[go: up one dir, main page]

Academia.eduAcademia.edu
Impact of Hot Mix Asphalt Properties on its Permanent Deformation Behaviour Author Imran Hafeez 05-UET/PhD-CE-22 Supervisor Dr. Mumtaz Ahmed Kamal Professor, Department of Civil Engineering DEPARTMENT OF CIVIL ENGINEERING FACULTY OF CIVIL & ENVIRONMENTAL ENGINEERING UNIVERSITY OF ENGINEERING AND TECHNOLOGY TAXILA October 2009 Impact of Hot Mix Asphalt Properties on its Permanent Deformation Behaviour Author Imran Hafeez 05-UET/PhD-CE-22 A thesis submitted in partial fulfillment of the requirement for the degree of PhD Civil Engineering Thesis Supervisor Dr. Mumtaz Ahmed Kamal Professor, Department of Civil Engineering Thesis Supervisor’s Signature: __________________________________________ External Examiner’s Signature: Prof. Dr. Aziz Akbar CED, UET, Lahore External Examiner’s Signature Dr. Shahab Khanzada NHA, Islamabad DEPARTMENT OF CIVIL ENGINEERING FACULTY OF CIVIL & ENVIRONMENTAL ENGINEERING UNIVERSITY OF ENGINEERING AND TECHNOLOGY, TAXILA October 2009 II Abstract Over the past twenty years; road traffic (both passenger and freight) has grown significantly and loading is progressively getting worse due to the introduction of newer and more powerful trucks with heavier wider bodies in Pakistan. Consequently, premature rutting in the form of shear flow in flexible pavements has been observed during high ambient temperatures. National Highway Authority (NHA), Pakistan has been facing serious threats like, frequent pavement failures, poor riding quality and high maintenance cost. Modifications of Asphalt cement with polymers and rigid pavement design trends, being the end solutions, have increased the construction cost, even more than four times than that of conventional design. In order to cater for the growing axle load demand and to increase the performance of asphalt concrete mixes, true prediction and accurate estimation of probable behavior of mixes need to be investigated. A comprehensive laboratory study was carried out using NHA aggregate gradation Class “A” for asphaltic wearing course, which has commonly been used in the field. The main objective of the research work was to evaluate the effects of temperature and loading on the permanent deformation behavior of mixes, designed with the same aggregate gradation and three commonly available asphalt cement types (A.C). Two gradations i.e. “01” and “02”, within the envelope of the same gradation, were chosen for this study. Three asphalt cement (A.C.) types i.e. two neat A.C with penetration grade “60/70”, “40/50” and one modified A.C. (base “60/70” pen. grade with Elvaloy Terploymer) were selected. Six mixes ranging from finer to coarser aggregate gradation were therefore designed at optimum filler contents, in order to get better mix cohesion, resistant to rutting and to improve serviceability. Three percentages of mineral fillers (i.e. 2.4%, 3.4%. & 4.4%) were trialed in order to determine the optimum filler content for asphalt mixes. Mix design properties i.e. optimum asphalt content, percentage air voids, voids in mineral aggregates, voids filled with asphalt, flow and stability were determined using Marshall Method of Mix Design. Asphalt cement consistency i.e. penetration grade, ductility, softening point and rheological properties i.e. phase angle & complex shear modulus were III also measured. The temperature influence on rheological properties were determined at constant frequency of 10 Hz using Dynamic Shear Rheometer. Two performance tests i.e. uniaxial repeated loading strain & wheel tracking tests were chosen to measure the permanent deformation of asphalt mixes. Relationship between rutting and factors effecting rutting were developed. Domains of intercept and slope coefficient and permanent deformation coefficients alpha (α) and mu (µ) were also determined. New relationships have been developed to predict the impact of Hot Mix Asphalt properties on its permanent deformation behaviour. It has been revealed that mixes with coarser gradation and polymer modified asphalt showed better resistance against rutting. Permanent deformation coefficients, such as α and µ indicate influence of mix properties on mix rut potential. It was further observed that temperature has no effect on α but influences significantly on µ. This study presents relationships developed to correlate temperature and rut values of mixes, tested under wheel tracker (WT). It also presents a comparison between uni-axial repeated load strain tests (or repeated creep test) and wheel tracker tests in terms of regression constants, shift factors and a correlation between both the test methods. Intercept coefficient varies in a narrow range. Based on intercept coefficient, uniaxial repeated load strain test (repeated creep test) did not provide clear ranking of mixes at specified temperatures and stress conditions. The slope of data line between both the tests reduces with increase in temperature and stress levels, which means permanent deformation increases with an increase in temperature and stress. An average value has been determined for approximate shift of creep test data to WT test data. Furthermore a statistical estimation of Mechanistic-Empirical Model has been formulated that relates plastic to elastic strain ratio with different variable that influenced in the mix rut development. Using 54 variables, comprising six mixes designed at low asphalt contents, three temperatures (25, 40 and 550C) and three stress levels (100, 300 and 500 kPa) under repeated load test, a mathematical model was developed to assess the magnitude of plastic to elastic strain ratio. It was observed that plastic to elastic strain ratio is a function of number of load repetitions, temperature, stress levels, shear complex modulus of asphalt cement and IV aggregate gradations respectively. Despite few limitations i.e. single source of aggregate, single method for mix design, two test methods, certain parameters that help to estimate the permanent deformation in the asphalt layers of flexible pavements were successfully captured in the model. Keywords: Flexible Pavement, Hot Mix Asphalt, Uniaxial Repeated Loading, Wheel Tracker, Permanent Deformation V Undertaking I certify that research work titled “Impact of Hot Mix Asphalt Properties on its Permanent Deformation Behaviour” is my own work. The work has not been presented elsewhere for assessment. Where material has been used from other sources it has been properly acknowledged / referred. Signature of Scholar Imran Hafeez 05-UET/PhD-CE-22 VI Acknowledgement I would like to thank those individuals who were instrumental in the completion of this research. Specially, I would like to thank my supervisor, Professor Dr. Mumtaz Ahmed Kamal, for willingly sharing his knowledge, guidance, encouragement and patience. His support and enthusiasm are greatly appreciated. I look forward to future endeavors. I would also like to thank Professor Dr. Waseem Mirza for sharing his experience and for his advice and fruitful comments. Special thanks to Professor Dr. Tanveer Iqbal Qayyum for his help and advice. Thanks to my examining committee members, Professor Dr. Hashim Nisar and Prof. Dr. Qaiser Uz-Zman. My special thanks to Taxila Institute of Transportation Engineering staff for their support and help in conducting this research and their patience and understanding. I would like to express my appreciation to persons working in Transportation Engineering Laboratory. I would like to give my deepest thanks to my wife, my daughter and my son for their understanding and unlimited patience. Without their support this work would have not been possible. VII Table of Contents Abstract III Undertaking VI Acknowledgement VII VIII Table of contents List of Figures XII List of Tables XV List of Frequently used symbols and abbreviations Chapter One: XVII Introduction 1.1 Background 02 1.2 Problem Statement 03 1.3 Research Objectives 03 1.4 Research Methodology 04 1.5 Organization 06 Chapter Two: Literature Review 2.1 Introduction 09 2.2 Permanent deformation (rutting) of Asphalt Mixes 09 2.3 Influence of Asphalt Mixture Properties on Rutting 12 2.4 Influence of Aggregates Properties on Rutting 13 2.5 Influence of Mineral Filler’s Properties on Rutting 15 2.6 Influence of Asphalt Cement Properties on Rutting 16 2.7 Stress Strain Behavior of Asphalt Materials 21 2.8 Resistance to Permanent Deformation of HMA 24 2.9 Permanent Deformation Prediction Models 27 VIII Chapter Three: Material Characterization 3.1 Introduction 34 3.2 Aggregates for Asphalt Mixtures 34 3.2.1 Size and Grading 35 3.2.2 Particle Shape 36 3.2.3 Absorption 38 3.2.4 Clay Content (Sand Equivalent) 38 3.2.5 Toughness 39 3.2.6 Soundness 39 3.2.7 Deleterious Materials 40 3.3 3.4 Chapter Four: Characteristics of Asphalt Cement 41 3.3.1 Consistency of Asphalt Cement 41 3.3.2 Elastic Recovery 43 3.3.3 Torsional Recovery 44 Dynamic Shear Rheometer (DSR) 45 Mix Design Methods 4.1 Introduction 50 4.2 Combined Grading of Aggregates 50 4.3 Asphalt Mixtures 52 4.4 Mix Design Properties 52 4.5 Mineral Filler Optimization 54 4.6 Asphalt Mixtures Volumetric 54 4.6.1 Specific Gravity 55 4.6.2 Voids in Mineral Aggregates, VMA 56 4.6.3 Air Voids, Va 57 4.6.4 Voids Filled with Asphalt, VFA 57 IX Chapter Five: Uniaxial Repeated Load Strain Test (URLST) 5.1 Introduction 60 5.2 Universal Testing Machine (UTM-5P) 60 5.3 Uniaxial Repeated Load Strain Test (URLST) 61 5.4 Test Conditions 62 5.4.1 Load Conditions 62 5.4.2 Temperature Conditions 62 5.5 Testing Methodology 62 5.6 Discussion of URLST Results 63 5.7 Resilient Strain 65 5.8 Summary of Results 70 Chapter Six: Wheel Tracking Test 6.1 Introduction 72 6.2 Wheel Tracking Device 72 6.3 Specimen Preparation on Roller Compactor 75 6.4 Discussion of Results 78 Chapter Seven: Results and Discussions 7.1 Introduction 83 7.2 Regression Coefficients 83 7.2.1 Intercept Coefficient (a) 83 7.2.2 Slope Coefficient (b) 85 7.3 Permanent Deformation Coefficient 86 7.3.1 Alpha (α) 86 7.3.2 Mu (µ) 88 7.4 Comparisons of Results to Other Researchers 90 7.5 Regression Analysis 92 7.6 Ranking of Mixes using Intercept Coefficient 93 7.7 Shift Factor Computations 94 7.8 Correlations Between URLST and WT Test 95 7.9 Summary 100 X Chapter Eight: Modeling the Permanent Deformation Properties of Hot Mix Asphalt 8.1 Introduction 102 8.2 Data Analysis and Model Development 103 8.2.1 General Trends of Variables 103 8.2.2 Development of Prediction Model 106 8.2.3 Comparison of Computed Model with 108 MEPDG Model 8.2.4 Chapter Nine: Sensitivity Analysis 109 Conclusions and Recommendations 9.1 Introduction 112 9.1 Conclusions 112 9.2 Recommendations for Future Study 113 REFERENCES 114 ANNEXURES 126 Annexure –A: Plots of Accumulative Strain 128 Annexure –B: Trends of Accumulated Strains in Mixes 137 Annexure –C: Plots of Resilient Strain 146 Annexure –D: Influence of Wheel Tracker Load Cycle on Rut Depth 149 XI List of Figures Fig. No. Description Page No. 1.1 Flow Chart Diagram (Scope of Work) 5 2.1 Accumulated Plastic Strains in Pavement 9 2.2 Typical Repeated Load Permanent Deformation Behaviors of 11 Pavement Materials 2.3 Rutting Caused by Shear Displacement of Weak Asphalt Layer 12 2.4 Temperature Shift Behavior of Asphalt Binder 17 2.5 Microscopic View of Liquid Flow Properties 18 2.6 Stress Strain Behavior of Bituminous Material 19 2.7 Shear Loading Behavior of Aggregate 22 2.8 Contrasting Stone Skeletons 22 2.9 Permanent Strain under Load Pulse 23 2.10 Accumulated Permanent Strain Under Repeated Load 23 2.11 26 2.12 Effect of Aggregate and Mortar on the failure Behavior of Asphalt Mixtures Log-Log Form of Power Model 29 3.1 Aggregate Gradation ‘01’ and ‘02’ under NHA Class ‘A’ 36 3.2 Briquette Specimens 43 3.3 Trimming of Specimens 43 3.4 Torsional Recovery Test Apparatus 44 3.5 Dynamic Shear Rheometer 46 3.6 Test Specimen on DSR 46 3.7 Influence of Temperature on Linear Visco-Elastic Properties 48 5.1 Universal Testing Machines (UTM-5P) 61 5.2 Influence of Temperature and Stress Levels on Accumulative Strain 64 5.3 Influence of Stress Levels on Resilient Strain, Mix 1a 67 5.4 Influence of Stress Levels on Resilient Strain, Mix 1b 67 5.5 Influence of Stress Levels on Resilient Strain, Mix 1c 68 5.6 Influence of stress levels on resilient strain, Mix 2a 68 XII Fig. No. Description Page No. 5.7 Influence of Stress Levels on Resilient Strain, Mix 2b 69 5.8 Influence of Stress Levels on Resilient Strain, Mix 2c 69 6.1 Wheel Tracker 72 6.2 Wheel Tracker Solid Rubber Tyre 73 6.3 Wheel Tracker out put Display 74 6.4 Slab Compaction on Roller Compacter 75 6.5 Rut Development in Mixes 77 6.6 Relationship of Load Cycle & Rut Depth, Mix 1a 78 6.7 Relationship of Load Cycle & Rut Depth, Mix 1b 79 6.8 Relationship of Load Cycle & Rut Depth, Mix 1c 79 6.9 Relationship of Load Cycle & Rut Depth, Mix 2a 80 6.10 Relationship of Load Cycle & Rut Depth, Mix 2b 80 6.11 Relationship of Load Cycle & Rut Depth, Mix 2c 81 7.1 Influence of Temperature and Stress Levels on Alpha 87 7.2 Influence of Temperature and Stress Levels on Mu 89 7.3 Domains of Alpha 91 7.4 Domains of Mu 91 7.5 Influence of Test Type on Intercept Coefficient 92 7.6 Shift of URLST Data to Wheel Tracker Data 94 7.7 Correlations between URLST and WT Test Data at 250C & 100kPa 95 7.8 Correlations between URLST and WT Test Data at 250C & 300kPa 96 7.9 Correlations between URLST and WT Test Data at 250C & 500kPa 96 7.10 Correlations between URLST and WT Test Data at 400C & 100kPa 97 7.11 Correlations between URLST and WT Test Data at 400C & 300kPa 97 0 7.12 Correlations between URLST and WT Test Data at 40 C & 500kPa 98 7.13 Correlations between URLST and WT Test Data at 550C & 100kPa 98 7.14 Correlations between URLST and WT Test Data at 550C & 300kPa 99 7.15 Correlations between URLST and WT Test Data at 550C & 500kPa 99 8.1 General Trends of Plastic Strains 104 XIII Fig. No. Description Page No. 8.2 General Trends of Elastic Strains 105 8.3 Measured Versus Predicted εp/εr, 10 in/in 107 8.4 MEPDG Versus Computed Model εp/εr, 106 in/in 108 8.5 Sensitivity of εp/εr Model to Predictor Variables 110 6 XIV List of Tables Table Description Page No. No. 3.1 Adopted Aggregate Gradations & Specifications 35 3.2 Physical Properties Of Aggregates 40 3.3 Consistency of Different Types of Ac 42 3.4 Specifications of Asphalt Cement 42 3.5 Elastic Recovery Test Results 44 3.6 Torsional Recovery Test Results 45 3.7 Rheological Properties of Binders 47 4.1 Summary of Sieve Analysis Results 51 4.2 Adopted Gradations & Specifications 51 4.3 Selection of Mix Types 52 4.4 Hot Mix Asphalt Design Properties 53 4.5 Mineral Filler to Asphalt Cement Ratio 54 4.6 Hot Mix Asphalt Design Volumetric Properties 55 5.1 Specimens at Each Test Condition 63 5.2 Mean Accumulated Strain (εp) of Mixes 63 5.3 Resilient Strain of Mixes 66 6.1 Details of Specimens 75 6.2 Compaction Pressures 76 6.3 Rut depth of Mixes Measured on Wheel tracker 76 6.4 Intercept Coefficient “a” of Mixes 81 6.5 Slope Coefficient “b” of Mixes 81 7.1 Intercept Coefficient ‘a’ at 250C 84 7.2 Intercept Coefficient ‘a’ at 400C 84 7.3 Intercept Coefficient ‘a’ at 550C 84 7.4 Slope Coefficient ‘b’ at 250C 85 0 7.5 Slope Coefficient "b” at 40 C 85 7.6 Slope Coefficient "b” at 550C 85 XV Table Description Page No. No. 7.7 Permanent Deformation Parameter "α” at 250C 86 7.8 Permanent Deformation Parameter "α” at 400C 86 7.9 Permanent Deformation Parameter "α” at 550C 86 7.10 Permanent Deformation Parameter “µ” at 250C 88 7.11 Permanent Deformation Parameter “µ” at 400C 88 7.12 Permanent Deformation Parameter “µ” at 550C 88 7.13 Ranking of Mixes for Wheel Tracker 93 7.14 Ranking of Mixes for URLST 93 8.1 Influence of Test Conditions and Mix Parameters on Strain 106 XVI List of Frequently used symbols and abbreviations Symbols Meaning a Intercept Coefficient α Alpha b Slope Coefficient εr Resilient Strain εp Plastic Strain G Complex Shear Modulus µ Mu s Normal Stress T Temperature Gmm Maximum Theoretical Specific Gravity Gmb Bulk Specific Gravity Abbreviation Meaning AASHTO American Association of State Highway and Transportation Officials ASTM American Society for Testing Materials AC Asphalt Cement NHA National Highway Authority MEPDG Mechanistic-Empirical Pavement Design Guide PMA Polymer Modified Asphalt HMA Hot Mix Asphalt WT Wheel Tracker URLST Uniaxial Repeated Load Strain Test UTM Universal Testing Machine LVDT Linear Variable Displacement Transformers VMA Voids in Mineral Aggregate VFA Voids Filled with Asphalt DSR Dynamic Shear Rheometer XVII Chapter One 1 Chapter One Introduction 1.1 Background Inconsistent asphalt-concrete properties together with illegal Axle loads, high tire pressures and high temperatures are considered to be the most commonly observed causes behind pavement frequent failures, in Pakistan. National Highway Authority (NHA), Pakistan has continuously been modifying the aggregate gradations and penetration grade of asphalt cements, without prior investigation of the mix behaviour under the prevailing axle load and environmental conditions of the country. Traffic growth rate on major highways in Pakistan, in particular the truck traffic has been increasing day by day due to its geostrategic locations and international trade corridor for the countries like, China, Afghanistan and many states of Russia. Indiscriminate use of materials in the construction of flexible pavements, especially in asphalt- concrete layers, has become one of the major causes of failures, in areas, having high temperatures and vehicle-load profile. Premature rutting in the form of shear flow of asphalt concrete, being the consequences, has directly been effecting the pavements service life, riding quality and their economic life cycles cost. True prediction of asphaltic material behaviours and their precise selection on the basis of performance can be one of the solutions towards this chaotic problem. At higher temperatures i.e. 40oC and above, the rutting susceptibility of asphalt mixes needs to be studied in the laboratory before its laying at site. Comprehensive laboratory investigation is required, to study the influence of physical and mechanical properties of aggregates on rutting resistance or the permanent deformation behaviour of asphalt mixes. 2 1.2 Problem Statement The accumulation of permanent deformation in the asphalt surfacing layer appears to be the major cause of rutting. To minimize this, it is necessary to pay extra attention to material selection, mixture design and rutting measurement techniques. The questions of how to measure rutting resistance of asphalt mixtures, what parameters to use as a measure of resistance, and how to model and predict the development of permanent deformation need to be addressed. In particular, measurement of rutting with simple performance test has become the focus of current research. A number of studies were carried out to investigate the permanent deformation behaviour of asphalt concrete materials using various testing procedures and mainly based on uniaxial tests. There was a need to make a more detailed study of the permanent deformation response of asphalt concrete mixtures. An attempt has been made in this study to tackle the issues raised in the preceding paragraphs. Based on laboratory tests that are judged to be simulative of field loading conditions, the study attempts to provide more knowledge on the effect of volumetric composition, loading, and temperature conditions on permanent deformation response of asphalt concrete mixtures. In particular substantial effort has been made to correlate two performance test procedures in terms of permanent deformation or rutting. Modeling the permanent deformation behaviour of asphalt concrete mixes forms the other major part of this study. 1.3 Research Objectives The objectives of this research were: 1. To design asphalt concrete mixes for NHA, Class-A, aggregate gradation at optimum filler content, using three commonly used AC types and to determine the effect of asphalt cement penetration grade on HMA properties. 3 2. To determine the influence of stress levels, temperatures on permanent deformation coefficients (α) and mu (µ) and their domains of asphalt mixes. 3. To evaluate the resistance to permanent deformation of asphaltic mix using wheel tracker 4. To define a relationship in terms of plastic to elastic strain ratio from laboratory data using unconfined repeated creep test and to measure its validation by comparing with AASHTO Mechanistic-Empirical Pavement Design Guide (MEPDG) Model 1.4 Research Methodology The methodology adopted to meet the objectives of this study involves a review of literature and a laboratory investigation. The literature review is conducted to identify important component material properties that influence the permanent deformation response of mixtures, and available permanent deformation models and their theoretical basis. Testing methods that are used to characterize permanent deformation property of asphalt mixtures are also reviewed. The laboratory investigation is conducted using two testing procedures; the uniaxial repeated load strain test (repeated creep test) and the wheel tracker. Specimens made with six mixes are tested in both procedures. The uniaxial repeated load strain test (repeated creep test) results are used both for modeling purposes and evaluation of the effect of various factors on the permanent deformation response. The wheel tracker test results are used to study the rutting potential of mixes under different temperature conditions or to define a measure of resistance to rutting (permanent deformation). Scope of experimental work has been shown in Fig 1.1. 4 Scope of Experimental Work Study of Properties of Aggregates & AC Hot Mix Asphalt Design (Six Mixes) Performance testing on HMA Specimens Physical & Mechanical Properties of Aggregate Shape Test Consistency and Rheological Properties of Asphalt Cement Softening Point Tests Water Absorption Test Penetration Tests LAA Tests Ductility Tests Soundness Test Sp. Gravity Tests Sand Equivalent Test Phase Angle Deleterious Test Uni-axial Repeated Load Strain Test: (At 25, 40, 550C temperature) & (100, 300, 500 kPa, stress levels) Wheel Tracker Test (At 25, 40, 550C temperature) & Standard 720 Newton Load Shear Complex Modulus Analysis of results Figure 1.1: Flow Chart Diagram (Scope of Work) 5 1.5 Organization The work carried out to achieve these objectives is described in the following chapters: Chapter Two provides a background and literature review on permanent deformation of asphalt mixtures. The chapter entails a brief discussion of the various issues related to permanent deformation of asphalt mixtures. The general concept is also offered in the context of the stress-strain behaviour and resistance to permanent deformation of asphalt mixtures. Particular attention was paid to the influence of the different components in the asphalt mixture on the permanent deformation behaviour. Permanent strain models have also been reviewed carefully. Chapter Three explains material characterization, which includes physical and mechanical properties of coarse aggregates, mineral filler, consistency of two neat binders and polymer modified asphalt. Modification of PMA and its formulation have also been reported in this chapter. Results of rheological characterization of PMA, carried out on Dynamic Shear Rheometer at different temperatures have also been provided. Chapter Four explicates procedure adopted for the design of six HMA mixes, using two aggregate gradations, three AC types and optimum filler content. Results of volumetric properties, stability, stiffness index of mixes and effects of mineral filler percentages to design properties of HMA have also been reported in this chapter. Chapter Five elucidates the measure of resistance to permanent deformation of HMA, under Uni-axial Repeated Load strain testing at different temperatures and stress levels. Mixes behaviour in terms of elastic response i.e. resilient strain, resilient modulus and plastic response i.e. accumulative strains were studied. The results were then plotted among different variables for an in depth study of the behaviour of different mixes. Chapter Six reports mixes behaviour under accelerated loading using wheel Tracker. This chapter also provides relationships between load cycles and rut depth at different temperatures. 6 Chapter Seven provides and discusses summary of results in terms of mixes (with different aggregate grading and AC types) behaviour under different conditions of loadings and temperatures. Relationships among different variables and the outputs in terms of permanent deformation behaviour of mixes have been presented graphically. Chapter Eight convolutes modeling the permanent deformation behaviour of apshlt mixes using the uniaxial load strain (creep) test and a comparison of Mechanistic-Empirical Pavement Design Guide Model with the proposed model. It also concludes the most influential parameter in the permanent deformation of mixes. Chapter Nine concludes the findings of the research work and suggests recommendations. 