 oz/cwt Glenium 3030 NS Slump Flow (inches) 21 31.5 27 30.5 VSI 031.52. T50 (sec.) 2.65 0.68 1.1 0.91 Slump Flow (inches) 14 23.75 21 25 VSI 0 1.5 1 1.5 T50 (sec.) >30 1.09 1.31 1.22 Air Content (%) 33 * * Temp. ( o F) 77 77 75 81 Unit Weight (lb/ft 3 ) 144.9 145.6 144 146.9 Pl a n t J ob S i t e Batch ID Item *The air content meter was not available for use. 152 0 6 12 18 24 30 36 10 15 20 25 30 35 40 Concrete Age (minutes) S l um p Fl ow ( i n c h es) 8 oz/cwt Glenium 3030 NS 8 oz/cwt Glenium 3000 NS 10 oz/cwt Glenium 3030 NS 12 oz/cwt Glenium 3030 NS Proposed Quality Control Limits Figure 6.1 ? Slump Flow vs. Concrete Age for Phase I 6.3 PHASE II ? EFFECT OF RETARDER DOSAGE In this phase, the effect of the retarding admixture on the retained workability was examined by varying the Delvo dosage from 0 oz/cwt to 8 oz/cwt in increments of 4 oz/cwt. Table 6.2 presents the fresh concrete properties obtained for Phase II. Figure 6.2 demonstrates how the slump flow varied with concrete age for each concrete batch, while Figure 6.3 shows the results obtained from the setting test. From Figure 6.2, it is clear that retarding admixtures have an effect on the retained workability on the concrete mixtures. It can be seen that the concrete mixture that contained no retarding admixtures lost its workability at an incredibly rapid rate from a slump flow of 16 inches to a slump flow of 8 inches, which corresponds to a slump of 0 153 inches, within 3 hours after the first batch would have been placed. This rapid loss is undesirable for a drilled shaft concrete mixture in applications where retained workability is required. Figure 6.2 further reveals that the addition of 4 oz/cwt of Delvo was quite effective in increasing the time in which the concrete mixture remained workable. Figure 6.3 shows that 4 oz/cwt of Delvo extended the initial set of the concrete mixture from 364 minutes to 552 minutes (~ 3 hours). This corresponded to a slump increase from 0 inches after 3 hours to 5 inches (slump flow ? 9.5 in.) after 4 ? hours. By looking at Figure 6.3, the increase of Delvo dosage from 4 oz/cwt to 8 oz/cwt extended the initial set time from 552 minutes to 783 minutes (3 hours and 50 minutes). This increase in initial set corresponded to an increase of slump from 5 inches (slump flow ? 9.5 in.) to 7 inches (slump flow ? 11 in.) at approximately the same concrete age. Although the retained workability was increased, the change was less compared to the change from 0 oz/cwt to 4 oz/cwt or from 0 oz/cwt to 8 oz/cwt of Delvo. It must be emphasized that the retained workability of concrete is not only a function of the extended initial set, but it is also a function of other factors such as temperature effects, long-term effectiveness of the HRWRA, and the effect of different supplementary cementitious materials on the hydration rate. Therefore, the dosage of retarding admixture should be sufficient to compensate for these effects and provide the extended initial necessary for the duration of the pour so that retained workability and shaft completion can be achieved. This phase has shown that the use of retarding admixtures can be very effective in extending the time in which the concrete mixture will remain workable. 154 Table 6.2 ? Fresh Concrete Properties for Phase II 0 oz/cwt Delvo 4 oz/cwt Delvo 8 oz/cwt Delvo Slump Flow (inches) 25 28 29 VSI 122 T50 (sec.) 1.59 0.53 0.68 Slump Flow (inches) 16 17.5 19 VSI 011 T50 (sec.) >30 >30 >30 Air Content (%) 2.7 2.8 2.8 Temp. ( o F) 71 72 71 Unit Weight (lb/ft 3 ) 144 144.6 144 Batch ID Pl a n t Jo b S i t e Item 0 6 12 18 24 30 0 60 120 180 240 300 360 Concrete Age (minutes) S l um p F l ow ( i nc h e s ) 0 oz/cw t Delvo 4 oz/cw t Delvo 8 oz/cw t Delvo Proposed Quality Control Limits Figure 6.2 ? Slump Flow vs. Concrete Age for Phase II 155 0 1,000 2,000 3,000 4,000 5,000 6,000 200 300 400 500 600 700 800 900 1000 Concrete Age (minutes) P e n e t r at i o n R e si st an ce ( p si ) 0 oz/cw t Delvo 4 oz/cw t Delvo 8 oz/cw t Delvo INITIA L SET FINAL SET Figure 6.3 ? Penetration Resistance vs. Concrete Age for Phase II 6.4 PHASE III - APPROPRIATE SCC MIXING PROCEDURE In order to determine an appropriate SCC mixing procedure to be used throughout this research project, two different mixing procedures were compared in this phase. Table 6.3 presents the fresh concrete properties, while Figure 6.4 illustrates how the slump flow varied with concrete age for Phase III. It can be seen from these results that addition of the HRWRA before (mixing process 1) or after (mixing process 2) the addition of cementitious materials had no significant affect on the short or long term concrete workability. Based on these results, it was determined that mixing procedure 2 was the appropriate mixing procedure to be used throughout the laboratory portion of this research project. This decision was based on several factors that included the following: 156 ? The determination of the water slump was relatively easy and quick to perform under laboratory conditions. ? By obtaining a water slump any excess free water due to inaccurate moisture corrections could be avoided. ? The water slump was an indicator of the concrete mixture?s consistency due to the mixing water alone. Later observations found that the stability of the SCC mixtures were sensitive to free water. Due to this fact, the water slump assisted in determining the appropriate mixture proportions to be used for several SCC mixtures. Table 6.3 ? Fresh Concrete Properties for Phase III Mixing Process 1 Mixing Process 2 Wet Slump (inches) NA 4 3/4 Slump Flow (inches) 27 26 VSI 1.5 1.5 T50 (sec.) 1.23 1.44 Slump Flow (inches) 16 16.5 VSI 00.5 T50 (sec.) >30 >30 Air Content (%) 3.3 3.4 Temp. ( o F) 73 79 Unit Weight (lb/ft 3 ) 145.2 144.5 Item Pl a n t J ob S i te Batch ID 157 0 6 12 18 24 30 0 60 120 180 240 300 360 Concrete Age (minutes) S l um p Fl ow ( i nc he s ) Mixing Process 1 Mixing Process 2 Proposed Quality Control Limits Figure 6.4 ? Slump Flow vs. Concrete Age for Phase III 6.5 PHASE IV - SELECTION OF SCC PROPERTIES 6.5.1 Fresh Concrete Properties and Workability The fresh concrete properties for both the ordinary drilled shaft concrete (ODSC) and SCC mixtures are presented in Tables 6.4 and 6.5. The data in Table 6.4 indicates that the ODSC mixtures were found to be within the specified range of slump values at placement in addition to demonstrating typical workability characteristics for wet-hole construction. As seen in Tables 6.4 and 6.5 the SCC mixtures typically contained more entrapped air than the ODSC mixtures. It is hypothesized that the HRWR admixture Glenium 3030 NS increased the entrapped air compared to the MRWR admixture PolyHeed N. Table 6.5 shows that no SCC mixtures possessed appropriate workability characteristics at the batch plant location. The SCC mixtures exhibited signs of 158 segregation by evidence of the VSI rating at higher slump flow values. However, after 50 minutes of continuous mixing the SCC mixtures were found to be within the specified quality control limits set forth in this research. The results further indicate that the stability of the SCC mixtures tends to increase as the water slump was reduced, as indicated by the differences among the VSI ratings for the SCC mixture, and especially for the silica fume and GGBFS mixtures at higher slump flow values. Tables 6.4 and 6.5 reveal that the ODSC mixtures did not show an excessive loss of workability due to continuous mixing compared to the SCC mixtures. For instance, after 50 minutes of continuous mixing, the highest slump loss experienced by the ODSC mixtures was ? inch compared to the least amount of slump flow loss by any SCC mixture of 6.5 inches. Thus, it is obvious that the SCC mixtures were more sensitive and experienced more workability loss when subjected to the same mixing conditions. Despite this issue, the SCC mixtures were capable of providing an increase in workability at placement compared to the ODSC mixtures. This enhanced workability is apparent in Figures 6.5 and 6.6. Table 6.4 ? Fresh Concrete Properties for ODSC Mixtures 159 1:ODSC 2:ODSC Slump (inches) 8.5 8.75 Air Content (%) 2.4 2 Temp. ( o F) 73 73 Unit Weight (lb/ft 3 ) 146.7 146.5 Item Pl a n t Jo b S i t e ODSC Mixtures Slump (inches) 99 Table 6.5 ? Fresh Concrete Properties for SCC Mixtures 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Slump Flow (inches) 29 30 30 30 27.5 26 28 VSI 2.533221.52 T50 (sec.) 0.97 1.41 1.13 1.31 1.59 1.25 1.60 Wet Slump (inches) 4 5 6 1.5 3/4 3/4 1.0 Slump Flow (inches) 16.5 19.0 18.0 20.5 19.0 19.5 20.0 VSI 1.0 1.0 1.0 0.5 0.5 0.5 0.5 T50 (sec.) >30 >30 >30 2.31 >30 >30 2.47 Air Content (%) 3.8 3.5 4.2 3.5 4.5 5.1 3.8 Temp. ( o F) 73 71 74 77 77 79 76 Unit Weight (lb/ft 3 ) 143.6 144.4 144.6 146.8 146 143.7 145.3 Pl ant Jo b Si t e Item Self-Consolidating Concrete Mixtures 160 160 Figure 6.5 ? Workability of ODSC Mixture (approximately 8.25 inches) Figure 6.6 ? Workability of SCC Mixture (approximately 20 inch slump flow) 161 Figures 6.7 and 6.8 present the slump or slump flow loss versus concrete age for the ODSC and SCC mixtures. Figure 6.7 shows that the ODSC mixtures displayed desirable slump retention characteristics in which the slump slowly diminished and exceeded 5.5 inches after 6 hours. This corresponded to a slump loss of 3 inches for 1:ODSC and 2:ODSC after placement. On the other hand, the slump flow loss for the SCC mixtures ranged from 6.5 to 10 inches after placement, which corresponds to a slump loss of more than 5 inches. Therefore, the SCC mixtures were more inclined to have a larger change in workability for the same amount of time compared to the ODSC mixtures. Although the SCC mixtures experienced larger changes in workability, the workability of the SCC mixtures was generally similar or higher than those of the ODSC mixtures after 5.5 to 6.5 hours as shown in Figure 6.9. It can be concluded from Figures 6.7 through 6.9 that at a concrete age of 5.5 to 6.5 hours both the ODSC and SCC mixtures would have complied with the recommendation provided by O?Neill and Reese (1999) that states that the drilled shaft concrete should have at least 4 inches of slump after 4 hours, but neither mixture would have met the recommendation provided by Brown (2004) that suggests that the drilled shaft concrete mixture should not experience a slump loss of no more than 2 inches for the duration of the pour. This is considering the fact that the duration of the pour was over 6 hours. The data on Figure 6.8 indicate that the SCC mixtures that incorporated 4 oz/cwt of the mid-range water reducing admixture PolyHeed 1025 seem to maintain their workability better than those that contained only the Glenium 3030 NS. The SCC mixture 6:36-40-FA seems to be an outlier among this trend. Two primary factors are thought to have contributed to this outcome. Firstly, it is believed that the decrease in set 162 time for 6:36-40-FA mixture seen in Figure 6.10 may have increased the rate of slump flow loss. Moreover, after the mixing for the 6:36-40-FA mixture was complete, it was found that slump flow of the 6:36-40-FA mixture was very low. The mixture was re- dosed with 3 oz/cwt of Glenium 3030 NS to achieve a slump flow 20.5 inches, and as reported in Section 2.3.2 the workability regained from the re-dosage by HRWRA may decrease at a faster rate. 1 3 5 7 9 11 0 60 120 180 240 300 360 Concrete Age (minutes) S l um p ( i nc he s ) 1:ODSC 2:ODSC SCDOT Quality Control Limits Figure 6.7 ? Slump vs. Concrete Age 163 0 6 12 18 24 30 0 60 120 180 240 300 360 420 Concrete Age (minutes) S l u m p F l ow ( i nc he s ) 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Proposed Quality Control Limits Figure 6.8 ? Slump Flow vs. Concrete Age 0 1 2 3 4 5 6 7 8 9 300 310 320 330 340 350 360 370 380 Concrete Age (minutes) S l um p ( i nc he s ) 1:ODSC 2:ODSC 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-FA 9:36-44-FA Figure 6.9 ? Slump vs. Concrete Age for all Concrete Mixtures 164 0 500 1000 1500 2000 2500 3000 3500 4000 4500 600 700 800 900 1000 1100 1200 1300 1400 1500 1600 1700 1800 1900 2000 2100 Concrete Age (minutes) P e n e t r at i o n R e sist an ce ( p s i ) 