- “Hi'fii‘ :. ”3:25;~ «.“N‘... “1-way? v .N. 1.3:;- -§*-\ . at. ‘13-‘11 47L- .....«‘.‘»_ thmxrwm _ 7'<:- 1;}: wfifiifw- 5‘? _ -_ .—~ a”. xéxz'flrr 3:: any}; I n 4. I“?! 9‘34} . ¢ '1 \s' ”\ohrfu u..$’.:__ ‘31:! -_ _._ W- u. '~r~.~— V‘m-q. ‘e I » I >.r.',-;«..< .y. Wm... nae: I. s. r t "73'; \l.‘:'_‘_- -_. . . V ‘ ‘. ‘A «0 #2; F 'rsv-u 1-“ 5mw a ,. .4...-:x§£-~ ._. 11:25:45? ' ’ :4’m>:5:;:u;w r‘ w m . ‘ ~ ak- (IV-:1" . '53“. w Lu 3.7.2... .5 .3 ‘ -J "hug; "’ ‘5‘: I N” '22:” n . '51: . , - ‘&:M at. M ' '7’“ W123: meg-Aw: .. w “K “HEEL; “$5312; " .3"? . A‘ ‘.‘ “V t v H. x ' as- ‘gw 7-9":~;,. ,I 04y ; m“ mzli ‘3”... LIBRARY Michigan State University PLACE IN RETURN BOX to remove this checkom from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. %EE,DUE DATE DUE DATE DUE WE / JUL 2 12062 1 1100 W.“ -. _.__— ..-_.....——- This is to certify that the dissertation entitled - BACKCALCULATION OF PAVEMENT LAYER PROPERTIES FROM DEFLECTION DATA presented by Tariq Mahmood has been accepted towards fulfillment of the requirements for Doctor of Philosophy degreein Civil Engineering Ronald Harichandr nW‘CL'L Gilbert BaladiMM‘ Majorprofes r Date 11/17/93 man an Ajfmm‘w Action/Equal Opportunity Imitation 042771 BAC BACKCALCULATION OF PAVEMENT LAYER PROPERTIES FROM DEFLECTION DATA By Tariq Mahmood A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Civil and Environmental Engineering 1993 ‘X ABSTRACT BACKCALCULATION OF PAVEMENT LAYER PROPERTIES FROM DEFLECTION DATA BY TARIQ MAI-IMOOD A computer program named MICHBACK has been developed for the backcalculation of pavement layer properties. The program uses a modified Newton algorithm to backcalculate pavement layer moduli and thicknesses from measured surface deflections. It is shown that the newly developed algorithm to predict the roadbed soil modulus at the cost of only a single call to a mechanistic analysis program is accurate. An iterative modified Newton method to calculate the stiff layer depth from deflection data is presented and its accuracy is discussed. The ability of MICHBACK to predict any one of the layer thicknesses along with layer moduli from deflection data is presented. An extensive sensitivity analysis of the backcalculated results to the initial seed moduli is included. Some of the user-friendly features of the MICHBACK program are also presented. The program has been extensively tested using theoretical deflection basins generated by using a multilayer linear elastic program, as well as field data obtained by using a falling weight deflectometer (FWD). The results of these tésts which validate the accuracy and robustness of the program are included. The sensitivity of the results to many factors known to affect the backcalculation results are also explored. The capabilities of the program have been compared with other leading backcalculation programs and the results indicate the superiority of the MICHBACK program in many aspects. The effect of temperature on the backcalculated layer moduli has also been examined. TO MY PARENTS AND FAMILY ACKNOWLEDGEMENTS First of all I want to thank Allah Almighty for providing me with the opportunity to undertake and complete this task. Thanks are also due to the Pakistan Army Corps of Engineers for sponsoring my study program. I also want to thank the Michigan Department of Transportation and the University of Michigan Transportation Research Institute for sponsoring this study. I want to express my gratitude and sincere thanks to my two co—chairpersons, Dr. R.S. Harichandran and Dr. G.Y. Baladi for their continuous professional guidance and moral support without which it would not have been possible to pursue this study. Dr. Harichandran guided the development of the backcalculation algorithm and the programming; Dr. Baladi guided the field data collection and validation, and was the principal investigator for the project. Dr. W.C Taylor and Dr. R.V Erickson deserve special thanks for serving on the advisory committee and for their comments and interest. The Technical Advisory Committee for this project from Michigan Department of Transportation, including Dave Smiley, Jack DeFoe, Jerry Sweeny, and Ishvarlal Patel, also deserve special thanks for their guidance and comments. The dedicated effort of Kurt Bancroft in conducting the falling weight deflection tests and supervising coring operation is also recognized and appreciated. I acknowledge with gratitude the dedicated and professional assistance provided by Cynthia Ramon and Weijun Wang in writing the MICHBACK program. Hamid Mukhtar deserves special thanks for sharing the data prepared by him, and his advise during the course of this study. Efforts of Stuart 8., Guy N., John A., iv :4». l well Erich T., and others in data acquisition and reduction is appreciated. The patience, understanding, and continuous prayers of my wife and my family deserve a special mention and thanks. I am grateful to my elder brother Mumtaz Ahmad Javaid for never letting my absence from home being felt. My children Iqr'a and Umar deserve special thanks for their understanding and for providing much needed moments of pleasure and relief. . My wife deserves special thanks for her sacrifices, her help in data reduction, proof reading, and preparation of the thesis. She prayed at difficult moments and was a source of encouragement and inspiration. Finally, I am grateful to my late parents; . their prayers and the strength of their education is felt every step of the way. May Allah be happy with them. TABLE OF CONTENTS LIST OF TABLES ...................................... xii LIST OF FIGURES .................................... xvii CHAPTER 1 W ...................................... 