7 Chapter Two 8 Chapter Two Literature Review 2.1 Introduction Rutting or permanent deformation is considered one of the major distress mechanisms in flexible pavements, which may occur in all layers of the pavement structure and results from lateral distortion and densification. In this chapter, rutting specifically in asphalt layers of flexible pavements, which is mainly due to Hot Mix Asphalt (HMA), are defined and its possible causes and distress mechanisms in particular have been discussed. The consideration of permanent deformation at mixes design and its measurement are also discussed. 2.2 Permanent Deformation (Rutting) of Asphalt Mixes Permanent deformation in asphalt (flexible) pavements, commonly referred to rutting, usually consists of longitudinal depressions in the wheel paths, which are an accumulation of small amounts of unrecoverable deformation caused by each load application as shown in Figure 2.1 (Asphalt Institute,1996). If an asphalt mixture ruts, it is normally because the mixture has insufficient shear strength to support the stresses to which it is submitted. (Sousa et al. 1991) Figure 2.1: Accumulated Plastic strains in Pavements (After Asphalt Institute, 1996) 9 Researches in the history showed that rutting in the Hot Mix Asphalt (HMA) layer will generally occur within the top 3- to 5-in. If a poor quality HMA mixture is being used, increasing the thickness of this poor quality layer will not decrease the rutting in the HMA layer. In fact, improving the material properties and mix characteristics will be significant in decreasing the rut depth (Kennedy, et al, 1996). There are several wheel path rutting classifications, one of which was provided in 1979 by the Federal Highway Administration, which classified rutting into three levels of severity: 1. Low, from 6 to 12.5 mm (0.25 to 0.5 inches), 2. Medium, from 12.5 to 25 mm (0.5 to 1.0 inches), and 3. High, over 25 mm (1 inch). For normal cross slope values, a rut depth of 12.5 mm (0.5 inch) is typically accepted as the maximum allowable rut depth (Huang, 1993 & Kennedy, et al, 1996). Mechanistic-Empirical Design Guide (MEPDG) has defined three distinct stages for the permanent deformation behaviour of pavement asphalt materials under a given set of material, load and environmental conditions. Primary stage has high initial level of rutting, with a decreasing rate of plastic deformations, predominantly associated with volumetric change. Secondary stage has small rate of rutting exhibiting a constant rate of change of rutting that is also associated with volumetric changes; however, shear deformations increase at increasing rate. While the tertiary stage has a high level of rutting predominantly associated with plastic (shear) deformations under no volume change conditions as shown in Figure 2.2 (AASHTO Design Guide, 2002). If an asphalt material is loaded with a stress that is above the flow strength of the material, at that temperature the material will start to deform (Stumpf, 2007). First the material will deform rapidly, then, after some strain hardening has taken place, the material gets to a stage with a lower creep rate as shown in Figure 2.2. This stage is known as secondary creep, or steady state creep. In the third stage the material becomes unstable and rapid collapse is the result. 10 Permanent Strain εp Primary Secondary Tertiary Flow Point Load Repetitions Figure 2.2 Typical Repeated Load Permanent Deformation Behaviours of Pavement Materials (AASHTO Design Guide, 2002) The proportion of permanent deformation taking place as shown in Figure 2.2 in the different creep phases is important. The critical rut depth is generally set at 10 mm, if this depth is reached in the primary phase or in the first part of the secondary phase, the functional life of the Hot Mix Asphalt (HMA) layer is reduced drastically. In secondary phase, the rate of deformation slows down considerably. The dominant mode of deformation is caused by shear stress that overcomes flow strength of the material. The flow strength consists of two components, i.e. friction and cohesion. The deformation takes place in small iterations with each load application. Eventually the void condition and the level of permanent strain will cause the HMA to enter the tertiary phase and rapid unstable shear failure occurs (Carpenter, 1993). Traffic-associated permanent deformation, rutting in particular, results from a rather complex combination of densification and plastic flow mechanisms. Plastic flow involves essentially no volume change, and gives rise to shear displacements in which both depression and heave are usually manifested as shown in Figure 2.3. 11 Figure 2.3: Rutting caused by Shear Displacement of Weak Asphalt Layer (After Asphalt Institute 1996) Plastic flow occurs when the shear stresses imposed by traffic exceed the inherent strength of the pavement layers mostly at high temperatures. Multiple studies have identified this mechanism as a primary cause of rutting problems in North America (Huber and Heiman 1989, Sousa, et al. 1991, Brown and Cross 1992). 2.3 Influence of Asphalt Mixtures Properties on Rutting Brown and Pell (1974) concluded that a dense graded asphalt mixture exhibits less deformation than a gap graded mixture due to less aggregate interlocking in gap graded mixture. Evidences show that the effects of rutting can be reduced by use of dense aggregate gradations. On proper compaction, mixtures with dense or continuous aggregate gradations are more closely spaced than open or gap graded mixtures and therefore have fewer voids. Also at higher temperatures, the aggregate interlocking becomes more prominent so gap graded mixtures are more susceptible to rutting at higher temperature which was later on confirmed by test track results [Huang 1995, White 2003, Sivasubramaniam 2004]. Surface texture of the aggregate is particularly important 12 for good rutting resistance in thicker asphalt-bound layers and hotter climates where a rough surface texture is required. Shape of the particle is also an important factor. Mathews and Monismith (2003) studied the effect of excessive asphalt cement content, excessive fine grained aggregate and high percentages of natural, rounded aggregate particles and concluded that this excessiveness can be a common material-related causes of permanent deformation. 2.4 Influence of Aggregates properties on Rutting Rutting resistance of asphalt concrete under traffic and environmental loads depends on the aggregate structure in the asphalt mix. Aggregate gradation and aggregate shape properties or morphology of aggregate materials have been recognized by the Strategic Highway Research Program (SHRP) among the top factors that influence the stability of hot mix asphalt (HMA). In dense graded asphalt mixes, coarse aggregate size and shape properties are believed to some extent contribute to the rutting resistance of asphalt concrete. Previous research studies realizes the important role, the coarse aggregate plays in the rutting behaviour of HMA related aggregate structure stability to coarse aggregate morphologies (Klaus, 2003). Conversely, instead of locking together, smooth, rounded aggregate particles tend to slide past each other. If the aggregate provides a high degree of internal friction (Ø), the shear strength of the asphalt mixture will be increased and, therefore, the resistance to rutting. This is accomplished by selecting an aggregate that is angular, cubical, has a rough surface texture, and is graded in a manner to develop particle to particle contact Mc Gennis et al, (1994). Fred (1967) reported that aggregate gradation appeared to have more influence than aggregate type (at constant asphalt content) and at longer time of loading. He also concluded that the temperature susceptibility characteristics of the asphalt appear to have more influence at longer time of loading. 13 Crawford (1989) concluded from a study related to tender mixtures that particle shape and the amount of material passing No. 4 sieve (4.75-mm) were major factors contributing to the tenderness of an asphalt concrete mixture. He also stated that rounded, uncrushed aggregates are more likely to contribute to tender mixtures and, therefore, more rutting susceptible, especially as the amount of uncrushed material passing No. 4 sieve increases. Kim et al. (1992) demonstrated that aggregate type has significant effects on fatigue resistance and permanent deformation of asphalt concrete. Gradation only had minor effects on permanent deformation. Interactions of aggregate type with gradation, asphalt type, air voids, and temperature were found to be significant for the permanent deformation of asphalt concrete Krutz and Sebaaly (1993) evaluated the effects of aggregate gradation on permanent deformation of HMA mixtures for the Nevada Department of Transportation and concluded that the best aggregate gradation is dependent on the type and source of aggregate and the coarse aggregate gradations (bottom of band) performed the worst and fine aggregate gradations (middle and top band) produced better performing mixtures. Yeggoni et al. (1994) conducted a laboratory study to evaluate the influence of coarse aggregate shape and texture on permanent deformation characteristics of HMA mixtures. The authors concluded that an increase in the percentage of crushed coarse aggregate resulted in increased Hveem stability, Marshall Stability, and resistance to permanent deformation. They also found a strong correlation between rutting potential and the shape of the coarse aggregate particles as measured using image analysis. Kennedy et al. (1996) stated that, in order to prevent permanent deformation of HMA pavements, one should avoid gradations near the maximum density because, although they theoretically produces the strongest HMA mixtures, due to their relatively low voids in the mineral aggregate, these types of mixtures would be very sensitive to asphalt content and might presented the risk of flushing due to inevitable variations during construction. It would be better to use aggregates with angular particles because they exhibited greater interlock and internal friction and, hence, resulted in greater mechanical stability than rounded particles. It would be 14 better to use aggregates with rough surface texture because they might tend to form stronger mechanical bonds when compared to smooth-textured aggregates and provided higher VMA in a compacted mass. Mallick and Kandhal (2001) summarized their observations by stating that the statistical analyses of Asphalt Pavement Analyzer (APA) rut depth data obtained on all mixtures indicated significant differences in performance among different gradations. They observed that, for granite and limestone, Below Restricted Zone (BRZ) generally exhibited the highest and Through Restricted Zone (TRZ) exhibited the lowest rut depths, and above restricted zone (ARZ) showed intermediate rut depths. For river gravel mixtures, the order from highest to lowest rut depth was ARZ, BRZ, and TRZ. The BRZ limestone mixture yielded the highest peak shearing strain for both wearing and binder courses. It is commonly understood that larger, more angular aggregates, with rough surface texture increases rutting resistance (Button et al. 1990, Sousa et al. 1991, Brown and Bassett 1990, Kandhal and Mallick 2001). Angular rocks are considered to provide better stone on stone interlock than rounded aggregate, there by reducing the susceptibility to rutting (Asphalt Institute 1996, Ahlrich 1996, Marks et al. 2001). 2.5 Influence of Mineral Fillers Properties on Rutting Herrin and Goetz (1954) observed from a laboratory evaluation that the strength of the asphalt mixture, regardless of the type of coarse aggregate, increased substantially when the fine aggregate was changed from rounded sand to crushed fine aggregates. Uge and Van de Loo (1974) found that at moderate or high temperature, the relative displacements of mineral particles occurring during laying or compaction under prolonged loading may accelerate rutting. Consequently, they recommended the use of mixtures with low workability in order to improve the arrangement of mineral skeleton and internal friction and hence to minimize rutting potential. They also concluded that harsh mixtures that are well compacted after laying will be highly resistant to rutting. 15 Asphalt Institute (1995) in Marshall Method of mix design reported that mix durability has also been related to the amount of fine “dust or dirt” particles in the mixture. Excessive fine lowered the quality of the asphalt film on the aggregate. Depending on the size of these particles, the mix may be stiffer or tenderer. Shashidhar et al (1999) worked on the maximum packing fraction and generalized Einstein coefficient to characterize the stiffening potential of mineral filler in Hot Mix Asphalt (HMA) and proposed that those parameters had contributed towards stiffening and physio-chemical improvement of asphalt concrete. Ahmad et al. (2004) studied the effect of filler type especially Portland cement and lime stone on Marshall stability and retained strength of asphalt concrete in order to compare the effect of different types and ratios of mineral filler on the strength properties of asphalt mixes. They concluded that cement filler resulted higher values of retained strength. National Cooperative Highway Research Program (2005) has reported that mineral filler were originally added to dense-graded asphalt concrete to fill the voids in the aggregate skeleton and to reduce the voids in the mixture. The additions of fines to the asphalt binder can have three main effects; extend the asphalt binder, or stiffness the asphalt binder, or both. Huang et al (2007) investigated the effect of mineral filler on mix design and performance characteristics of HMA mixtures by selecting three types of mineral fillers and four filler contents in order to study the relationship between filler contents and rut depth. They concluded that filler with rough texture and high percentage usually increases the stiffness and decreases the rut depth. They further recommended that filler contents ranges would be required to be investigated in order to ensure the performance of the mixture. 2.6 Influence of Asphalt Cement Properties on Rutting Asphalt Cement (bitumen) AC is a viscoelastic material with suitable mechanical/rheological properties for traditional paving and roofing applications because of their good adhesion 16 properties to aggregates (Akmal and Usmani, 1999). Bitumen is a material characterized by a time of loading and temperature dependence of the mechanical response to loading [Burger et al, 2001]. However, the increase in traffic load requires improving the mechanical properties of conventional asphalt mixtures. The chemical composition of the AC has a significant effect on its viscoelastic properties and hence on its performance as road paving material in asphalts. The behaviour of asphalt at high temperature conditions for short time spans is equivalent to its performance at low temperature conditions for longer time durations. This concept floated by McGennis, et al. (1995) and is called temperature shift or in other words the superposition theory of asphalt binder, which has been explained by Asphalt Institute (2003) as in Figure 2.4. Figure 2.4: Temperature Shift Behaviour of Asphalt Binder (After Asphalt Institute, 2003) In hot climatic conditions or under slow moving trucks, asphalt behaves like a viscous liquid and only aggregates are the contributing element of hot mix asphalt that bear the traffic loads. At micro level, the contiguous layers of molecules seem sliding pass each other. This phenomenon has been presented by Asphalt Institute (2003) as shown in Figure 2.5. Whereas in cold climatic conditions or under fast moving trucks (rapidly applied loads), asphalt behaves like an elastic solid and deforms when loaded, but returns to its original shape when unloaded. If it is stressed beyond its strength, it may rupture. 17 Figure 2.5: Microscopic View of Liquid Flow Properties (After Asphalt Institute, 2003) At intermediate temperature conditions, asphalt binder exhibits the characteristics of both viscous liquids and elastic solids. Due to this property of asphalt, it is considered to be an excellent adhesive material for use in paving. When heated, asphalt acts as a lubricant, allowing the aggregate to be mixed, coated, and tightly-compacted to form a smooth and dense surface. After cooling, it acts as a glue to hold the aggregate together in a solid matrix. In its finished state, the behaviour of the asphalt is termed as visco-elastic i.e., it has both elastic and viscous characteristics, which depends on the temperature and rate of loading. The elastic response or recoverable part, viscoelastic response and plastic response or non-recoverable which appears in the form of permanent deformation have been illustrated in Figure 2.6 (Gibb, 1996). Rutting at high temperatures, crack initiation and propagation in the low temperature region and other forms of pavement defects are not only due to traffic loads but also due to the capability of the asphalt concrete to sustain temperature changes. Increased traffic factors such as heavier loads, higher traffic volume, and higher tire pressure demand higher performance pavements. High performance pavement requires AC that is less susceptible to high temperature rutting or low temperature cracking, and has excellent bonding to stone aggregates (Ait-kadi et al, 1996). 18 Figure 2.6: Stress Strain Behaviour of Bituminous Material (After Gibb, 1996) William, et al., (1967) studied the influence of rheological properties of asphalt on rate of deformation and strength of asphalt concrete and reported a direct relation between them. The asphalt viscosity directly affects the strength of asphalt concrete in compression (rutting) for the practical range of temperatures. The log of pavement resistance and of cohesion varies directly with the log of asphalt viscosity. Modulus of elasticity in compression was influenced by the type of asphalt, temperature and amount of lateral confinement. Repeated loading produced a marked decrease in flexural strength of asphalt concrete. 19 Brown and Snaith (1974) suggested that the increase in deformation is related to the decrease in binder viscosity at high temperatures (400C), thereby leading to a lower interlock between the aggregates. The contribution of the aggregate skeleton towards the behaviour of the mixture becomes more significant at higher temperatures. Pellinen and Witczak (2002) found the aggregate influence to be more dominant than the influence of the binder on the modulus at high temperatures and the binder influence to be more dominant over the aggregate influence at low temperatures. At high temperatures the effects of confining stresses play a significant role in the permanent deformation of the mixture. Awad (1972) and Morris (1973) found that the effect of confining stresses on elastic and permanent deformation strains respectively was more important at high temperatures than at low temperatures. Mahboub and Little (1988) found that mixtures containing less viscous asphalts are less stiff and are more prone to rutting. Monismith et al. (1999) also recommended more viscous asphalt cements in thicker pavements and hotter climates on the basis of similar observations. Monismith and Tayebali (1988) examined the relative behaviour of mixtures with and without modifiers. They found that mixtures containing modified asphalt cement showed better resistance to rutting at high temperatures than the mixture containing the neat asphalt cement. They also reported that resistance to rutting may be improved by the use of modifiers (polymers, micro fillers, etc.) which make asphalt binder more viscous at higher temperatures without any adverse effect at low temperature Anderson, et al. (1995) found that asphalt binder behaves both as viscous liquids and elastic solids at normal pavement temperatures. Heukelom (1999) developed charts to study the effects of temperatures on the mechanical behaviour of asphaltic bitumen. He performed the standard laboratory consistency tests and appraised the data used for entering Van der Poel’s stiffness nomograph. He concluded that ordinary laboratory tests at different temperatures could show relationships and stiffness modulus of bitumen & could be obtained from Van der Poel’s nomograph. 20 Robert (2000) reported that the rutting tendency of a pavement, greatly influenced by the ratio of the complex modulus (G*) to the phase angle ( ). High value of G*and low values of of bitumen are required for more rut resistance and low value of “G*” and “ ” are required for more fatigue resistance. Tarefder, et al. (2003) investigated the most important factors affecting rutting and performance grade (PG) of bitumen and determined that specimen type, test temperature and moisture has significant influence on binder performance. Kanitong et al. (2005) compared the rutting performance of polymer modified bitumen with unmodified and concluded that the overall performance of polymer modified binder was better than those of unmodified. As an alternative to larger top size and coarser mixes, polymer-modified asphalt cement and other modified asphalt cement products are also being investigated by many agencies to increase the resistance to permanent deformation (Ponniah and Kennepohl 1996, Prowell 2001). 2.7 Stress Strain Behaviour of Asphaltic Materials Neat asphalt shows very complex stress-strain behaviour. It behaves as an elastic solid at low temperatures and high frequencies of loading, and as a viscous fluid at high temperatures and low loading frequencies. It has been mentioned that rutting occurs in mixtures with low shear resistance or strength compared to the repeated stress it is subjected to. One can get an insight into the effect of aggregate properties on shear strength of mixtures by considering their effect on cohesion and angle of internal friction as illustrated in Figure 2.7. For a given level of stress, temperature and rate of loading, the shear strength depends on the cohesion and angle of internal friction. The cohesion is affected by the viscosity of asphalt binder and the proportion of fines. The angle of internal friction is obtained from aggregate interlocking. Higher values of angle of internal fraction can be obtained by using rough textured, angular and well graded aggregates. The mechanical interlock of the aggregate particles thus plays a key role in shearing resistance. The binder content is also known to affect angle of internal fraction because it changes the 21 degree of mechanical interlock between the particles, i.e., the higher the proportion of binder in the mix, the further apart the aggregate particles are spread (Cheung, 1995). Aggregate properties and aggregate gradation play a major role in the potential for rutting of an asphalt pavement. Cubical, rough-textured aggregates are more resistant to the shearing action of traffic than rounded, smooth-textured aggregates. Cubical aggregates also tend to interlock better, resulting in a more shear resistant mass of material as shown in Figure 2.8. In addition, increased compaction during construction or the use of higher percentages of coarse aggregate fractions in the aggregate gradation provides more stone-to-stone contact in the asphalt mix which, in turn, helps to reduce pavement rutting. Figure 2.7: Shear Loading Behaviour of Aggregate (After Asphalt Institute, 2003) Figure 2.8: Contrasting Stone Skeletons (After, Asphalt Institute, 2003) 22 As the response of the asphalt mixture is linear visco-elastic under the application of the load, this element of the total strain is irrecoverable and with the repeated load application it accumulates, leading to the formation of surface ruts (Gibb, 1996) as shown in figure 2.9 & 2.10. Figure 2.9: Permanent Strain under Load Pulse (After Gibb, 1996) Figure 2.10: Accumulation of Permanent Strain under Repeated Load (After Gibb, 1996) 23 2.8 Resistance to Permanent Deformation of HMA Resistance to permanent deformation or shearing stress is a stability-related phenomenon. HMA mixtures are designed with adequate stability to ensure adequate performance and are considered to be the core aspect of HMA mixture design with respect to rutting. Stability is affected by type/grade and amount of asphalt binder, aggregate properties (such as absorption, texture, and shape of particle), gradation, compaction level, and temperature. Higher stability is promoted by using hard cubical aggregates with rough surface textures, dense gradations, comparatively low asphalt binder contents, harder (stiffer) asphalts, and well-compacted mixtures as long as the air voids do not fall below a certain level. The asphalt binder used in the mix also affects the rut resistance of an asphalt mix but to a lesser degree than the aggregate characteristics. A mix made with a soft grade of asphalt cement will be less resistant to rutting at high temperatures than a comparable mix that contains a harder (more viscous) asphalt grade. Rutting in an asphalt mix normally occurs in the early years of a pavement’s life when the asphalt binder is relatively low in viscosity. Whereas, rutting is less likely to occur in a pavement after the asphalt binder has aged or oxidized with exposure to the elements to a higher viscosity. Findley, et al. (1976) studied the resistance to deformation of HMA under creep loading and concluded that under constant static or repeated loading, a visco-elastic material undergoes flow or ‘creep’, which includes recoverable and irrecoverable, time dependent and time independent components of deformation. They also concluded that instantaneous elasticity, creep under constant stress, instantaneous recovery, delayed recovery, and permanent strain can be used to characterize viscoelastic materials including HMA. Cooper, et al. (1985) concluded that low voids in mineral aggregates (VMA) results in good resistance to permanent deformation. They also reported that rutting resistance of asphalt mixture can also be improved by reducing air voids up to a certain limit & higher compaction energy results in a low air void content in the field. 