1:ODSC 2:ODSC 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Initial Set Final Set Figure 6.10 ? Penetration Resistance vs. Concrete Age 6.5.2 Segregation and Bleeding Results The results from the segregation column and bleeding tests are summarized in Table 6.6. The data on Tables 6.5 and 6.6 indicate that the SCC mixtures were placed into the column mold at the lower end of the proposed quality control limits and at a VSI rating of 1 or less. Under such conditions the SCC mixtures were stable and exhibited minimal segregation. However, the observed behavior of the SCC mixtures suggests that at higher values of slump flows and VSI ratings above 1 that the SCC mixtures would demonstrate a higher degree of segregation. For example, Table 6.5 shows that the fly ash (FA) SCC mixtures tested at the batch plant showed signs of segregation at higher 165 slump flow values. This segregation was evident by a thick mortar layer on the surface of the tested concrete in the mixing drum as well as clear evidence of segregation in the flow patty. Thus, it is reasonable to believe that if segregation column tests were performed under these conditions, the coarse aggregate concentration in the bottom sections would be higher, and the values of the segregation index would be increased. Regarding the bleeding test, the ODSC mixtures demonstrated a higher degree of bleeding compared to the SCC mixtures prepared at the same water-to-cementitious materials ratio. It is thought that this higher degree of bleeding could possibly be due to the fact that the initial set for the ODSC mixtures was extended at least 9 hours compared to the SCC mixtures as shown in Figure 6.10. This allowed a much longer time frame in which the ODSC mixtures could bleed. Furthermore, the SCC mixtures typically contained a higher percentage of entrapped air, which is very effective in reducing bleeding. Lastly, the incorporation of the VMA for the SCC mixtures may have assisted in the reduction of bleed water. However, it should be noted that the influence of the new polyethylene glycol based VMAs on bleeding is not well known. Further evaluation of Table 6.6 reveals that the SCC mixtures prepared at a water-to-cementitious materials ratio of 0.36 limited the amount of bleeding. 166 Table 6.6 ? Segregation and Bleeding Results Mixture Slump or Slump Flow (inches) Segregation Index (%) Bleeding (%) 1:ODSC 8.50 * 2.05 2:ODSC 8.75 * 2.34 3:41-48-FA 16.5 0.0379 0.33 4:41-44-FA 19 0.51 0.54 5:41-40-FA 18 0.41 0.77 6:36-40-FA 20.5 0.53 0 ** 7:36-40-SG 19 0.15 0 ** 8:36-40-SF 19.5 0.56 0 ** 9:36-44-FA 20 0.54 0 ** * Not Conducted for ODSC Mixtures **Bleeding was difficult to measure. Any water present at the surface was not clear and consisted primarily of cementitious materials. 6.5.3 Passing Ability: J-Ring and L-Box The test results of the J-Ring and those of the L-Box are summarized in Table 6.7. Regarding the (FA) SCC mixtures, those prepared with sand-to-aggregate ratios of 0.44 and 0.48 exhibited greater passing ability among closely spaced reinforcement using the J-Ring than those prepared with a sand-to-aggregate ratio of 0.40. This could be due to the fact that the (FA) mixtures prepared at a sand-to-aggregate ratio of 0.40 contained a higher amount of coarse aggregate that increased the collision and interaction among the solid particles at the vicinity of the reinforcement that resulted in a greater tendency of blockage. The (SF) and (SG) mixtures prepared with a sand-to-aggregate ratio of 0.40 demonstrated a passing ability similar to the fly ash (FA) mixtures prepared with sand-to- aggregate ratios of 0.44 and 0.48. However, no other (SF) and (SG) mixtures were prepared at varying sand-to-aggregate ratio. Therefore, no conclusion can be drawn in 167 regards to if (SF) and (SG) mixtures prepared at higher sand-to-aggregate ratios would show greater passing ability at similar slump flows. It can be seen from Table 6.7 that all SCC mixtures demonstrated very low passing ability using the L-Box apparatus. This is due to the fact that the maximum size aggregate size of ? inch used for this research project was simply too large for the clear spacing between the reinforcement in the L-Box apparatus. For example, the clear spacing between the reinforcement for the L-Box as recommended by the PCI (2003) was 1.375 inches, which corresponds to less than 2 times the maximum aggregate size. This spacing is unrealistic for most drilled shaft applications, but if very congested reinforcement cages exist the L-Box could be used to ensure high passing ability of the SCC mixture. In case of this research, the L-Box apparatus was found to be ineffective in determining the passing ability of SCC mixtures designed for drilled shaft applications. As a result, the blocking ratios determined from the L-Box should be disregarded for this research. Table 6.7 ? Passing Ability Results for SCC Mixtures L-Box Diameter (inches) Inverted T 50 (sec) Inverted VSI J-Ring Flow (inches) Inverted Ratio of J-Ring Flow to Slump Flow Passing Ability Rating h 2 /h 1 3:41-48-FA 20 1.28 1 17.5 0.88 2 0.27 4:41-44-FA 20.5 2.13 1 18 0.88 2 0.3 5:41-40-FA 20 1.91 1 16 0.8 2 0.078 6:36-40-FA 20 3.9 0.5 16 0.8 2 0.059 7:36-40-SG 20.5 5.03 1 18 0.88 2 0 8:36-40-SF 22 1.69 0.5 20.5 0.93 1 0.087 9:36-44-FA 20.5 2.25 0.5 18 0.88 2 0.036 Slump Flow J-Ring 168 6.5.4 Compressive Strength The results for the compressive strength testing for the ordinary drilled shaft concrete (ODSC) and SCC mixtures are given in Figure 6.11. Regarding the (FA) SCC mixtures, Figure 6.11 indicates that regardless of the water-to-cementitious materials ratio, it appears that the sand-to-aggregate ratio did not influence the strength development. Figure 6.11 further reveals that the SCC mixtures prepared at a water-to- cementitious materials ratio of 0.41 demonstrated slightly lower compressive strengths compared to those of the ODSC mixtures. However, the average difference among the compressive strengths is reduced as the SCC mixtures continue to hydrate. The average difference between the compressive strengths are 15.5%, 12%, 10.5%, 9%, and 6.5% at ages of 3, 7, 14, 28, and 56-day, respectively. It is thought that the slightly lower compressive strengths can be attributed to the following: ? The SCC mixtures contained 8% higher replacement percentage of cement by fly ash compared to the ODSC mixtures. Therefore, the amount of early heat evolution is decreased and in turn reduces the early age strength, but not the long term strength. ? Secondly, Mindess et al. (2003) reports that there is only enough calcium hydroxide in the paste in which the Class F fly ash can react to form calcium silicates. This would suggest higher replacement percentages of fly ash may result in unreacted ash causing a slight reduction in compressive strengths. Further evaluation of Figure 6.11 shows that the reduction in water-to- cementitious materials ratio from 0.41 to 0.36 for the (FA) SCC mixtures increased the 169 compressive strength, on average, by 1600 psi at 28-days. This is expected since it is a well known fact that the compressive strength is increased as the water-to-cementitious materials ratio decreases. Furthermore, the use of the silica fume or ground granulated blast furnace slag (GGBFS) was found to increase the compressive strength compared to the (FA) SCC mixtures. In addition to the continuing pozzolanic reaction between the amorphous silica in the silica fume and the calcium hydroxide, the high fineness of the silica fume allows the particles to pack densely between the cement particles and improves the interfacial transition zone. As a result, the silica fume greatly reduces the void spaces within the cement paste, and the bond of the cement paste with the aggregate is improved allowing the aggregate to better participate in stress transfer (Neville 1996). These contributions provided by the silica fume can be capable of generating higher compressive strength compared to the use of fly ash alone as in the case of this research. Unlike fly ash, high replacements of GGBFS, with values ranging from 25-65%, can be utilized since GGBFS have cementitious properties of their own and only 10-20% of cement is needed for activation. The hydration of GGBFS produces primarily calcium silicates and produces less calcium hydroxide than portland cement alone in addition to showing some pozzolanic behavior (Neville 1996). Neville (1996) reports that progressive release of alkalis by the GGBFS along the formation calcium hydroxide by portland cement results in a continuing reaction of GGBFS over a long period of time. Thus, there is a long term strength gain associated with the GGBFS. For these reasons, high compressive strengths at higher replacement values of portland cement by GGBFS can be achieved as in the case of this research. 170 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000 0 7 14 21 28 35 42 49 56 Concrete Age (Days) C o m p re s s i v e S t re n g t h ( p s i ) 1:ODSC 2:ODSC 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Figure 6.11 ? Compressive Strength vs. Concrete Age for Phase IV 6.5.5 Modulus of Elasticity The modulus of elasticity was determined for all concrete mixtures in conjunction with the compressive strength tests. Figure 6.12 presents the results obtained for the ordinary drilled shaft concrete (ODSC) and SCC mixtures. Generally speaking, the development for the modulus of elasticity is similar to the strength development provided in Figure 6.11. For example, the SCC mixtures prepared at a water-to- cementitious materials ratio of 0.41 demonstrated slightly lower elastic modulus values at early ages compared to those for the ODSC mixtures, and as the SCC mixtures continue to hydrate the difference among these mixtures are decreased. This should be expected 171 since the modulus of elasticity is a function of the compressive strength. This relationship between compressive strength and modulus of elasticity can further be seen by examining the silica fume and GGBFS mixtures. Figure 6.12 shows that the silica fume and GGBFS mixtures that exhibited higher compressive strengths also show higher modulus of elasticity values compared to the fly ash (FA) SCC mixtures at the same water-to-cementitious materials ratio. Furthermore, it appears from Figure 6.12 that the modulus of elasticity for the fly ash SCC mixtures was not significantly affected by the varying sand-to-aggregate ratio. In order to determine if the calculated elastic modulus for the concrete mixtures is of typical sound concrete; the calculated modulus of elasticity in this research was compared with the ACI 318 (2002) Building Code ( ccc fWpsiE '33)( 5.1 ?= ) and the ACI Committee 363 (2002) ?State-of the-Art Report on High-Strength Concrete? ( )000,000,1'000,40)( += cfpsiE c models (see Chapter 2). These results are presented in Figures 6.13 and 6.14. Figure 6.13 reveals that the ACI 318 (2002) overestimated the modulus of elasticity for both the ODSC and SCC mixtures. Furthermore, the ACI 318 (2002) was found to increasingly overestimate the modulus of elasticity as the compressive strength increased. This finding coincides with ACI Committee 363 (2002) that found that the equation used by the ACI 318 (2002) was only valid for compressive strengths up to 6,000 psi. On the other hand, the ACI Committee 363 (2002) equation provided an improved and typically a conservative estimate for the modulus of elasticity for both the ODSC and SCC mixtures as shown in Figure 6.14. It is believed that the improved and conservative estimation for the modulus of elasticity lies in the fact that the 172 equation provided by the ACI Committee 363 (2002) is valid for compressive strengths from 3,000 to 12,000 psi. 2 3 4 5 6 0 7 14 21 28 35 42 49 56 Concrete Age (Days) M odul us o f E l a s t i c i t y ( 1 . 