1 1.1 GENERAL ...................................... 1 1.2 PROBLEM STATEMENT ............................ 2 1.3 RESEARCH OBJECTIVES ........................... 3 1.4 THESIS LAYOUT ................................. 3 CHAPTER 2 W ................................. 5 2.1 GENERAL ...................................... 5 2.2 NONDESTRUCTIVE DEFLECTION TESTING DEVICES ....... 8 2.3 SURFACE WAVE TESTING ........................... 10 2.3.1 Impulse Load Testing .......................... 11 2.3.2 Steady-State Vibration ......................... 12 2.3 .3 Advantages and Limitations ...................... 12 2.4 NONDESTRUCTIVE DEFLECTION TESTING .............. 13 2.4.1 Static Force-Deflection ......................... 14 2.4.2 Dynamic Steady State Vibrator .................... 15 2.4.3 ImmctLoading ........ 20 2.5 DEFLECTION RESPONSE OF PAVEMENTS ............... 21 2.6 INTERPRETATION OF DEFLECTION DATA .............. 23 2.6.1 Empirical Analysis ............................ 23 vi W B. 2.6.2 Rational Methods .............. ' ............... 25 2.6.3 Mechanistic Analysis ........................... 25 2.7 SPECIAL PROPERTIES OF PAVING MATERIALS .......... 28 2.7.1 Material Non-Linearity ......................... 30 2.7.2 Temperature Dependency ....................... 35 2.8 MECHANISTIC ANALYSIS MODELS .................... 36 2.8.1 Layered Elastic Model ......................... 36 2.8.2 Hogg’s Model ............................... 40 2.8.3 Equivalent Thickness Model ...................... 41 2.8.4 Finite Element Method ......................... 42 2.8.5 Dynamic Analysis ............................ 42 2.8.6 Nonlinear Elasticity ........................... 46 2.8.7 Viscoelastic Model ............................ 49 2.8.8 Pavement Material Type and Choice of Analysis Model . . . . 50 2.9 BACKCALCULATION METHODS ..................... 51 2.9.1 Iterative Methods .............................. 51 2.9.2 Database Approach ............................ 58 2.9.3 Statistical Analysis ............................. 59 2.9.4 Conversion of Backcalculated Layer Moduli to Standard Conditiom .................................. 60 2.9.5 Sources of Error In Backcalculated Layer Moduli ......... 68 2.9.6 Error Measures and Convergence Criteria ............. 7O 2.10 SUMMARY ...................................... 71 CHAPTER 3 W W ....................... 74 3- 1 GENERAL ...................................... 74 3-2 IMPROVED ESTIMATION OF ROADBED MODULUS ......... 75 3-3 NEWTON’S METHOD .............................. 77 3-4 NEWTON’S METHOD WHEN THE DERIVATIVES OF THE FUNCTION ARE NOT AVAILABLE ..................... 79 3-5 MUL'TI-DIMENSIONAL FORM OF NEW'TON’S METHOD ...... 81 3.6 USE OF NEWTON’S METHOD TO BACKCALCULATE vii 3.7 3.8 3.10 4.3 4.4 PAVEMENT LAYER MODULI ......................... 82 LAYER THICKNESS ESTIMATION ..................... 85 THE MODIFIED NEWTON METHOD .................... 87 STIFF LAYER EFFECTS AND DEPTH TO STIFF LAYER ...... 87 LOGARITHMIC TRANSFORMATION ................... 90 3.10.1 Relationship between Surface Deflection and Layer Moduli ............................... 90 3.10.2 Implementation of Logarithmic Transformation ...... 99 101 101 TYPICAL PAVEMENT SECTIONS AND TEST PARAMETERS USED ......................................... 102 NUMERICAL ILLUSTRATION: IMPROVED ESTIMATION OF ROADBED MODULUS ............................. 105 ESTIMATION OF LAYER THICKNESSES AND THEIR EFFECT ON THE BACKCALCULATED LAYER MODULI ........... 105 4.4.1 AC Thicknom .............................. 107 4.4.2 Base Thickness ............................. 115 STIFF LAYER EFFECTS AND DEPTH ESTIMATION . ....... 121 4.5.1 Backcalculation of Layer Moduli and Stiff Layer Depth . . . 121 4.5.2 Sensitivity of Backcalculated Moduli to Stiff Layer Characteristics .............................. 125 CONVERGENCE CHARACTERISTICS .................. 137 4.6.1 Three-Layer Flexible Pavements .................. 138 4.6.2 Four-Layer Flexible Pavement ................... 138 4.6.3 Four-Layer Composite Pavements ................. 140 4.6.4 Three-Layer Pavements Over a Stiff layer ............ 140 4.6.5 Five-Layer Flexible Pavement .................... 143 4.6.6 Performance Comparison ...................... 143 UNIQUENESS OF THE BACKCALCULATED RESULTS ...... 148 viii 4.8 EFFECT OF INACCURACIES IN DEFLECTIONS AT SIMULATED SENSOR LOCATIONS ON BACKCALCULATED RESULTS . . . . 150 4.9 EFFECT OF POISSON’S RATIO ON THE BACKCALCULATED LAYER MODULI ................................. 