24 Huber, et al. (1987) examined 9 test sites in Saskatchewan to evaluate the mix design characteristics and performance and concluded that higher asphalt contents and voids filled with asphalt were the basic parameters that had increased the rutting potential of mix. Brown (1990) reported that HMA pavements constructed at approximately 7.0 to 8.0 percent air voids are further compacted to approximately 4.0 percent air voids under traffic loads, if the mix is properly designed. Molenaar (1993) reported that plasticity models such as the Mohr-Coulomb yield criterion describe the response of a material in relation to an ultimate surface beyond which no stresses are permitted to occur. If the state of stress is inside the ultimate surface, the deformations are purely elastic and plastic deformation occurs at states of stress on the ultimate surface. The Mohr-Coulomb failure criterion expressed in Equation 2.1 provides an elegant means of demonstrating the effect of both aggregates and the mortar on the permanent deformation behaviour (FHWA). Figure 2.11 illustrates the effect of the aggregates and the mortar on the failure behaviour of asphalt mixtures. For simplicity purposes, the cohesive strength is considered to be wholly influenced by the mortar while the angle of internal friction is influenced entirely by the aggregates. If the binder has a high cohesion at high temperatures, then it will provide better resistance to permanent deformation than a binder with a lower cohesion. Similarly aggregates with a higher internal angle of friction can be expected to provide a better resistance to permanent deformation than similar aggregates with a lower angle of internal friction. However, it must be emphasized that this is a very simple idealization of the asphalt mixture. τ = c + σ tan φ Where: t = shear strength, c = cohesion, σ= normal stress, Φ = angle of internal friction. 25 (2.1) Figure 2.11: Effect of Aggregate and Mortar on the Failure Behaviour of Asphalt Mixtures Cominsky, et al (1994) studied the resistance of HMA to rutting and reported that both the mineral aggregate and asphalt cement contributes towards the shear strength of HMA. Collop, et al. (1995) treated permanent deformation or rutting in bituminous pavements as a linear visco-elastic flow phenomenon and found that permanent deformation per wheel pass is directly proportional to the static axle load and inversely proportional to the vehicle speed. The phenomenon is quite useful for investigating the road damaging potential of heavy vehicles and evaluating important trends. Ford, et al. (1988) investigated that pavements with air voids lower than 3.0 percent tend to rut while those with higher air voids do not, as long as the aggregate quality is satisfactory. They also concluded that pavements with air voids lower than 3.0 percent have a tendency to rut severely. Kamal, et al (2005) studied the insitu behaviour of asphalt concrete with and without PMA under same temperature and loading conditions and compared resilient modulus and creep stiffness of both type of mixes, using the indirect tensile strength test (ASTM D4123) and repeated load uniaxial stain test. They reported a drastic reduction of about 85% in resilient modulus was observed for an increase in temperature from 250C to 400C. 26 Ziari et al (2007) studied the effects of temperature and different percentage of bitumen on the resistance to permanent deformation of HMA mixture and concluded that significant degree of confidence that the mix will not fail on the roadway due to permanent deformation can be achieved by simulating the laboratory test findings with field performance of mixes. The distribution of stresses in the pavement depends on the stiffness, Poisson’s ratio and thickness of the pavement layers. A typical asphalt pavement structure consists of a top layer, base, subbase and subgrade. Similar to granular materials, the resistance to permanent deformation of asphalt mixtures is to a large extent dependent on the magnitude of the confining stresses that result from the wheel load. Generally, an increase in confinement enhances the resistance to permanent deformation. By smartly stacking the pavement layers, significant confinement levels can be generated leading to enhanced resistance to permanent deformation. Such a stacking mostly involves a layer with a high modulus below the asphalt top layers (Huang 1993). 2.9 Permanent Deformation Prediction Models The permanent deformation response of asphalt concrete mixes to loading, temperature and other influential parameters may have to be characterized by a number of models in the previous research work. Accordingly, recommendation and limitations have been reported. A few models have been briefed in this section. Garba (2002) reported the effects of material properties on permanent deformation of asphalt mixtures, mechanisms of the permanent deformation, and methods of its prediction. The main objectives of research work included the review and evaluation of available models for permanent deformation of asphalt concrete mixtures, investigation of the effect of volumetric composition, loading and temperature conditions on the permanent deformation of asphalt concrete, and the identification and definition of simple measures of resistance to permanent deformation. In order to meet that objectives, repeated load triaxial creep and recovery tests were conducted at 250C and 500C under varying stress conditions. Binder content, void content, confining stress and deviator stress were found to influence the permanent 27 deformation characteristics significantly. It was reported that most of these parameters are not sensitive to changes in volumetric composition and therefore are not suitable for comparison of mixtures made from the same materials but with varying proportion of the components. The bounding surface plasticity approach was found to be a convenient method to model the accumulation of permanent deformation. It was demonstrated that deformations calculated using cyclic hardening model based on bounding surface plasticity fits the measured deformation quite well. The elasto-viscoplastic model, which is based on strain decomposition approach, provides a suitable method for analysis of creep and recovery test results. Deformations of asphalt mixtures, calculated using this model also fit the measured deformation quite well. The NCHRP 1-26 study (Barenberg and Thompson, 1990) recommended the use of the permanent strain accumulation model developed at Ohio State University (Majidzadeh et al 1981). This strain model predicts total rutting, considers the rutting rate of the pavement as indicated by the following equation: Єp /N=A(N)m Where (2.2) Єp = permanent strain N= number of load application A= Experimental constant (depends on material type and stress state) m= Experimental constant (depends on material type) Equation 2.2 is valid for describing the progression of rutting in pavement layers, asphalt surface and base courses, granular base and subbase courses, and subgrade soils (Majidzadeh, et al 1981). Several material permanent strain accumulation models have been developed so far to predict the permanent deformation in AC pavement layers. Pavement system rutting models were also evaluated in the NCHRP 1-26 (Barenberg and Thompson 1990). The study have revealed that those models which are related to log of permanent strain to the log of load repetition appear to be the most appropriate and versatile for practical use. 28 This power model is often fitted to the accumulated permanent deformation curve. It is probably the most commonly used permanent deformation equation. The power model plots as straight line on log-log scale. It has also been thought that the slope and intercept of this model when plotted on log-log scale may be used as indicators of rutting resistance (Garba, 2004). The basic permanent strain to load repetition model expressed as Єp =aNb (2.3) It was initially proposed for subgrade and unbound materials by Monismith (1976); and initially used for asphalt concrete mixes by (Khedr, 1986). Later on, various researchers used the same model for asphaltic concrete (Diylajee and Raymond, 1982), (Vuong and Amstrong 1991), (Behzadi and Yandell, 1996). Where a & b are intercept and slope coefficients and N is the load repetition. The curve of power model on log-log scale between load repetition and permanent strain can be expressed in graphical form in Figure 2.12. Figure 2.12: Log-Log Form of Power Model (After Khedr, 1986) The log relationship of Єp =aNb, can be expressed in the following form; 29 Log Єp = b Log N + a, (2.4) They further states that this laboratory relationship between permanent strain and load application can be plotted and extrapolated for general analysis purpose. A typical value for the exponent “b” varies between 0.1 and 0.2; while “a” is highly dependent upon the magnitude of the repeated stress state. It was further noted that limiting the stress ratio (repeated deviator stress/maximum deviator stress) is a good approach to minimize the permanent deformation, and that the stress ratios (repeated stress/maximum strength) less than approximately 60 to 70 percent will generally help to limit the accumulation of permanent deformation. Permanent deformation is also somewhat dependent on the previous stress history to which the material has been subjected. The initial application of large stress repetitions is much more damaging than the initial application of low stresses (Hassan, 1998). Using equation 2.3, intercept coefficients and slope coefficients have been determined for mixes in chapter Six & Seven. According to VESYS, 1991 progressing of rutting with load repetition can be measured using layer elastic theory, where in all layers can be modeled using a constitutive model in the form given in equation 2.5 (partial differentiation form). ∂ε p ∂N = ε pn = ( ∂ aN b ∂N ) (2.5) ε pn = abN ( b−1) The resilient strain (εr) is assumed to be independent of load repetition. The ratio of plastic to resilient strain can thus be defined as: ε pn ⎛ ab ⎞ b−1 = ⎜⎜ ⎟⎟ N = µN −α εr ⎝ εr ⎠ (2.6) From Equation 2.6, the rate of plastic strain (1-b) can be defined with a permanent deformation coefficient alpha (α) & plastic to elastic strain ratio can be defined with a coefficient mu (µ) as shown; µ= εr ab α = 1− b (2.7) 30 Where Єp = Permanent Strain (rut value) N = Number of Load Application a = Intercept Coefficient; and b = Slope Coefficient µ = mu (ratio of plastic to elastic response) α = alpha (rate of change of the plastic response) Where µ is the permanent deformation parameter representing the constant of proportionality between permanent strain and resilient strain (i.e. plastic strain at N=1) and α is a permanent deformation parameter indicating the rate of decrease in incremental permanent deformation as the number of load applications increases. Thus two set of parameters a &b and µ& α are closely related. Rita and Matthew (1991) determined alpha and mu using log10Єp-log10N as base relationship and proposed that these parameters could be used to assess plastic strain to elastic strain ratio. Alpha and mu according to Sullivan (2002) are the stress and temperature dependent non linear parameters and can be used for modelling permanent deformation of mixes. The current Mechanistic Empirical Pavement Design Guide (MEPDG) incorporates a power model for generating rutting predictions for asphalt concrete (Stephen et. al. 2007). Rutting model developed from laboratory uniaxial repeated load strain tests as provided in MEPDG in the following form has been used as basis to estimate the relationships between the predictor variables and the permanent deformation parameters: εp = a1T a N a εr 2 3 (2.8) Where, εp, εr, are the plastic and elastic strains respectively, at N repetitions of load and ai are the non linear regression coefficient. This model was proposed using the uniaxial repeated 31 haversine pulse loading and measuring the strains with on-sample Linear Variable Displacement Transformers (LVDT). An attempt is made in this study to model the permanent deformation behaviour of asphalt concrete mixes using equation 2.8 as base guide line. In laboratory testing, off-sample LVDT’s were used to measure the strain response under application of square shape pulse loading. Modeling the permanent deformation behaviour of asphalt concrete mixes is thus the main part of this study. 32 Chapter Three 33 Chapter Three Material Characterization 3.1 Introduction Hot Mix Asphalt generally consists of combination of different size of aggregates with mineral fillers, uniformly mixed and coated with asphalt cement, each having its own particular characteristics, which will be more suitable to specific design and construction purposes. Before designing asphalt paving mixes, selection, proportioning and characterization of individual material are imperative to obtain the desired quality and properties of finished mix. For the current study, aggregates were obtained from the local source i.e., Margalla Crush Quarry and asphalt cement from Attock Refinery, Rawalpindi, & National Refinery, Karachi, Pakistan. 3.2 Aggregate for Asphalt Mixtures The largest portion of the resistance to permanent deformation of the mixture is provided by the aggregate structure. Aggregate is expected to provide a strong stone skeleton to resist repeated load applications. Gradation, shape, and surface texture have a great influence on HMA properties. Angular, rough-textured aggregates provide more shear strength than rounded, smooth-textured aggregates. When a load is applied to the aggregate in an asphalt mixture, the angular, cubical, rough-textured aggregate particles lock tightly together and function as a large, single elastic mass, thus increasing the shear strength of the asphalt mixture. To set up a laboratory study and investigate effect of coarse aggregate properties on the HMA aggregate structure and the stability of asphalt concrete (AC) pavements, aggregate consensus and source properties have to be properly taken into account together with the aggregate sizes or gradations. 34 The suitability of aggregates from Margalla Crush Quarry for use in asphalt construction was determined by evaluating the material in terms of the followings; ̇ ̇ ̇ ̇ ̇ ̇ 3.2.1 Size and grading Particle shape Absorption Clay Contents (Sand Equivalent) Toughness Soundness Size and Grading Maximum size and aggregate grading are directly controlled by NHA aggregate grading “Class-A” for wearing coarse (NHA General Specification, 1998). Maximum size of aggregate is related with the typical lift thickness used on National Highways in Pakistan. NHA in its general specifications has specified two aggregate gradations, namely class ‘A’ and ‘B’, the coarser and finer respectively. Two gradations i.e. “01” and “02”, within the envelope of NHA class ‘A’ gradation (commonly used in the field) have arbitrarily been selected by the author for this study as reported in Table 3.1. Table 3.1: Adopted Aggregate Gradations & Specifications Sieve Size Combined grading (Asphaltic Wearing Course Class-A) Gradation “1” Inch mm Gradation Gradation “2” NHA Targeted Targeted Specifications Gradation ‘Class-A’ value value Asphalt Institute Gradation (1994) 1 25.00 100 100 100 100 100 100 3/4 19.00 90-100 90 95-100 100 90-100 90-100 1/2 12.50 - - - 3/8 9.50 56-69 56 59.1-69.1 69.1 56-70 56-80 #4 4.75 38-46 38 38.2-48.2 48.2 35-50 35-65 #8 2.36 25-33 25 24.3-30.3 30.3 23-35 23-49 #50 0.300 5-12 5 4.5-10.5 10.5 5-12 5-19 #200 0.075 3.4-5.3 3.4 3.3-5.3 5.3 2-8 2-8 - 35 The targeted values of aggregate gradation “01” were on the coarser side of the limits, while that of aggregate gradation “02” were on the finer side of the gradation limits. Both the gradations have been shown graphically in Figure 3.1. Aggregate Gradation Limit 100 90 Gradation "02" 80 Percentage Passing 70 60 Gradation"01" 50 40 30 20 10 0 0.01 0.1 1 10 Seive Size (mm) Gradation "02" 100 Gradation "01" Figure 3.1: Aggregate Gradation ‘01’ and ‘02’ under NHA Class ‘A’ 3.2.2 Particle Shape Particle shape changes the workability of the paving mix as well as the compactive efforts, necessary to obtain the required density. Particle shape has an effect on the strength of the asphalt mix. Irregular or angular particles, such as crushed stone and gravels, and some natural gravel and sands, tends to interlock when compacted and resist displacement. Best interlock is generally obtained with sharp-cornered cubical shaped particles. Round particles, such as most natural gravels and sands from stream beds, can be used successfully in asphalt paving mixes. 36 Hot Mix Asphalt can generally be made more stable and resistant to rutting by requiring that a significant portion i.e. 50% of the coarse aggregate be angular. (Asphalt Institute, MS-4, 2003). Coarse aggregate angularity can be defined as “The percentage by mass of the aggregates larger than 4.75 mm with one or more fractured faces” (Asphalt Institute SP-2, 2003). High shear strength for rutting resistance and a high degree of internal friction can be achieved by specifying this property (ASTM D5821). The type of aggregate surface texture depends on hardness, grain size, and pore characteristics of the parent rock as well as the extent to which forces acting on the particle surface have smoothened or roughened the surface (Su 1996). The type of aggregate surface texture is categorized by shape features such as angular, rounded, smooth or rough. The aggregate surface texture affects the aggregate interlock. Durable angular aggregates with a rough surface texture are normally considered to offer good aggregate interlock. Fine aggregate angularity can be defined as “The percent air voids present in loosely compacted aggregates smaller than 2.36 mm” (Asphalt Institute SP-2, 2003). Fractured faces are indicated by the void content measured as un-compacted void contents of fine aggregates (AASHTO T304). Greater the void contents more will be the fractured faces. High degree of internal friction and high shear strength for rutting resistance can be achieved by specifying this property. Particle shape, surface texture and grading influence fine aggregate angularity. Aggregates shapes were measured using flakiness and elongated index gauges. Flat and elongated particles can be defined as “The percentage by mass of coarse aggregates that have a maximum to minimum dimension ratio greater than five” (Asphalt Institute SP-2, 2003). Flakiness of aggregate particles are classified as flaky when they have thickness (small dimension) of less than 0.6 of their mean sieve size, this size being taken as the mean of the limiting sieve apertures used for determining the size fraction in which the particle occurs. While aggregate particles are classified as elongated when they have a length (greatest dimension) of more than 1.8 of their mean sieve size, this size being taken as the mean of the limiting sieve apertures used for determining the size fraction in which the particles occurs 37 (ASTM D4791). Results of flakiness index and elongation index have been reported in Table 3.2. 3.2.3 Absorption The porosity of an aggregate is generally indicated by the amount of water it absorbs when soaked in water. A porous aggregate will also absorb asphalt which will tend to make asphalt mixtures dry or less cohesive. An extra amount of asphalt must be incorporated into the paving mix to compensate for the absorption of asphalt by the aggregate. Moreover, aggregate that are very porous tend to require a significant amount of extra asphalt to make up for the high absorption rate. Highly porous aggregates normally are not used unless they possess certain other qualities or properties that make them desirable in spite of the high absorption rate (Asphalt Institute MS-4, 2003). If absorptive aggregates that have high water content are used, will be required in the production of HMA to ensure that the moisture in the pores can evaporate. Otherwise, the asphalt may not be properly absorbed, leading to compaction difficulties (Transportation Research Board, 2000). Results of absorption of aggregates have been reported in Table 3.2. 3.2.4 Clay Content (Sand Equivalent) Asphalt Institute defines the clay content as “clay content is the percentage of the clay material contained in the aggregate fraction that is finer than a 4.75 mm sieve” (Asphalt Institute SP-2, 2003). Aggregate cleanliness may often be determined by visual inspection, but a washed-sieve analysis generally provides positive proof. The sand equivalent test, developed by the California Division of Highway and described in ASTM D2419 (AASHTO T176) is a method of determining the relative proportion of detrimental fine dust or clay-like materials in the portion of the aggregate passing the 4.75mm sieve(Asphalt Institute MS-4, 2003). equivalent of fine aggregate have been reported in Table 3.2. 38 Sand 3.2.5 Toughness Aggregates are subjected to additional crushing and abrasive wear during manufacture, placing and compaction of asphalt paving mixes. Aggregates are also subjected to abrasion under traffic loads. They must exhibit, to a certain degree, ability to resist crushing, degradation, and disintegration; Asphalt Institute’s defines the toughness as “toughness is the percent loss of material from an aggregate blend during the Los Angeles Abrasion Test (ASTM C131 or AASHTO T96)”. The toughness property test estimates the coarse aggregate resistance to abrasion and mechanical degradation that occurs during handling, construction and service. To perform the test, coarse aggregates larger than 2.36 mm are subjected to impact and grinding by steel spheres. Due to this mechanical degradation, the mass percentage of the coarse material lost gives the toughness. Typically the maximum loss values range from 35 – 45 percent. Aggregates at or near the pavement surface require greater toughness than aggregate in the lower layers where loads have dissipated or are not as concentrated. Relatively high resistance to wear, as indicted by a low percent of abrasion loss, is desirable characteristics of aggregates to be used in asphalt pavement surface layers. Aggregates having higher abrasion losses, within limits, may generally be used in lower pavement layers where they will not be subjected to the high stresses caused by traffic. Los Angles Abrasion (LAA) values determined in the laboratory for Margalla Quarry aggregates have been reported in Table 3.2. 3.2.6 Soundness Aggregates must be resistant to breakdown and disintegration from weathering (wetting/drying and freezing/thawing) or they may break apart and cause premature pavement distress. Durability and soundness are terms, typically given to an aggregate’s weathering resistance characteristic. Aggregates used in HMA are dried in the production process and therefore should contain almost no water. Thus, for aggregate used in HMA, freezing/thawing should not be a significant problem. According to Asphalt Institute’s Superpave Series No. 2, soundness is defined as “The percent loss of material from an aggregate blend during the 39 sodium or magnesium sulfate soundness test (AASHTO T104)”. The resistance of aggregate to in-service deterioration is determined by using this test. The test is applicable to both coarse and fine aggregates. Results of soundness of aggregates have been reported in Table 3.2. 