0 x 1 0 6 ps i ) 1:ODSC 2:ODSC 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Figure 6.12 ? Modulus of Elasticity vs. Concrete Age for Phase IV 173 0 1 2 3 4 5 6 0123456 Measured Elastic Modulus (1.0 X 10 6 psi) P r e d ic t e d E l a s t ic M odulus ( 1 .0 X 1 0 6 p s i) 1:ODSC 2:ODSC 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Line of Equality -10%Error +10% Error Figure 6.13 ? Predicted vs. Measured Elastic Modulus according to ACI 318 (2002) Equation for Phase IV 0 1 2 3 4 5 6 0123456 Measured Elastic Modulus (1.0 X 10 6 psi) P r e d ic t e d E l a s t ic M odulus ( 1 .0 X 1 0 6 p s i) 1:ODSC 2:ODSC 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Line of Equality +10% Error -10%Error Figure 6.14 ? Predicted vs. Measured Elastic Modulus According to ACI 363(2002) Equation for Phase IV 174 6.5.6 Drying Shrinkage The drying shrinkage results for the ordinary drilled shaft concrete (ODSC) and SCC mixtures are presented in Figure 6.15. The data on Figure 6.15 indicate that the water-to-cementitious materials ratio appears to be the main factor influencing the amount of drying shrinkage. The results show that the reduction in water-to-cementitious materials ratio from 0.41 to 0.36 decreased the specimen?s tendency to shrink. This trend should be expected since the drying shrinkage is known to be reduced as the water-to- cementitious materials ratio is decreased. However, it appears that the mixture prepared with ground granulated blast furnace slag (SG mixture) produced slightly higher drying shrinkage values compared to the other mixtures prepared at a water-to-cementitious materials ratio of 0.36. Among the SCC mixtures prepared at a water-to-cementitious materials ratio of 0.41, the results indicate the SCC mixture prepared at sand-to-aggregate ratio of 0.48 shows evidence of higher drying shrinkage than those prepared at 0.40 and 0.44 as well as the highest drying shrinkage overall. However, the difference in drying shrinkage values among these SCC mixtures is actually quite minimal and operator or equipment error could very easily alter this outcome. This data suggest that the SCC mixtures prepared at a water-to-cementitious materials ratio of 0.41 exhibited drying shrinkage values similar to the ODSC mixtures. This outcome is reasonable given the fact that the paste volume fraction and aggregate volume fraction did not significantly vary compared to the SCC mixtures. Furthermore, Figure 6.15 indicates that the two ODSC mixtures showed practically the same drying shrinkage characteristics throughout the test. This 175 indicates that use of the #789 coarse aggregate gradation for 1: ODSC had no significant effect on the drying shrinkage values. Unfortunately, no absolute conclusion can be made at this time concerning the effect of the sand-to-aggregate ratio on the drying shrinkage, due to the lack of drying shrinkage data for 9:36:44-FA. It must be noted that this research will be updated as soon as additional test results are available. 0 100 200 300 400 500 600 0 7 14 21 28 35 42 49 56 63 70 77 84 91 Concrete Age (Days) D r yi n g S h r i n kag e ( m i r o s t r ai n ) 1:ODSC 2:ODSC 3:41-48-FA 4:41-44-FA 5:41-40-FA 6:36-40-FA 7:36-40-SG 8:36-40-SF 9:36-44-FA Figure 6.15 ? Drying Shrinkage vs. Concrete Age 176 6.5.7 Permeability Results of the rapid chloride permeability tests (RCPT) for the ordinary drilled shaft concrete (ODSC) and SCC mixtures are given in Figures 6.16 and 6.17. The RCPT values at 91-days for the ordinary drilled shaft concrete mixtures were 4530 and 4562 as compared to 2584, 2773, and 3067 for the SCC mixtures at the same water-to- cementitious materials ratio. This reduction in RCPT values for the SCC mixtures can be attributed to the fact the SCC mixtures contained a higher replacement percentage of fly ash (FA). For example, the SCC mixtures with a water-to-cementitious materials ratio of 0.41 consisted of 33% replacement of cement by fly ash, where as the ODSC mixtures only contained 25% replacement of cement by fly ash. Since fly ash is notably more spherical than cement, the additional replacement of fly ash may have allowed the particles to pack more tightly within the pore spaces creating a denser microstructure. Figures 6.16 and 6.17 indicate that the RCPT values were found to decrease from 91 to 365-days. As the hydration process proceeds, the interconnected pores that were present at 91-days are being filled by the continuous formation of C-S-H and the continuous growth of the calcium hydroxide within the capillaries pores. As a result, the porosity of the paste will decrease with time lowering the RCPT values. It is also important to notice that the trends from Figure 6.16 are also present in Figure 6.17. At 365-days the ODSC mixtures still exhibited higher RCPT values compared to the SCC mixtures at the same water-to-cementitious materials ratio. As discussed in Section 2.4.4, the coefficient of permeability decreases as the water-to-cementitious materials ratio is reduced. The reduction in RCPT values for the fly ash (FA) SCC mixtures (particularly 3:41-48-FA, 6:36-40-FA, and 9:36-44-FA) due 177 to the decrease in water-to-cementitious materials ratio is evident on Figure 6.16. The reduction in water-to-cementitious materials ratio from 0.41 to 0.36 decreased the RCPT values, on average, 1500 coulombs at the same concrete age. It is also important to emphasize the effect of the supplementary cementitious material on the RCPT values at a water-to-cementitious materials ratio of 0.36. Figure 6.16 indicates that the introduction of silica fume considerably reduced the RCPT values compared to the (FA) mixtures. This reduction in RCPT values comes from the ability of the silica fume to pack tightly between pore spaces creating a very dense microstructure. However, the introduction of ground granulated blast furnace slag (GGBFS) increased the RCPT values compared to the fly ash (FA) SCC mixtures. The increase in RCPT values for the (SG) mixture may stem from the angular shape of the GGBFS compared to the spherical nature of the fly ash. Due to this fact, the particles may have not been able to pack as closely causing an increase in the interconnected capillary pores. 178 0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000 1 : OD S C 2 :O D S C 3 :4 1 -4 8- FA 4 :4 1 -4 4 -F A 5 :4 1 -4 0- FA 6 :3 6 -4 0 -F A 7 : 36 - 4 0 -S G 8 :3 6 -4 0 -S F 9 : 36 - 4 4- FA C oul om b s Figure 6.16 ? 91-Day Permeability Results 0 200 400 600 800 1000 1200 1400 1: O D S C 2: O D S C 3: 41- 48 - FA 4: 41- 44 - FA 5: 41- 40 - F A C o ul om b s Figure 6.17 ? 365-Day Permeability Results 179 6.5.8 Comparison between Laboratory and Field Conditions A small scale field study was conducted at APAC Ready Mix Concrete in Marion, South Carolina. The primary objectives of this field study were three fold: 1. To ensure that the chemical admixture dosages determined for laboratory conditions remain sufficient for field conditions. 2. To evaluate the effect of mixing imposed by a ready mix truck on the slump flow after 50 minutes of continuous mixing. 3. To compare the compressive strength and modulus of elasticity values obtained from the field specimens to that of the laboratory specimens. The two SCC mixtures selected to be used for this field study were 3:41-48-FA and 6:36- 40-FA. The SCC mixtures were selected based upon the fact that at the time of this field study these mixtures were considered to be the most likely utilized for the full-scale field project. The only modification to these mixtures was that an additional 2 oz/cwt of Delvo was added to account for the effect of hot weather conditions on slump retention and time of setting. No ODSC mixtures were selected to be tested for this field study considering the fact that the ODSC mixtures used for laboratory purposes have been routinely accepted in South Carolina. The raw materials and chemical admixtures used for this small scale field study were provided by the same suppliers as those from the laboratory materials. Batching and Mixing: Each SCC mixture consisted of two cubic yards of concrete that was batched and mixed according to normal operations of the plant with the exception of the chemical admixtures. Figures 6.18 through 6.21 demonstrate how the raw materials were obtained and batched at this ready mix concrete plant. The chemical admixtures 180 were added to the SCC mixtures after the ready mix truck exited the material hopper and wash down was performed. This was done for two primary reasons. Firstly, the chemical admixtures needed for the SCC mixtures were not available for automated dispensing since they were not used in everyday operation. Secondly, the chemical admixtures were added after the wash down process was performed in order to obtain a water slump. Generally, 3 to 5 gallons of water is typically used for the wash down process. This extra water must be taken into account in the batching process since the stability of the SCC mixtures was found to be sensitive to excessive free water. The researcher requested that the plant operator withhold 5 gallons of batching water to account for the wash down process. A water slump was taken after the batching and wash down process was completed. It was found that obtaining a water slump under field conditions required a significant increase in time compared to laboratory conditions. Under laboratory conditions the water slump could be obtained no more than 5 minutes after water-to- cementitious materials contact; however, under field conditions obtaining the water slump required no less than 30 minutes after water-to-cementitious materials contact. This was primarily due to the time required to perform the wash down process and obtaining a concrete sample for testing. Unlike laboratory conditions where the concrete sample is very accessible and easy to obtain from the mixer, obtaining the concrete sample from the ready mix truck required more equipment, people, time, and attentiveness to detail due to safety precautions. Alternatives to this approach could be employed in order to reduce the time in which the accuracy of the moisture corrections can be determined. The following procedure could be utilized to accomplish this goal: 181 1. The use of wash down water could be prohibited in cases where SCC mixtures are used in order to control unwanted water. By implementing this suggestion no water will have to be withheld to account for the wash down process, and the time required to check the accuracy of the moisture corrections can be reduced. 2. It may be possible to check the accuracy of the moisture corrections without having to obtain a concrete sample. This can be accomplished by rotating the concrete mixture toward the concrete chute and make a water slump estimate based on the observation of the wetness or dryness of the concrete mixture. Most experienced quality control individuals can provide a reasonable estimate of the water slump, typically within +/- 1 inch of the actual water slump. Figure 6.18 ? Attaining Raw Materials from Stock Piles 182 Figure 6.19 ? Unloading Raw Materials onto Conveyer Belt Figure 6.20 - Raw Materials being Delivered to Hopper via Conveyer Belt 183 Figure 6.21 ? Raw Materials being Mixed by Ready Mix Truck Fresh Concrete Properties: The fresh concrete properties for the field study are presented in Table 6.8 and Figure 6.22. The fresh concrete properties for the plant location were very similar to those of the laboratory, suggesting that the batch sizes utilized for laboratory evaluation were sufficient to simulate the performance of the chemical admixtures in large batches. However, the slump flow characteristics for the job site location between the laboratory and field conditions produced very different results. It must be noted that in Figure 6.22 the mixing time represents the time of mixing after the first slump flow was taken. This was necessary due to the substantial difference in times for obtaining the water slump. The data in Table 6.8 and Figure 6.22 indicate that the slump flow loss due to continuous mixing was much less under field conditions compared to the laboratory. The primary reason for this outcome is likely due to the rotational speed of the mixing drum for the ready mix truck compared to laboratory 184 mixer. The laboratory mixer was rotated at a higher speed, providing heavy agitation of the concrete mixture. Where as, the mixing drum for the ready mix truck was set to 4-5 rotations per minute. The larger batch size used for the field study may have also contributed to lower slump loss for the field study. After 50 minutes of continuous mixing, the slump flow was found to be outside the proposed quality control limits. The SCC mixtures were allowed to rotate at an increased rotational speed for a short duration until the slump flow was found to be within the proposed quality control limits. The slump flow at which cylindrical specimens were made for the hardened concrete properties was 20 inches for 3:41-48-FA and 21 inches for 6:36-40-FA. Table 6.8 - Fresh Concrete Properties for both Laboratory and Field Conditions 3:41-48-FA (Lab) 3:41-48-FA (Field) 6:36-40-FA (Lab) 6:36-40-FA (Field) Slump Flow (inches) 29 30 30 28 VSI 2.5 2.5 2 2.5 T50 (sec.) 0.93 1.53 1.31 2.53 Wet Slump (inches) 4 ? 