154 4.10 COMPARISON OF DEFLECTION OUTPUT OF DIFFERENT ELASTIC LAYER PROGRAMS ....................... 158 4.10.1 Comparison of Deflection Output ................ 159 4.10.2 Comparison of Backcalcuiated Results ............. 162 4.11 COMPARISON OF MICHBACK RESULTS WITH SHRP STUDY ....................................... 164 171 173 173 173 5.3 PROCESSING A FWD DEFLECTION DATA FILE .......... 174 5.3.1 Reviewing and Preprocessing the Deflection Data ....... 175 5.3.2 Data Analysk Options ......................... 178 5.4 PRIEENTATION OF THE BACKCALCULATED RESULTS . . . . 180 5-5 PROGRAM STRUCTURE ........................... 180 5-6 BACKCALCULATION OF LAYER PROPERTIES ........... 182 CHAPTER 6 ........................ 187 6- 1 GENERAL ..................................... 187 6.2 PAVEMENT SELECTION ........................... 187 6-3 MARKING, CODING, CORING AND NDT ............... 195 5.4 BACKCALCULATION OF LAYER MODULI FOR SELECTED ix ——*—__.— 6.4 6.4 6.4 6.4 6.4 6.4 4.6 4.6 6.4 6.4 6.4 ”84 PAVEMENTSECTIONS............... ............. 197 6.4.1 Flexible Pavement Section MSU07F - Variable Deflections ................................. 202 6.4.2 Flexible Pavement Section MSU19F - Uniform Deflections ................................. 209 6.4.3 Flexible Pavement Section MSUIOF - Uniform Deflections ................................. 215 6.4.4 Flexible Pavement Section MSU13F - Variable Deflectiom ................................. 219 6.4.5 Flexible Pavement Section MSU14F - Variable Deflection .................................. 219 6.4.6 Flexible Pavement Section MSU29F - Variable Deflections ................................. 223 6.4.7 Flexible Pavement Section MSU35F - Variable Deflections ................................. 226 4.6.8 Flexible Pavement Section MSU43F - Variable Deflections ................................. 231 4.6.9 Composite Pavement Section MSU01C - Variable Deflections ................................. 231 6.4.10 Composite Pavement Section MSU05C - Variable Deflections ................................. 237 6.4.11 Composite Pavement Section MSU08C - Variable Deflectiom ................................. 241 6.4.12 Composite Pavement Section MSU04C - Variable Deflections ................................. 241 6.5 BACKCALCULATION OF LAYER MODULI AT DIFFERENT LOAD LEVELS .................................. 246 6.5.1 Flexible Pavement Section MSU19F - Variable Load Level ..................................... 249 6.5.2 Flexible Pavement Sections MSU35F - Variable Load Level ..................................... 259 6.5.3 Composite Pavement Section MSU01C - Variable Load Level ..................................... 265 6.5.4 Discussion ................................. 265 7.2 TEMPERATURE EFFECTS .......................... 278 7.2.1 Flexible Pavement Section MSU52F - Temperature Variation .................................. 280 7.2.2 Flexible Pavement Section MSU13F - Temperature Variation ................................. 288 7.2.3 flexible Pavement Section MSU19F - Temperature Variation .................................. 292 7.3 THE ASPHALT INSTTTUTE TEMPERATURE CORRECTION PROCEDURE .................................... 295 7.4 THE AASHTO TEMPERATURE CORRECTION PROCEDURE . . 299 7.5 SUMMARY AND CONCLUSIONS ._ ..................... 303 CHAPTER 8 W .............. 304 8.1 SUMMARY ..................................... 304 8.2 ACCOMPLISHMENTS ............................. 305 8.3 CONCLUSIONS .................................. 306 8.4 RECOMMENDATIONS FOR FUTURE RESEARCH .......... 308 xi _‘—-—_—’ 12662.1. Table 2.2. Table 2.1. Table 2.2. Table 4.1. Table ‘4.2. Table 4.3. Table 4.4. Table 4.5. Table 4.6. Table 4.7. Table 4.8. Table 4.9. Table 4.10. Table 4.11. Table 4.12. LIST OF TABLES Summary of deflection basin parameters (Mahoney, et al., 1991). ..................................... 27 Backcalculation programs compiled by SHRP as of November 1990 (modified afler ref. Mahoney, 1991). .............. 52 A typical three layer flexible pavement ................. 103 A typical four layer flexible pavement. ................ 103 Typical four layer composite pavements. ............... 103 A typical five layer flexible pavement. ................ 103 Improvement of the roadbed modulus for a three layer flexible pavement. .................................. 106 Improvement of the roadbed modulus for a four layer flexible pavement. .................................. 106 Improvement of the roadbed modulus for a five layer flexible pavement. .................................. 106 The percent error in the backcalculated layer moduli due to percent errors in the AC thickness. ....................... 108 Errors in the backcalculated layer moduli due to errors in the specified AC thickness for different AC thicknesses. ........ 113 Errors in the backcalculated layer moduli due to errors in the specified AC thickness for different AC modulus. ......... 116 The percent errors in the backcalculated layer moduli due to percent errors in the base thickness. .................. 117 Backcalculation of stiff layer depth and moduli for a three layer flexible pavement. .................. 