3.2.7 Deleterious Materials Asphalt Institute’s Superpave Series No. 2 defines the deleterious materials as “The mass percentage of contaminants such as clay lumps, shale, wood, mica, and coal in the blended aggregates (AASHTO T112)”. The test is applicable to both coarse and fine aggregates. Wet sieving aggregate size fractions over the specified sieves were performed and the percentage by mass of the material lost as a result of it yielded the percent of the clay lumps and the friable particles. Maximum allowable percentage of these materials ranges from as low as 0.2 percent to as high as 10 percent and depends upon the exact composition of the contaminant (Asphalt Institute, SP-2, 2003). Results of deleterious materials have been reported in Table 3.2. Table 3.2: Physical Properties of Aggregates Sr. Test Description No. Specification Results Recommended Values (NHA General Specification) Reference 1 Flakiness Index (FI) BS 812, Part 1 4.75% 15(Max) 2 Elongation Index (EI) BS 812, Part 1 15(Max) 3 Aggregate Absorption AASHTO T 166 4.4 % 0.88% 4 Sand Equivalent AASHTO T 176 72 45(Min) 23% 30 (Max) 3.32% 12 (Max) 1 0.2-10 5 6 7 Loss Angles Abrasion value (LAA) AASHTO T96 ASTM C 131 Sodium Sulphate AASHTO T104, Soundness value ASTM C88 Deleterious materials Asphalt Institute, SP-2, 40 3.3 Characteristics of Asphalt Cement For engineering and construction purposes, three properties or characteristics of asphalt are important i.e. consistency, purity and safety (Asphalt Institute MS-4, 2003). Asphalts are thermoplastic materials, since they gradually liquefy when heated. They are characterized by their consistency or ability to flow at different temperatures. Consistency is the term used to describe the viscosity or degree of fluidity of asphalt at any particular temperature. The consistency of asphalt cement varies with temperature; therefore, it is necessary to use a standard temperature when comparing the consistency of one asphalt cement with another. Asphalt cements are graded based on ranges of consistency at a standard temperature. Consistency of paving asphalt is commonly specified and measured by a viscosity test or a penetration test (Asphalt Institute MS-4, 2003). Other tests i.e. ductility, softening point provides additional information and confidence towards consistency. 3.3.1 Consistency of Asphalt Cement The penetration test determines the penetration of semi-solid and solid asphaltic materials. The needles, containers and other conditions described in this test method provide for the determinations of penetrations up to 500. Higher values of penetration indicate softer consistency. Where the conditions are not specifically mentioned, the temperature, load, and time are understood to be 250C, 100 g, and 5s, respectively (AASHTO T 49-03 or ASTM D5). Results of penetration tests for three AC types have been reported in Table 3.3. Softening point of asphalt cement was performed by using ring and ball apparatus confirming to AASHTO T53. This test method covers the determination of the softening point of bitumen in the range from 30 to 1570C (86 to 3150F) using the ring-and-ball apparatus immersed in distilled water (30 to 800C), glycerin (above 80 to 1570C), or ethylene glycol (30 to 1100C). Bitumen’s are visco-elastic material without sharply defined melting points; they gradually become softer and less viscous as the temperature rises. Due to this property, softening points must be determined by an arbitrary and closely defined method for the reproducibility of the test. Results of softening point test for three AC types have been reported in Table 3.3. 41 The ductility of a asphalt cement can be defined as the “distance to which it will elongate before breaking when two ends of a briquette specimen of the material, are pulled apart at a specified speed (5cm/min ±5.0%) and at a specified temperature (25±0.50C) (ASTM 113-99 or AASHTO T 51-00). This test method provides measure of tensile properties of bituminous materials and may be used to measure ductility for specification requirements. Ductility is an indicator of how flexible behaviour of bitumen under various temperatures. Results of ductility test for three AC types have been reported in Table 3.3. Specific gravity of two neat bitumen with penetration grade “60/70” & “modified bitumen”, obtained from Attock Refinery Limited, Rawalpindi, and penetration grade “40/50” from National Refinery Limited, Karachi, Pakistan were measured as per AASHTO T 228-2, and reported in Table 3.3. Table 3.4 shows the specifications of these grades. Table 3.3: Consistency of Different Types of AC Sr. No. Description Type 1 2 Ring & Ball Softening Point Penetration 0 3 Ductility @ 25 C 4 Specific gravity PMA (60/70 base) Modified penetration grade “60/70” neat penetration grade “40/50” neat 58 49 56 46 65 44 45 100 67 1.023 1.03 1.032 Table 3.4: Specifications of AC Test Description Test Methods cm ASTM T113 100 C (0F) AASHTO T92 232 (450) 232 (450) Min 1/10 mm ASTM D 5 60-70 40-50 - AASHTO T 53 50 60 Min Ductility @ 250C Flash Point Penetration @ 250C Softening Point AASHTO M-20 Units 0 0 C 42 P.G 60/70 P.G 40/50 Max/Min Min Polymer modified bitumen used for this research work was obtained from Attock Refinery. The modification of bitumen was carried out by Mathy Technology, USA (Gerald, 2001). It was based on the climatic as well as the traffic conditions of one of the flexible pavement trial section of NHA. The base bitumen i.e., 60-70 penetration AC was modified with 1.6% Elvaloy® 4170 in the presence of superphsophoric acid. The addition of the 1.6% Elvaloy® 4170 caused the tensile strength of the modified binder to increase significantly relative to the unmodified binder (Gerald, 2001). The only reason for selecting the PMA is to compare the resistance to rut of modified bitumen with neat asphalt. The properties of PMA determined in the laboratory have been discussed in the proceeding sections. 3.3.2 Elastic Recovery It is defined as “the recovered strain, as a percentage of the original strain attained at the break point (end of loading phase)” (AASHTO T 301). Elastic recovery test was performed on AC by means of a ductilometer with briquette specimens as shown in Figure 3.2. The specimens were pulled apart at 5cm/min and held after a specified elongation as shown in Figure 3.3. This procedure was used to evaluate the elastic recovery of PMA by the percent of recoverable strain measured after the elongation of a sample. Figure 3.2 : Briquette Specimens Figure 3.3: Trimming of Specimens (After Asphalt Institute, 2003) (After Asphalt Institute, 2003) 43 Elastic recovery was calculated using the following equation and has been reported in Table 3.5. Elastic Recovery (%) = 100 × a b (3.1) Where “a” is the recovered displacement (or strain), in mm & “b” is original displacement (or strain), in mm. Table 3.5 Elastic Recovery Test Results at 25 oC Types PMA 60/70 40/50 Loading Rate 5 cm/min 5 cm/min 5 cm/min Elongation Recovery (%) 3.3.3 71% 77.5% 74% 66.5% 70% 68% 72.5% 75% 71% Torsional Recovery Austroads defines torsional recovery as the measure of the extent of recovery of the originally applied rotation which is usually 180 degrees, in percentage. A flat-bottomed, cylindrical, seamless tin 55mm in diameter and 35 mm in depth, Disc made of aluminum and spindle assembly, bolt, and pointer are made of steel. The recommended device provides a scale, of 80 mm radius and graduated in degrees around at least half of its circumference (Austroads T-122). Figure 3.4 shows the assembly of torsional recovery apparatus. Graduation scale up to 1800 Pointer Spindle Figure 3.4 : Torsional Recovery Test Apparatus 44 Torsional recovery in %age was calculated for each type of asphalt cement at 250C, using the following equation; Torsional Recovery (%) = 100 × A 180 (3.2) Where, A is the recovered angle, in degrees. The results have been reported in Table 3.6. Table 3.6 Torsional Recovery Test Results. Type of AC 40-50 Pen. grade PMA 60-70 Pen. grade Recovery Time 30 sec 30 sec 30 sec Recovery (degrees) 90 Recovery (%) 5% 3.4 140 120 240 300 220 200 180 170 7.7% 6.6% 13.3% 16.6% 12.2% 11.11% 10% 9.44% Dynamic Shear Rheometer (DSR) This test method covers the determination of the dynamic shear modulus and phase angle of AC binders when tested in dynamic (oscillatory) shear using parallel plate geometry. It is applicable to AC binders having dynamic shear modulus values in the range from 100 Pa to 10 mPa (AASHTO T 315-05). This test method is intended for determining the linear viscoelastic properties of AC binders as required for specification testing and is not intended as a comprehensive procedure for the full characterization of the viscoelastic properties of AC binders. A dynamic shear rheometer test system consists of parallel metal plates, an environmental chamber, a loading device, and a control and data acquisition system and shown in Figure 3.5. 45 Figure 3.5: Dynamic Shear Rheometer Metal test plates of 8.00 ±5.00 mm in diameter smooth polished surface. A chamber for controlling the test specimen temperature, by heating (in steps or ramps), or cooling (in steps or ramps), to maintain a constant specimen environment. A temperature controlling device capable of maintaining the specimen temperature from 5 to 120 0C ± 0.1 0C test temperature. The loading device which applies a sinusoidal oscillatory load to the specimen is shown in Figure 3.6, to the frequency of 10.0 + 0.1 rad/sec. Specimen Figure 3.6: Test Specimen on DSR 46 Rheology is the study of deformation and flow of bitumen [Thomase, 2002] that explains the elastic and viscous behaviour of bitumen, when subjected to a stress [Barnes et al 1989]. Complex modulus (G*) and phase angle ( ) are considered to be the principal rheoligcal parameters, normally measured from a device known as Dynamic Shear Rheometer (DSR) [Huang, and Zeng, 2007]. The complex modulus (G*) is the peak-to-peak shear stress to absolute value of peak-to-peak shear strain and the phase angle ( ) is the angle in radian between a simultaneously applied stress and the resulted sinusoidal stress in a controlled strain testing mode (Asphalt Institute, SP-2, 2003). Following rheological properties can be determined using DSR; • • • Strain amplitude, nearest 0.01 percent. Complex modulus (G*) for the 10 measurements, kPa to three significant figures. Phase angle ( ) for the second 10 cycles, nearest 0.1 degrees. Table 3.7 shows the results of the tests being performed on DSR. The relationships between different rheological properties are shown graphically in Figure 3.7. Table 3.7: Rheological Properties of Binders Sample Type. 40-50 60-70 PMA Frequency Temperature Phase Angle (ºC) ( ) (rad/s) (Hz) 10 1.59 25 10 1.59 40 10 1.59 55 10 1.59 25 10 1.59 40 10 1.59 55 10 1.59 25 10 1.59 40 10 1.59 55 47 Mean Complex Modulus (G*) G*/sin( ) (kPa) 72.5 462 484 80.8 37 37.8 86.6 3 3.22 74.7 995 1031.56 83.8 75 75.4 87.7 6 5.56 66.9 875 951 69.8 60 63.72 70.7 7 7.97 (Phase Angle & Complex Shear Modulus) G* (kPa) 2000 40/50 60/70 100 δ [ ο ] 1800 90 1600 80 G* for 60/70 1400 Phase Angle PMA 70 60 1200 1000 800 G* for 40/50 G* G* for PMA PMA 50 40/50 40 Phase Angle for 40/50 30 600 60/70 400 20 200 10 Phase Angle for 60/70 Phase Angle for PMA 0 0 25 40 55 0 Temperature ( C) Figure 3.7: Influence of Temperature on Linear Visco-Elastic Properties Rheological characteristics like phase angle, complex modulus and G*/sin( ) of three asphalt cement types i.e. 60/70 & 40/50 penetration grade AC and PMA are studied in dynamic shear rheometer at three temperatures i.e. 25, 40 and 550C. Study revealed that temperature has significant effect on binder rheology, specifically at intermediate to high temperatures. Complex modulus and G*/sin ( ) reduces significantly at higher temperature (550C). Phase angle in case of 60/70 and 40/50 penetration grade bitumen was observed to be more effected than PMA. 48 Chapter Four 49 Chapter Four Mix Design Method 4.1 Introduction A mix design method is used for determining the gradation of course and fine aggregates to be combined to achieve a predetermined percentage of air void volume and voids in mineral aggregate for a given quantity of asphalt cement. Mixes are designed for heavy-duty asphalt pavements, keeping in view the specifications and requirements of National Highway Authority, Pakistan in terms of stability and durability. The aggregate gradations meet NHA gradation requirements. The materials used in mix design practice conformed to specifications and testing procedures as per Asphalt Institute (MS 2), American Association of State Highway and Transportation Officials (AASHTO) and American Society for Testing and Material (ASTM). This chapter presents HMA design using Marshall Method of Mix Design for the preparation of six mixes. 4.2 Combined Grading of Aggregate Local aggregates sources and asphalt cement as described in section 3.2, 3.3, & 3.4 were chosen for mix design preparation. Three aggregate fractions at source were sieved and mean passing size have been reported in Table 4.1. Two aggregate gradations with one-inch (1”) maximum size as shown in Table 4.2 were developed from mean sieve sizes. Coarse aggregate, which is the material, retained on an AASHTO NO.4 Sieve, consist of 100% crushed rock, having at least two faces mechanically fractured. The filler consisted of final divided mineral matter such as rock dust. 50 Table 4.1: Summary of Sieve Analysis Results Sieve size Percentage Passing (%) (Inches) (mm) 0---5mm Size 5---13mm size 13---25mm size 1” 25 100 100 100 3/4 19 100 100 74.10 1/2 12.50 - - - 3/8 9.50 100 99.20 1.40 #4 4.75 99.50 25.20 0.60 #8 2.36 75.70 2.20 0.30 #50 0.300 23.20 1.60 0.20 #200 0.075 10.60 1.10 0.10 Table 4.2: Adopted Gradations & Specifications Sieve Size Combined grading (Asphaltic Wearing Course Class-A) Gradation “1” Inch mm Gradation Gradation “2” NHA Targeted Targeted Specifications Gradation ‘Class-A’ value value 1 25.00 100 100 100 100 3/4 19.00 90-100 90 95-100 100 1/2 12.50 - 3/8 9.50 56-69 56 59.1-69.1 69.1 #4 4.75 38-46 38 38.2-48.2 48.2 #8 2.36 25-33 25 24.3-30.3 30.3 #50 0.300 5-12 5 4.5-10.5 10.5 #200 0.075 3.4-5.3 3.4 3.3-5.3 5.3 - 51 Asphalt Institute Gradation (1994) 100 100 90-100 90-100 - - 56-70 56-80 35-50 35-65 23-35 23-49 5-12 5-19 2-8 2-8 4.3 Asphalt Mixtures The Job Mix Formula (JMF) identifies the gradation of the combined aggregate materials and the selected asphalt cement (Asphalt Institute, 2003). Six mixes were designed using both the aggregate gradations and three AC types. Mixes designated as 1a, 1b, 1c and 2a, 2b, 2c with gradation “1” and gradation “2” respectively and reported in Table 4.3. Table 4.3: Selection of Mix Types Gradation “1” 1a PMA (60/70+Elvaloy) 4.4 1b 60/70 Gradation “2” 1c 40/50 2a PMA (60/70+Elvaloy) 2b 60/70 2c 40/50 Mix Design Properties An Asphalt Institute Marshall Method of Mix Design has adopted for the preparation of mixes (Asphalt Institute, 2003). The Marshall method as presented here is applicable only to hot-mix asphalt cements and containing aggregates with maximum size of 25mm (1 in.) or less. The method is intended for laboratory design of asphalt hot-mix paving. The Marshall method uses standard cylindrical test specimens of 64mm (2 ½ in.) height x 102 mm (4 in.) diameter. These were prepared using a specified procedure for heating, mixing, and compacting the asphalt aggregate mixtures. The two principal features of the Marshall method of mix design are stability-flow test and density-voids analysis of the compacted test specimens. Following criteria was adopted for the design of mixes (NHA General Specification, 1998); Compaction, number of blows 75 Stability (kg) (minimum) 1000 Flow, 0.25mm 8-14 Percent air voids in mix (Va) 4-7 Loss in Stability (%), 20 (Max) 52 The stability of the test specimens is the maximum load resistance in Newton’s (lb.) that the standard test specimen will develop at 600C (1400F). The flow value is the total movement or strain, in units of 0.25-millimeter (1/100 in.) occurring in, the specimen between no load and maximum load during the stability test (Asphalt Institute, 2003). Stability by immersion of specimen in water at sixty (600C) for twenty four (24) hours as compared with the stability measured after immersion in water at sixty degree (600C) for twenty (20) minutes is the loss of stability value. Stiffness index is an empirical relationship which is the ratio of stability to flow of mixes at 600C (NHA General Specification, 1998). Optimum asphalt contents, stability, loss of stability, flow and stiffness index of mixes has been reported in Table 4.4. Table 4.4: Hot Mix Asphalt Design Properties Mix Types 1a (PMA) 1b (60/70) 1c (40/50) 2a (PMA) 2b (60/70) 2c (40/50) Optimum AC Contents (%) 3.83 Stability (Kg) Loss of Stability (%) Flow (0.25mm) Stiffness index (Stability/flow) 1378 11.00 10.80 128 3.87 1305 14.30 11.00 119 3.95 1356 12.70 10.50 129 4.29 1335 8.90 9.80 136 4.31 1298 11.00 11.30 115 4.33 1314 10.50 10.50 125 Stiffness index of mix ‘1c’ and ‘1a’has the same value. Mix ‘2a’ showed higher value; while mixes ‘1b’ & ‘2b’ showed the lowest. 53 4.5 Mineral Filler Optimization Mineral fillers are the part of mineral aggregates, they fill interstices and provide contact points between larger aggregates particles and thereby strengthen the mixture. Utmost efforts were made to ensure less organic material passing sieve no. 200 (75µm) having plasticity index (PI) less than 4. Three percentages of filler i.e. 2.4%, 3.4%. & 4.4% in Marshall Method of mix design were used in order to determine the optimum filler content for asphalt mixes. Mineral filler to AC ratio determined through trials have been reported in Table 4.5, which shows that the filler to asphalt content ratio ranges from 0.9 to 1.0 in asphalt mixes. Table 4.5: Mineral Filler to Asphalt Cement Ratio Sr. No. Description Mixes with aggregate Mixes with aggregate gradation “01” gradation “02” Mix Mix Mix Mix Mix Mix “1a” “1b” “1c” “2a” “2b” “2c” 1 Optimum Asphalt Content (%) 3.83 3.87 3.95 4.29 4.31 4.33 2 Filler Contents (%) 3.45 3.48 3.56 4.29 4.31 4.33 3 Filler to AC ratio (%) 0.9 0.9 0.9 1.0 1.0 1.0 4.6 Asphalt Mixture Volumetric Mixes probable durability and in-service performance were determined by analyzing a compacted paving mixture for air voids (Va), voids in mineral aggregates (VMA) and voids filled with asphalt (VFA), commonly known as the volumetric properties of a compacted paving mixture and represented in Table 4.6. The consequences of a change in any of the volumetric factors or any detour in the procedure that offsets the total process will be a loss of performance or service life. Mixtures that ultimately consolidate to less than three percent air voids can be expected to rut and shove if placed at heavy traffic locations (Asphalt Institute, 2003). 54 Table 4.6: Hot Mix Asphalt Design Volumetric Properties Gsb Gmm Gmb 1a (PMA) Optimum AC Contents (%) 3.83 2.650 2.522 1b (60/70) 3.87 2.65 1c (40/50) 3.95 2a (PMA) 2.373 Va (%) 5.90 VMA (%) 13.90 VFA (%) 58.00 2.515 2.371 5.70 13.99 59.00 2.65 2.514 2.370 5.70 14.10 60.00 4.29 2.662 2.520 2.395 4.96 13.90 64.00 2b (60/70) 4.31 2.662 2.516 2.386 5.17 14.23 63.70 2c (40/50) 4.33 2.662 2.515 2.384 5.20 14.32 63.70 Mix Types 4.6.1 Specific Gravity According to Asphalt Institute (2003) “The ratio of the mass in air of a unit volume of a permeable material (including both permeable and impermeable voids in the aggregates) at a stated temperature to the mass in air of equal density of an equal volume of gas-free distilled water at a stated temperature is known as bulk specific gravity (Gsb) for the total aggregates”. The ratio of the mass in air of a unit volume of an impermeable material at a stated temperature to the mass in air of equal density of an equal volume of gas-free distilled water at a stated temperature is known as apparent specific gravity (Gsa). While, the ratio of the mass in air of a unit volume of a permeable material (excluding voids permeable to asphalt) at a stated temperature to the mass in air of equal density of an equal volume of gas-free distilled water at a stated temperature is the effective specific gravity (Gse) of aggregates. The effective specific gravity of the aggregate is assumed constant because absorption does not vary appreciably with change in asphalt contents. The maximum specific gravity (Gmm) at different asphalt contents was measured to calculate air voids as shown in Table 4.6. The volume of asphalt binder absorbed by an aggregate would be less than the volume of water absorbed. Consequently, the value for the effective specific gravity of an aggregate is between its bulk and apparent specific gravities. (Asphalt Institute, 2003). 55 4.6.2 Voids in Mineral Aggregates, VMA According to Asphalt Institute (2003), the voids in the mineral aggregates, are defined as the intergranular void space between the aggregates particles in a compacted paving mixture that includes the air voids and the effective asphalt content, expressed as a percent of the total volume of the sample. The VMA are calculated based on the bulk specified gravity of the aggregates and is expressed as a percentage of the bulk volume of the compacted paving mixture, given as follow and reported in Table 4.6. VMA = 100 − G mb G sb × Ps (4.1) Where VMA = voids in the mineral aggregates, percent of bulk volume. Gsb = bulk specific gravity of total aggregates Gmb = bulk specific gravity of compacted mixture Ps = aggregates content, percent by total mass of mixture. The most difficult mix design property to achieve is a minimum amount of voids in the mineral aggregates. The goal was to furnish enough space for the asphalt cement so it could provide adequate adhesion to bind the aggregates particles, but without bleeding when temperatures rise and the asphalt expands. Normally, the VMA decreases to a minimum value with increase in asphalt content. At some point as the asphalt content increases, the VMA begins to increase because relatively more dense material (aggregates) is displaced and pushed apart by the less dense material (asphalt content). The asphalt content on the “wet” side of VMA curve was avoided, even the minimum air void and VMA criteria met. Design asphalt contents in this range have a tendency to bleed and /or exhibit plastic flow when placed in the field. Any amount of additional compaction from traffic leads to inadequate room for asphalt expansion, loss of aggregates-to-aggregates contact, and eventually, rutting and shoving in high traffic areas Asphalt Institute (2003). 56 4.6.3 Air Voids, Va. According to Asphalt Institute (2003), the total volume of the small pockets of air between the coated aggregates particles throughout a compacted paving mixture, expressed as percent of the bulk volume of the compacted paving mixture. The air voids, Va, in the total compacted paving mixture consists of the small air spaces between the coated aggregates particles. The mathematical relationship has shown as; V a = 100 × G mm − G mb G mm (4.2) Where, Va = air voids in compacted mixture, percent of total volume. Gmm = maximum specific gravity of paving mixture (as determined in previous article or as determined directly for a paving mixture by ASTM D2041/AASHTO T209) Gmb = bulk specific gravity of compacted mixture The design range of air voids of mixes were kept from 4 to 7 percent for heavy traffic. The overall objective was to limit adjustments of the design asphalt content to less than 0.5 percent air voids from the median of the design criteria (four percent), especially on the low side of the range to minimize chances of rutting in the field (Asphalt Institute, 2003). 4.6.4 Voids Filled with Asphalt, VFA. According to Asphalt Institute (2003), the percentage portion of the volume of intergranular void space between the aggregates particles that is occupied by the effective asphalt. It is expressed as the ratio of (VMA-Va) to VMA. The voids filled asphalt, VFA is the percentage of the integral void space between the aggregates particles (VMA) that are filled with asphalt. The mathematical relationship has shown as; 57 VFA = 100 × VMA − V a VMA (4.3) Where, VFA = voids filled with asphalt, percent of VMA VMA = voids in the mineral aggregates, percent of bulk volume Va = air voids in compacted mixture, percent of total volume. Volumetric analysis i.e. specific gravity, maximum theoretical specific gravity, air voids, voids in mineral aggregates and voids filled with asphalt, determined from each mix type have been reported in Table 4.6. due to its high fluidity, PMA produced relatively high-density mixes than neat AC for the same gradation. The amount of effective asphalt cement, which ultimately affects the amount of air voids, utilized within the asphaltic mixture has closely been controlled to produce desired results. Void volume control in the mixture is critical. Insignificant variations in the effective proportion of asphalt cement resulting from variations in aggregate absorption characteristics may substantially affect the life and quality of the pavement. Additionally, because aggregate from different rock quarries and pit locations have varying absorptive qualities, representative samples during the testing process were carefully selected and analyzed in order to produce an even uniform pavement mix. After successful design of HMA mixtures, performance testing was carried out to investigate their resistance to permanent deformation under repeated loading. 