3 1.5 ? 1 Slump Flow (inches) 16.5 27.5 20.5 26 VSI 1 1.5 0.5 1.5 T50 (sec.) >30 1.62 2.31 3.5 Air Content (%) 3.8 3.1 3.46 3.1 Temp. ( o F) 73 96 75 100 Jo b S i t e Item Self-Consolidating Concrete Mixture Pl a n t 185 0 5 10 15 20 25 30 0 5 10 15 20 25 30 35 40 45 50 Mixing Time (minutes) S l u m p Flow ( i nc h e s ) 3:41-48-FA (Lab) 3:41-48-FA (Field) 6:36-40-FA (Lab) 6:36-40-FA (Field) Figure 6.22 ? Slump Flow vs. Mixing Time under Laboratory and Field Conditions The concrete temperatures in the field were higher than those for the laboratory as indicated in Table 6.8. Prolonged exposure to high temperatures can be detrimental to the slump retention and can significantly decrease setting times. In order to monitor the temperatures for the concrete samples for the slump retention and setting test, an I-Button was placed beside the samples until testing was complete. The temperature profile attained from the I-Button is presented in Figure 6.23. The times specified on Figure 6.23 are of those in which the first SCC mixture was made until the last setting test was complete. The SCC mixture 3:41-48-FA was batched on 7/27/04 at 10:45 AM and 6:36- 40-FA on 7/27/04 at 2:35 PM. Figure 6.23 further reveals that the concrete samples were exposed to temperatures ranging from 76 to 113 degrees o F. The corresponding setting test results for the field specimens are presented in Figure 6.24. From this figure it can be determined that the SCC mixture 3:41-48-FA under field conditions experienced much 186 faster setting times as compared to the laboratory conditions. The SCC mixture 6:36-40- FA demonstrated similar set times under field conditions compared to laboratory conditions. The faster set time for 3:41-48-FA may be due to longer exposure to high temperatures compared to 6:36-40-FA that experienced high temperatures for a short duration followed by a sharp decrease in temperature. The slump retention for 3:41-48-FA was also affected by the high temperatures and faster set times. The slump flow retention for 3:41-48-FA was determined to be 8 inches after four hours, which corresponded to a slump of approximately 3 ? inches. This slump was found to be much less than the laboratory specimen at approximately the same concrete age. On the other hand, the slump flow for 6:36-40-FA was determined to be 10 inches after 4 hours, which corresponded to a slump of 6 inches. This slump flow was found to be slightly less than under laboratory conditions even with similar set times. These results would suggest that even with similar set times the rate of slump loss is increased when exposed to high temperatures. These results further indicate the need for testing to determine the amount of additional retarder needed to account for high temperatures to ensure that proper retained workability and set times are achieved. 187 70 80 90 100 110 120 12:00 AM 12:00 AM 12:00 AM 12:00 AM 12:00 AM 12:00 AM 12:00 AM 12:00 AM 12:00 AM Tem p er a t ur e ( o F) 7/27/04 10:18AM 7/27/04 12:42 PM 7/27/04 3:06 PM 7/27/04 5:30 PM 7/27/04 7:30 PM 7/27/04 9:54 PM 7/28/04 12:18 AM 7/28/04 2:42 AM Figure 6.23 ? Temperature Profile Obtained from I-Button 0 500 1000 1500 2000 2500 3000 3500 4000 4500 0 200 400 600 800 1000 1200 Concrete Age (minutes) P e n e t r a t i o n R esi stan ce ( p si ) 3:41-48-FA Lab 3:41-48-FA Field 6:36-40-FA Lab 6:36-40-FA Field Initial Set Final Set Figure 6.24 ? Penetration Resistance vs. Concrete Age under Laboratory and Field Conditions 188 Hardened Concrete Properties: The results for the compressive strength and modulus of elasticity for both SCC mixtures are given in Figures 6.25 and 6.26. The data collected from the compressive strength and modulus of elasticity allowed for a comparison between field and laboratory conditions. It was found that the field specimens exhibited slightly higher compressive strengths compared to the laboratory specimens. It is thought that the higher compressive strengths are due to possibly withholding extra water for the wash down process, which may have caused a reduction in water-to-cementitious materials ratio. Nevertheless, the compressive strength and modulus of elasticity values obtained from the field specimens corresponded well to those obtained from the laboratory. 0 2000 4000 6000 8000 10000 12000 0 7 14 21 28 35 42 49 56 Concrete Age (Days) C o m p r e ssi v e S t r e n g t h ( p si ) 3:41-48-FA (Lab) 3:41-48-FA (Field) 6:36-40-FA (Lab) 6:36-40-FA (Field) Figure 6.25 ? Compressive Strength vs. Concrete Age under Laboratory and Field Conditions 189 0 1 2 3 4 5 6 0 7 14 21 28 35 42 49 56 Concrete Age (Days) M odul us of E l a s t i c i t y ( 1 . 0 x 1 0 6 ps i ) 3:41-48-FA (Lab) 3:41-48-FA (Field) 6:36-40-FA (Lab) 6:36-40-FA (Field) Figure 6.26 ? Modulus of Elasticity vs. Concrete Age under Laboratory and Field Conditions 6.6 PHASE V - METHODS TO MODIFY THE VISCOSITY OF SCC MIXTURES 6.6.1 Fresh Concrete Properties Phase V was formulated to clarify the effects of different methods to modify the viscosity of SCC mixtures. Among the SCC mixtures prepared with VMA as the method to modify the viscosity, the results indicate at similar slump flow values the viscosity of the SCC mixtures was not increased even at high dosage amounts of VMA. However, to some extent the results suggest that the VMA may provide a slight increase in stability at higher values of slump flow with increasing VMA dosage. This is according to the VSI ratings provided in Table 6.9 for those mixtures at higher slump flow values. At lower slump flow values the VMA appeared to have no profound effect on the stability of the SCC mixtures. 190 The results for the SCC mixtures prepared with high dosages of fly ash as the method to modify the viscosity reveal that an increase in fly ash percentage did not correspond to an increase in viscosity. In fact, the water slump of the SCC mixtures increased as the percentage of fly ash was increased, which resulted in a higher degree of segregation at the batch plant locations. This higher degree of segregation is not evident by the VSI ratings provided in Table 6.9 since 3 is the highest VSI rating, but it is an observational behavior made by the researcher. Neither the stability nor the viscosity of these SCC mixtures seemed to be significantly affected by the higher percentages of fly ash at lower slump values such as those that represent job site testing. The results in Table 6.9 show that the incorporation of the silica fume provided no increase in viscosity by evidence of the T 50 times. However, the utilization of the silica fume was found to improve the stability of the SCC mixtures as can be seen by the decreased VSI ratings. It is believed that this stability can be attributed to the reduction of the water slump and absorption of free water due to the high fineness of the silica fume. Replacement percentages of silica fume above 8% were found not to provide a considerable increase in stability at similar slump flow values nor a considerable reduction in water slump. Thus, replacement percentages above 8% would not be necessary and only result in increased cost. Furthermore, Table 4.5 indicates that the required dosage of HRWRA was increased as the replacement percentage of silica fume was increased. The incorporation of GGBFS provided results similar to the silica fume mixtures. For example, the GGBFS provided a reduction in the water slump and showed a higher degree of stability at the batch plant compared to 6:AL-36-33 FA. In addition, the 191 incorporation of the GGBFS seems to slightly increase the T 50 times compared to the silica fume and fly ash mixtures at similar slump flow values. As with the silica fume mixtures, the required dosage of HRWRA was increased with the incorporation of the GGBFS. The amount of entrapped air for the GGBFS mixtures was increased with higher replacement percentages of GGBFS. The amount of entrapped air was 4.1% for 13:AL- 36-40 SG, 10.5% for 14:AL-36-50 SG, and 17% for 15:AL-36-60 SG. These results indicate that the amount of entrapped air nearly doubled for every 10% increase in GGBFS. An experiment was conducted to determine if the Glenium 3030 NS was in fact the primary cause of this entrapped air. In this experiment, mixture 14:AL-36-50 SG was remade with exactly the same mixture proportions with the Glenium 3030 NS being replaced by PolyHeed 1025. Table 4.5 shows that the required dosage amount of PolyHeed 1025 to achieve similar slump flow values at the batch plant was 20 oz/cwt. This high dosage amount was necessary due to the fact that the PolyHeed 1025 is a mid- range water reducing admixture. The results of this experiment can be seen by examining mixture 14:AL-36-50 SG (LA) in Table 6.9, where (LA) stands for low air. The results show that when only the PolyHeed 1025 was utilized the amount of entrapped air was reduced from 10.5% to 2.6%, respectively. Based upon these results it was determined that the Glenium 3030 NS was the main reason for the increase in entrapped air. These results in Table 6.9 show that the Micron 3 did not provide an increase in stability or viscosity compared to the fly ash mixtures even at high replacement percentages. This outcome was unexpected since the Micron 3 was considerably finer 192 than both the cement and the traditional fly ash. However, it is believed that this outcome is due to the fact that the Micron 3 was in replacement of the cement and not the fly ash. Since the Micron 3 is notably more spherical than the cement it provided an increase in workability. This can be seen by observing the increase in water slump for 16:AL-36-8 M3 compared to 6:AL-36-33 FA. Earlier results illustrated that as the water slump was increased the stability of the mixture was decreased, especially at higher slump flow values. The stability of the Micron 3 mixtures was reduced in spite of the fact that the Micron 3 was finer. Furthermore, the results in Table 6.9 show that the introduction of the limestone filler provided no considerable increase in stability, but exhibited a reduction in water slump for a constant dosage of HRWRA. This reduction in water slump at the batch plant is most likely due to the reduction of water content from 284 lb/yd 3 for the fly ash mixtures to 270 lb/yd 3 for the limestone filler mixtures as shown in Table 4.5. 