122 xii _———-———— —-—-——-'-—. 11118 41 12614.1 14164.1 W. 1441114.] 1314.1 1411: 4.1 . 116224; Table 4.13. Table 4.14. Table 4.15. Table 4.16. Table 4.17. Table 4.18. Table 4.19. Table 4.20. Table 4.21. Table 4.22. Table 4.23. Table 4.24. Table 4.25. Table 4.26. Table 4.27. Backealculation of stiff layer depth and moduli for a four layer flexible pavement. ......................... Backcalculation of stiff layer depth and moduli for a four layer composite pavement. ................. Error in the the backcalculated layer moduli due to percent error in the depth to the stiff layer. The effect of stiff layer modulus on the backcalculated layer moduli (stiff layer depth fixed) ...................... Effect of stiff layer modulus on the backcalculated layer moduli and stiff layer depth. ................... Comparison of the results of three programs for a three layer pavement. .................................. Comparison of the results of three programs for a four layer pavement. ._ ................................. Comparison of the backcalculated results for a composite pavement section. 4 Comparison of the backcalculated results for a composite pavement section consisting of a granular separation layer. .......... Comparison of the results of three programs for a three layer pavement over stiff layer. ........................ The MICHBACK backcalculation results for a five layer pavement. .................................. Comparison of the performance'of MICHBACK and EVERCALC programs. Uniqueness of the MICHBACK solution for a three layer flexible pavement. .................................. Uniqueness of the MICHBACK solution for a four layer flexible pavement. .................................. Uniqueness of the MICHBACK solution for a five layer flexible pavement. .................................. xiii 124 126 134 136 139 139 141 142 144 149 149 149 _—-——-. _—.— —_—— —— ——-—- _————— _— _— 1111: 4. lab}: 4. 1: 114.. Table 4.1 M 111114.; 1111: 4.3 141143 1411 4.3 141216 4.31 Table 4.28. Table 4.29. Table 4.30. Table 4.31. Table 4.32. Table 4.33. Table 4.34. Table 4.35. Table 4.36. Table 4.37. Table 4.38. Table 4.39. Table 4.40. Table 6.1. Table 6.2. Table 6.3. Table 6.4. Percent errors in the backcalculated layer moduli due to errors in deflections - arithmetic scale. ...................... 152 Percent errors in the backcalculated layer moduli due to errors in deflections - logarithmic scale. ..................... 153 Percent error in the backcalculated layer moduli due to :l: 0.5 mil error in deflections - logarithmic scale. ..... 155 Percent errors in the backcalculated layer moduli due to error in Poison’s ratio for flexible pavements. ....... 157 Percent errors in the backcalculated layer moduli due to error in Poison’s ratio for composite pavement. ...... 157 Deflections from different elastic layer programs ........... 160 Percent differences in the deflections of elastic layer programs. . 161 Backealculated results of MICHBACK for deflection data generated by different elastic layer programs. .................. 163 Deflection and cross sectional data for nine test sections (after Rada, et al., 1992). ........................ 165 Comparison of the true and backcalculated moduli values (after Rada, et al., 1992). .................... 166 MICHBACK results for nine pavement sections. .......... 167 Comparison of errors in the modulus values. ............ 168 Statistics of the maximum relative error for four computer programs. ..... _ ....................... 170 Criteria for final section selection. ................... 189 Flexible pavement sections selected for study. ............ 191 Composite pavement sections selected for study. .......... 193 Summary of the layout and conditions of the cored flexible pavement section. ............................. 199 xiv Table 6.5. Table 6.6. Table 6.7. Table 6.8. Table 6.9. Table 6.10. Table 6.11. Table 6.12. Table 6.13. Table 6.14. Table 6.15. Table 6.16. Table 6.17. Table 6.18. Table 6.19. Table 6.20. Table 6.21. Table 6.22. Table 6.23. Summary of the layout and condition of the cored composite pavement sections .............................. Backealculated moduli for pavement section MSU07F. Backcalculated moduli for pavement section MSUl9F. Backcalculated moduli for pavement section MSUIOF. 000000 Backcalculated moduli for pavement section MSU13F. Backealculated moduli for pavement section MSUl4F. Backcalculated modulus for pavement section MSU29F ....... Backcalculated modulus for pavement section MSU35F ....... Backcalculated moduli for pavement section MSU43F. Backcalculated moduli for pavement section MSU01C. Backcalculated moduli for pavement section MSUOSC. Backcalculated moduli for pavement section MSU08C. Roadbed characteristics of three pavement sections. Backealculated moduli at various loads for pavement section MSUl9F. .................................. Differences between measured and linearly extrapolated deflection basins for pavement section MSUl9F. ................ Backealculated moduli at various loads for pavement section MSU35F. .................................. Differences between measured and linearly extrapolated deflection basins for pavement section MSU35F. ................ Backealculated moduli at various loads for pavement section MSU01C. .................................. Differences between measured and linearly extrapolated deflection basins for pavement section MSU01C. 201 228 230 250 263 266 Table 6.24. Table 6.25. Table 7.1. Table 7.2. Table 7.3. Table 7.4. Table 7.5. Table 7.6. Table 7.7. Table 7.8. Table 7.9. Backcalculation of layer moduli at different Poisson’s ratios for pavement section MSU19F (test No. 19113011). .......... Backcalculation of layer moduli at different Poisson’s ratios for pavement section MSU35F (test No. 35114211). .......... Deflection variation with temperature for pavement section MSU52F. .................................. Percent variation in deflection from lowest temperature recorded for pavement section MSU52F. ..................... Backealculate layer moduli for selected sections. .......... Deflection variation with temperature for pavement section MSU13F. .................................. Percent variation in deflections from lowest temperature recorded for pavement section MSUl3F. ..................... Deflection variation with temperature for pavement section MSU019F. ................................. Percent variation in deflections from lowest temperature recorded for pavement section MSUl9F. ..................... Observed and Asphalt Institute AC modulus temperature correction factor. .................................... Observed and AASHTO temperature correction factors for Do. xvi 272 273 282 282 285 289 289 293 I . 0. 1 .l 11" ‘ 0‘ ‘IW :16 Al .t 61/.- ar 6 2 2 L m. ”4...... mu. m a. m a... a tie . 4 1% .. .1 PM“?! I 4 E AIL. a." ~ MM. .3. mm.- Puma... 54412.5 «I I. .34 the.» Mp n :15. $12.] Figure 2.1. Figure 2.2. Figure 2.3. Figure 2.4. Figure 2.5. Figure 2.6. Figure 2.7. Figure 2.8. Figure 2.9. Figure 2.10. Figure 2.11. Figure 2.12. Figure 2.13. Figure 2.14. LIST OF FIGURES Typical output of a vibrating steady state force generator (Moore, et al., 1978). ........................... 16 Typical load-deflection curves for frequencies of 10, 15, and 40 Hz. (Green, et al., 1974). ......................... 17 Mass-spring-dashpot representation of a pavement structure subjected to a forced dynamic vibration (Lorenz, et al., 1953). . . l9 Well-designed pavement deflection history curve (Moore, et al., 1978). ..................................... 22 Influence of season on pavement deflections (Izada, 1966). ..... 24 Use of deflection basin parameters to analyze pavement structural layers, from Utah overlay design procedure (Molenaar, et al.,1982). ................................... 26 Component versus system analysis of pavement structures. ..... 29 Typical variation of resilient modulus with repeated stress for cohesive roadbed soil (Ming-Shah, 1989). ............. 33 Multi-layered elastic system (Yoder, et al., 1975). .......... 37 Assumed periodicity of FWD impulses (Sebaaly, et al., 1986). ..................................... 45 Schematic showing measured and predicted surface deflection basins (Stolle, et al., 1989). ................. 47 Typieal flow chart for an iterative program (Lytton, 1989) ...... 54 Basic process for matching deflection basins (Van Cauwelaert, 1989). .......................... 57 Predicted pavement temperatures (Asphalt Institute, 1977). ..... 62 Figure 2.15. Figure 2.16. Figure 3.1. Figure 3.2. Figure 3.3. Figure 3.4. Figure 3.5. Figure 3.6. Figure 3.7. Figure 3.8. Figure 3.9. Figure 3.10. Figure 3.11. Figure 4.1. Figure 4.2. Figure 4.3. Temperature prediction graphs for pavements equal to or less than 2 inches thick (Asphalt Institute, 1977) .................. 63 General hyperbolic stress-strain curve for base, subbase, and roadbed materials (Richard, et al., 1975). ............... 66 General illustration of Newton’s method (Dennis, 1983). ...... 78 Seeant approximation of Newton’s method (Dennis, 1983). ..... 80 Graphieal illustration of Newton’s method iteration to find pavement layer moduli. ...................... 83 Effect of AC layer modulus on pavement surface deflections. . . . 91 Effect of AC layer modulus on pavement surface deflections (logarithmic scale). ............................. 92 Effect of base layer modulus on pavement surface deflections. . . . 93 Effect of base layer modulus on pavement surface deflections (logarithmic scale). ............................. 94 Effect of roadbed layer modulus on pavement surface deflections. ............................. 95 Effect of roadbed layer modulus on pavement surface deflections (logarithmic scale). ............................. 96 Graphic illustration of surface deflection variation with AC and roadbed modulus. .............................. 97 Graphic illustration of surface deflection variation with AC and roadbed modulus (logarithmic scale) ................. 98 Errors in the backealculated AC modulus due to incorrect AC thickness specification. ....................... 