58 Chapter Five 59 Chapter Five Uniaxial Repeated Load Strain Test 5.1 Introduction Uniaxial Repeated Load Strain Test (URLST) conforms to the requirements of the design draft issued by the British Standard Institute (BSI) as a method of measurement of resistance of asphalt mixes to permanent deformation behaviour, subjected to unconfined uniaxial cyclic loading on Universal Testing Machine (UTM-5P). A comprehensive laboratory investigation was carried out on six specified mixes to study the permanent deformation of asphalt concrete mixes at different temperatures and stress levels. The main objective of this chapter is to investigate the performance of HMA material in terms of resilient & permanent strains under cyclic loading conditions. 5.2 Universal Testing Machine (UTM-5P) UTM as shown in Figure 5.1, is a series of closed loop servo control material testing machine designed to transmit energy to the HMA specimens using high pressure air acting on double sided piston (actuator). The actuator is mechanically coupled to the specimen through a reaction loading frame. Air pressure is controlled by a servo valve where in small electric currents are used to open and close the control spool of the valve. These transducers convert mechanical movement into standard electronic signals and via Control and Data Acquisition System (CDAS) displays output on the personal computer. The transducers signals are also used to control the system. 60 Figure 5.1: Universal Testing Machines (UTM-5P) 5.3 Uniaxial Repeated Load Strain Test (URLST) In UTM-5P, Controlled strain test applied 1800 block (square) repeated load pulses, with a pulse width and pulse period of 500 milli second and 2000 milli second respectively, to the specimens to make maximum damage to the specimens. With these stress levels (100 kPa, 300 kPa, 500 kPa) and temperatures (250C, 400C, & 550C), shear mode of failure for some specimens were observed beyond 1800 cycles (K.B. De Vose and Feeley, 2002). A static conditioning time of 100 sec and conditioning stress of 10kPa was also applied prior to commencement of actual test. Following the conditioning period, a fixed twenty seconds rest period was programmed where the applied stress was set to zero. Data is collected from the loading pulses at a linear (i.e. equidistant) time interval and then stored in a buffer. Digital filtering is applied to the data at various stages of its capture and processing to ensure smoothness in the reported and displayed data. As test proceeds, plotted 61 data is displayed with linear vertical and horizontal axis in terms of accumulated strain, derivation of accumulated strain (slope), resilient strain, resilient modulus and creep stiffness. 5.4 Test Conditions 5.4.1 Load Conditions Specimens were subjected to repeated pulse loading of 1800 cycles at 100, 300 and 500 kPa stress levels. Pulse width and pulse period were kept 500 milli second and 2000 milli second respectively. A range of stresses from 100 kPa to 500 kPa were selected for this study to cover almost all the domains of loads plying on the road. The magnitude of the loading stress and the timing width of the pulse applied to the specimen during a test were directly controlled by the “test loading stress” and “pulse width” parameters in the test set up and control edit screen. A minimum force of 20N ensured that the loading actuator would not be lifted off the specimen between pulse applications. As pulse loading continued, the permanent deformation in terms of accumulated strain was measured using two Linear Variable Displacement Transformers (K.B. De Vose and Feeley, 2002). 5.4.2 Temperature Conditions The typical test temperatures i.e. 250C, 400C, & 550C, are adopted in the laboratory for research study to accommodate the influence of temperature conditions in summer expected on National Highways. It is the representative of typical temperature conditions observed on flexible pavements in Pakistan 5.5 Testing Methodology For each stress level, three specimens were tested at 250C, 400C and 550C. Samples are evaluated for 54 test conditions (27 for each type of gradation) and total 162 specimens are prepared. Test matrix for different conditions, used in the experimental study has been shown in Table 5.1. 62 Table 5.1: Specimens at each Test Condition Stress (kPa) Temperature (0C) Mixes with Gradation 01 1a 1b 1c 100 25 300 500 100 300 500 100 300 500 40 55 5.6 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 Mixes with Gradation 02 2a 2b 2c 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 3 Discussion of URLST Results Influence of pulse counts (1800 cycles) on accumulated strains for six types of mixes (1a, 1b, 1c, 2a, 2b, 2c) at three temperatures (250C, 400C, & 550C) and three stress levels (100 kPa, 300 kPa, & 500 kPa) have been shown in Annexure-A and summarized in Table 5.2. Table 5.2: Mean Accumulated Strain (εp) of Mixes Sr. Temperature Stress Percentage Accumulated Strain (εp) of Mixes No. (0C) (kPa) 1a 1b 1c 2a 2b 2c 1 25 100 0.193 0.281 0.183 0.286 0.403 0.315 2 25 300 0.399 0.564 0.375 0.516 0.572 0.493 3 25 500 0.616 0.686 0.5926 0.907 0.958 0.926 4 40 100 0.332 0.424 0.389 0.540 0.590 0.536 5 40 300 0.547 0.742 0.666 0.609 0.774 0.676 6 40 500 0.881 0.990 0.946 0.989 1.057 0.995 7 55 100 0.438 0.881 0.526 0.833 0.941 0.747 8 55 300 1.114 1.32 1.124 1.052 1.487 1.058 9 55 500 1.242 1.692 1.247 1.320 1.951 1.375 63 Table 5.2 show that permanent strain increases with the increase in stress levels and temperature. The effect of temperature variation is significant as compared to stress levels. Results of Table 5.2 have been shown graphically in Figure 5.2. Mixture with Gradation "01" & PMA (1a) 1.30 Mixture with Gradation "02" & PMA (2a) 1.50 1.30 1.10 500kPa 1.10 0.90 500kPa 0.90 Єp (% ) Єp (%) 0.70 300kPa 0.70 300kPa 0.50 0.50 100kPa 100kPa 0.30 0.30 0.10 0.10 10 25 0 Temperature ( C) 40 100kPa 55 300kPa 10 500kPa Mixture with Gradation "01" & 60/70 Pen. Grade (1b) 1.90 25 0 Temperature ( C) 1.90 1.50 1.70 55 500kPa 300kPa Mixture with Gradation "02" & 60/70 Pen. Grade (2b) 2.10 1.70 40 100kPa 1.50 1.30 500kPa 0.90 300kPa 0.70 0.50 500kPa 1.30 Єp (%) Єp (%) 1.10 100kPa 1.10 0.90 300kPa 0.70 100kPa 0.50 0.30 0.30 0.10 10 25 40 0.10 55 10 0 Temperature ( C) 300kPa 500kPa Mixture with Gradation "01" & 40/50 Pen. Grade (1c) 1.50 25 0 Temperature ( C) 1.30 1.30 1.10 1.10 40 100kPa 300kPa 55 500kPa Mixture with Gradation "02" & 40/50 Pen. Grade (2c) 1.50 500kPa 0.90 0.90 500kPa Єp (%) Єp (%) 100kPa 0.70 300kPa 0.50 300kPa 0.70 100kPa 0.50 100kPa 0.30 0.30 0.10 10 25 0 Temperature ( C) 40 100kPa 0.10 55 300kPa 10 500kPa 25 0 Temperature ( C) 40 100kPa 300kPa Figure 5.2: Influence of Temperature and Stress Levels on Accumulated Strain 64 55 500kPa Figure 5.2 shows that permanent strain (εp) in all mixes increases from low stress level (100 kPa) to high stress level (500kPa) at low temperature (250C). At low temperature and high stress level, densification of mixes may lead to a decreasing trend. At medium temperature (400C), εp has a constant rate of increase with the increase in stress. The stiffness offered by the mix at this specific temperature may be combined effects of both the gradation and the binder. At higher temperature (550C), the trends are likely the same as at low temperatures i.e. εp increases at the decreasing rate, but higher than low temperature. The reasons may be the shift of resistance offered by binder to aggregate skeleton. Results of percentage accumulated strain drawn against pulse counts were multiplied by one million to obtain positive values on logarithmic scale according to Figure 2.12 in section 2.9. Straight line trends were developed as shown in Annexure-B. Equations for each stress level have been shown on the respective figures. Where in percentage accumulated strain is denoted “y”, pulse counts with “x” and regression coefficients with “a” and “b”, which are namely intercept coefficients and slope coefficients respectively. Detail discussion of regression coefficients have been given in chapter Seven. 5.7 Resilient strain Resilient strain is the recoverable strain phase after a loading event. Resilient strain is mainly affected by the density of mix, moisture content, temperature, particle size and shape, aggregate gradations, lateral confinement, loading condition and bulk stress. Resilient strain can be calculated using the following relation; Єr = ∆r / havg. Єr = resilient strain ∆r = resilient deformation = ∆r,h - ∆r,I havg. = average height of test specimen ∆r,h = high deformation reading for the LVDT and ∆r,I = low deformation reading for the LVDT (5.1) Where 65 The results of resilient strain of asphalt mixes under applied uniaxial loads have been plotted graphically and shown in Annexure-C, in order to make a comparison among gradations, asphalt types, temperatures and stress levels on resilient strain values. Resilience is the property of a material to absorb energy when it is deformed elastically and then, upon loading to have this energy recovered. In other words, it is the maximum energy per volume that can be elastically stored. Results of resilient strain have been shown in Table 5.3, which shows strong correlations among all the four variables. Resilient strain increases with the increase in stress levels and temperatures similar to accumulative strain as shown in Figure 5.3 through 5.8. It was observed that PMA and 40-50 penetration grade asphalt mixes has almost equal resilient strain with gradation “01”. Table 5.3: Resilient Strain of Mixes Sr. Temp. Stress No. (0C) (kPa) 1 25 100 2 25 300 3 25 500 4 40 100 5 40 300 6 40 500 7 55 100 8 55 300 9 55 500 Resilient Strain of Mixes 1a 1b 1c 2a 2b 2c 0.0364 0.0574 0.0346 0.0231 0.0454 0.0217 0.0450 0.0728 0.0423 0.0553 0.0750 0.0601 0.0698 0.0768 0.0635 0.0884 0.1096 0.0928 0.0461 0.0620 0.0509 0.0308 0.0680 0.0254 0.0487 0.0792 0.0620 0.0647 0.0826 0.0672 0.0757 0.0809 0.0732 0.1076 0.1312 0.1109 0.0568 0.0629 0.0580 0.0394 0.0826 0.0524 0.0622 0.1185 0.0734 0.0885 0.1287 0.1018 0.1011 0.1177 0.1044 0.1157 0.1677 0.1339 66 Table 5.3 shows that at low stress level (100 kPa) mixes with coarser gradation exhibits more resilient strain and at high stress level (500 kPa) mixes with finer gradation expressed more strain. The main reason behind is the total strain developed under same stress conditions. Mixes with fine gradation has more value of total strain, hence the amount of recoverable strain has also higher value than coarse mixes. It has also been observed in Table 5.3 that at higher temperatures (400C & 550C) the percentage difference in rate of recoverable strain or Resilient Strain( %) resilient strain reduces. These results have further been elaborated in Figure 5.3 through 5.8. 0 Mixture with Gradation "01" & PMA (1a) 0.11 0.10 0.09 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0 100 200 300 400 500 55 C 0 40 C 0 25 C 600 Stress (kPa) Figure 5.3: Influence of Stress Levels on Resilient Strain, Mix 1a Resilient Strain( %) Mixture with Gradation "01" & 60/70 Pen. Grade (1b) 0.13 0 55 C 0.12 0.11 0.10 0 0.09 40 C 0.08 0 25 C 0.07 0.06 0.05 0.04 0 100 200 300 400 Stress (kPa) 500 Figure 5.4: Influence of Stress Levels on Resilient Strain, Mix 1b 67 600 Resilient Strain( %) Mixture with Gradation "01" & 40/50 Pen. Grade (1c) 0.11 0 55 C 0.10 0.09 0.08 0.07 0.06 0.05 0.04 0.03 0.02 0 40 C 0 25 C 0 100 200 300 400 500 600 Stress (kPa) Figure 5.5: Influence of Stress Levels on Resilient Strain, Mix 1c Mixture with Gradation "02" & PMA (2a) 0.14 0 55 C Resilient Strain( %) 0.12 0 40 C 0.1 0.08 0 25 C 0.06 0.04 0.02 0 0 100 200 300 400 500 Stress (kPa) Figure 5.6: Influence of Stress Levels on Resilient Strain, Mix 2a 68 600 Mixture with Gradation "02" & 60/70 Pen. Grade (2b) 0.18 0 55 C Resilient Strain( %) 0.16 0 40 C 0.14 0.12 0 25 C 0.1 0.08 ` 0.06 0.04 0.02 0 0 100 200 300 400 500 600 Stress (kPa) Figure 5.7: Influence of Stress Levels on Resilient Strain, Mix 2b Resilient Strain( %) Mixture with Gradation "02" & 40/50 Pen. Grade (2c) 0.16 0 55 C 0.14 0 0.12 40 C 0.10 0 25 C 0.08 0.06 0.04 0.02 0.00 0 100 200 300 400 500 Stress (kPa) Figure 5.8: Influence of Stress Levels on Resilient Strain, Mix 2c 69 600 5.8 Summary of Results A comprehensive laboratory investigation was carried out on six mixes ranging from finer to coarser. Specimens were subjected to cyclic loading on UTM-5P to study the resistance against permanent deformation of the mixes at 250C, 400C and 550C. At low temperatures and stress levels, both coarse and fine graded mixes showed less accumulated strain, whereas at higher temperatures and stress levels, coarse graded mix with PMA showed good resistance to permanent deformation. Mixes with coarser gradations offered more resistance to rutting than mixes with finer gradations under URLST. Overall performance of mixes with penetration grade “60/70” bitumen were lower as compared to mixes with penetration grade “40/50” and PMA. 70 Chapter Six 71 Chapter Six Wheel Tracking Test 6.1 Introduction Wheel Tracker typically measures the rut, created by repeated passage of a wheel over prismatic asphalt concrete samples. It was used to assess the resistance to rutting of the asphaltic material, under standard defined conditions of load and temperatures i.e. 25, 40 and 55 0C. This chapter presents the relationship established between rutting and factors effecting rutting. Ranges of intercept and slope coefficient were also determined. 6.2 Wheel Tracking Device A loaded wheel shown in Figure 6.1 tracks a sample under specified conditions of speed and temperature while the development of the rut is monitored continuously during the test. Figure 6.1: Wheel Tracker (After Cooper, 2006) 72 The rut resistance can be quantified as the rate of rutting during the test or the rut depth at the conclusion of the test, measured with Linear Variable Displacement Transformers (LVDT) 25mm (min). Slab specimens were prepared in the laboratory for research study. The susceptibility of an asphaltic material to deform is assessed by measuring the rut formed by repeated passes of a loaded wheel at specified temperatures. The wheel tracking apparatus consists of loaded wheel which bears on a sample held on a moving table as shown in Figure 6.2. The moving table reciprocate with simple harmonic motion with a frequency of 26.5 passes per minutes (European Standard- PrEN 13108/12697-22, 2002). Figure 6.2: Wheel Tracker Solid Rubber Tyre (After Cooper, 2006) The wheel is fitted with solid rubber tyre of outside diameter 200 mm. The tyre is a rectangular section 50 ± 1 mm wide and 10mm to 13 mm thick. The wheel tracker is fitted with a temperature controlled cabinet with a maximum temperature up to 650C ± 10C. Square slab specimens (305x305mm) of asphalt mixes with typical asphalt wearing course thickness of 50mm thick, fitted with wheel tracker (WT) table, clamps for securing specimen holders. Mixes were evaluated under a loaded wheel (700 ± 20 N) tracked with simple harmonic motion 73 through a distance of 305mm on specimens under specified conditions i.e. 53 passes per minute at temperatures 250C, 400C and 550C (European Standard- PrEN 13108/12697-22, 2002). The operational software run under Windows 95 to start and stop the WT, control speed and acquire deformation and temperature data. An on-screen display provides a continuously updated graph of time versus deformation as shown in Figure 6.3. The test data are stored in a text file for subsequent analysis using a spreadsheet. Figure 6.3: Wheel Tracker out put Display (After Cooper, 2006) 74 6.3 Specimen Preparation on Roller Compactor Specimens of each type of asphalt concrete mix were prepared with the help of roller compacter at specific temperatures and number of specimen for each condition is given in Table 6.1. Weight of each specimen was taken 12.50 kg (approximately). Mixing of aggregate with binder was carried out at a working temperature of 1550C + 50C. Table 6.1: Details of Specimens Number of Specimens for Each Mix Type Gradation 01 Gradation 02 1b 1c 2a 2b Temperature (0C) 1a 25 3 3 3 3 3 3 40 3 3 3 3 3 3 55 3 3 3 3 3 3 2c Roller Compactor compacts the specimen uniformly with its specially designed press system to a specified pressure as shown in Figure 6.4. Figure 6.4: Slab Compaction on Roller Compacter (After Cooper, 2006) 75 For each temperature, compaction of slab specimens at roller compactor was undertaken in four stages. Pressure at each stage and number of passes corresponding to each pressure have been reported in Table 6.2. Table 6.2: Compaction Pressures 1 Compaction Stages 2 3 4 Pressure (bar) 2 5 4 3 No. of Passes 10 10 5 5 Wheel tracker tracked for 18000 pulses or 25mm rut depth, which ever happens first, at standard conditions of applied load (720N), frequency (26.5rpm) on compacted confined slab specimens for six mixes. Rut development of mixes under wheel tracking test was analyzed and summary of rut depth, measured at three temperatures have been reported in Table 6.3. Table 6.3: Rut Depth of Mixes Measured on Wheel Tracker Temperature Sr. (0C) No. Measured Rut Depth (mm) of Mixes “1a” “1b” “1c” “2a” “2b” “2c” 1 25 2.74 3.90 2.82 4.53 5.99 4.60 2 40 6.20 9.99 6.62 10.86 14.60 12.08 3 55 8.53 15.20 11.61 17.80 23.40 19.00 Table 6.3 shows that mixes with coarser aggregate gradation has lesser value of rut depth as achieved by mixes with finer gradation, irrespective of AC type. At 250C & 400C temperatures, mixes with PMA and 40/50 penetration grade AC have shown almost similar rut potential, but at 550C mixes with PMA has the least value of rut depth among all the mixes. Percentage difference in rut potential of mixes with aggregate gradation “01” with three AC types is more than mixes with gradation “02”. The development of rut in asphalt concrete specimens under the wheel pass of the Wheel Tracker has been shown in Figure 6.5. 76 Figure 6.5: Rut Development in Mixes 77 6.4 Discussion of Results Influence of load cycles on rut depth was measured and plotted using the data out put file and shown graphically in Annexure-D. Results of WT show that permanent deformation is a function of number of load repetitions. At high temperature (550C), PMA mix shows better resistance to rutting as total rut depth achieved is only 8.5 and 17.8 mm in mixes with gradation “01” and “02” respectively. Similarly, rut depth of 60/70 pen. grade & 40/50 pen grade asphalt mixes was 15.2mm & 11.61mm with aggregate gradation “01”, while 23.4mm & 19.00mm with aggregate gradation “02” respectively. Rut development on WT is further plotted on log-log scale as discussed earlier in section 2.9 and shown in Figure 2.12 and the results are expressed in Figure 6.6 through 6.11. Straight line trends have shown good agreement (R2). Relationships between load repetitions and rut depth can be measured at any value of N using these plots. Mixture with gradation "01" & PMA (1a) 7.5 y 25 = 0.2694Ln(x) + 3.7261 R2 = 0.9524 6 Rut Depth (mmx10 ) 7.0 6.5 6.0 5.5 5.0 y 55 = 0.2231Ln(x) + 4.7724 R2 = 0.9944 y 40 = 0.2037Ln(x) + 4.8079 R2 = 0.9965 4.5 4.0 10 1a, 25C 100 1a, 40C 1000 Load Cycle (LOG N) 1a, 55C Log. (1a, 25C) 10000 Log. (1a, 40C) Figure 6.6: Relationship of Load Cycle & Rut Depth, Mix 1a 78 100000 Log. (1a, 55C) Mixture with gradation "01" & 60/70 Pen. Grade (1b) 7.5 y 25= 0.2737Ln(x) + 3.9131 R2 = 0.9838 6 Rut Depth (mmx10 ) 7.0 6.5 6.0 5.5 5.0 y 55 = 0.2174Ln(x) + 5.1163 R2 = 0.9591 y 40 = 0.2295Ln(x) + 4.8107 R2 = 0.9452 4.5 4.0 10 100 1b, 25C 1b, 40C 1b, 55C 1000 Load Cycle (LOG N) 10000 Log. (1b, 25C) Log. (1b, 40C) 100000 Log. (1b, 55C) Figure 6.7: Relationship of Load Cycle & Rut Depth, Mix 1b Mixture with gradation "01" & 40/50 Pen. Grade (1c) 7.5 y 25 = 0.2957Ln(x) + 3.4551 R2 = 0.3927 7.0 6 Rut Depth (mmx10 ) 6.5 6.0 5.5 5.0 4.5 y 55 = 0.2844Ln(x) + 4.3368 R2 = 0.9809 y 40 = 0.2111Ln(x) + 4.7834 R2 = 0.9909 4.0 10 1c 25C 100 1c, 40C 1c, 55C 1000 Load Cycle (LOG N) Log. (1c 25C) Log. (1c, 40C) Figure 6.8: Relationship of Load Cycle & Rut Depth, Mix 1c 79 10000 100000 Log. (1c, 55C) Mixture with gradation "02" & PMA (2a) 7.5 y 25= 0.1317Ln(x) + 5.3749 R2 = 0.9926 6 Rut Depth (mmx10 ) 7.0 6.5 6.0 5.5 y 40 = 0.2161Ln(x) + 4.9767 R2 = 0.9676 5.0 10 2a, 25C 100 2a, 40C y 55 = 0.2217Ln(x) + 5.1177 R2 = 0.9875 1000 Load Cycle (LOG N) 2a, 55C Log. (2a, 25C) 10000 Log. (2a, 40C) 100000 Log. (2a, 55C) Figure 6.9: Relationship of Load Cycle & Rut Depth, Mix 2a Mixture with gradation "01" & 60/70 Pen. Grade (2b) 8.0 y 25 = 0.1633Ln(x) + 5.1857 R2 = 0.9869 6 Rut Depth (mmx10 ) 7.5 7.0 6.5 6.0 5.5 y 55 = 0.2118Ln(x) + 5.3468 R2 = 0.9691 y 40 = 0.1968Ln(x) + 5.2741 R2 = 0.9801 5.0 10 2b, 25C 100 2b, 40C 1000 Load Cycle (LOG N) 2b, 55C Log. (2b, 25C) 10000 Log. (2b, 40C) Figure 6.10: Relationship of Load Cycle & Rut Depth, Mix 2b 80 100000 Log. (2b, 55C) Mixture with gradation "01" & 40/50 Pen. Grade (2c) 8.0 7.5 y 25 = 0.2606Ln(x) + 4.1251 R2 = 0.9942 6 Rut Depth (mmx10 ) 7.0 6.5 6.0 5.5 5.0 y 55 = 0.239Ln(x) + 5.0268 R2 = 0.9165 y 40 = 0.199Ln(x) + 5.1743 R2 = 0.9824 4.5 4.0 10 2c, 25C 100 2c, 40C 1000 Load Cycle (LOG N) 2c, 55C Log. (2c, 25C) 10000 100000 Log. (2c, 40C) Log. (2c, 55C) Figure 6.11: Relationship of Load Cycle & Rut Depth, Mix 2c For quick review, summary of intercept coefficients and slope coefficients from Figure 6.6 through 6.11 has been reported in Table 6.4 & 6.5 respectively. It shows that ranges of intercept coefficient “a” and slope coefficient “b” for all the six mixes were observed from 3.46 to 5.38 and 0.16 to 0.30 respectively. Table 6.4: Intercept Coefficient “a” of Mixes Sr. No. Temp. (0C) 1 25 2 3 40 55 Intercept Coefficient “a” of Mixes “1a” “1b” “1c” “2a” “2b” “2c” 3.73 4.81 4.77 3.91 4.81 5.12 3.46 4.78 4.34 5.38 4.98 5.12 5.19 5.27 5.35 4.13 5.17 5.03 Table 6.5: Slope Coefficient “b” of Mixes Sr. No. 1 2 3 Temp. (0C) 25 40 55 “1a” 0.27 0.21 0.22 “1b” 0.27 0.23 0.22 Slope Coefficient “b” of Mixes “1c” “2a” “2b” 0.30 0.13 0.16 0.21 0.22 0.20 0.28 0.22 0.21 81 “2c” 0.26 0.20 0.24 Chapter Seven 82 Chapter Seven Results and Discussions 7.1 Introduction Comprehensive performance based testing on several specimens were carried out after complete characterization of materials and determinations of HMA properties as described in the previous chapters. One of the objectives of the testing was to investigate the effect of HMA properties on its permanent deformation behaviour. The evaluation of the resistance to permanent deformation of six mixes was carried out using both the test methods. This chapter comprised of calculating the regression coefficients i.e. intercept and slope coefficient and Permanent Deformation Coefficients alpha and mu. A brief introduction of these coefficients has already been provided in chapter Two. Also, computation of domains of “α” & “µ” and its comparison with other researcher has been reported in this chapter. In addition, a comparison of both the test have also been made using the shift factors techniques, ranking of mixes (based on intercept coefficients) and a correlation between both the test methods have also been made a part of this study. 7.2 Regression Coefficients The derivation of regression coefficients i.e. intercept coefficients “a” & slope coefficients “b” are based on log-log scale between load repetition and permanent strain as already explained in section 2.9. Relationships have been plotted for six mixes under each temperature and stress conditions as reported in Annexure-D. This section provides a comparison of results obtained for each coefficient. 7.2.1 Intercept Coefficient (a) Intercept coefficient (a) of mixes for each temperature and stress condition have given in 83 Table 7.1, 7.2, & 7.3. One can observe the accuracy of computed data from the Standard Deviation (SD) & Coefficient of Variance (CV) Table 7.1: Intercept Coefficient ‘a’ at 250C Stress Level 100 300 500 Mixes with gradation "01" Mixes with gradation "02" 1a 1b 1c 2a 2b 2c Standard Deviation 4.07 4.74 5.13 4.41 4.91 5.29 4.55 4.91 5.52 4.06 4.66 5.08 4.53 4.92 5.25 4.86 5.21 5.25 0.06 0.06 0.02 Coefficient of Variance 1.3 1.2 0.4 Table 7.2: Intercept Coefficient ‘a’ at 400C Stress Level 100 300 500 Mixes with gradation "01" Mixes with gradation "02" 1a 1b 1c 2a 2b 2c Standard Deviation 4.77 5.15 5.25 4.73 4.89 5.07 4.93 5.15 5.25 5.06 5.14 5.36 5.15 5.13 5.31 4.91 5.07 5.10 0.06 0.04 0.04 Coefficient of Variance 1.22 0.74 0.78 Table 7.3: Intercept Coefficient ‘a’ at 550C Stress Level 100 300 500 Mixes with gradation "01" Mixes with gradation "02" 1a 1b 1c 2a 2b 2c Standard Deviation 4.73 4.82 4.98 4.82 5.06 5.32 4.82 4.97 5.09 4.78 4.83 5.05 5.02 5.07 5.48 4.83 4.99 5.32 0.04 0.04 0.07 Coefficient of Variance 0.74 0.83 1.39 From Table 7.1, 7.2 and 7.3, it can be observed the intercept coefficient increases with an increase in stress level, irrespective of aggregate gradations, A.C types, mix types, and test temperatures. Domains of ‘a’ for temperature 250C, 400C and 550C are 4.07 to 5.52, 4.73 to 5.36, and 4.733 to 5.481 respectively. Irrespective of test temperature and stress levels, domain of mixes with aggregate gradation “01”was observed to be 4.06 to 5.52 and for mixes with gradation “02” was 4.06 to 5.481. Standard deviation measured for ‘a’ has a range from 0.02 to 0.0723. 