193 Table 6.9 ? Fresh Concrete Properties for Phase V (cont. on next page) 194 1:AL-41-0 VMA 2:AL-41-2 VMA 3:AL-41-10 VMA 4:AL-41-18 VMA 5:AL-36-33 (2) FA 6:AL-36-33 FA 7:AL-36-40 FA 8:AL-36-50 FA Slump Flow (inches) 29 31 29 29 28 32 33 32 VSI 3 3 2.5 2 2 3 3 3 T50 (sec.) 1.1 0.98 0.85 0.9 0.88 0.76 0.62 1.5 Wet Slump (inches) 7 6 6 6.5 2.25 3.75 4.00 5 Slump Flow (inches) 18.0 23.0 18.0 22.0 17.5 21.0 20.0 21 VSI 0.0 1.0 0.0 1.0 0.0 0.5 0.0 0.5 T50 (sec.) >30 1.52 >30 1.4 >30 1.86 2.10 1.09 Air Content (%) 3 3.2 4 3.4 4.4 3.4 3.0 3.7 Temp. ( o F) 72 73 76 72 79 74 74 75 Unit Weight (lb/ft 3 ) 146.8 146.6 146.1 145.3 145.68 145.5 146.4 145 Item Pl a n t J ob Si t e Self-Consolidating Concrete Mixtures 194 Table 6.9 - Fresh Concrete Properties for Phase V (cont. on next page) 195 9:AL-36-6 SF 10:AL-36-8 SF 11:AL-36-10 SF 12:AL-36-15 SF 13:AL-36-40 SG 14:AL-36-50 SG 14:AL-36-50 SG (LA) 15:AL-36-60 SG Slump Flow (inches) 29 29.5 29 29 26 29 28 30 VSI 1.5 1.5 1.5 1.5 1.5 2.0 1.5 2 T50 (sec.) 1.16 0.76 0.94 1.16 2.29 1.50 1.56 1.12 Wet Slump (inches) 1 0.25 0.25 0 0 0.25 0.50 0.75 Slump Flow (inches) 22.0 21.5 20.75 22.0 16.0 18.0 21.0 20 VSI 0.0 0.0 0.5 0.5 0.0 1.0 0.5 0.5 T50 (sec.) 1.2 1.01 1.22 1 >30 >30 2.62 2.37 Air Content (%) 3.4 3.7 3.5 5.0 4.1 10.5 2.6 17 Temp. ( o F) 75 76 74 75 83 75 76 76 Unit Weight (lb/ft 3 ) 145.6 145.2 145.68 144.2 147.8 137.4 148.0 136.92 Item Pl a n t Jo b S i t e Self-Consolidating Concrete Mixtures 195 Table 6.9 ? Fresh Concrete Properties for Phase V (cont.) 196 16:AL-36-8 M3 17:AL-36-12 M3 18:AL-36-16 M3 19:AL-36-8 LS 20:AL-36-15 LS 21:AL-36-20 LS Slump Flow (inches) 32 34 31 27.5 27 28 VSI 3 3 3 2 2 2.5 T50 (sec.) 1.15 0.72 0.9 1.44 1.69 1.56 Wet Slump (inches) 6 7 6 2 2.5 3.25 Slump Flow (inches) 22.0 24.0 22.0 17.5 19.5 21.0 VSI 1.0 1.5 1.0 0.5 0.5 0.5 T50 (sec.) 1.06 0.75 1.09 >30 >30 1.16 Air Content (%) 2.9 2.6 2.8 4.5 5 3.7 Temp. ( o F) 76 77 78 77 76 77 Unit Weight (lb/ft 3 ) 147.16 145.68 147.4 144.8 144.2 145.6 Item Pl ant Jo b S i t e Self-Consolidating Concrete Mixtures 196 6.6.2 Hardened Concrete Properties Compressive Strength: The hardened concrete properties determined for Phase V can be seen by looking at Figures 6.27 through 6.38. Each SCC mixture was compared to a base line mixture in order to determine the effect of each method to modify the viscosity on the hardened concrete properties. SCC mixture 1:AL-41-0 VMA was determined to be the appropriate base line mixture for all SCC mixtures prepared at a water-to- cementitious materials ratio of 0.41, while 6:AL-36-33 FA will be used as the base line mixture for all SCC mixtures prepared at a water-to-cementitious materials ratio of 0.36. Figure 6.27 indicates that the base line mixture with no VMA showed higher values of compressive strength at all ages, while the SCC mixture 2:AL-41-2 VMA and 3:AL-41-10 VMA produced the lowest compressive strength results. However, this is most likely due to the influence of other external factors rather than the incorporation of the VMA. The compressive strengths for these mixtures ranged from 6,800 to 7,500 psi at 56-days. By comparing the SCC tures 1:AL-41-0 VMA and 2:AL-41-2 VMA to 5:AL- 36-33 (2) FA and 6:AL-36-33 FA in Figure 6.29 it can be seen that the reduction in water-to-cementitious materials ratio from 0.41 to 0.36 increased the compressive strengths, on average, 2,300 ps 6-days. This is expected since it is a well known fact that the strength is increased as water-to-cementitious materials ratio decreases. Furthermore, the results indicate that compressive strength decreased as the percentage of fly ash increased. This is prim to the fact there is only enough calcium hydroxide in the paste in which the Class F fly ash can react to form calcium silicates. Therefore, higher replacement percentages of fly ash may result in unreacted ash that mix i at 5 the arily due 197 may cause a reduction in the compressive strength. Additionally, Figure 6.29 shows that gth, the typical e, cussed in Section 6.6.1, unusually high amounts of entrapped air of 10.5% and 17 th nd n . the replacement percentages of fly ash above 50% should not be utilized due to the fact the critical 28-day compressive strength of 5,200 psi would not be achieved. The results obtained from the compressive strength tests for both the silica fume and GGBFS mixtures can be seen in Figures 6.31 and 6.33. Unlike Phase IV where the incorporation of the silica fume produced a clear increase in compressive stren incorporation of the silica fume for these materials did not indicate a significant increase in compressive strength at 56-days. This was unexpected since silica fume is known to ly improve the microstructure and increase compressive strengths. Furthermor Figure 6.31 reveals that the SCC mixture with 15% replacement of silica fume showed a decrease in compressive strength compared to the base line mixture. This decrease in compressive strength can be possibly attributed to the fact that the SCC mixture contained a higher percentage of entrapped air compared to the base line mixture. As dis % were obtained at replacement percentages of 50% and 60% GGBFS, respectively. It was determined that as the porosity increased the compressive streng was decreased as shown in Figure 6.33. For example, mixtures 14:AL-36- 50 SG a 14:AL-36-50 SG (LA) was comprised of exactly the same mixture proportions with the exception of the water reducing admixture. According to Figure 6.33, the reduction i entrapped air from 10.5% to 2.6% corresponded to an increase of compressive strength of at least 4,800 psi at 28 and 56-day. This is a remarkable increase in compressive strength, and it shows the significance of the porosity-compressive strength relationship By examining mixtures 13:AL-36-40 SG and 14:AL-36-50 SG (LA) it appears that as 198 replacement of GGBFS increased from 40% to 50%, the compressive strength was increased by 1,400 psi and 1,800 psi at 28 and 56-days, respectively. This is considering the fact that both mixtures contained reasonably low air contents. Moreov er, the utilizat ve of e amount However, ion of GGBFS was found to increase the compressive strength compared to the base line mixture of no less than 1,300 psi at 56-days. Figure 6.35 indicates that the use of Micron 3 produced a decrease in compressi strengths compared to the base line mixture. The SCC mixtures with Micron 3 demonstrated an average compressive strength loss of 1,000 psi at 28 and 56-days. However, it is thought that this decrease in compressive strength is not a result of the Micron 3, but is due to the fact that the Micron 3 was in replacement of cement instead of fly ash. As the replacement percentage of Micron 3 was increased the total percentage fly ash within the SCC mixture increased accordingly. As higher replacement percentages of Micron 3 was introduced into the SCC mixtures, it is probable that th of unreacted fly ash was increased resulting in a decrease in compressive strength. This is analogous to the high fly ash mixtures shown in Figure 6.29. all Micron 3 mixtures were well above the required critical 28-day compressive strength of 5,200 psi despite the reduced compressive strength. Figure 6.37 shows that the introduction of the limestone filler caused a reduction in compressive strength compared to the base line mixture. Figure 6.37 further reveals that the compressive strength of the SCC mixtures is lower as the percentage of limestone filler increases. The reduction in compressive strength is a direct result of the limestone filler being inert. The replacement of cementitious materials by limestone filler is in 199 effect raising the water-to-cementitious materials ratio resulting in lower compressiv strengths. e Modulus of Elasticity: Figures 6.28, 6.30, 6.32, 6.34, 6.36, and 6.38 present the calculated modulus of elasticity values obtained for all SCC mixtures for Phase V. The development for the modulus of elasticity is similar to th e strength development provided in Figu of this res 6.27, 6.29, 6.31, 6.33, 6.35, and 6.37. Furthermore, Figures 6.27 through 6.38 indicate that higher modulus of elasticity values were achieved for higher values compressive strengths. In order to determine if the calculated elastic modulus for the concrete mixtures is of typical sound concrete; the calculated modulus of elasticity in phase of the research was compared with the ACI 318 (2002) Building Code ( ccc fWpsiE '33)( 5.1 ?= ) and the ACI Committee 363 (2002) ?State-of the-Art Report on High-Strength Concrete? ( )000,000,1'000,40)( += cfpsiE c models. Figure 6.39 reveals that the ACI 318 (2002) overestimated the modulus of elasticity for both the ODSC and SCC mixtures. Furthermore, the ACI 318 (2002) was found to increasingly overestimate the modulus of elasticity as the compressive strength increased. This finding coincides with ACI Committee 363 (2002) that found that the equation used by the ACI 318 (2002) was only valid for compressive strengths up to 6,000 psi. On the other hand, the ACI Committee 363 (2002) equation provided an improved and typically conservative estimate for the modulus of elasticity for both the ODSC and SCC mixtures as shown in Figure 6.40. It is believed that the improved and conservative estimation for the modulus of elasticity lies in the fact that the equation provided by the ACI Committee 363 (2002) is valid for compressive strengths from 3,000 to 12,000 psi. 200 Figure 6.27 ? Compressive Strength vs. Concrete Age for VMA Mixtures Figure 6.28 ? Modulus of Elasticity vs. Concrete Age for VMA Mixtures 0 1000 6000 56 Co r e ve S 9000 10000 2000 3000 4000 5000 7000 8000 0 7 14 21 28 35 42 49 Concrete Age (Days) m p s s i t r e ng th (psi ) 1:AL-41-0 VMA 2:AL-41-2 VMA 3:AL-41-10 VMA 4:AL-41-18 VMA 0 2 3 5 6 Concrete Age (Days) 1 .0 x 10 p s i) 1 4 0 7 14 21 28 35 42 49 56 Modu lus of Elast icity ( 6 1:AL-41-0 VMA 2:AL-41-2 VMA 3:AL-41-10 VMA 4:AL-41-18 VMA 201 Figure 6.30 ? Modulus of Elasticity vs. Concrete Age for Fly Ash Mixtu Figure 6.29 ? Compressive Strength vs. Concrete Age for Fly Ash Mixtures 0 1 2 3 4 5 6 0 7 14 21 28 35 42 49 56 Concrete Age (Days) Mod u l u s of E l as ti ci ty (1. 0 x 10 6 ps i) 5:AL:36-33 (2) FA 6:AL-36-33 FA 7:AL-36-40 FA 8:AL-36-50 FA 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000 0 7 14 21 28 35 42 49 56 Concrete Age (Days) Co mp re s s iv e Stre n g t h ( p s i ) 5:AL-36-33 (2) FA 6:AL-36-33 FA 7:AL-36-40 FA 8:AL-36-50 FA res 202 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 0 7 14 21 28 35 42 49 56 Concrete Age (Days) Com p ress i v e S t reng th (p s i ) 12000 6:AL-36-33 FA 9:AL-36-6 SF 10:AL-36-8 SF 11:AL-36-10 SF 12:AL-36-15 SF Figure 6.31 ? Compressive Strength vs. Concrete Age for Silica Fume Mixtures 0 1 2 3 4 5 6 0 7 14 21 28 35 42 49 56 Concrete Age (Days) Mo du l u s of El asti c i t y (1 . 0 x 1 0 6 psi) 6:AL-36-33 FA 9:AL-36-6 SF 10:AL-36-8 SF 11:AL-36-10 SF 12:AL-36-15 SF F s igure 6.32 ? Modulus of Elasticity vs. Concrete Age for Silica Fume Mixture 203 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 12000 13000 14000 0 7 14 21 28 35 42 49 56 Concrete Age (Days) Co mpress i ve Stren g th (p si) 6:AL:36-33 FA 13:AL-36-40 SG 14:AL-36-50 SG 14:AL-36-50 SG (LA) 15:AL-36-60 SG Figure 6.33 ? Compressive Strength vs. Concrete Age for GGBFS Mixtures 0 1 2 3 4 5 6 7 0 7 14 21 28 35 42 49 56 Concrete Age (Days) Modulus of E l asticity ( 1 .0 x 10 6 psi ) 6:AL-36-33 FA 13:AL-36-40 SG 14:AL-36-50 SG 14:AL-36-50 SG (LA) 15:AL-36-60 SG Figure 6.34 ? Modulus of Elasticity vs. Concrete Age for GGBFS Mixtures 204 Figure 6.36 ? Modulus of Elasticity vs. 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 11000 0 7 14 21 28 35 42 49 56 Concrete Age (Days) 12000 C o mpr e ss i ve St r e ngt h ( p si ) 6:AL-36-33 FA 16:AL-36-8 M3 17:AL-36-12 M3 18:AL-36-16 M3 Figure 6.35 - Compressive Strength vs. Concrete Age for Micron 3 Mixtures 0 1 2 3 4 5 6 0 7 14 21 28 35 42 49 56 Concrete Age (Days) M o d u lu s o f Ela s t ic ity (1 .0 x 1 0 6 p s i) 6:AL-36-33 FA 16:AL-36-8 M3 17:AL-36-12 M3 18:AL-36-16 M3 Concrete Age for Micron 3 Mixtures 205 0 1000 2000 3000 4000 5000 6000 7000 8000 9000 10000 0 7 14 21 28 35 42 49 56 Concrete Age (Days) C o mpr essi ve Str e ngt h (psi ) 6:AL-36-33 FA 19:AL-36-8 LS 20:AL-36-15 LS 21:AL-36-20 LS Figure 6.37 ? Compressive Strength vs. Concrete Age for Limestone Mixtures 0 1 2 3 4 5 0 7 14 21 28 35 42 49 56 Concrete Age (Days) Mo dulus of Elasticity (1 x 10 6 p s i) 6 6:AL-36-33 FA 19:AL-36-8 LS 20:AL-36-15 LS 21:AL-36-20 LS Figure 6.38 ? Modulus of Elasticity vs. Concrete Age for Limestone Mixtures 206 Figu Equation for Phase V Figure 6.40 ? Predicted vs. Measured Elastic Modulus According to ACI 363 (2002) Equation for Phase V re 6.39 ? Predicted vs. Measured Elastic Modulus According to ACI 318 (2002) 0 1 2 3 4 5 6 0123456 Measured Elastic Modulus (1.0 x 10 6 psi) P r e d ic te d E l a s t i c M odulus (1. 