109 Errors in the backcalculated base modulus due to incorrect AC thickness specification. ....................... 110 Errors in the backealculated roabed modulus due to incorrect AC thickness specifieation. ................. 111 xviii Figure 4.4. Figure 4.5. Figure 4.6. Figure 4.7. Figure. 4.8. Figure 4.9. Figure 4.10. Figure 4.11. Figure 4.12. Figure 4.13. Figure 5.1. Figure 5.2. Figure 5.3. Figure 5.4. Figure 5.5. Figure 5.6. Figure 6.1. Errors in the vertical stress at the top of the base layer due to incorrect AC thickness specification. ................ Errors in the backcalculated AC modulus due to incorrect base thickness specifieation. .......................... Errors in the backealculated base modulus due to incorrect base thickness specification. ....................... Errors in the backcalculated roadbed modulus due to incorrect base thickness specification. ................. Effect of the depth to stiff layer on surface deflections. ...... Errors in backcalculated AC modulus due to incorrect stiff layer depth specification. ............................ Errors in the backcalculated base modulus due to incorrect stiff layer depth specification. ...................... Errors in the backcalculated roadbed modulus due to incorrect stiff layer depth specification. ................ Comparison of the performance of MlCI-IBACK and EVERCALC programs. .................................. Maximum absolute relative percent error in layer moduli for different programs. ............................ Deflection profile for pavement section MSU07F. ......... Zoom-in function between stations 10 and 15 of pavement section MSU07F. .................................. Flow chart for MICHBACK program. ................ Details of backcalculataion procedure for cases A & B. ...... Details of backealculataion procedure for cases C & D. ...... Details of backcalculataion procedure for case B. .......... Distribution of selected pavement sections across the State of Michigan. .................................. 114 118 119 120 128 130 131 132 146 169 183 185 186 194 II 1‘ t .V .. w. .b... we... .w. .3. a; .r a... ...... P y b .7 D 5 .' .. a Z. fin E El. F :1 E .5 2% a a .t the CUR. Figure 6.2. Figure 6.3. Figure 6.4. Figure 6.5. Figure 6.6. Figure 6.7. Figure 6.8. Figure 6.9. Figure 6.10. Figure 6.11. Figure 6.12. Figure 6.13. Figure 6.14. Figure 6.15. Figure 6.16. Figure 6.17. Figure 6.18. Figure 6.19. Deflection profile for pavement section MSU07F. ......... 203 Variation of backcalculated layer moduli, the AC thickness, and the deflection of sensor 1 and 7 along pavement section MSU07F. .................................. 206 Variation of deflections for all seven sensors along pavement section MSUO7F. ............................. 208 Deflection profile for pavement section MSUl9F. ......... 210 Variation of backcalculated layer moduli, the AC thickness, and the deflection of sensor 1 and 7 along pavement section MSUlQF. 213 Variation of deflections for all seven sensors along pavement section MSUl9F.. ............................. 214 Deflection profile for pavement section MSUlOF. ......... 216 A typical deflection basin at test location No.2 of pavement section MSUlOF. .................................. 217 Deflection profile for pavement section MSU13F. ......... 220 Deflection profile for pavement section MSUl4F. ......... 222 Deflection profile for pavement section MSU29F. ......... 225 Typical deflection basins at test loeations 18, 27, 41, 44, and 52 for pavement section MSUZ9F. ..................... 227 Deflection profile for pavement section MSU35F. ......... 229 Deflection profile for pavement section MSU43F. ......... 232 Deflection profile for pavement section MSU01C. ......... 234 Typieal deflection basins at test locations 4, 35, and 37 for pavement section MSU01C. ....................... 235 Deflection profile for pavement section MSUOSC. ......... 238 Typical deflection basins at test locations 28, 43, 45, and 46 for pavement section MSUOSC. ....................... 240 xx 11,411 4 figure 1 1111: 1 12m 6 111116 11441: 6. . 71444 6. 64426.: F *l ‘43? 6.2 141*: 6.2 Figure 6.20. Figure 6.21. Figure 6.22. Figure 6.23. Figure 6.24. Figure 6.25. Figure 6.26. Figure 6.27. Figure 6.28. Figure 6.29. Figure 6.30. Figure 6.31. Figure 6.32. Deflection profile for pavement section MSU08C. ......... 242 Typieal deflection basins at test locations 2, 7, 9, and 33 for pavement section MSU08C. ....................... 244 Deflection profile for pavement section MSU04C. ......... Typical deflection basins at test locations 4, 10, 11, and 12 for pavement section MSU04C. ....................... 247 Comparison of measured and expected linear response deflection basins for pavement section MSU19F (FWD test code 19113011). ................................. 252 Comparison of measured and expected linear response deflection basins for pavement section MSU19F (FWD test code 191 1831 1). ................................. 253 Comparison of measured and expected linear response deflection basins for pavement section MSU19F (FWD test code 19121611). ................................. 