84 7.2.2 Slope Coefficient (b) Slope coefficient shows a reverse trend unlike intercept coefficient as tabulated in Table 7.4, 7.5 and 7.6. Table 7.4: Slope Coefficient ‘b’ at 250C Mixes with gradation "01" Stress Level 1a 1b 1c 100 300 500 0.20 0.12 0.10 0.15 0.10 0.08 0.15 0.09 0.07 Mixes with gradation "02" 2a 2b 2c Standard Deviation 0.20 0.14 0.12 0.15 0.11 0.10 0.10 0.08 0.08 0.02 0.01 0.01 Coefficient of Variance 9.33 7.52 7.41 Table 7.5: Slope Coefficient ‘b’ at 400C Stress Level 100 300 500 Mixes with gradation "01" Mixes with gradation "02" 1a 1b 1c 2a 2b 2c Standard Deviation 0.12 0.08 0.08 0.13 0.10 0.09 0.18 0.12 0.10 0.10 0.09 0.09 0.11 0.10 0.09 0.17 0.12 0.11 0.01 0.01 0.01 Coefficient of Variance 8.76 6.18 4.20 Table 7.6: Slope Coefficient ‘b’ at 550C Stress Level 100 300 500 Mixes with gradation "01" Mixes with gradation "02" 1a 1b 1c 2a 2b 2c Standard Deviation 0.16 0.15 0.12 0.16 0.14 0.10 0.19 0.16 0.13 0.16 0.13 0.11 0.13 0.11 0.10 0.16 0.12 0.11 0.004 0.007 0.005 Coefficient of Variance 8.76 6.18 4.61 It decreases with the increase in stress level, irrespective of aggregate gradations, A.C types, mix types, and test temperature. For all the mixes, the domain is higher and percentage difference of ‘b’ is more at low temperature (250C) and stress level (100 kPa) than higher temperature (550C) and stress levels (500 kPa). Mixes with “PMA” and “40/50” penetration grade have higher value of ‘b’ as compared to “60/70” penetration grade. Overall domain of ‘b’ for mixes with gradation “01” and “02” is 0.07 to 0.20 and 0.08 to 0.20 respectively. Standard deviation observed for “b” has a range from 0.004 to 0.02. 85 7.3 Permanent Deformation Coefficients Two very use full coefficients i.e. Alpha (α) & Mu (µ) have been computed in this section from straight line trends reported in Annexure-D, by using the derivation of base log-log scale of power model as explained earlier in section 2.9. This section provides a comparison of results obtained for each coefficient. 7.3.1 Alpha (α) Permanent deformation coefficient (α) as given in Tables 7.7, 7.8 and 7.9 and graphically shown in Figure 7.1 has similar trends as intercept coefficient i.e. with an increase in stress level, alpha increases. Table 7.7: Permanent Deformation Parameter ‘α’ at 250C Stress Level 100 300 500 Mixes with gradation "01" 1a 1b 1c 0.80 0.85 0.85 0.88 0.90 0.91 0.90 0.92 0.93 Mixes with gradation "02" 2a 0.80 0.86 0.88 2b 0.85 0.89 0.90 2c 0.90 0.92 0.92 Table 7.8: Permanent Deformation Parameter ‘α’ at 400C Mixes with gradation Mixes with gradation "02" Stress "01" Level 1a 1b 1c 2a 2b 2c 100 0.88 0.87 0.82 0.91 0.89 0.85 300 0.92 0.90 0.88 0.91 0.90 0.88 500 0.92 0.91 0.90 0.92 0.91 0.89 Standard Deviation 0.02 0.01 0.01 Coefficient of Variance 1.79 0.86 0.74 0.01 0.01 0.01 Coefficient of Variance 1.32 0.70 0.42 Standard Deviation Coefficient of Variance 0.004 0.007 0.005 0.45 0.85 0.57 Standard Deviation Table 7.9: Permanent Deformation Parameter ‘α’ at 550C Stress Level 100 300 500 Mixes with gradation "01" 1a 1b 1c 0.84 0.84 0.84 0.85 0.87 0.85 0.88 0.91 0.87 Mixes with gradation "02" 2a 2b 2c 0.84 0.87 0.84 0.87 0.90 0.88 0.89 0.90 0.90 Overall value of ‘α’ for coarse as well as fine graded mixes has a range from 0.80 to 0.92, thus showing minor effect of temperature and stress levels. However trends show that ‘α’ decreases 86 with increase in temperature and increases with increase in stress levels. Standard deviation measured for ‘α’ has a range from 0.004 to 0.02. Mixture with gradation "01" & PMA (1a) 0.92 0.92 0.90 0.90 500 kPa 0.88 0.86 300 kPa 0.84 0.82 300 kPa 0.86 0.84 100 kPa 0.82 100 kPa 0.80 0.78 0.78 20 30 0 Temperature ( C) 40 100 kPa 50 60 300 kPa 20 0 0.92 500 kPa 300 kPa 0.86 0.84 A lpha 0.90 0.88 100 kPa 0.82 20 30 0 Temperature ( C) 40 100 kPa 50 300 kPa 60 0.86 300 kPa 0.84 0.82 100 kPa 0.80 20 30 0 Temperature ( C) 40 100 kPa 50 60 300 kPa 500 kPa 30 40 50 60 100 kPa 300 kPa 500 kPa 500 kPa 300 kPa 100 kPa ` 20 30 0 Temperature ( C) Figure 7.1: Influence of Temperature and Stress Levels on Alpha 87 500 kPa Mixture with gradation "02" & 40/50 (2c) 0.93 0.92 0.91 0.90 0.89 0.88 0.87 0.86 0.85 0.84 0.83 Alpha 0.88 300 kPa 100 Temperature ( C) 500 kPa 0.90 100 kPa 60 300 0 0.92 50 500 20 500 kPa 40 Mixture with gradation "02" & 60/70 (2b) 0.92 0.91 0.90 0.89 0.88 0.87 0.86 0.85 0.84 Mixture with gradation "01" & 40/50 (1c) 0.94 30 Temperature ( C) 500 kPa Mixture with gradation "01" & 60/70 (1b) 0.94 A lp h a 500 kPa 0.88 0.80 Alpha Mixture with gradation "02" & PMA (2a) 0.94 Alpha A lpha 0.94 40 100 kPa 50 300 kPa 60 500 kPa 7.3.2 Mu (µ) Mu, as given in Table 7.10, 7.11 and 7.12 and graphically shown in Figure 7.2 has similar trends as slope coefficient. Table 7.10: Permanent Deformation Parameter ‘µ’ at 250C Stress Level 100 300 500 Mixes with gradation "01" Mixes with gradation "02" 1a 1b 1c 2a 2b 2c Standard Deviation 20.061 13.534 8.037 20.45 9.793 6.315 19.854 11.066 6.017 17.953 9.091 6.796 16.171 10.423 6.382 18.813 9.432 5.678 0.601 0.707 0.307 Coefficient of Variance 3.12 6.87 4.70 Table 7.11: Permanent Deformation Parameter ‘µ’ at 400C Stress Level 100 300 500 Mixes with gradation "01" Mixes with gradation "02" 1a 1b 1c 2a 2b 2c Standard Deviation 13.865 9.190 6.393 12.276 9.283 5.820 21.235 11.636 7.083 17.078 8.250 5.817 10.499 6.731 4.603 20.369 10.45 6.003 1.640 0.634 0.304 Coefficient of Variance 10.29 6.85 5.11 Table 7.12: Permanent Deformation Parameter ‘µ’ at 550C Stress Level 100 300 500 Mixes with gradation "01" 1a 1b 1c 15.406 12.412 19.862 13.124 7.868 12.401 7.145 4.491 6.897 Mixes with gradation "02" 2a 2b 2c 21.573 8.828 17.670 9.185 4.572 6.902 5.135 3.316 4.627 Standard Deviation 1.579 1.224 0.553 Coefficient of Variance 10.55 13.58 10.50 It decreases with an increase in stress levels and temperature, irrespective of aggregate gradations, binder types and mix types. Overall domain of ‘µ’ for all the mixes was observed to be is 21.57 to 3.31. Standard deviation measured for ‘µ’ has a range from 0.304 to 1.64. 88 Mixture with gradation "01" & PMA (1a) 25.00 Mixture with gradation "02" & PMA (2a) 25.00 100 kPa 20.00 20.00 15.00 100 kPa 300 kPa 10.00 Mu Mu 15.00 5.00 300 kPa 10.00 5.00 500 kPa 500 kPa 0.00 0.00 20 30 40 0 Temperature ( C) 50 100 kPa 60 300 kPa 20 35 0 100 kPa 40 45 100 kPa 50 55 300 kPa 500 kPa 15.00 100 kPa 300 kPa Mu Mu 15.00 10.00 60 Mixture with gradation "02" & 60/70 (2b) 20.00 20.00 30 Temperature ( C) 500 kPa Mixture with gradation "01" & 60/70 (1b) 25.00 25 10.00 300 kPa 5.00 5.00 500 kPa 0.00 20 25 30 35 40 45 50 55 500 kPa 0.00 60 20 Temperature ( C) 100 kPa 300 kPa 35 40 45 100 kPa 50 55 300 kPa Mixture with gradation "02" & 40/50 (2c) 25.00 60 500 kPa 100 kPa 20.00 20.00 100 kPa 15.00 15.00 300 kPa Mu Mu 30 Temperature ( C) 500 kPa Mixture with gradation "01" & 40/50 (1c) 25.00 25 0 0 300 kPa 10.00 10.00 5.00 5.00 500 kPa 500 kPa 0.00 0.00 20 25 30 35 40 45 50 55 60 20 30 0 0 Temperature ( C) 100 kPa 300 kPa 500 kPa Temperature ( C) Figure 7.2: Influence of Temperature and Stress Levels on Mu 89 40 100 kPa 50 300 kPa 60 500 kPa 7.4 Comparisons of Results to Other Researchers Alpha and mu reported in this research ranges from 0.8 to 0.92 and 3.31 to 21.57 respectively. Average values of alpha and mu from all the results are 0.881 and 10.812 respectively. Specimens, prepared at 5 to 7 % air voids, were subjected to 1800 load repetitions at constant load frequency 60 cycles/minutes (0.1 sec load duration and 0.9 sec. rest period), under unconfined conditions, at different stress levels (100, 300, 500 kPa) and temperatures (25, 40, 550C). Domains of alpha and mu with respect to temperature and stress level have been shown in Figure 7.3 & 7.4 respectively. Aalpha and mu reported by Rita et al. (1991) ranges from 0.254 to 0.81 and 0.006 to 3.10 and the average values for the same from all the results were 0.56 and 0.58 respectively. In their study, a constant load frequency of 60 cycles/minutes (0.1 sec load duration and 0.9 sec. rest period) was used for all test specimens, prepared at 3.5%, 6.5% and 9.5% air voids. Specimens were subjected upto 30,000 load repetitions with a deviator stress range from 68 kPa to 200 kPa and three temperatures i.e. 65, 80, and 95 0F. Results of alpha and mu, derived by Monismith and Rauhut are 0.56, and 0.28 to 0.35 (Monismith et al 1976), 0.45 to 0.9 and 0.1 to 0.48 (Rauhut et al 1978) respectively. Slope parameter report by Garba (2002) are in the range of 0.1617 to 0.1806 for HMA specimens with 4%, 4.7% and 5.4% asphalt cement contents and from 0.135 to 0.1836 for specimen with 3%, 5% and 8% air voids. This parameter had been found within narrow range and suggested not a good measure of resistance to measure the permanent deformation of mixes prepared with same materials. Intercept parameter was concluded to be a sensitive parameter in the power model that showed variation with both the percentage asphalt contents and air voids. Intercept parameter varies from 2.031 to 4.52 for specimens with 4%, 4.7% and 5.4% asphalt contents and from 1.649 to 6.91 for specimens with 3%, 5% and 8% air voids. Permanent Deformation Coefficients computed in this study shows similar domains as in previous researchers has reported. 90 Domains of Alpha 0.95 Alpha 0.90 0.85 0.80 0.75 0 200 400 Stress level 600 25C,1a 25C,1b 25C,1c 25C,2a 25C,2b 25C, 2c 40C,1a 40C,1b 40c, 1c 40C,2a 40c,2b 40C,2c 55C,1a 55C,1b 55C,1c 55C,2a 55C,2b 55C,2c Figure 7.3: Domains of alpha Domains of Mu 25.00 Mu 20.00 15.00 10.00 5.00 0.00 0 100 200 300 400 500 600 Stress levels 25C, 1a 25C,1b 25C,1c 25C,2a 25C,2b 25C,2c 40C,1a 40C,1b 40C,1c 40C,2a 40C,2b 40C,2c 55C,1a 55C,1b 55C,1c 55C,2a 55C,2b 55C,2c Figure 7.4: Domains of Mu 91 7.5 Regression Analysis Further, mixes were ranked using intercept coefficient ‘a’ in descending order, which is a representative of permanent strain at N=1, on a log-log scale, obtained from the base power model as explained earlier in section 2.9. The results of Intercept coefficient for uniaxial load strain test and WT have been reported in Table 6.4, 7.1, 7.2 & 7.3. It can be observed that ‘a’ increases with an increase in stress level, irrespective of aggregate gradations, bitumen types, mix types, and test temperatures. The intercept coefficient ‘a’ of mixes under all temperature and stress conditions in creep and WT test ranges from 4.73 - 5.48 and 4.34 - 5.35 respectively. Results of these Tables have been shown in Figure 7.5, which clearly shows that ‘a’ varies in a narrow range of 4.34 to 5.48 for both the tests. 40C,100kpa Comparison of WT & URLST 40C,300kpa 8 1a Intercept Coefficient (a) 7 1b 1c 2a 2b 40C,500kpa 2c 55C,100kpa 6 5 55C,300kpa 4 55,500kpa 3 40C,WT 2 55C,WT 1 25C, 100kpa 0 0 1 2 3 4 5 6 7 25C, 300kpa 25C, 500kpa Mixes(1a, 1b, 1c, 2a, 2b, 2c) Figure 7.5: Influence of test type on intercept coefficient Abscissa in Figure 7.5 shows six mixes (1a, 1b, 1c, 2a, 2b, 2c) ascending order. 92 25C, WT 7.6 Ranking of Mixes using Intercept Coefficient Further, mixes were ranked based on ‘a’ value to observe the best possible options of mix performance under given temperatures for both the tests and tabulated in Table 7.13 and 7.14. Table 7.13: Ranking of Mixes for Wheel Tracker Sr. Temperature No. (0C) 1st 2nd 3rd 4th 5th 6th 1 25 1c 1a 1b 2c 2b 2a 2 40 1c 1a 1b 2a 2c 2b 3 55 1c 1a 2c 2a 1b 2b Ranking of Mixes based on Intercept Coefficient “a” Table 7.13 shows that increase in temperature from 40-550C has affected only the position of mix ‘2c’ and ‘1b’ from rank 5 to 3 respectively. The reasons may be that 40/50 pen. grade is more harder grade than 60/70 pen. grade and it showed lower intercept value at 550C. Intercept coefficients of mixes with coarser gradation (1a, 1b & 1c) have lower value than finer mixes (2a, 2b &2c) in wheel tracker test. Table 7.14: Ranking of mixes for URLST Sr. Temperature Stress No. (0C) (kPa) 1st 2nd 3rd 4th 5th 6th 1 25 100 2a 1a 1b 1c 2b 2c 2 25 300 2a 1a 1b 1c 2b 2c 3 25 500 2a 1a 2b 2c 1b 1c 4 40 100 1b 1a 2c 1c 2a 2b 5 40 300 1b 2c 1a 2a 1a 1c 6 40 500 1b 2c 1c 1c 2b 2a 7 55 100 1a 2a 1c 1b 2c 2b 8 55 300 1a 2a 1c 2c 1b 2b 9 55 500 1a 2a 1c 2c 1b 2b Ranking of Mixes 93 It is difficult to conclude mixes ranks using intercept coefficient in the repeated creep test. However, one can draw a conclusion from Table 7.13 & 7.14 that mixes with PMA can be ranked at high temperature (550C). Increase in stress levels has relatively minor effects on permanent strains and hence ‘a’ value. However, significant influence of temperature on the regression constant has been observed. Although, confinement of slab on WT may has some more stiffness effects on the mix, which ultimately reduces the rate of rut depth. 7.7 Shift Factor Computations Permanent strain obtained from repeated creep test can be converted to rut depth using layer strain method expressed as; [ RutDepth = ∑ ε p (i )× hi N i =1 ] (7.1) Where i remain as one, N is the total number of load repetitions; εp is the permanent strain and hi, the thickness of Marshall Specimens (63mm). Rut depth obtained from the above method, was plotted on log-log scale after multiplying εp with one million in order to obtain straight line trends. Data obtained from both the tests were plotted graphically in Figure 7.5, and shift factors were determined. w t-A Wheel Tracker versus Uniaxial Repeated Load Test w t-B w t-C A-500kpa B-500kpa 9 C-500kpa D-500kpa 8 Log (Rut Depthx10) (mm) E-500kpa 7 F-500kpa A-300kpa 6 6 B-300kpa 5 C-300kpa D-300kpa 4 E-300kpa F-300kpa 3 Creep Test Data 2 A-100kpa WT Test Data B-100kpa C-100kpa 1 D-100kpa - E-100kpa 1 10 100 1000 Load Cycles LOG (N) Figure 7.6: Shift of URLST Data to Wheel Tracker Data 94 10000 100000 F-100kpa Figure 7.6 shows that plots of both the tests don’t have the same trend lines, but to compare repeated creep test with WT test an average shift factor of 0.48 have been used. 7.8 Correlations between URLST and WT Test A comparison of rut on log-log scale between both types of test data has been made to develop relationships at each temperature and stress levels as shown in Figure 7.6. Rut data of URLST have been taken on coordinate and that of WT test on abscissa. It can be observed from Figure 7.6 through 7.15 that the slope of plots gradually reduces by increasing stress levels and temperatures. At the same time, increase in temperature and stress levels has also reduces the range of data variation. Wheel tracker specimens being confined in the test mould showed less rate of increase of permanent strain (rut depth) than unconfined repeated creep test. One can draw a conclusion from the preceding discussion that rutting can be prediction from any of the method and can be compared reasonably with one another. 0 Comparison at 25 C & 100 kPa 8.50 WT (Rut Depth)mm 8.00 7.50 7.00 y1a = 85.785x2 - 397.21x + 465.17 R2 = 0.976 y2a = 15.825x2 - 69.623x + 82.453 R2 = 0.99 y1b = 125.29x2 - 585.55x + 689.66 R2 = 0.996 6.50 y2b = 81.2x2 - 380.72x + 452.4 R2 = 0.992 6.00 5.50 5.00 4.50 4.00 2.30 y2c = 272.96x2 - 1295.1x + 1541.8 R2 = 0.996 y1c = 280.42x2 - 1335.1x + 1594.8 R2 = 0.963 2.32 2.34 2.36 2.38 2.40 UTM (Rut De pth)mm 2.42 2.44 2.46 1a 1b 1c 2a 2b 2c Poly. (1c) Poly. (2c) Poly. (1b) Poly. (1a) Poly. (2b) Poly. (2a) Figure 7.7: Correlations between URLST and WT Test Data at 250C & 100kPa 95 0 Comparison at 25 C & 300 kPa 8.50 y1a = 298.39x2 - 1418.6x + 1691.4 R2 = 0.9792 WT (Rut Depth)mm 8.00 y2a = 39.932x2 - 184.87x + 219.99 R2 = 0.9977 7.50 7.00 6.50 y1b = 331.11x2 - 1577.9x + 1885.3 R2 = 0.9936 y2b = 171.4x2 - 816.93x + 979.53 R2 = 0.9896 6.00 5.50 5.00 4.50 y2c = 703.71x2 - 3390x + 4088.3 R2 = 0.9981 y 1c= 443.38x2 - 2127.4x + 2557.4 R2 = 0.9205 4.00 2.30 2.35 2.40 2.45 2.50 UTM (Rut Depth)mm 1a Poly. (2c) 1b Poly. (1c) 1c Poly. (1b) 2a Poly. (1a) 2b Poly. (2b) 2c Poly. (2a) Figure 7.8: Correlations between URLST and WT Test Data at 250C & 300kPa 0 Comparison at 25 C & 500 kPa 8.50 8.00 y2a = 62.036x2 - 293.36x + 352.84 R2 = 0.9963 y1a = 431.89x2 - 2081.2x + 2512.7 R2 = 0.9707 y2b = 267.36x2 - 1290.6x + 1563.7 R2 = 0.9941 WT (Rut Depth)mm 7.50 7.00 6.50 y1b = 659.79x2 - 3189.4x + 3859.9 R2 = 0.9965 6.00 5.50 5.00 4.50 y2c = 362.89x2 - 1747.4x + 2109 R2 = 0.9984 y1c = 8017x2 - 39289x + 48141 R2 = 0.9663 4.00 2.30 2.35 2.40 2.45 2.50 UTM (Rut Depth)mm 1a Poly. (1c) 1b Poly. (1a) 1c Poly. (2c) 2a Poly. (1b) Series5 Poly. (2b) 2b Poly. (2a) Figure 7.9: Correlations between URLST and WT Test Data at 250C & 500kPa 96 2c 0 Comparison at 40 C & 100 kPa 8.50 y2a = 0.7355x2 - 6.4169x + 19.854 R2 = 0.9905 y1a = 0.6168x2 - 5.0671x + 16.073 R2 = 0.9977 8.00 WT (Rut depth)mm 7.50 y1b = 0.57x2 - 4.7054x + 15.449 R2 = 0.9979 7.00 6.50 y2b = 0.6085x2 - 4.9251x + 15.582 R2 = 0.9973 6.00 5.50 5.00 y2c = -0.2671x2 + 3.5816x - 4.4757 R2 = 0.9986 y1c = 1.1523x2 - 11.566x + 34.706 R2 = 0.9988 4.50 4.00 4.00 4.50 5.00 5.50 6.00 6.50 UTM-5P (Rut De pth) mm 1a Poly. (2c) 1b Poly. (2b) 1c Poly. (2a) 2a Poly. (1b) 2b Poly. (1a) 2c Poly. (1c) Figure 7.10: Correlations between URLST and WT Test Data at 400C & 100kPa 0 Comparison at 40 C & 300 kPa 8.50 8.00 y1a = 1.1835x2 - 11.121x + 31.842 R2 = 0.9982 y2a = 0.6856x2 - 5.5859x + 16.72 R2 = 0.989 WT (Rut depth) mm 7.50 7.00 6.50 6.00 5.50 5.00 4.50 y2b = 0.5542x2 - 4.6522x + 15.612 R2 = 0.9949 y1b = 0.466x2 - 3.7412x + 13.028 R2 = 0.9982 y2c = 0.2145x2 - 1.0869x + 6.2694 R2 = 0.9977 y1c = 0.4675x2 - 4.1183x + 14.55 R2 = 0.9978 4.00 4.50 5.00 5.50 6.00 6.50 UTM-5P (Rut Depth) mm 1a 1b 1c 2a 2b Poly. (2c) Poly. (1a) Poly. (2a) Poly. (1b) Poly. (2b) 2c Figure 7.11: Correlations between URLST and WT Test Data at 400C & 300kPa 97 Poly. (1c) 0 Comparison at 40 C & 500 kPa 8.50 WT (Rut Depth)mm 8.00 7.50 y2a = 1.2742x2 - 12.696x + 37.66 R2 = 0.9943 y1a = 0.7785x2 - 6.9721x + 20.988 R2 = 0.9978 7.00 6.50 6.00 5.50 5.00 4.50 4.00 5.00 y2b = 0.0168x2 + 1.5601x - 2.6272 R2 = 0.9903 y1b = 0.6082x2 - 5.2359x + 16.726 R2 = 0.9974 y2c = -0.0175x2 + 1.6625x - 1.9362 R2 = 0.9963 y1c = 0.7005x2 - 6.2399x + 19.11 R2 = 0.9952 5.20 5.40 5.60 5.80 6.00 6.20 UTM-5P (Rut Depth)mm 1a Poly. (2c) 1b Poly. (2b) 1c Poly. (2a) 2a Poly. (1b) 2b Poly. (1a) 2c Poly. (1c) Figure 7.12: Correlations between URLST and WT Test Data at 400C & 500kPa 0 Comparison at 55 C & 100 kPa 8.50 WT (Rut Depth)mm 8.00 7.50 7.00 y2a = 0.5813x2 - 4.9708x + 16.739 R2 = 0.9897 y1a = 0.6854x2 - 5.7827x + 18.043 R2 = 0.9975 y1b = 0.3647x2 - 2.4325x + 9.3175 R2 = 0.9937 y2b = -0.1539x2 + 3.4111x - 6.9468 R2 = 0.9926 6.50 6.00 5.50 5.00 4.50 4.00 5.00 1a Poly. (1c) y2c = -1.6635x2 + 19.995x - 52.588 R2 = 0.9796 y1c = 2.0859x2 - 20.177x + 54.608 R2 = 0.9893 5.20 1b Poly. (2c) 5.40 5.60 UTM (Rut Depth)mm 1c Poly. (2b) 2a Poly. (2a) 5.80 6.00 2b Poly. (1a) Figure 7.13: Correlations between URLST and WT Test Data at 550C & 100 kPa 98 2c Poly. (1b) 0 Comparison at 55 C & 300 kPa 8.50 WT (Rut Depth)mm 8.00 y2a = 0.3999x2 - 3.224x + 12.556 R2 = 0.9931 y1a = 0.3732x2 - 2.939x + 11.467 R2 = 0.9969 7.50 7.00 6.50 y2b = -0.5394x2 + 7.5384x - 18.263 R2 = 0.993 6.00 y1b = -0.0623x2 + 1.7455x - 0.7971 R2 = 0.989 5.50 5.00 y2c = -0.6765x2 + 8.6389x - 20.027 R2 = 0.9802 y1c = 0.5508x2 - 4.5584x + 15.06 R2 = 0.986 4.50 4.00 5.00 5.20 1a Poly. (2c) 5.40 1b Poly. (2a) 5.60 5.80 UTM (Rut Depth)mm 1c Poly. (2b) 2a Poly. (1b) 6.00 2b Poly. (1c) 6.20 2c Poly. (1a) Figure 7.14: Correlations between URLST and WT Test Data at 550C & 300 kPa 0 Comparison at 55 C & 500 kPa 8.00 WT (Rut Depth)mm 7.50 y2a = 0.794x2 - 6.4726x + 17.549 R2 = 0.9884 y1a = 0.4169x2 - 3.4149x + 12.603 R2 = 0.9955 7.00 6.50 y2b = 0.4236x2 - 3.7442x + 14.44 R2 = 0.9981 6.00 5.50 y1b = -3.8074x2 + 46.259x - 133.23 R2 = 0.9931 5.00 4.50 4.00 5.20 y2c = -5.9797x2 + 71.54x - 206.66 R2 = 0.9878 y1c = 0.6175x2 - 5.3107x + 16.976 R2 = 0.9827 5.40 5.60 5.80 6.00 6.20 6.40 UTM (Rut Depth)mm 1a Poly. (2c) 1b Poly. (1b) 1c Poly. (2b) 2a Poly. (1c) 2b Poly. (1a) Figure 7.15: Correlations between URLST and WT Test Data at 550C & 500 kPa 99 2c Poly. (2a) 7.9 Summary Straight line trends of URLST data files were plotted between Cycles (N) and Rut depth, using log techniques, for the determinations of regression constants i.e. ‘intercept coefficients’ and ‘slope coefficients’. Intercept coefficient ‘a’ increases with an increase in stress levels, but slope coefficient ‘b’ decreases under the same conditions. Within the same class of aggregate grading (NHA Class “A”), it is difficult to study the influence on slope coefficient. Ranges of intercept coefficient “a” and slope coefficient “b” for all the six mixes were observed from 3.46 to 5.38 and 0.16 to 0.30 respectively. Mixes initially designed for URLST test were also tested under wheel tracker using the same three temperatures and measurements of rut depth against number of load cycles were plotted. Mixes with PMA & 40/50 penetration grade bitumen have almost same rut potential at lower temperature (250C).PMA performed better than that of neat binders at higher temperatures (550C). Intercept coefficient (a) and slope coefficient (b) were also computed using the log-log scale on the test data files. It was observed that with increase in temperature, magnitude of “a” increase, indicating rut susceptibility of mixes. Similarly, alpha (α) increases with the increase in stress levels, but mu (µ) decreases under same conditions. Effect of temperature observed on value of alpha was minor, while mu varies significantly. Domains of alpha and mu observed for low asphalt content mixes are 0.80 to 0.92 and 3.32 to 21.57 respectively. Results of both the test methods as described in the previous chapters were further correlated using shift factor, ranking of mixes and correlation developments. It was observed that rutting can be predicted from any of the method and can be compared reasonably with one another. Uniaxial load strain (creep) test did not provide a clear indication of mixes ranks than wheel tracker test. Shift factor is useful number in order to compare the results and plots of repeated creep tests and it can be shifted to that of WT test with an approximate average of 0.48. However, it requires further investigation for precision. The plots drawn between both the tests, clearly indicates a reduction in the slope of line with the increase in temperature and stress levels. 100 Chapter Eight 101 Chapter Eight Modeling the Permanent Deformation Properties of Hot Mix Asphalt 8.1 Introduction This chapter mainly focused on the statistical estimation of Mechanistic-Empirical Model that relates plastic to elastic strain ratio with different variable that influenced in the mix rut development. Using 54 variables, comprising six asphalt mixes, three temperatures (250C, 400C and 550C) and three stress levels (100, 300 and 500 kPa) under repeated load test, mathematical model have been developed to assess the magnitude of plastic to elastic strain ratio. Despite certain limitations, critical parameters have been captured in the model that helps estimating the permanent deformation in the asphalt layers of flexible pavements. This chapter provides an alternative test methodology with similar accuracy of AASHTO Mechanistic Empirical Pavement Design Guide (MEPDG) model for the prediction of permanent deformation. The current MEPDG incorporates a power model for generating rutting predictions for asphalt concrete. Rutting model developed from laboratory uniaxial repeated load strain tests as provided in MEPDG in the following form has been used as basis to estimate the relationships between the predictor variables and the permanent deformation parameters (Stephen et. al. 2007) : εp = a1T a N a εr 2 3 (8.1) Where, εp, εr, are the plastic and elastic strains respectively, at N repetitions of load and ai are the non linear regression coefficient. 102 8.2 Data Analysis and Model Development 8.2.1 General Trends of Variables Graphical study of characteristics of the data sample help to observed general trends [ McCuen, 1985]. The analysis of data initiates with sorting individual effects of stresses, temperatures and mix types on the plastic and elastic properties. For example, the effects of temperatures and stress levels are evaluated by calculating the mean values of elastic and plastic parameters of asphalt mixes as given in Table 5.2 & 5.3. The effects of binder type can be seen in the form of mix type in alphabetical order of mix designations i.e. ‘a’, ‘b’ and ‘c’. The general trends of ‘εp,’& ‘εr’ have been shown graphically in Figure 8.1. On the abscissa are the stress levels and on the ordinate are the plastic & elastic strain components. One can observe the effects of different variables on plastic and elastic strains. Plots for each temperature (250C, 400C, 550C) have been drawn. Figure 8.1 shows that accumulative plastic strain or the permanent strain increases with increase in stress as well as the temperature each asphalt mix. Figure 8.2 shows an increase in resilient strain with increases in the temperature. 