0 x 1 0 6 psi ) -10% Error +10 Error Line of Equality 0 1 2 3 4 5 6 0123456 Measured Elastic Modulus (1.0 x 10 6 psi) P r edic t e d Ela s tic Modulus (1.0 x 10 6 psi) +10% Error -10% Error Line of Equality 207 6.6.3 Effect of VMA on Fresh Concrete Properties The results in Table 6.9 indicate that the polyethylene glycol VMA was found not to significantly influence the viscosity of the SCC mixtures even at high dosage amounts. However, results may suggest that the incorporation of the polyethylene glycol VMA provided a slight increase in stability compared to the SCC mixtures with similar mixture proportions without VMA at higher values of slump flow. Furthermore, this VMA could possibly be used to overcome deficiencies in mixture constituents in everyday batch plant operations. To study the effects of this viscosity agent on the stability of SCC mixtures when subjected to inaccurate moisture corrections, mixing water was added to or deducted from a base line mixture. This base line mixture was 5:AL-36-33 2 . FA with the following modifications. The first modification is the incorporation of the VMA. Each testing point consisted of one m oz/cwt to 8 oz/cwt for all concrete mixtures. The results in Table 6.10 and Figure 6.41 indicate that at the water content equal to the base line mixture the VMA appears to have minimal effect on the slump flow, stability, or viscosity. However, as additional water was added to the base line mixture, the stability of the mixture was found to be increased with the addition of VMA. The increase in stability can be attributed to the lower slump flow values for the SCC mixtures with VMA at +10 and +20 lb/yd 3 . This decrease in slump flow values for the SCC mixtures that incorporated VMA is most likely due to the fact the VMA is rendering the mixture more robust as excess water is released by the HRWRA. At lower slump flow values and lower free water contents the VMA was found to be ineffective in increasing the stability of the SCC mixtures. The results from ixture that incorporated the 2 oz/cwt of VMA and one mixture without the VMA. Secondly, the HRWR admixture dosage was reduced from 10 208 the study indicate that the use of this type of VMA can decrease the sensitivity of the SCC mixtures due to inaccurate moisture corrections. However, at lower slump flow values the use of the polyethylene glycol VMA may not provide an increase in stability. When low water-to-cementitious materials ratios are used and the SCC mixtures are placed at lower slump flow values, the incorporation of this type of VMA may not be necessary. On the other hand, if moisture variability can not be properly controlled and the SCC mixtures are placed at higher slump flow values the use of a VMA can be beneficial. Table 6.10 ? Tabulated Results of the Effect of VMA on Fresh Concrete Properties Wet Slump (inches) Weight (lb/yd 3 ) Conte (%) Unit Air Fresh Properties Slump Flow nt Temp. ( o F) Diameter (inches) T50 (sec) VSI 20 lb/y 3 +10 lb/yd without VMA 2 3 +0 lb/yd with VMA 3 -20 lb/yd without VMA + d 3 without VMA 9 147.9 2.7 75 31 0.44 3 +20 lb/yd with VMA 8 146.8 4 75 28 0.9 2 3 5 143.64 5 76 27.5 0.96 2.5 +10 lb/yd 3 with VMA 5 142.2 3.5 74 26 1.38 +0 lb/yd without VMA 2 1/2 144.6 5.1 75 25.25 1.56 1.5 3 2 145.4 3.9 75 25 1.69 1 -10 lb/yd 3 without VMA 1 1/4 145.5 3.9 73 24.5 1.47 1 -10 lb/yd with VMA 1/4 144.9 4.2 74 23 2.35 1 3 0 143.7 5 74 21.5 4.75 0 -20 lb/yd 3 with VMA 0 145 4 76 20.5 3.97 0 209 Figure 6.41 ? Graphical Results of the Effect of VMA on Fresh Concrete Properties 6.7 SUMMARY OF RESEARCH FINDINGS ? Based on the results of this research, it is concluded that the early or delayed addition of the HRWRA has no considerable effect on the workability characteristics of the SCC mixtures. ? It was 0 25 30 Unit Water Content (lb/yd 3 ) S l ump F l o w ( i nc he s) 5 10 15 20 30 35 -30 -20 -10 0 10 20 VMA Used VMA Not Used B a s e M i x ur Line t e Additional Water Deducted Water determined that obtaining a water slump before the addition of the HRWRA was suitable for not only ensuring correct moisture corrections, but also ? It was found that obtaining a water slump under real field conditions required a considerable increase in time compared to laboratory conditions. in determining appropriate mixture proportions for the SCC mixtures. 210 ? The ODSC mixtures prepared for this research were found to be within the specified range of slump values at placement in addition to demonstrating typical workability characteristics for wet-hole construction. ? The incorporation of the 4 oz/cwt mid-range water reducing admixture PolyHeed 1025 seems to help maintain the workability of the SCC mixtures better than those that contained only Glenium 3030 NS. ? The ODSC mixtures demonstrated a higher degree of bleeding compared to the SCC mixtures prepared at the same water-to-cementitious materials ratio. ? It was concluded that the HRWRA Glenium 3030 NS was the primary cause of the increased air content for the SCC mixtures. ? The results from this research indicate that at similar slump flow values the viscosity of the SCC mixtures was not increased at high dosage amounts of VMA. However, the results suggest that the VMA may provide a slight increase in stability at higher values of slump flow with increasing VMA dosage. ? At lower slump flow values such as those representing the job site testing; the VMA appears to have no effect on the stability of the SCC mixtures. ? The results show that the use of VMA could possibly be used to overcome deficiencies in mixture constituents in everyday batch plant operations, which mainly consist of inaccurate moisture corrections. ? Based on the materials and mixtures proportions used for Phase V, the incorporation of high amounts of fly ash, Micron 3, or limestone filler as the method to modify the viscosity was found not to increase the viscosity of the SCC mixtures by evidence of the T 50 times. 211 ? The incorporation of the silica fume provided no increase in viscosity by evidence of the T 50 times. However, the utilization of the silica fume was found to improve the stability of the mixtures. ? The incorporation of GGBFS provided a reduction in the water slump and showed a higher degree of stability at the batch plant compared to mixtures prepared with fly ash and at the same water-to-cementitious materials ratio. In addition, the incorporation of the GGBFS seems to slightly increase the T 50 times compared to the silica fume mixtures and the fly ash mixtures at similar slump flow values. ? The SCC mixtures prepared at a water-to-cementitious materials ratio of 0.41 demonstrated slightly lower compressive strengths compared to those of the ODSC mixtures. It is concluded that this is a result of the higher replacement percentage of cement by fly ash for those mixtures. ground granulated blast furnace slag (GGBFS) can provide an increase the ? er caused a reduction in compressive duction ising the ? At a constant water-to-cementitious materials ratio the use of the silica fume or compressive strength. ? High replacement percentages of fly ash may result in unreacted ash that may cause a reduction in the compressive strength. The introduction of the limestone fill strength compared to the base line mixture. The primary reason for the re in compressive strength is a direct result of the limestone filler being inert. The replacement of cementitious materials by limestone filler is in effect ra water-to-cementitious materials ratio resulting in lower compressive strengths. 212 ? The SCC mixtures prepared at the same water-to-cementitious materials rati the ODSC mixtures demonstrated a reduction in RCPT values at 91 and 365-days. The RCPT values obtained at 91-days for the SCC and o as ? ODSC mixtures were to the formation of C-S-H and the CC mixtures. found to decrease with time. This is because as the hydration process proceeds, the interconnected pores that were present at 91-days are being filled by the continuous and overall increase in volume due continuous grow of the calcium hydroxide within the capillaries pores, which lowered the RCPT values at 365-days. ? The introduction ground granulated blast furnace slag (GGBFS) was found to increase the RCPT values compared to the (FA) S 213 CHAPTER 7 PROPOSED EXPERIMENTAL FIELD STUDY 7.1 INTRODUCTION The primary purpose of the field study is to evaluate the use of SCC as a viable material to be used in drilled shaft construction. This field study will provide a means of comparison between self-consolidating concrete and ordinary drilled shaft concrete for both fresh and hardened properties under simulated field conditions. A brief discussion of the proposed field study is shown below. This discussion includes test shafts, fresh concrete property testing, hardened concrete property testing, placement monitoring, testing of non-exhumed shafts, testing of exhumed shafts, and instrumentation. The proposed site for this field study is located approximately 1.25 miles southeast of Nichols, South Carolina as shown in Figure 7.1. Unfortunately, due to long construction delays and design set backs the researcher will have very limited involvement in the field study. As a result, changes to this proposed field study below will almost certainly occur without the researcher?s knowledge. It is recommended that the actual construction details and concrete testing that occurred for this field study be obtained from the research advisors. Furthermore, all testing procedures listed in this chapter should be conducted using current ASTM standards. 214 Project Location Figure 7.1 ? Proposed Field Site (courtesy of Mapquest 2005) 7.2 TEST SHAFTS ? Day 1 Experimental Casting: 2 - 6.0 ? X 30 ft test shafts made with ordinary drilled shaft concrete. One test shaft consisting of ordinary drilled shaft concrete shall be exhumed at 28 days or later after placement for visual inspection and testing. ? Day 2 Experimental Casting: 2 - 6.0 ? X 30 ft test shafts made with self- consolidating concrete (SCC). One test shaft consisting of SCC shall be exhumed at 28 days or later after placement for visual inspection and testing. 7.3 FRESH CONCRETE PROPERTY TESTI ? Slump Test ASTM C 143 (1998) o Performed on all ordinary drilled shaft and time of placement. 215 NG concrete mixtures directly after mixing o The slump of the ordinary drilled shaft concrete mixtures, at the time of placement, shall be 8 ? 1 inches. o Performed periodically on all ordinary drill shaft concrete mixtures for a duration of no less than 5 hours after batching (slump retention). ? Slump Flow Test o Performed on SCC mixtures directly after mixing and time of placement. o The slump flow of the SCC mixtures, at the time of placement, shall be 21 ? 3 inches. o Performed periodically on all self-consolidating concrete mixtures for a duration of no less than 5 hours after batching (slump flow retention). ? Total Air Content and Unit Weight ASTM C 138 (1998) ? J-Ring Test ? Segregation Column ? Bleeding Test ASTM C 232 (1998) ? Setting by Penetration Resistance ASTM C 403 (1998) o 6 ? x 6 inch cylindrical specimens of mortar shall be obtained by wet sieving. 7.4 HARDENED CONCRETE PROPERTY TESTING ? Compressive strength, (f c ): ASTM C 39 (1998) and Elastic Modulus, (E c ): ASTM C 469 (1998) o 3 ? 6 ? x 12 inch molded specimens shall be cast per testing age. o The curing of the specimens shall be done in accordance with ASTM C 31 (1998). o The specimens are to be demolded no earlier than 2 x initial set. o The specimens should be tested at ages of 3, 7, 14, 28, and 56 days. ? Drying Shrinkage: ASTM C 157 (1998) o 3 ? 3 x 3 x 12 inch molded specimens shall be cast per mixture. o The shrinkage bars shall be placed in a lime saturated bath for the first 28 days. Afterwards, the specimens shall be removed from the lime bath and placed in air storage. 216 o The specimens are to be demolded no earlier than 2 x initial set. o The specimens should be tested at 1, 2, 3, 7, 14, 28, 56, 91, 180, and 365 days after removal from lime saturated bath. ? Permeability: ASTM C 1202 (1998) o 3- 4 ? x 8 inch molded specimens shall be cast per testing age. o The specimens are to be demolded no earlier than 2 x initial set. o The curing of the specimens shall be done in accordance with ASTM C 31 (1998). o The specimens should be tested at ages of 91 and 365 days. 7.5 PLACEMENT MONITORING ? The elevation difference between the inside and outside of the rebar cage shall be determined by the use of plumb-bobs. ? Coloring of various concrete loads for the exhumed shafts shall be as follows: o 1 st load shall be of normal color followed by red then black loads. o The red load shall be placed normally followed by a 60 minute delay simulating construction difficulties. 