254 Comparison of measured and expected linear response deflection basins for pavement section MSU19F (FWD test code 19128221). ................................. 255 Variation of backealculated layer moduli and the deflection of sensor 1, 2, and 7 along pavement section MSU19F at low load level. ..................................... 257 Variation of backcalculated layer moduli and the deflection of sensor 1, 2, and 7 along pavement section MSU19F at high load level. ..................................... 258 Comparison of measured and expected linear response deflection basins for pavement section MSU35F (FWD test code 35118931). ................................. 260 Comparison of measured and expected linear response deflection basins for pavement section MSU35F (FWD test code 351 14222). ................................. 261 Comparison of measured and expected linear response deflection basins for pavement section MSU35F (FWD test code 35123311). .................................. 262 xxi figure 1 115118 1 Figure 6.33. Figure 6.34. Figure 6.35. Figure 7.1. Figure 7.2. Figure 7.3. Figure 7.4. Figure 7.5. Figure 7.6. Figure 7.7. Figure 7.8. Figure 7.9. Figure 7.10. Comparison of measured and expected linear response deflection basins for pavement section MSU01C (FWD test code 01284611). ................................. Comparison of measured and expected linear response deflection basins for pavement section MSU01C (FWD test code 01272411). ................................. Comparison of measured and expected linear response deflection basins for pavement section MSU01C (FWD test code 01261611). ................................. Variation of maximum deflection with temperature observed for pavement section MSU52F. ....................... Percent variation in deflections with temperature for pavement section MSU52F. ............................. Variation of backealculated AC modulus with temperature for pavement section MSUS2F. ....................... Variation of backcalculated base and roadbed moduli with temperature for pavement section MSU52F. ............. Percent variation in deflections with temperature for pavement section MSU13F. . . ........................... Variation of backcalculated AC modulus with temperature for selected pavement sections. ....................... Percent variation in deflections with temperature for pavement section MSUl9F. ............................. Comparison of observed and Asphalt Institute for selected pavement sections .............................. Variation of peak deflections with temperature for selected pavement sections .............................. Comparison of observed and AASHTO temperature conversion factors for peak deflections. ....................... mi 268 269 270 281 286 287 301 301 302 CHAPTER 1 INTRODUCTION 1.1 GENERAL A pavement system subjected to a vehicular or other load input produces a measurable output response in the form of surface deflections. Hence, pavement deflections represent an overall "system response" of the paving layers and the roadbed soil to an applied load. Pavement surface deflections have traditionally been used as an indicator of its structural capacity. Emphasis on mechanistic design and analysis led to the search for efficient schemes to backcalculate layer properties needed for the analysis. The assessment of layer properties and their variation along the road is essential for accurate evaluation of existing pavements and for the design of the asphalt overlays. A review of existing backcalculation schemes and their characteristics suggest that, in general, they can be grouped into one of three categories: those based on regression equations, those based on pre-calculated database of deflections basins, and those based on iterative methods. The accuracy of the methods based on statistical equations is generally low and problem dependent. Methods based on a database of deflection basins received a better acceptance because of the MODULUS program (Scullion, et al. , 1990). However the accuracy of the backcalculated results are affected by the seed values of the layer moduli especially that of the roadbed soil. Another shortcoming is that the estimation of the stiff layer depth is not very accurate which in turn can contribute considerable error to the backcalculated results. Existing iterative methods are relatively slow, share the disadvantage of dependence on seed modulus values, and are unable to mechanistically predict the stiff layer depth. In many cases the accuracy of these methods decreases with the increasing number of pavement layers. Most existing iterative programs seek to minimize an objective function for the estimation of layer moduli. The objective function is normally the weighted sum of squares of the difference between calculated and measured surface deflections (Uzan, et al., 1989). One of the problems of this approach is that the multi-dimensional surface represented by the objective function may have many local minima. The minimum