103 1.50 0 55 C 0 40 C 1.00 0 25 C 0.50 Mixture with Gradation "02" & PMA (2a) 2.00 0 55 C 1.50 0 40 C 1.00 0 25 C 0.50 0.00 0 0.00 0 100 200 300 400 500 55 Deg. C 0 40 C 1.00 0 25 C 0.50 Mixture with Gradation "02" & 60/70 Pen. Grade (2b) 2.00 0 55 C 0.00 0 55 C 1.50 0 40 C 1.00 0.50 0 25 C 0.00 0 100 200 300 400 500 600 0 100 Stress (kPa) 25Deg. C 40 Deg. C Mixture with Gradation "01" & 40/50 Pen. Grade (1c) 1.50 0 55 C 0 40 C 1.00 0 25 C 0.50 200 300 400 Stress (kPa) 25 Deg. C 55 Deg. C 0.00 500 600 40 Deg. C 55 Deg. C Mixture with Gradation "02" & 40/50 Pen. Grade (2c) 2.00 Accumulated Plastic Strain (εp) % Accumulated Plastic Strain (εp) % 40 Deg. C 55Deg. C 1.50 2.00 600 600 Mixture with Gradation "01" & 60/70 Pen. Grade (1b) 2.00 200 400 Stress (kPa) 25 Deg. C Stress (kPa) 25 Deg.C 40 Deg.C Accumulated Plastic Strain (εp ) % Accumulated Plastic Strain (εp) (%) 2.00 Accumulated Plastic Strain (εp) % Accumulated Plastic Strain (εp) (%) Mixture with Gradation "01" & PMA (1a) 0 55 C 1.50 0 40 C 1.00 0 25 C 0.50 0.00 0 100 200 300 400 Stress (kPa) 25 Deg. C 500 40 Deg. C 600 0 55 Deg. C 100 200 300 400 Stress (kPa) 25 Deg. C Figure 8.1: General trends of Plastic Strains 104 40 Deg. C 500 600 55 Deg. C Mixture with Gradation "01" & PMA (1a) 0 0.14 55 C 0.10 40 C 0 0 25 C 0.06 0.02 25 Deg. C 400 600 0 40 C 0.1 0 0.06 25 C 55 Deg. C 40 Deg. C 0 0.18 Mixture with Gradation "01" & 60/70 Pen. Grade (1b) 0.14 55 C 0.10 40 C 0 0 0 25 C 0.06 200 0.02 400 25 Deg. C Stress (kPa) Resilient Strain( %) 200 Stress (kPa) Resilient Strain( %) 0 55 C 0.14 0.02 0 600 55 Deg. C 40 Deg. C 0.18 Mixture with Gradation "02" & 60/70 Pen. Grade (2b) 0 55 C 0.14 40 C 0.1 25 C 0 0 0.06 0.02 200 Stress (kPa) 25 Deg. C 400 55 Deg. C 600 40 Deg. C 0.18 Mixture with Gradation "01" & 40/50 Pen. Grade (1c) 0.14 0 55 C 0.10 0 40 C 0 0.06 0 25 C 0.02 100 Stress (kPa) 0.18 Resilient Strain( % ) 0 Resilient Strain( %) Mixture with Gradation "02" & PMA (2a) 0.18 Resilient Strain( %) Resilient Strain( %) 0.18 200 300 400 500 600 25 Deg. C 55 Deg. C 40 Deg. C Mixture with Gradation "02" & 40/50 Pen. Grade (2c) 0 55 C 0.14 0 40 C 0.10 0 25 C 0.06 0.02 0 Stress (kPa) 100 200 300 400 500 600 25 Deg. C 55 Deg. C 40 Deg. C 105 0 Stress (kPa) 100 200 300 25 Deg. C 400 500 55 Deg. C 600 40 Deg. C Figure 8.2: General Trends of Elastic Strains To conclude the influence of data trends as shown in Fig. 8.1 and 8.2, Table 8.1 provides a qualitative view of the effects of test conditions on the elastic and plastic strains. Table 8.1: Influence of Test Conditions and Mix Parameters on Strain Predictor Variables Criterion Variables Plastic Strain (εp) Elastic Strain (εr) Temperature Very Strong Strong Asphalt Type (G*) Strong Strong Stress Strong to Moderate Very Strong Aggregate Type (Gradation) Low Moderate As shown in Table 8.1, the elastic and plastic stains are highly dependent on the temperature, and stress conditions and moderately dependent on the mean shear complex modulus of asphalt cement. Very little effect has been observed with the change in the gradation. Also, plastic strain component is more sensitive as compared to elastic strain, especially at 400C and 550C. 8.2.2 Development of Prediction Model The MEPDG power model was initially adopted in this study to observe the agreement between the predicted and the measured plastic to elastic ratio under each test condition. The simplest technique is to combine all the variables and perform ordinary least-square regression on the entire data set. Predictor variables as given in Table 6 were then incorporated in step wise step iteration process using the Solver tool. The Microsoft Excel Solver (MES) can find an optimal value of coefficients for a formula using “what-if analysis” tools. Non liner regression coefficients (a1, a2, a3, a4) were then obtained at least-square of difference between the predicted and measured ‘εp/εr’ ratio. The equation that provided the minimum least square and at the same time highest coefficient of determination (R2) was selected as final model. The plots of all predictor variables versus measured variables using the same model at 54 test conditions and 162 data files have been shown in Figure 8.3 106 Agreement of Model with Measured data Predicted εp/εr 20.0 15.0 10.0 5.0 nonlinear least square 0.0 0.0 5.0 10.0 15.0 Measured εp/εr 20.0 Figure 8.3: Measured versus Predicted εp/εr, 106 in/in The plot in Figure 8.3 shows a variation in the trends, especially at higher numbers, but there is no bias observed in the model. One observes from Figure 8.3 that the procedure adopted for mathematical development of model, yields a reasonable agreement between the predicted and measured values of strain ratio (R2 = 0.573). The mathematical form of the model is presented here; εp σ a Ta Na = εr G* 1 2 3 a4 Where: εp = Plastic Strain εr = Elastic Strain σ = Stress Levels (kPa), T = Temperature (0C) N * = Load Repetitions G = Mean Complex Shear Modulus of Asphalt Cement and, a1, a2, a3, a4 = Non linear regression coefficient. 107 (8.2) The above model can be used to determine the permanent deformation of mixes under repeated loading in terms of plastic to elastic strain ratio. Taking the logarithm of above model and placing values of non liner regression coefficients, the above model can be shown as; ⎛ε ⎞ log⎜⎜ p ⎟⎟ = 0.018 log σ + 0.121log T + 0.277 log N − 0.031log G * ⎝ εr ⎠ (8.3) The above model is an implicit form to predict permanent deformation as a function of stress, temperature, load cycle and shear modulus of binder. 8.2.3 Comparison of Computed Model with MEPDG Model Computed model was then compared with the MEPDG model using test data to check the relevant accuracy. The trends plotted between both the models have been shown graphically in Figure 8.4, which produced an excellent agreement (R2 =0.99). This analysis enhanced the confidence in creating the alternatives for estimating the rutting predictions. MEPDG Model Versus Computed Model 16 R2 = 0.9935 14 Computed Model 12 10 8 6 4 2 0 0 5 10 MEPDG MODEL Figure 8.4: MEPDG Versus Computed Model εp/εr, 106 in/in 108 15 8.2.4 Sensitivity Analysis Following the model development, a sensitivity analysis was conducted to ascertain the comparative importance of predictor variables in the recommended model. Figure 8.5 provides a summary of the relative sensitivity of εp/εr model to predictor variables and a picture that temperature is the most sensitive parameter, followed by stress levels and complex shear modulus of asphalt (descending order). However, regression constant against number of load repetition has been observed higher than others, which indicates its relative importance in any of the test condition. In summary, the above research work presents an assessment of influence of the mix parameters and test conditions on the permanent deformation characteristics of asphalt concrete in terms of εp/εr. Based on the comprehensive laboratory study on six mixes, it was observed that plastic to elastic strain ratio is a function of number of load repetitions, temperature, stress levels, shear complex modulus of asphalt cement and aggregate gradations respectively. Also, the sensitivity analysis indicates that the temperature has influential effects on permanent deformation than other variables as considered in this experimental study. Plastic to elastic strain can be computed mathematically, using the above model with a relatively good reliability. The proposed model has a best compatibility with the MEPDG model. 109 Mixture with Gradation "01" & PMA (1a) Temperature (T) Mixture with Gradation "02" & PMA (2a) Stress (σ) Temperature (T) Stress (σ) Percentage change in strain ratio Percentage change in strain ratio 260 260 160 60 -40 0 100 200 300 400 500 -140 Percentage Change in temp & stress 60 -40 0 Stress (σ) T(1a,100kPa) σ (1a,40C) σ (1a,55C) T (1a,300kPa) T (1a,500kPa) Temperature (T) Stress (σ) 160 60 -40 0 100 200 300 400 500 60 -40 0 100 200 300 400 500 -140 -140 Percentage Change in temp & stress Percentage Change in temp & stress -240 -240 σ(1a,25C) T(1a,100kPa) σ (1a,40C) σ(1a,55C) T(1a,300kPa) T(1a,500kPa) σ (1a,25C) T(1a,100kPa) σ(1a,40C) σ (1a,55C) T(1a,300kPa) T (1a,500kPa) Mixture with Gradation "02" & 40/50 Pen. Grade (2c) Percentage change in strain ratio Mixture with Gradation "01" & 40/50 Pen. Grade (1c) Temperature (T) Stress (σ) 260 Percentage change in strain ratio 500 σ (1a,25C) 260 160 160 60 -40 400 Mixture with Gradation "02" & 60/70 Pen. Grade (2b) Percentage change in strain ratio Percentage change in strain ratio Temperature (T) 300 Percentage Change in temp & stress Mixture with Gradation "01" & 60/70 Pen. Grade (1b) 260 200 -140 σ(1a,40C) T(1a,500kPa) T (1a,100kPa) T (1a,300kPa) 100 -240 -240 σ (1a,25C) σ (1a,55C) 160 0 100 200 ` 300 400 500 -140 -240 Percentage Change in temp & stress 260 Temperature (T) Stress (σ) 160 60 -40 0 100 200 300 400 500 -140 -240 Percentage Change in temp & stress σ(1a,25C) T(1a,100kPa) σ (1a,40C) σ (1a,25C) T(1a,100kPa) σ (1a,40C) σ (1a,55C) T(1a,300kPa) T(1a,500kPa) σ(1a,55C) T(1a,300kPa) Temp. (1a,500) Figure 8.5: Sensitivity of εp/εr Model to Predictor Variables 110 Chapter Nine 111 Chapter Nine CONCLUSIONS AND RECOMMENDATIONS 9.1 Introduction This chapter presents the findings of study carried out on six asphalt mixes by using two test methods. The main objectives of the research work were to assess the influence of the mix parameters and test conditions on the permanent deformation characteristics of asphalt concrete. Two test procedures were conducted on six mixes in this research study, comprising two aggregate gradations and three binders. The effects of variables on the mix performance were studied using regression constants (intercept and slope coefficients), permanent deformation coefficients (alpha and mu) and plastic to elastic strain ratios (εp/εr). Relationship has also been developed between two commonly known tests (uniaxial creep and wheel tracker test), to predict resistance of permanent deformation or rutting of mixes at elevated temperatures. Relationship between HMA properties to its permanent deformation behaviour in the form of a model has been proposed as the key findings of the study. 9.1 Conclusions Based on the literature review, testing and analysis of test results, and modeling efforts under taken in this research work, following conclusions have been drawn; • Plastic to elastic strain ratio is a function of number of load repetitions, temperature, stress levels, shear complex modulus of asphalt cement and aggregate gradations. However, AASHTO Mechanistic Empirical Pavement Design Guide (MEPDG) model which has only two parameters (load repetitions and temperature) results in a similar accuracy with different sensitivities of the independent variables on permanent deformation prediction. 112 • Repeated creep testing showed equal potential of predicting permanent deformation compared to repeated load testing. Haversine repeated load testing carried out for the MEPDG requires more complicated system as compared to running simple repeated creep testing. The proposed model provides an alternative test methodology with similar accuracy of MEPDG model for the prediction of permanent deformation. MEPDG model works for the mixes used in this study. • Mixes with coarser gradation offered more resistance to rutting than mixes with finer gradations under uniaxial repeated load strain test. • Intercept coefficient ‘a’ increases with an increase in stress levels, but slope coefficient ‘b’ decreases under the same conditions. Ranges of ‘a’ & ‘b’ for mixes in study were observed from 3.46 -5.38 & 0.16-0.3 respectively. Permanent Deformation Coefficient alpha ‘α’ has similar trends as ‘a’ and mu ‘µ’ as ‘b’. Domains of α & µ for mixes in study were 0.8-0.92 & 3.32-21.57 respectively. • Master curve can be plotted between Uniaxial Repeated Load Strain Test & Wheel Tracker test by an average shift factor of 0.48. 9.3 Recommendations for Future Study • Comprehensive experimental study may be conducted on the moisture sensitivity as well as on the fatigue limits of the six mixes and the aging effects of binder on mixes behaviour. Similarly, uses of hydraulic cement, lime, crumb rubber and other locally available materials to enhance the performance of mixes can be used to accommodate their effects in a complete model development. At the same time, full scale accelerated testing should be carried out to investigate the performance of these mixes in the field • under local environmental conditions. Limits of permanent deformation in terms of plastic to elastic strain ratio should be specified under the current flexible pavement design, material properties, axle loading and temperature conditions of Pakistan. 113 References 114 REFERENCES Ahlrich, R.C. 1996. Influence of Aggregate Properties on Performance of Heavy-Duty HotMix Asphalt Pavements. Transportation Research Record No. 1547, pp. 7-14. Ahmad, H. Alijassar, Sayed Metwali, Mohammad A.Ali, 2004. “Effect of Filler Types on Marshall Stability and Retained Strength of Asphalt Concrete” International Journal of Pavement Engineering, 5:1 pp 47-51. Ait-kadi A., B. Brahimi, M. Bousmina, 1996. Polymer Engineering Sciences, 36 (12), 17241733. Akmal N., and Usmani A. M., 1999. “Polymers”, News, 24, 136-140. American Association of State Highway and Transportation Officials (AASHTO). 2004. Determining the Soundness of Aggregates by Sodium Sulphate or Magnesium Sulphate, AASHTO Designation T104. American Association of State Highway and Transportation Officials (AASHTO). 2004. “Determining the Percentage Deleterious Materials present in Aggregates Blend”, AASHTO Designation T112. American Association of State Highway and Transportation Officials (AASHTO). 2004. Plastic Fines in Graded Aggregate and Soils by Use of the Sand Equivalent Test, AASHTO Designation T176. American Association of State Highway and Transportation Officials (AASHTO). 2004. Elastic Recovery Test of Bituminous Materials by Means of a Ductilometer, AASHTO Designation T 301. American Association of State Highway and Transportation Officials (AASHTO). 2004. Uncompacted Void Content of Fine Aggregate, AASHTO Designation T304. 115 American Association of State Highway and Transportation Officials (AASHTO). 2004. Determining the Rheological Properties of Asphalt Binder Using a Dynamic Shear Rheometer (DSR), AASHTO Designation T315. American Association of State Highway and Transportation Officials (AASHTO). 2002. Standard Test Method for Density of Semi-Solid Bituminous Materials, AASHTO Designation T228. American Association of State Highway and Transportation Officials (AASHTO). 2002. “Design of New and Reconstructed Flexible Pavements” Chapter 3, Design Guide. American Association of State Highway and Transportation Officials (AASHTO). 2000. Softening Point of Bitumen (Ring and Ball Apparatus), AASHTO Designation T53. American Society for Testing and Materials (ASTM). 2004. Standard Test Method for Penetration of Bituminous Materials. ASTM Designation D5. American Society for Testing and Materials (ASTM). 2004. Test Method for Theoretical Maximum Specific Gravity and Density of Bituminous Paving Mixtures. ASTM Designation D2041 American Society for Testing and Materials (ASTM). 2004. Test Method for determining the Percentage of Plastic Fines in Graded Aggregate and Soils by Use of the Sand Equivalent Test. ASTM Designation D2419. American Society for Testing and Materials (ASTM). 2004. Flat or Elongated Particles in Coarse Aggregates, ASTM Designation D4791. American Society for Testing and Materials (ASTM). 2004. Test Method for determining the Percentage of Fractured Faces in Coarse Aggregate. ASTM Designation D5821 American Society for Testing and Materials (ASTM). 2004. Test Method for Indirect Tension Test for Resilient Modulus of Bituminous Mixtures. ASTM Designation D4123 American Society for Testing and Materials (ASTM). 2004. Standard Test Method for Soundness of Aggregates by Use of Sodium Sulfate or Magnesium Sulfate. ASTM Designation C88 116 American Society for Testing and Materials (ASTM). 2004. Standard Test Method for Resistance to Degradation of Small-Size Coarse Aggregate by Abrasion and Impact in the Los Angeles Machine. ASTM Designation C131 American Society for Testing and Materials (ASTM). 2004. Standard Test Method for Ductility of Bituminous Materials. ASTM Designation D113-99 Anderson, R. M., McGennis, R.B., Kennedy, T. W. and Solaimanian, M. 1995. Background of Superpave Asphalt Mixture Design & Analysis. Publication No. FHWA-SA-95-003. National Asphalt Training Center Demonstration Project 10. Asphalt Institute. 1996. “Superpave TM Mix Design” Superpave Series SP-2, Lexington, Kentucky, USA. Asphalt Institute, 2003. “Mix Design Method for Asphalt Concrete and other Hot-Mix Types,” Manual series No. 02, Six Edition, Kentucky, USA. Asphalt Institute, 2003. “Asphalt Hand Book,” Manual series No. 04, Six Edition, Kentucky, USA. Asphalt Institute. 2003. “Superpave Mix Design” Superpave Series SP-2, Third Edition, 2001.Printing. Revised 2003, Lexington, Kentucky, USA. Australian Roads Standards, 2006. “Torsional Recovery of Polymer Modified Binders” AUSTROAD Designation T-122. Awad, I.S., 1972. Characterisation of the Stress-Strain Relationships of Asphalt Treated Base Material, Graduate Report, University of Washington. Barengberg, E.J. and M.R. Thompson. 1990. Calibrated Mechanistic Structural Analysis Procedures for Pavements Phase 1 & II of NCHRP Project 1-26. National Cooperative Highway Research Program. Transportation Research Board. Washington, DC. 117 Barnes, H.A., J.F. Hulton and K. Walters, 1989. “Introduction to rheology”. Elsvier, Barking. Behzadi, G. and Yandell, W.O., 1996. “Determination of elastic and plastic subgrade soil parameters for asphalt cracking and rutting prediction.” Transp. Res. Rec. 1540, Transportation Research Board, Washington, D.C., 97-104. Brown, S. F. and Pell, P. S. 1974. Repeated Loading of Bituminous Materials. CAPSA 74. National Institute for Road Research, Republic of South Africa, Durban. Brown, S.F. and Snaith M.S., 1974. The Permanent Deformation of Dense Bitumen Macadam Subjected to Repeated Loading. Proceedings of the Association of Asphalt Paving Technologists Technical Sessions, vol. 43, Virginia, United States of America. Brown, E. R., Cross, Stephen A. 1990. A Study of In-Place Rutting of Asphalt Pavements. The Association of Asphalt Paving Technologists v. 59. Brown, E. R. 1990. Density of Asphalt Concrete – How much is needed? Transportation Research Record 1282, Transportation Research Board, Washington D. C Brown, R.R., and Bassett, C.E. 1990. Effects of Maximum Aggregate Size on Rutting Potential and Other Properties of Asphalt –Aggregate Mixtures. Transportation Research Record 1259, pp. 107-119. Brown, E.R., and Cross, S.A. 1992. A National Study of Rutting in Hot Mix Asphalt (HMA) Pavements. NCAT Report 92-5, National Centre for Asphalt Technology. Burger, A.F., Van de Van, M.F.C, J. Muller, K.J. Jenkins., “Rheology of Polymer Modified Bitumen: A Comparison Study of Three Binders and Three Binder/Filler Systems”, 20th South African Transport Conference, Meeting the Transport Challenges in Southern Africa, 16-20 July, 2001. Button, J.W., Perdomo, D., Lytton, R.L. 1990. Influence of Aggregate on Rutting in Asphalt Concrete Pavements. Transportation Research Record No. 1259, pp.141-152. 118 Carpenter, S. H., 1993. Permanent deformation: Field evaluation. Transport research record 1417 Materials and Construction, Asphalt concrete mixtures. Transportation Research Board, National Academy Press, Washington Cheung, C.Y. 1995. Mechanical behaviour of bitumen and bituminous mixes. Ph.D. Thesis, University of Cambridge. Collop, A. C., Cebon, D. and Hardy, S. A. 1995. Viscoelastic Approach to Rutting in Flexible Pavements. Journal of Transportation Engineering v. 121, Issue 1, p. 82-93. Cominsky, R., R. B. Leahy, and E. T. Harrigan. 1994. Level One Mix Design: Material Selection, Compaction, and Conditioning. Report SHRP-A-408. Strategic Highway Research Program, National Research Council. Cooper, K. E., Brown, S. F., and Pooley, G. R. 1985. The Design of Aggregate Gradings for Asphalt Base courses. The Association of Asphalt Paving Technologists v. 54. Cooper Technology, 2006. “ Wheel Tracker Small Device”, Product Catalogue, Issue 1, CRTWTEN1& CRT-WTEN2 in conformance to E13108. Crawford, C. 1989. Tender Mixes: Probable Causes, Possible Remedies. Quality Improvement Series 108-3186. National Asphalt Pavement Association, Lanham, MD. Diylajee, V.A., and Raymond, G.P., 1982. “Repetitive load deformation of cohesionless soil,” J.Geotech. Engrg. Div., ASCE, 108(10), 1215-1229. European Standard; prEN 12697-22, 2002. “Test methods for hot mix asphalt” part-22, Wheel Tracking. Findley, W. N., Lai, J. S., and Onaran, K. 1976. Creep and Relaxation of Non linear Viscoelastic Materials. North Holland Publishing, Amsterdam. 119 Ford, Miller C. 1988. Pavement Densification Related to Asphalt Mix Characteristics. Flexible Pavement Construction, Transportation Research Record 1178. Transportation Research Board, Washington D. C. Fred, N. Finn., 1967. “Factors involved in the design of Asphaltic Pavement Surfaces,” National Cooperative Highway Research Program Report 39, Highway Research Board, Washington, D.C. Garba, R., 2002. “Permanent deformation properties of asphalt concrete mixtures, PhD Thesis, Norwegian University of Science and Technology”. Gerald, Reinke; 2001. “Elvaloy Formulation with Attock Asphalt” Mathy Technology & Engineering Services, Inc; Gibb, J.M., 1996.Evaluation of resistance to permanent deformation in the design of bituminous paving mixtures. Ph.D. Thesis, University of Nottingham. Hassan, S. M., 1998. “Performance Prediction Models for Asphalt Concrete Pavements”, Treatise on Flexible Pavement Design, pp, 15-28. Herrin, M., and W. H. Goetz. 1954. Effect of Aggregate Shape on Stability of Bituminous Mixes. In Proceedings, Vol. 33, Highway Research Board. Hertzberg, R.W., 1996. Deformation and fracture mechanics of engineering materials. 4th edition, John Wiley and sons, New York. Heukelom, W. 1999. “A bitumen test data chart for showing the effect of temperature on the mechanical behaviour of asphaltic bitumen”, Journal of the institute of Petroleum, vol 55. Huang, Y. H. 1993. Pavement Analysis and Design. Prentice-Hall, Englewood Cliffs, NJ. Huang, Y.H., 1993. Pavement Analysis and Design, Prentice Hall, New Jersey, United States of America. 120 Huang, Shin-Che and Zeng Menglan, 2007. “Characterization of aging effect on rheological properties of asphalt-filler systems”, International Journal of Pavement Engineering, 8:3, 213223. Huang, H., 1995. “Analysis of Accelerated Pavement Tests and Finite element Modeling of Rutting Phenomenon”, Ph.D. Thesis, Purdue University. Huang, B., Xiang Shu, Xinwei Chen. 2007. “Effects of mineral fillers on hot-mix asphalt laboratory measured properties”, International Journal of Pavement Engineering 8:1, pp1-9. Huber, G. A. and Heiman, G. H. 1987. Effects of Asphalt Concrete Parameter on Rutting Performance: A Field Investigation. Association of Asphalt Paving Technologists v. 56. Huber, G.A., and Heiman, G.H. 1989. Effect of Asphalt Concrete Parameters on Rutting Performance: A Field Investigation. Pavement Rutting Seminar- Roads and Transportation Association of Canada. K. B. De Vose. and A.J. Feeley. 2002. “Feed back controlled repeated uniaxial loading strain test” Hardware Reference. Manual, Universal Testing Machine (UTM-5P), pp11-128, 2002. Kamal, M.A., Shahzeb F., Baber Y., 2005. “Resilient Behaviour of Asphalt Concrete under repeated loading & effects of temperature” Journal of the Eastern Asia Society for Transportation Studies, Vol. 6, pp 1329-1343. Kandhal, S. P., Mallick B. R. and Brown R. E., 1998. ”Hot Mix Asphalt For Intersections in Hot Climates”, National Center for Asphalt Technology, NCAT Report No. 98-06, Auburn, Alabama. Kandhal, S. P. and Mallick R.B., 2001. “Effects of Mix Gradation on Rutting Potential of Dense-Graded Asphalt Mixtures”, Transportation Research Board, National Research Council, Transportation Research Record No. 1767, Washington, D.C., pp. 146-151. Kanitong, K., Bahia, H., 2005. Relating adhesion and cohesion of asphalts to the effect of moisture on laboratory performance of asphalt mixtures, Transportation Research Record, 1901, 33-43. 121 Kennedy, T.W., P. S. Kandhal, S. F. Brown, D. Y. Lee, and F. L. Roberts. 1996, Hot Mix Asphalt Materials, Mixture Design and Procedure. NAPA Research and Education Foundation, Lanham, MD. Khedr, S.A., 1986. “Deformation mechanism in asphaltic concrete.” J. Transp. Engrg. , ASCE, 112(1), 29-45 Kim, Y. R., N. Yim, and N. P. Khosla. 1992. Effects of Aggregate Type and Gradation on Fatigue and Permanent Deformation of Asphalt Concrete. In American Society for Testing and Materials, STP 1147, West Conshohocken, PA, Klaus D., 2003. Asphalt flow improvers – A new technology for reducing mixing temperature of asphalt concrete mixes with high-resistance against permanent deformation, Proceedings of the 22nd PIARC World Road Congress, pp7. Krutz, N. C., and P. E. Sebaaly. 