7.6 TESTING OF NON-EXHUMED SHAFTS ? Cores o Location: Cores should be taken at locations of 5, 15, and 25 feet below the surface of the non-exhumed test. ? Compressive strength, (f c ): ASTM C 39 (1998), Elastic Modulus, (E c ): ASTM C 469 (1998), o Testing Age: 56 days (Coring should occur at 51 days) o Size: 4 inch core diameter o Specimens: 4 x 8 inch o Treatment: In accordance with ACI 318 (2002) Section 5.6.5 ? Permeability ASTM C 1202 (1998) o 3 specimens o Cored at 51 days and tested at 91 day 217 o Size: 4 ? x 2 inch disks 7.7 TESTING OF EXHUMED SHAFTS ? Cut Cross-Sections: o Cross sections shall be cut at locations of 5, 15, and 25 below the surface of the exhumed shaft by means of a specialized wire saw. o Visual Assessment of aggregate, coloring of loads, and void distribution shall be conducted for all cut cross-sections. o Impact-echo mapping of cross-section?s density shall be performed for all cut cross-sections. o Extract one 4 ? x 8 inch core from each cut face for calibration of impact- echo. 7.8 INSTRUMENTATION ? Temperature Profiles using I-Buttons (non-exhumed shafts) o I-Buttons should be fixed firmly to steel dowels and placed in 5 feet intervals vertically. o The location of the steel dowels shall be 2.5, 7.5, 12.5, 17.5, 22.5, and 27.5 feet below the surface of the shaft. o There shall be 9 I-Buttons per dowel and 54 I-Buttons per shaft (Total = 108). o Data Collection: The sample interval for the I-Buttons shall be 15 minutes for the first 28 days after placement. After the data has been collected for the first 28 days, the I-Buttons are to be reset at a sample interval of 4 hours for a year. ? Pressure Profile using Load Cells (non-exhumed shafts) o There shall be 6 load cells per shaft located at 2.5, 7.5, 12.5, 17.5, 22.5, and 27.5 feet below the surface of the shaft. ? Cross-Hole Sonic Logging (non-exhumed shafts) o Cross-hole sonic logging should be conducted at locations of 5, 15, and 25 feet below the surface of the shaft. 218 CHAPTER 8 SUMMARY, CONCLUSIONS, AND RECOMMENDATIONS 8.1 INTRODUCTION The laboratory testing program was executed to determine if self-consolidating concrete can be used as a viable material for drilled shaft construction. The primary objectives of this research were to identify appropriate testing techniques, identify characteristics for this specific application, and potential problems or concerns with the use of SCC in drilled shaft construction. Furthermore, the laboratory testing program examined the difference between ordinary drilled shaft concrete and SCC for both fresh and hardened properties. The fresh properties include filling ability, passing ability, segregation resistance, workability over time, bleeding characteristics, and controlled setting, while the hardened properties included the comparison of the compressive strength, elastic modulus, and drying shrinkage. It is hopeful that this research will lead to additional interest in this topic from state and national transportation agencies so that further research in this area can be conducted. This chapter will present the summary and conclusions drawn from this research while offering recommendations based on the summary and conclusions. 8.2 SUMMARY AND CONCLUSIONS The laboratory testing of the fresh and hardened properties for both the ODSC and SCC mixtures provided insight into the use of SCC for drilled shaft applications. 219 The results provided in Chapter 6 were thoroughly examined and conclusions were drawn from these results. The following section presents the summary conclusions drawn from this research testing program. ? The use of set-retarding admixtures can significantly increase the time in which a concrete mixture will remain workable. For drilled shaft applications, the dosage of set-retarding admixture should be adjusted to provide an extended initial set necessary to ensure that the workability of the concrete mixture is maintained for the duration of the pour and to allow for any construction delays in concreting and removal of the temporary casing after concreting is completed. ? For a given dosage of set-retarding admixture, concrete mixtures can experience faster initial set times and increased slump loss when exposed to higher temperatures compared to the laboratory conditions. ? This research shows that merely estimating the additional amount of set-retarding admixture needed in hot weather conditions is not a sufficient measure to ensure that proper retained workability and set times are achieved. The amount of retarding admixture should be based on trial mixes under simulated conditions. ? Results indicate that SCC mixtures can experience an increased workability loss compared to ODSC mixtures when subject to similar mixing conditions. The rate and amount of workability loss depends on the initial workability conditions, degree of agitation, rotational speed of the mixing drum, and duration of mixing. ? The SCC mixtures prepared for this research provided a considerable increase in workability at placement compared to the ODSC mixtures. This enhanced 220 workability may be capable of overcoming placement difficulties associated with tremie placing concrete and congested rebar cages. ? SCC mixtures are more inclined to have a larger change in workability for the same amount of time compared to the ODSC mixtures. ? The Slump Flow, T 50 , and J-Ring tests were deemed acceptable quality control procedures for both laboratory and field conditions for drilled shaft applications. ? It is a must that the static stability of the SCC mixture be part of the determination of the VSI rating. This should be done by observing the SCC mixture in the wheelbarrow or mixing drum directly after the completion of the slump flow test. ? It is concluded that the segregation column can be used to provide a quick and valuable testing procedure for laboratory purposes to determine the probability of a SCC mixture to segregate, especially at higher values of slump flow. ? The critical sand-to-aggregate ratio for these materials for passing ability and segregation resistance is 0.44 or higher. The possibility of blockage and segregation may be increased at sand-to-aggregate ratios below 0.44. ? The SCC mixtures exhibited increased stability as the water slump and water-to- cementitious materials ratio was reduced, which is more apparent at higher values of slump flow. ? The ODSC mixtures demonstrated a higher degree of bleeding compared to the SCC mixtures prepared at the same water-to-cementitious materials ratio. ? If the moisture variability at the batch plant can be properly controlled and since the SCC mixtures for this research are placed at lower values of slump flow (18 to 221 24 inches), it is the opinion of the author that very workable and stable SCC mixtures can be achieved without the use of the polyethylene glycol VMA. ? It was concluded for this research that since the fine and coarse aggregates were high quality and the aggregate volume fraction was not drastically different, the modulus of elasticity of the SCC mixtures was not significantly affected by the varying sand-to-aggregate ratio, which ranged from 0.40 to 0.48. ? The results from this research indicate that with these materials and mixture proportions, the equation provided by ACI 318 (2002) typically overestimated the modulus of elasticity especially at higher values of compressive strengths. Conversely, the equation provided by the ACI Committee 363 (2002) was found to provide an improved and conservative estimate of the modulus of elasticity. ? The water-to-cementitious materials ratio appeared to be the main factor influencing the drying shrinkage. The results indicate that the reduction in water- to-cementitious materials ratio decreased the concrete specimen?s tendency to shrink. ? It can be concluded that the coefficient of permeability decreases as the water-to- cementitious materials ratio is reduced. ? It can be concluded that the introduction of silica fume can considerably reduce the permeability of concrete. 8.3 RECOMMENDATIONS ? Due to the time-dependent effects of the HRWR admixture, it is recommended that laboratory mixing procedures account for transportation time and the HRWR 222 admixture be adjusted to account for the transportation time so that SCC mixture will meet the specified quality control limits upon arrival to the job site. ? It is recommended that field adjustments to the chemical admixtures at the job site be avoided; rather all efforts should be made to correctly batch the chemical admixtures at the source of mixing. ? It is highly suggested that trial mixes be conducted under simulated conditions in order to determine the appropriate set-retarding admixture dosage and proper workability retention is achieved. ? It is suggested that the slump flow and J-Ring test be used as quality control test for the SCC mixtures in both the laboratory and field settings. The segregation column test can be used for laboratory purposes. ? In order to overcome placement problems associated with ODSC mixtures it is recommended that the SCC mixtures be placed at a slump flow ranging from 18 to 24 inches with a VSI rating of 1 or less. It is the opinion of the author that these quality control limits will provide a SCC mixture with sufficient flowability, workability, and stability for both dry and wet-hole construction. ? The following suggestions are offered to overcome issues associated with congested rebar cages. Firstly, the use of small well-graded rounded river gravel should be utilized in applications where the rebar cages are congested. In regards to the SCC mixtures, it is suggested that the use of a small well-graded rounded river gravel with a #7 gradation be used to ensure high passing ability and reduced possibility of segregation. 223 ? It is recommended that water-to-cementitious materials ratios of 0.36 to 0.40 be utilized for drilled shaft applications. ? In order to help reduce the amount of bleed water generated the following suggestions are offered: 1. Use large amounts of supplementary cementitious materials to reduce the amount of free water 2. Air entrainment 3. By reducing the free water content or water-to-cementitious materials ratio 4. Utilizing of new polycarboxylate ester based mid and high range water reducing admixtures 5. Presence of adequate proportions of very fine aggregate particles, which can consist of raising the sand-to-aggregate ratio 6. The use of a binding type VMA ? If a low permeability-high durability drilled shaft concrete mixture is required in areas prone to chemical attack, a low water-to-cementitious materials ratio and/or the incorporation of fine material, such as silica fume, should be used to provide the necessary durability. ? It is recommended that the SCC mixture 9:36-44 FA be utilized for the field study to be conducted in South Carolina. ? Establish a ?high performance concrete criteria? for drilled shaft construction based on the following: 1. Workability requirements 2. Shaft size 3. Congestion of the rebar cages 4. Typical aggregate size and type 5. Retained workability requirements 6. Avoidance of segregation & bleeding 224 REFERENCES ACI Committee 209. 2002. Prediction of Creep, Shrinkage, and Temperature Effects in Concrete Structures. In ACI Manual of Concrete Practice 2002: Part 1. 209R- 1-209R-47. Farmington Hills, Missouri: American Concrete Institute. ACI Committee 318. 2002. Building Code Requirements for Structural Concrete. In ACI Manual of Concrete Practice 2002: Part 4. 363R-1-363R-55. Farmington Hills, Missouri: American Concrete Institute. ACI Committee 363. 2002. State-of-the-Art Report on High-Strength Concrete. In ACI Manual of Concrete Practice 2002: Part 5. 363R-1-363R-55. Farmington Hills, Missouri: American Concrete Institute. ASTM C 31. 1998. Standard Practice for Making and Curing Concrete Test Specimens in the Field. Annual Book of ASTM Standards 1998 Section 04, V. 4.02: Concrete and Aggregates. ASTM C 33. 2002. Standard Specification for Concrete Aggregates. Annual Book of ASTM Standards 2002. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 39. 1998. Standard Test Method for Compressive Strength of Cylindrical Concrete Specimens. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. 225 ASTM C 138. 1998. Standard Test Method for Unit Weight, Yield, and Air Content. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 143. 1998. Standard Test Method for Slump of Hydraulic-Cement Concrete. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 157. 1998. Standard Test Method for Length Change of Hardened Hydraulic-Cement Mortar and Concrete. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 172. 1998. Standard Practice for Sampling Freshly Mixed Concrete. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 192. 1998. Standard Practice for Making and Curing Concrete Test Specimens in the Laboratory. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 204. 1998. Standard Test Method for Fineness of Hydraulic Cement by Air Permeability Apparatus. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 232. 1998. Standard Test Method for Bleeding of Concrete using Method A. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 360. 1998. Standard Test Method for Ball Penetration in Freshly Mixed Hydraulic Cement Concrete. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. 226 ASTM C 403. 1998. Standard Test Method for Time of Setting of Concrete Mixtures by Penetration Resistance. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 469. 1998. Standard Test Method for Static Modulus of Elasticity and Poisson?s Ratio Strength of Concrete in Compression. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 494. 1998. Standard Specification for Chemical Admixtures for Concrete. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 618. 1998. Standard Specification for Coal Fly Ash and Raw or Calcined Natural Pozzolan for Use in Concrete. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates ASTM C 989. 1998. Specification for Ground Granulated Blast Furnace Slag for Use in Concrete and Mortars. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 1202. 1998. Standard Test Method for Electrical Indication of Concrete?s Ability to Resist Chloride Ion Penetration. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. ASTM C 1231. 1998. Standard Practice for use of Unbonded Caps in Determination of Compressive Strength of Hardened Concrete Specimens. Annual Book of ASTM Standards 1998. Section 04, V. 4.02: Concrete and Aggregates. 227 ASTM C XXXX/CXXX. 2004. Standard Test Method for Segregation of Hydraulic- Cement Self-Consolidating Concrete using Column Technique. Proposed ASTM Draft submitted to ASTM for acceptance. ASTM C XXXX/CXXX. 2004. Standard Test Method for Passing Ability of Self- Consolidating Concrete by J-Ring and Slump Cone. Proposed ASTM Draft submitted to ASTM for acceptance. Attiogbe E.K., H.T. See, and Joseph Daczko. 2003. Engineering Properties of Self- Consolidating Concrete. In First North American Conference on the Design and Use of Self-Consolidating Concrete 12-13 November 2002, edited by Surendra Shah, Joseph Daczko, and James Lingscheit, 331-336. Addison, Illinois: Hanley- Wood. Berke, N.S., C.R. Cornman, A.A. Jeknavorian, G.F. Knight, and O. Wallevik. 2003. The Effective Use of Superplasticizers and Viscosity-Modifying Agents in Self- Consolidating Concrete. In First North American Conference on the Design and Use of Self-Consolidating Concrete 12-13 November 2002, edited by Surendra Shah, Joseph Daczko, and James Lingscheit, 165-170. Addison, Illinois: Hanley- Wood. Bonen, David and Surendra Shah. 2004. The Effects of Formulation on the Properties of Self-Consolidating Concrete. In Concrete Science and Engineering: A Tribute to Arnon Bentur: Proceedings of the International RILEM Symposium held in Evanston, USA 24 March 2004, edited by K. Kovler, J. Marchand, S. Mindess, and J. Weiss, 43-56. Cachan Cedex, France: RILEM Publications s.a.r.l 228 Brown, Dan A. 2004. Zen and the Art of Drilled Shaft Construction. In Geotechnical Special Publication 124, ASCE, Vol. 124: 19-33. Brown, Dan A. 2004. Concrete. Presentation presented as part of the ADSC Short Course for Drilled Shaft Foundations. Brown, Dan A. 2004. Personal Commutation. Bury, M.A. and B.J. Christensen. 2003. The Role of Innovative Chemical Admixtures in Producing Self-Consolidating Concrete. In First North American Conference on the Design and Use of Self-Consolidating Concrete 12-13 November 2002, edited by Surendra Shah, Joseph Daczko, and James Lingscheit, 137-140. Addison, Illinois: Hanley-Wood. California Department of Transportation (Caltrans). 2004. Manual and Forms. http://www.dot.ca.gov California Department of Transportation (Caltrans). 2004. Photographs taken by Caltrans (Chapter 3) were provided by Dr. Dan Brown with permission. http://www.dot.ca.gov Camp, W.M, Dan A. Brown, and P.W Mayne. 2002. Construction Methods Effects on Axial Drilled Shaft Performance. Geotechnical Special Publication 116, ASCE: 193-208. Degussa Construction Materials. 2004. Master Builders Admixture Systems. http://www.masterbuilders.com Emborg, Mats. 1999. Rheology Tests for Self-Compacting Concrete-How Useful are they for the Design of Concrete Mix for Full Scale Production? In Self- Compacting Concrete: Proceedings of the First International RILEM Symposium 229 held in Stockholm, Sweden 13-14 September 1999, edited by A. Skarendahl and O. Petersson, 95-105. Cachan Cedex, France: RILEM Publications s.a.r.l Ferraris, Chiara F., Karthik H. Obla, and Russell Hill. 2001. The Influence of Mineral Admixtures on the Rheology of Cement Paste and Concrete. Cement and Concrete Research 31: 245-255. Florida Department of Transportation (FDOT). 2004. Photographs taken by FDOT (Chapter 3) were provided by Dr. Dan Brown with permission. http://www.dot.state.fl.us Gerwick Jr., Ben C. 2004. Placement of Tremie Concrete. Concrete International (January): 41-52. Hodgson, Donald (Trey) Nelson. 2003. Laboratory and Field Investigations of Self- Consolidating Concrete. M.S. Thesis., University of Auburn. Jones, Krista. 2004. The Age-Dependent Stiffness Development of Concrete. M.S. Thesis., University of Auburn. Khayat, Kamal, C. Hu, and H. Monty. 1999. Stability of Self-Consolidating Concrete, Advantages, and Potential Applications. In Self-Compacting Concrete: Proceedings of the First International RILEM Symposium held in Stockholm, Sweden 13-14 September 1999, edited by A. Skarendahl and O. Petersson, 143- 152. Cachan Cedex, France: RILEM Publications s.a.r.l Khayat, Kamal, Joseph Assaad, and Joseph Daczko. 2004. Comparison of Field- Oriented Test Methods to Assess Dynamic Stability of Self-Consolidating Concrete. ACI Materials Journal v. 101, n. 2, (March-April): 168-172. 230 Khayat, Kamal and Joseph Daczko. 2003. The Holistic Approach to Self- Consolidating Concrete. In First North American Conference on the Design and Use of Self-Consolidating Concrete 12-13 November 2002, edited by Surendra Shah, Joseph Daczko, and James Lingscheit, 9-14. Addison, Illinois: Hanley- Wood. Khayat, Kamal and Somnuk Tangtermsirikul. 2000. Fresh Concrete Properties. In State-of-the-Art Report (23) of RILEM Technical Committee 174-SCC, edited by A. Skarendahl and O. Petersson, 17-22. Cachan Cedex, France: RILEM Publications s.a.r.l. Kim, Jin-Keun, Sang Hun Han, Yon Dong Park, and Jae Ho Noh. 1998. Materials Properties of Self-Flowing Concrete. Journal of Materials in Civil Engineering. (July): 244-249. Mapquest. 2005. Mapquest.com Maps, Directions and more. http://www.mapquest.com Mindess, Sidney, J. Francis Young, and David Darwin. 2003. Concrete, 2 nd edition. Upper Saddle River, NJ: Prentice Hall. Mujtaba, Ghulam. 2004. Supplemental Technical Special Provisions for use of Flowing Concrete and Self-Consolidating Concrete for Manufacturing of Precast/Prestressed Concrete Products, Florida Department of Transportation. Gainesville, Florida: State Materials Office. Neville, A.M. 1996. Properties of Concrete, 4 th edition. New York: John Wiley & Sons, Inc. 231 O?Neill, Michael W. and Lymon C. Reese. 1999. Drilled Shafts: Construction Procedures and Design Methods, FHWA Report No. FHWA-IF-99-025, Vol. I, Chapters 1-8: Dallas, Texas: ADSC: The International Association of Foundation Drilling. Okamura, Hajime and Masahiro Ouchi. 1999. Self-Compacting Concrete. Development, Present Use and Future. In Self-Compacting Concrete: Proceedings of the First International RILEM Symposium held in Stockholm, Sweden 13-14 September 1999, edited by A. Skarendahl and O. Petersson, 3-14. Cachan Cedex, France: RILEM Publications s.a.r.l Petersson, Orjan. 1999. Workability. In Final Report of Task2: Brite-EuRam project BRPR-CT96-0366, http://scc.ce.luth.se/public/ report/workability Petersson, Orjan. 2000. L-Shape Box Test. In State-of-the-Art Report (23) of RILEM Technical Committee 174-SCC, edited by A. Skarendahl and O. Petersson, 126- 128. Cachan Cedex, France: RILEM Publications s.a.r.l. Precast/Prestressed Concrete Institute (PCI). 2003. Interim Guidelines for the Use of Self-Consolidating Concrete in Precast/Prestressed Concrete Institute Member Plants, 1 st edition. Chicago Illinois: Precast/Prestressed Concrete Institute. Raghavan, K.P., Sivarama Sarma, and D. Chattopadhyay. 2003. Creep, Shrinkage and Chloride Permeability Properties of Self-Consolidating Concrete. In First North American Conference on the Design and Use of Self-Consolidating Concrete 12- 13 November 2002, edited by Surendra Shah, Joseph Daczko, and James Lingscheit, 307-312. Addison, Illinois: Hanley-Wood. 232 Ramsburg, Paul. 2003. The SCC Test: Inverted or Upright. Concrete Producer v 21, n 7, (July): 34-38. Rols, Sebastien, Jean Ambroise, and Jean Pera. 1999. Effect of Different Viscosity Agents on the Properties of Self-Leveling Concrete. Cement and Concrete Research 29: 261-266. Sedran, T. 2000. Rheology. In Final Report of Task 3: Brite-EuRam project BRPR- CT96-0366, http://scc.ce.luth.se/public/ report/rheology Skarendahl, Ake. 2000. Definitions. Self-Compacting Concrete, In State-of-the-Art Report (23) of RILEM Technical Committee 174-SCC, edited by A. Skarendahl and O. Petersson, 3-5. Cachan Cedex, France: RILEM Publications s.a.r.l. Sonebi, M., and P.J.M. Bartos. 1999. Hardened SCC and its Bond with Reinforcement. In Self-Compacting Concrete: Proceedings of the First International RILEM Symposium held in Stockholm, Sweden 13-14 September 1999, edited by A. Skarendahl and O. Petersson, 275-289. Cachan Cedex, France: RILEM Publications s.a.r.l South Carolina Department of Transportation. 2004. Standard Specifications. http://www.dot.state.sc.us Su, J.K., S.W. Cho, C.C. Yang, and R. Huang. 2002. Effect of Sand Ratio on the Elastic Modulus of Self-Compacting Concrete. Journal of Marine Science and Technology, V.10, N.1: 8-13. Su, Nan, Kung-Chung Hsu, and His-Wen Chai. 2001. A Simple Mix Design Method for Self-Compacting Concrete. Cement and Concrete Research V. 31. (December 2001): 1799-1807. 233 Takada, Kazunori and Somnuk Tangtermsirikul. 2000. Testing of Fresh Concrete. Self-Compacting Concrete, In State-of-the-Art Report (23) of RILEM Technical Committee 174-SCC, edited by A. Skarendahl and O. Petersson, 25-39. Cachan Cedex, France: RILEM Publications s.a.r.l. Texas Department of Transportation (TxDOT). 2004. Photographs taken by TxDOT (Chapter 3) were provided by Dr. Dan Brown with permission. http://www.dot.state.tx.us Tragargh, J. 1999. Microstructural Features and Related Properties of Self- Compacting Concrete. In Self-Compacting Concrete: Proceedings of the First International RILEM Symposium held in Stockholm, Sweden 13-14 September 1999, edited by A. Skarendahl and O. Petersson, 175-186. Cachan Cedex, France: RILEM Publications s.a.r.l Turcry, P., A. Loukili, and K. Haidar. 2003. Mechanical Properties, Plastic Shrinkage and Free Deformation of Self-Consolidating Concrete. In First North American Conference on the Design and Use of Self-Consolidating Concrete 12-13 November 2002, edited by Surendra Shah, Joseph Daczko, and James Lingscheit, 301-306. Addison, Illinois: Hanley-Wood. Yao, Sam X. and Ben C. Gerwick, Jr. 2004. Underwater Concrete: Part 2 Proper Mixture Proportioning. Concrete International (February): 77-82. 234