1993. The Effects of Aggregate Gradation on Permanent Deformation of Asphalt Concrete. In Proceedings, Vol. 62, Association of Asphalt Paving Technologists. Mahboub, K. and Little, D. N. 1988. Improved Asphalt Concrete Design Procedure, Research Report 474-1F. Texas Transportation Institute Majidzadeh, K, S. Khedr, and H. Guirguis 1981. Implementation of a pavement design system, volume 1 and 2 of the final report. Research project EES 579. Ohio State Univeristy. Columbus, OH. Mallick, R. B. and P. S. Kandhal.2001. Effect of Mix Gradation on Rutting Potential of Dense Graded Asphalt Mixtures. A Paper Submitted for Presentation and Publication at the 80 th Annual Meeting of the Transportation Research Board, Washington, D. C. Marks, V. J., Monroe, R. W. and Adam, J. F. 2001. Relating Creep Testing to Rutting of Asphalt Concrete Mixes. For presentation at the Transportation Research Board 7-th Annual Meeting January 13-17. Washington, D.C. 122 Matthews, J. M. and Monismith, C. L. 2003. Effects of Aggregate Gradation on the Creep Response of Asphalt Mixtures and Pavement Rutting Estimates. ASTM Special Technical Publication, No 1147, ASTM, Philadelphia, PA., p p. 329–343. McCuen, R., “Statistical Methods for Engineers,” Prentice-Hall, Inc., (1985). McGennis, R. B., R. M. Anderson, T. W. Kennedy, and M. Solaimanian. 1994, Background of Superpave Asphalt Mixture Design and Analysis. Report SA-95-003. Federal Highway Administration, U.S. Department of Transportation. Monismith, C. L., and Tayebali, A. A. 1988. Permanent Deformation (Rutting) Considerations in Asphalt Concrete Pavement Sections. The Association of Asphalt Paving Technologists v. 57, Williamsburg, Virginia. Monismith, L. C., Brown F. S., 1999. “Developments in the Structural Design and Rehabilitation of Asphalt Pavement over Three Quarters of a Century”, Asphalt Paving Technology, 75th Anniversary Historical Review and Index of Journals 1975-1999, Volume 68A. Monismith, C.L., 1976. “Rutting prediction in asphalt concrete pavements.” Transp. Res. Rec. 616, Transportation Research Board, Washington, D.C., 2-8. Molenaar, A.A.A., 1993. Road materials Part III-Asphaltic Materials, Lecture Notes, Delft University of Technology, The Netherlands. Morris, J., 1973. The Prediction of Permanent Deformation in Asphalt Concrete Pavements. Ph.D. Thesis, University of Waterloo, Ontario, Canada. National Highway Authority, NHA., 1998, “General Specifications”, Surface Courses, Item No. 305, Pakistan. National Cooperative Highway Research Program; 2005. “Aggregate properties and the performance of superpave Designed Hot Mix Asphalt”, NCHRP Report No. 539. 123 Parker, Frazier, and Brown, E. R. 1992. Effects of Aggregate Properties on Flexible Pavement Rutting in Albama. Effects of Aggregate and Mineral Fillers on Asphalt Performance. ASTM STP 1147. Philadelphia. Pellinen, T.K. and Witczak, M.W., 2002. Stress Dependent Master Curve Construction for Dynamic (Complex) Modulus. Proceedings of the Association of Asphalt Paving Technologists Technical Sessions, vol. 71, CD Version. Ponniah, J., and Kennepohl, G. 1996. Polymer-Modified Asphalt Pavements in Ontario: Performance and Cost Effectiveness. Transportation Research Record No. 1545, pp. 151-160. Prowell, B.D. 2001. Design, Construction and early Performance of Virginia’s Hot-Mix Asphalt Stabilizer and Modifier Test Sections. Transportation Research Record No. 1767, pp. 7-14. Rauhut, J. and Jordhal J. 1978. “Effects on Flexible Highways of Increased Legal Vehicles Weights Using VESYS IIM, FHWA-RD-77-116. Rita, B. Leahy and Matthew W. Witczak, 1991. The Influence of Test Conditions and Asphalt Concrete Mix Parameters on Permanent Deformation Coefficient Alpha and Mu. Journal of the Association of Asphalt Paving Technologies, vol1, pp 333-363. Robert, N. Hunter 2000. “Asphalt in roads construction” American Society of Civil Engineering, Thomas Telford Publications, pp 75, London. Shashidhar, N., Needham, S.P., Chollar, B.H. and Romero, P. 1999. “Prediction of the performance of mineral fillers in stone matrix asphalt” J. Assoc. Asphalt Paving Technologists, 68, pp 222–251. Sivasubramaniam, S., Galal K. A., Noureldin, A. S., White T. D. and Haddock J. E., 2004. “Modeling Permanent Deformation Using Laboratory, Proto-type, and In-Service Accelerated Pavement Testing”, Transportation Research Board Annual Meeting, Washington, D.C. 124 Sousa, J.B., Craus J., Monismith, C.L. 1991. Summary Report on Permanent Deformation in Asphalt Concrete. Strategic Highways Research Program IR-91-104. Stephen, P., Yusu M., and Leslie, M. McCarthy., “ Evaluation and Modeling of Repeated Load Test Data of Asphalt Concrete for Mechanistic-Empirical Pavement Design”, Journal of Materials in Civil Engineering, ASCE, pp 993-999, Vol. 19, No. 11, November 2007. Stumpf, W., Van Rooyen, G.T., 2007. Mechanical Metallurgy NMM 700, Course notes, University of Pretoria, Pretoria. Su. Z., 1996. Mineral Aggregates; Their Classifications and Properties, ESHA Research Report, RT010-96-02, Department of Research and Technology, Smid & Holland Services B.V. Sullivan, B. Development of Flow Number and Flow time candidate simple performance test for asphalt mixtures. MS Thesis. Department of Civil and Environmental Engineering, Arizona State University, Tempe 2002. Tarefder, R.A., Zaman, M., Hobson, K., 2003. A Laboratory and statistical evaluation of factors affecting rutting, International Journal of Pavement Engineering. Volume 4 Number 1, 59-68. Thomase, G. Mezger, 2002. “The Rheology Handbook” pp13. Transporation Research Board, (TRB) 2000. Hot-Mix Asphalt Paving Hand Book, pp 18, USA. Uge, P., and Van de Loo, P. J. 1974. Permanent Deformation of Asphalt Mixes, Koninklijke/Shell-Laboratorium. Amsterdam. Vuong, B., and Amstrong, P., 1991. Repeated load Triaxial testing on the Subgrade from Mulgrace ALF site, Australian Road Research Board 125 White, T., Hua J. and Galal K. 2003. “Analysis of Accelerated Pavement Tests”, Transportation Research Board Annual Meeting, Paper CS8-7, Washington, D.C. William, L. Hewitt and Floyed, O. Slate, 1967. “The Effects of the Rheological Properties of Asphalt on Strength Characteristics of Asphalt Concrete”, Proceedings of 2nd International Conference on the structural design of asphalt pavements; University of Michigan, USA pp757-768. Yeggoni, M., J. W. Button, and D. G. Zollinger. 1994. Influence of Coarse Aggregate Size, Shape, and Surface Texture on Rutting of Hot Mix Asphalt Concrete. Report FHWATW94/1244-6. Federal Highway Administration and Texas Department of Transportation. Ziari, H., Mahmud Ameri˛ and Mohammad Mahdi Khabiri; 2007. “Resilient Behaviour of Hot Mixed and Crack Sealed Asphalt Concrete under Repeated Loading” Iran Science and Technology University, ISSN 1392-8619 print/ISSN 1822-3613 online, Vol XIII, pp 56-60. 126 Annexure 127 Annexure -A Plots of accumulative strain: at 250C o Permanent Strain~ Load Repetition at 25 C 1 500kPa 0.9 0.8 εp (%) 0.7 0.6 300kPa 0.5 0.4 0.3 100kPa 0.2 0.1 0 0 500 1000 1a,100kPa Load Repetition (N) 1500 1a,300kPa 2000 1a,500kPa Figure A.1: Influence of Pulse count on accumulative strain at 250C, Mix 1a o εp (%) Permanent Strain~ Load Repetition at 25 C 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 500 kPa 300 kPa ` 100 kPa 0 500 Load Repetition (N) 1000 1b,100kPa 1500 1b,300kPa 2000 1b,500kPa Figure A.2: Influence of Pulse count on accumulative strain at 250C, Mix 1b 128 o Permanent Strain~ Load Repetition at 25 C 1 0.9 500 kPa 0.8 εp (%) 0.7 0.6 300 kPa 0.5 0.4 0.3 100 kPa 0.2 0.1 0 0 500 Load Repetition (N) 1000 1500 1c,100kPa 1c,300kPa 2000 1c,500kPa Figure A.3: Influence of Pulse count on accumulative strain at 250C, Mix 1c o Permanent Strain~ Load Repetition at 25 C 500 kPa 1 0.9 0.8 εp (%) 0.7 300 kPa 0.6 0.5 0.4 0.3 100 kPa 0.2 0.1 0 0 500 1000 2a,100kPa Load Repetition (N) 1500 2a,300kPa Figure A.4: Influence of Pulse count on accumulative strain at 250C, Mix 2a 129 2000 2a,500kPa o Permanent Strain~ Load Repetition at 25 C 500 kPa 1 0.9 0.8 εp (%) 0.7 300 kPa 0.6 0.5 0.4 0.3 100 kPa 0.2 0.1 0 0 500 1000 2b,100kPa Load Repetition (N) 1500 2000 2b,300kPa 2b,500kPa Figure A.5: Influence of Pulse count on accumulative strain at 250C, Mix 2b o εp (%) Permanent Strain~ Load Repetition at 25 C 500 kPa 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 300 kPa 100 kPa 0 Load Repetition (N) 500 1000 2c,100kPa 1500 2c,300kPa Figure A.6: Influence of Pulse count on accumulative strain at 250C, Mix 2c 130 2000 2c,500kPa Plots of accumulative strain: at 400C 0 Permanent Strain~ Load Repetition at 40 C 500 kPa 1 εp (%) 0.8 300 kPa 0.6 0.4 100 kPa 0.2 0 0 Load Repetition (N) 500 1000 1a,100kPa 1500 2000 1a,300kPa 1a,500kPa Figure A.7: Influence of Pulse count on accumulative strain at 400C, Mix 1a 0 Permanent Strain~ Load Repetition at 40 C 1 500 kPa εp (%) 0.8 300 kPa 0.6 0.4 0.2 100 kPa 0 0 Load Repetition (N) 500 1000 1b,100kPa 1500 1b,300kPa 2000 1b,500kPa Figure A.8: Influence of Pulse count on accumulative strain at 400C, Mix 1b 131 0 Permanent Strain~ Load Repetition at 40 C 1 0.9 εp (%) 0.8 0.7 500 kPa 0.6 300 kPa 0.5 0.4 0.3 0.2 100 kPa 0.1 0 0 500 1000 1c,100kPa Load Repetition (N) 1500 2000 1c,300kPa 1c,500kPa Figure A.9: Influence of Pulse count on accumulative strain at 400C, Mix 1c 0 Permanent Strain~ Load Repetition at 40 C 500 kPa 1 0.9 0.8 300 kPa εp (%) 0.7 0.6 0.5 0.4 100 kPa 0.3 0.2 0.1 0 0 Load Repetition (N) 500 1000 2a,100kPa 1500 2a,300kPa Figure A.10: Influence of Pulse count on accumulative strain at 400C, Mix 2a 132 2000 2a,500kPa 0 Permanent Strain~ Load Repetition at 40 C 500 kPa 1.4 1.2 εp (%) 1 300 kPa 0.8 0.6 0.4 100 kPa 0.2 0 0 500 Load Repetition (N) 1000 2b,100kPa 1500 2b,300kPa 2000 2b,500kPa Figure A.11: Influence of Pulse count on accumulative strain at 400C, Mix 2b 0 Permanent Strain~ Load Repetition at 40 C 1 0.9 500 kPa 0.8 εp (%) 0.7 300 kPa 0.6 0.5 100 kPa 0.4 0.3 0.2 0.1 0 0 500 1000 2c,100kPa Load Repetition (N) 1500 2c,300kPa Figure A.12: Influence of Pulse count on accumulative strain at 400C, Mix 2c 133 2000 2c,500kPa Plots of accumulative strain development: at 550C o Permanent Strain~ Load Repetition at 55 C 1.4 1.2 500 kPa εp (%) 1 0.8 300 kPa 0.6 0.4 100 kPa 0.2 0 0 500 Load Repetition (N) 1000 1a,100kPa 1500 2000 1a,300kPa 1a,500kPa Figure A.13: Influence of Pulse count on accumulative strain at 550C, Mix 1a 0 Permane nt Strain~ Load Repetition at 55 C 1.4 1.2 500 kPa εp (%) 1 300 kPa 0.8 0.6 0.4 100 kPa 0.2 0 0 500 Load Repetition (N) 1000 1b,100kPa 1500 1b,300kPa 2000 1b,500kPa Figure A.14: Influence of Pulse count on accumulative strain at 550C, Mix 1b 134 o Permanent Strain~ Load Repetition at 55 C 1.4 1.2 500 kPa εp (%) 1 0.8 300 kPa 0.6 0.4 100 kPa 0.2 0 0 500 Load Repe tition (N) 1000 1c,100kPa 1500 1c,300kPa 2000 1c,500kPa Figure A.15: Influence of Pulse count on accumulative strain at 550C, Mix 1c 0 Permanent Strain~ Load Repetition at 55 C 500 kPa 1.40 1.20 300 kPa εp (%) 1.00 0.80 0.60 100 kPa 0.40 0.20 0.00 0 Load Repetition (N) 500 1000 2a,100kPa 1500 2a,300kPa Figure A.16: Influence of Pulse count on accumulative strain at 550C, Mix 2a 135 2000 2a,500kPa 0 Permanent Strain~ Load Repetition 55 C 1.6 1.4 500 kPa 1.2 εp (%) 1 300 kPa 0.8 0.6 0.4 100 kPa 0.2 0 0 500 Load Repetition (N) 1000 2b,100kPa 1500 2000 2b,300kPa 2b,500kPa Figure A.17: Influence of Pulse count on accumulative strain at 550C, Mix 2b 0 Permanent Strain~ Load Repetition at 55 C 1.4 500 kPa 1.2 εp (%) 1 300 kPa 0.8 0.6 100 kPa 0.4 0.2 0 0 500 Load Repetition (N) 1000 2c,100kPa 1500 2c,300kPa Figure A.18: Influence of Pulse count on accumulative strain at 550C, Mix 2c 136 2000 2c,500kPa Annexure -B Trends of Accumulated Strains in Mixes: at 250C ( 250C) % Acc. Strain x10 6 10.00 y100 = 0.202Ln(x) + 4.0717 R2 = 0.9582 y300 = 0.12Ln(x) + 4.7369 R2 = 0.966 y500= 0.1018Ln(x) + 5.1317 R2 = 0.9563 1.00 1 10 100 1000 10000 Pulse Count 1a, 25, 100 1a, 25,300 1a, 25, 500 Log. (1a, 25, 100) Log. (1a, 25,300) Log. (1a, 25, 500) Figure B.1: Trends of accumulated strain at 250C, Mix 1a o (25 C) 6 % Accumulative StrainX10 10.00 y 500 = 0.0764Ln(x) + 5.2903 R2 = 0.9734 y 100 = 0.1525Ln(x) + 4.4119 R2 = 0.9769 y 300 = 0.0977Ln(x) + 4.9116 R2 = 0.9424 1.00 1 10 100 1000 10000 Pulse Counts 1b,100 1b,300 1b,500 Log. (1b,500) Figure B.2: Trends of accumulated strain at 250C, Mix 1b 137 Log. (1b,300) Log. (1b,100) 0 (25 C) % Acc. Strain x10 6 10.00 y100 = 0.1463Ln(x) + 4.5462 R2 = 0.9779 y500 = 0.073Ln(x) + 5.522 R2 = 0.9791 y300 = 0.0938Ln(x) + 4.9078 R2 = 0.9463 1.00 1 10 1c,100 1c,300 100 Pulse Counts 1c,500 1000 Log. (1c,500) Log. (1c,300) 10000 Log. (1c,100) Figure B.3: Trends of accumulated strain at 250C, Mix 1c 0 (25 C) 6 % Accumulative Strainx10 10.00 y 100 = 0.2021Ln(x) + 4.0597 R2 = 0.977 y 500 = 0.119Ln(x) + 5.083 R2 = 0.8922 y 300 = 0.1384Ln(x) + 4.6635 R2 = 0.9104 1.00 1 10 100 1000 10000 Pulse Counts 2a,100 2a,300 2a,500 Log. (2a,100) Figure B.4: Trends of accumulated strain at 250C, Mix 2a 138 Log. (2a,300) Log. (2a,500) o (25 C) 6 % Accumulative Strainx10 10.00 y100 = 0.1503Ln(x) + 4.5297 R2 = 0.9662 y500 = 0.0973Ln(x) + 5.2469 R2 = 0.9881 y300 = 0.1083Ln(x) + 4.918 R2 = 0.9562 1.00 1 10 100 1000 10000 Pulse Counts 2b,100 2b,300 2b,500 Log. (2b,100) Log. (2b,500) Log. (2b,300) Figure B.5: Trends of accumulated strain at 250C, Mix 2b o (25 C) 6 % Accu mu lativ e Strain x 1 0 10.00 y 500 = 0.0792Ln(x) + 5.247 R2 = 0.9881 y 100 = 0.0967Ln(x) + 4.8638 R2 = 0.9431 y 300 = 0.0796Ln(x) + 5.2134 R2 = 0.9794 1.00 1 10 100 1000 10000 Pulse Counts 2c,100 2c,300 2c,500 Log. (2c,100) Log. (2c,300) Figure B.6: Trends of accumulated strain at 250C, Mix 2c 139 Log. (2c,500) Trends of Accumulated Strains in Mixes: at 400C (400C) % A ccumulative S train X 10 6 10.00 y 300 = 0.0808Ln(x) + 5.1522 R2 = 0.9387 y 100 = 0.12Ln(x) + 4.7371 R2 = 0.966 y 500 = 0.0792Ln(x) + 5.247 R2 = 0.9881 1.00 1 10 100 1000 10000 Pulse Counts 1a,100 1a,300 1a,500 Log. (1a,100) Log. (1a,300) Log. (1a,500) Figure B.7: Trends of accumulated strain at 400C, Mix 1a o (40 C) 6 % Accumulative Strain x 10 10.00 y 300 = 0.1233Ln(x) + 4.8935 R2 = 0.9786 y 100 = 0.1289Ln(x) + 4.7331 R2 = 0.9494 y 500 = 0.1217Ln(x) + 5.0696 R2 = 0.9934 1.00 1 1b,100 10 1b,300 1b,500 100 Pulse Counts Log. (1b,100) 1000 Log. (1b,300) Figure B.8: Trends of accumulated strain at 400C, Mix 1b 140 10000 Log. (1b,500) 0 (40 C) 6 % Accumulative Strain X10 10.00 y 100 = 0.1808Ln(x) + 4.9329 R2 = 0.9387 y 300 = 0.1191Ln(x) + 5.1522 R2 = 0.9618 y 500 = 0.0972Ln(x) + 5.247 R2 = 0.9881 1.00 1 10 100 1000 10000 Pulse Counts 1c,100 1c,300 1c,500 Log. (1c,100) Log. (1c,300) Log. (1c,500) Figure B.9: Trends of accumulated strain at 400C, Mix 1c O (40 C) 6 % Accumulative Strain x 10 10.00 y100 = 0.0945Ln(x) + 5.0601 R2 = 0.9028 y300 = 0.0858Ln(x) + 5.1443 R2 = 0.9591 y500 = 0.0847Ln(x) + 5.3571 R2 = 0.882 1.00 1 10 100 1000 10000 Pulse Counts 2a,100 2a,300 2a,500 Log. (2a,100) Figure B.10: Trends of accumulated strain at 400C, Mix 2a 141 Log. (2a,300) Log. (2a,500) 0 (40 C) % Accumulative Strain X10 6 10.00 y 100 = 0.108Ln(x) + 5.1522 R2 = 0.9387 y 300 = 0.101Ln(x) + 5.1318 R2 = 0.9563 y 500 = 0.0919Ln(x) + 5.3092 R2 = 0.995 1.00 1 10 100 1000 10000 Pulse Counts 2b,100 2b,300 2b,500 Log. (2b,300) Log. (2b,100) Log. (2b,500) Figure B.11: Trends of accumulated strain at 400C, Mix 2b 0 (40 C) 6 % Accumulative Strain X 10 10.00 y100 = 0.146Ln(x) + 4.911 R2 = 0.9984 y300 = 0.1217Ln(x) + 5.067 R2 = 0.9331 y500 =0.1088 Ln(x) + 5.096 R2 = 0.9934 1.00 1 2c,100 2c,300 10 100 Pulse Counts 2c,500 Log. (2c,100) Figure B.12: Trends of accumulated strain at 400C, Mix 2c 142 1000 Log. (2c,500) 10000 Log. (2c,300) 5.6.6 Trends of Accumulated Strains in Mixes: at 550C o (55 C) % Accumulative Strain X 10 6 10.00 y300 = 0.1525Ln(x) + 4.8194 R2 = 0.9748 y100 = 0.1595Ln(x) + 4.7329 R2 = 0.9496 y500 = 0.1234Ln(x) + 4.9792 R2 = 0.9824 1.00 1 10 100 1000 10000 Pulse Counts 1a,100 1a,300 1a,500 Log. (1a,500) Log. (1a,300) Log. (1a,100) Figure B.13: Trends of accumulated strain at 550C, Mix 1a o (55 C) 6 % Accumulative Strain X 10 10.00 y 300 =0.1353 Ln(x) +5.0591 R2 = 0.9751 y 100 = 0.1597Ln(x) + 4.8185 R2 = 0.9037 y 500 = 0.0946Ln(x) + 5.317 R2 = 0.9923 1.00 1 10 1b,100 1b,300 100 Pulse Counts 1b,500 Log. (1b,300) Figure B.14: Trends of accumulated strain at 550C, Mix 1b 143 1000 Log. (1b,500) 10000 Log. (1b,100) o (55 C) % Accu mu lativ e Strain X 1 0 6 10.00 y 300 = 0.1548Ln(x) + 4.9669 R2 = 0.975 y 100 = 0.1896Ln(x) +4.8189 R2 = 0.9448 y 500 = 0.1287Ln(x) +5.0907 R2 = 0.9865 1.00 1 10 100 1000 10000 Pulse Counts 1c,100 1c,300 1c,500 Log. (1c,500) Log. (1c,300) Log. (1c,100) Figure B.15: Trends of accumulated strain at 550C, Mix 1c o (55 C) % Accu m u lativ e Strain X 1 0 6 10.00 y300 =0.1256Ln(x) + 4.8263 R2 = 0.9781 y 100 = 0.1584Ln(x) + 4.7803 R2 = 0.9933 y 500 = 0.1082Ln(x) + 5.0544 R2 = 0.9994 1.00 1 10 100 1000 10000 Pulse Counts 2a,100 2a,300 2a,500 Log. (2a,500) Log. (2a,300) Figure B.16: Trends of accumulated strain at 550C, Mix 2a 144 Log. (2a,100) o (55 C) % Accumulativ e Strain X 10 6 10.00 y 300 = 0.1046Ln(x) + 5.0699 R2 = 0.998 y 100 = 0.1343Ln(x) + 5.0154 R2 = 0.9896 y 500 = 0.098Ln(x) + 5.4813 R2 = 0.7784 1.00 1 10 100 1000 10000 Pulse Counts 2b,100 2b,300 2b,500 Log. (2b,100) Log. (2b,300) Log. (2b,500) Figure B.17: Trends of accumulated strain at 550C, Mix 2b o (55 C) % Accumulative Strain X 10 6 10.00 y 300 = 0.1164Ln(x) +4.9869 R2 = 0.9793 y 100 = 0.1576Ln(x) + 4.8324 R2 = 0.9954 y 500 = 0.1053Ln(x) + 5.3167 R2 = 0.9921 1.00 1 10 100 1000 10000 Pulse Counts 2c,100 2c,300 2c,500 Log. (2c,500) Log. (2c,100) Figure B.18: Trends of accumulated strain at 550C, Mix 2c 145 Log. (2c,300) Annexure-C Plots of Resilient Strain: at 250C o o Resilient S train ~ Load Repetition at 25 C Resilient Strain~ Load Repetition at 25 c 0.14 0.18 0.12 0.16 0.14 500kPa 500 kPa 0.12 % Resilient Strain % Resilient Stra in 0.1 0.08 0.06 ` 300kPa 0.04 0.02 0.1 300 kPa 0.08 0.06 0.04 100 kPa 0.02 100kPa 0 0 0 500 1000 1a,100kPa Load Repetition (N) 1500 1a,300 kPa -0.02 2000 1a,500 kPa 0 500 2a,300kPa 2a,500kPa Resilient Strain ~ Load Repetition at 25 C 0.18 0.120 500 kPa 0.080 300 kPa 0.060 ` 0.040 100 kPa 0.020 % Resilient Strain 0.16 0.100 0.14 0.12 500 kPa 0.1 0.08 300 kPa 0.06 0.04 100 kPa 0.02 0.000 0 0 500 Load Repetition (N) 1000 1b,100kPa 1500 1b,300kPa 2000 0 1b,500kPa 500 Load Repetition (N) 1000 1500 2000 2b,100kPa 2b,300kPa 2b,500kPa o o Resilient Strain ~ Load Repetition at 25 C Resilient Strain~ Load Repetition at 25 C 0.12 0.140 0.1 0.120 500 kPa % Resilient Strain % Resilient Strain 2000 o Resilient Strain ~ Load Repetition at 25 C % Resilient Strain 1500 2a,100kPa Load Repetition (N) o 1000 0.08 0.06 300 kPa 0.04 0.02 100 kPa 0 500 1000 1c,100kPa 1500 1c,300kPa 0.080 0.060 300 kPa 0.040 0.020 0 Load Repetition (N) 500 kPa 0.100 100 kPa 0.000 2000 0 1c,500kPa Load Repetition (N) Figure C.1: Influence of Pulse count on resilient strain at 250C 146 500 1000 2c,100kPa 1500 2c,300kPa 2000 2c,500kPa Plots of Resilient Strain at 400C o 0 Resilient Strain ~ Load Repetition at 40 C Resilient Strain ~ Load Repetition at 40 C 0.100 500 kPa 0.080 0.060 300 kPa 0.040 100 kPa 0.020 % R esilien t S tra in % R esilient Strain 0.100 300 kPa 0.050 100 kPa 0.000 0.000 0 500 1000 Load Repetition (N) 1a,100kPa 1500 0 2000 1a,300kPa 1a,500kPa o 0.1 300 kPa 0.05 100 kPa 0 Load Repetition (N) 1b,100kPa 1500 % R esilien t S tra in % R esilient Strain 500 kPa 1000 2a,300kPa 0.2 0.18 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 2000 1b,300kPa 2a,100kPa 1500 2000 2a,500kPa Resilient Strain~ Load Repetition at 40 C 0.15 500 1000 o Resilient Strain~ Load Repetition at 40 C 0 500 Load Repetition (N) 0.2 500 kPa 300 kPa 100 kPa 0 1b,500kPa 500 Load Repetition (N) 1000 1500 2000 2b,100kPa 2b,300kPa 2b,500kPa o 0 Resilient Strain~ Load Repetition at 40 C Resilient Strain~ Load Repetition at 40 C 0.18 0.12 0.16 % R esilien t S tra in 500 kPa % R esilient Stra in 500 kPa 0.08 300 kPa 0.04 100 kPa 0.14 500 kPa 0.12 0.1 0.08 300 kPa 0.06 0.04 0.02 0 100 kPa 0 0 500 Load Repetition (N) 1000 1c,100kPa 1500 1c,300kPa 0 2000 1c,500kPa Load Repetition (N) Figure C.2: Influence of Pulse count on resilient strain at 400C 147 500 1000 2c,100kPa 1500 2c,300kPa 2000 2c,500kPa Plots of Resilient Strain: at 550C 0 0 Resilient Strain ~ Load Repetition at 55 C Resilient Strain ~ Load Repetition at 55 C 0.150 500 kPa 0.100 300 kPa 0.050 100 kPa % R esilient Strain % R esilient Strain 0.200 0.000 0 500 Load Repetition (N) 1000 1a,100kPa 1500 0.18 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 2000 1a,300kPa 1a,500kPa 500 kPa 300 kPa 100 kPa 0 500 Load Repetition (N) 1000 2a,100kPa 1500 2a,300kPa 0 Resilient Strain ~ Load Repetition at 55 C 0 Resilient Strain ~ Load Repetition 55 C 0.35 0.160 % Resilient Strain 0.120 0.100 300 kPa 0.080 0.060 0.040 100 kPa 0.3 % R esilient Strain 500 kPa 0.140 0.25 500 kPa 0.2 0.15 300 kPa 0.1 0.05 0.020 100 kPa 0 0.000 0 500 Load Repetition (N) 1000 1b,100kPa 1500 0 2000 1b,300kPa 1b,500kPa 500 Load Repetition (N) 1000 2b,100kPa 0 Resilient Strain ~ Load Repetition at 55 C 300 kPa 0.060 0.040 100 kPa 0.020 0 Load Repetition (N) 500 1000 1c,100kPa 1500 2000 1c,300kPa 1c,500kPa % R esilient Strain 500 kPa 0.080 2000 2b,500kPa 0 0.120 0.100 1500 2b,300kPa Resilient Strain ~ Load Repetition at 55 C 0.140 % Resilient Stra in 2000 2a,500kPa 0.2 0.18 0.16 0.14 0.12 0.1 0.08 0.06 0.04 0.02 0 0 500 Load Repetition (N) Figure C.3: Influence of Pulse count on resilient strain at 550C 148 500 kPa 300 kPa 100 kPa 1000 2c,100kPa 1500 2c,300kPa 2000 2c,500kPa Annexure-D Influence of Wheel Tracker Load Cycle on Rut Depth Mixture with gradation "01" & PMA (1a) 55 C 400C 5.000 4.000 3.000 250 C 2.000 0.000 40C,1a Load Cycles (N) 100 1300 2500 3700 4900 6100 7300 8500 9700 10900 12100 13300 14500 15700 16900 20000 18000 16000 14000 12000 10000 8000 6000 4000 2000 0 0.000 55C, 1a 25C,1a Mixture with gradation "01" & 60/70 pen grade (1b) 16.00 14.00 12.00 10.00 8.00 6.00 4.00 2.00 0.00 25C, 2a 40C,2A 55C,2A Load Cycles (N) Mixture with gradation "02" & 60/70 pen grade (2b) 25.000 0 55 C R ut D epth (m m ) 0 0 40 C 0 25 C 55 C 20.000 0 15.000 40 C 10.000 0 25 C 5.000 Load Cycles (N) Mixture with gradation "01" & 40/50 pen grade (1c) 25C, 2b 16900 15700 14500 13300 12100 9700 10900 8500 7300 6100 4900 3700 2500 100 25C, 1b 40C,1b 55C,1b Load Cycles (N) 14.000 1300 0.000 100 1500 2900 4300 5700 7100 8500 9900 11300 12700 14100 15500 16900 R u t D e p th (m m ) 400 C 10.000 8.000 6.000 4.000 250C 2.000 1.000 550 C 16.000 14.000 12.000 7.000 6.000 Mixture with gradation "02" & PMA (2a) 20.000 18.000 0 R u t D e p th (m m ) R ut D epth (mm) 9.000 8.000 40C,2b Mixture with gradation "02" & 40/50 pen. grade (2c) 20.000 12.000 0 55 C 0 55 C 18.000 16.000 10.000 R u t D ep th (m m ) R ut D epth (m m ) 55C, 2b 8.000 0 40 C 6.000 4.000 0 25 C 14.000 0 40 C 12.000 10.000 8.000 6.000 0 25 C 4.000 2.000 2.000 0.000 Load Cycles (N) 40C,1c 55C, 1c 25C,1c Figure D.1: Influence of Load Cycle on Rut Depth 149 Load Cycles (N) 13100 14100 15100 16100 17100 4100 5100 6100 7100 8100 9100 10100 11100 12100 100 1100 2100 3100 20000 18000 16000 14000 12000 10000 8000 6000 4000 2000 0 0.000 25C, 2c 40C,2c 55C, 2c Trends of Rut Depth Development in WT Mixture with gradation "01" & PMA (1a) 7.500 y25 = 0.2694Ln(x) + 3.7261 R2 = 0.9524 6 Rut Depth (mmx10 ) 7.000 6.500 6.000 5.500 5.000 y40 = 0.2037Ln(x) + 4.8079 R2 = 0.9965 4.500 y55 = 0.2231Ln(x) + 4.7724 R2 = 0.9944 4.000 10 100 1000 10000 100000 Load Cycle (LOG N) 1A, 25C Log. (1A, 25C) 1a, 40C Log. (1a, 40C) 1a, 55C Log. (1a, 55C) Figure D.2: Trends of Rut Depth Development, Mix 1a Mixture with gradation "01" & 60/70 pen. Grade (1b) 7.5 y25 = 0.2737Ln(x) + 3.9131 R2 = 0.9838 6 Rut Depth (mmx10 ) 7.0 6.5 6.0 5.5 y55 = 0.2174Ln(x) + 5.1163 R2 = 0.9591 5.0 y40 = 0.2295Ln(x) + 4.8107 R2 = 0.9452 4.5 4.0 10 100 1000 10000 100000 Load Cycle (LOG N) 1b, 25C Log. (1b, 25C) 1b, 40C Log. (1b, 40C) Figure D.3: Trends of Rut Depth Development, Mix 1b 150 1b, 55C Log. (1b, 55C) Mixture with gradation "01" & 40/50 pen. Grade (1c) y25 = 0.2957Ln(x) + 3.4551 R2 = 0.3927 7.0 6 Rut Depth (mm x 10 ) 7.5 6.5 6.0 5.5 5.0 y40 = 0.2111Ln(x) + 4.7834 R2 = 0.9909 4.5 4.0 10 100 1c 25C Log. (1c 25C) y55 = 0.2844Ln(x) + 4.3368 R2 = 0.9809 1000 10000 Load Cycle (LOG N) 1c, 40C Log. (1c, 40C) 100000 1c, 55C Log. (1c, 55C) Figure D.4: Trends of Rut Depth Development, Mix 1c Mixture with gradation "02" & PMA (2a) 7.5 6 Rut Depth (mmx10 ) 7.0 6.5 6.0 y 25 = 0.1317Ln(x) + 5.3749 R2 = 0.9926 5.5 5.0 ` y 40 = 0.2161Ln(x) + 4.9767 R2 = 0.9676 4.5 y 55 = 0.2217Ln(x) + 5.1177 R2 = 0.9875 4.0 10 100 1000 10000 100000 Load Cycle (LOG N) 2a, 25C Log. (2a, 25C) 2a, 40C Log. (2a, 40C) Figure D.5: Trends of Rut Depth Development, Mix 2a 151 2a, 55C Log. (2a, 55C) Mixture with gradation "02" & 60/70 pen. Grade (2b) 8.0 y25 = 0.1633Ln(x) + 5.1857 R2 = 0.9869 6 Rut Depth (mmx10 ) 7.5 7.0 6.5 6.0 5.5 5.0 y55 = 0.2118Ln(x) + 5.3468 R2 = 0.9691 4.5 y40 = 0.1968Ln(x) + 5.2741 R2 = 0.9801 4.0 10 100 2b, 25C Log. (2b, 25C) 1000 10000 Load Cycle (LOG N) 2b, 40C Log. (2b, 40C) 100000 2b, 55C Log. (2b, 55C) Figure D.6: Trends of Rut Depth Development, Mix 2b Mixture with gradation "02" & 40/50 pen. Grade (2c) 8.0 6 Rut Depth (mmx10 ) 7.5 y 25 = 0.2606Ln(x) + 4.1251 R2 = 0.9942 7.0 6.5 6.0 5.5 5.0 y 55 = 0.239Ln(x) + 5.0268 R2 = 0.9165 y 40 = 0.199Ln(x) + 5.1743 R2 = 0.9824 4.5 4.0 10 2c, 25C Log. (2c, 25C) 100 1000 Load Cycle (LOG N) 10000 2c, 40C Log. (2c, 40C) Figure D.7: Trends of Rut Depth Development, Mix 2c 152 100000 2c, 55C Log. (2c, 55C