to which a numerical solution may converge depends on the seed moduli supplied by the analyst and may yield unacceptable results. Also, for many cases, the method may fail to converge on a solution within a reasonable time. The exact layer thicknesses at the point of FWD testing are seldom known. Pavement coring is one direct method of measuring the layer thicknesses. A typical core, however, yields layer thickness information at a point. The FWD test provides deflection information over much wider distance (60 - inch for the MDOT FWD). Unfortunately, variation in construction and terrain make the variation in layer thicknesses inevitable. lnaccuracies in the layer thicknesses can contribute a signifieant error in the backcalculated layer properties. Therefore, backcalculation methods which can compute layer thicknesses along with the layer moduli from the deflection data will have the advantage of increased accuracy. Detection of the depth to the stiff layer is also essential for accuracy in the backcalculated layer moduli. Estimation of stiff layer depth by existing methods is not very accurate and ean induce considerable error in the backcalculated properties. Improvement of this estimate would represent a major contribution to this profession. 1.2 PROBLEM STATEMENT Most existing backcalculation of pavement layer moduli methods seek to minimize an objective function for the estimation of layer moduli. This approach makes the backcalculated results dependent on the values of the seed modulus. None .1 3 of the existing programs appear to determine the stiff layer depth mechanistically, they are capable of only providing a rough estimate which can adversely affect the backcalculated layer moduli. Hence, there is a need to produce a backcalculation algorithm such that its outputs are not affected by the values of the seed moduli, and is able to provide a mechanistic estimate to the stiff layer depth along with the layer moduli. 1.3 RESEARCH OBJECTIVES The main objective of this research is to develop a robust backcalculation program whose results are not sensitive to the seed values of the layer moduli. Also, the algorithm should have the capability to accurately compute the stiff layer depth. The program should be user-friendly, provide various options to the user to view and pre-process the deflection basins if necessary, to be of benefit to local State Highway Agencies (SHA) and able to directly process the format of the deflection output files of the Michigan Department of Transportation (MDOT) operated version of the KUAB Falling Weight Deflectometer. 1.4 THESIS LAYOUT This thesis is organized in eight chapters as follows: Chapter 2 - Literature review - Nondestructive deflection testing (NDT) methods and related equipment along with their limitatiOns and capabilities are briefly discussed. Various uses of the deflection data, backcalculation of layer moduli methods and their merits and limitations, and pavement material properties are introduced. Also some of the difficulties related to the backcalculation process, error sources, and various methods for converting the backcalculated properties to standard load and temperature conditions are presented. Chapter 3 - Efficient iterative methods for backcalculation of pavement layer BTU: a 3‘s .3! 4 properties - The Newton method and its application to the backcalculation of pavement layer properties is presented. A new method to estimate the modulus of the roadbed sail in a single call to an elastic layer program is introduced. Also, the modified Newton method and a logarithmic transformation to spwd up convergence is discussed. Chapter 4 - Verification of MICHBACK algorithm using theoretical deflection basins - Verification of the MICHBACK backcalculation algorithm using theoretical deflection basins are presented. The backcalculated results are compared to those obtained from MODULUS 4.0 and EVERCALC 3.0 computer programs. Important aspects of convergence characteristics and uniqueness of the solutions are tested. Sensitivity of backcalculated results to Poisson’s ratios, inaccuracies in deflections at different sensor locations and accuracy of elastic layer programs is also examined. Chapter 5 - Michback program structure and features - The features and the structure of the MICHBACK program are presented. Chapter 6 - Validation using field data - Validation of the MICHBACK program using FWD test data from pavements across the State of Michigan is presented. Chapter 7 - Temperature effects on the backcalculated layer moduli - The effects of the pavement temperature on the AC modulus are discussed. Chapter 8 - Summary, conclusions, and recommendations. CHAPTER 2 LITERATURE REVIEW 2.1 GENERAL A pavement system subjected to a vehicular or other load input produces an output response in the form of deflection. Hence, pavement deflections represent an overall ”system response" of the paving layers and the roadbed soil to an applied load. A load applied at a point (or an area) on the pavement surface will attenuate with depths and radial distances thereby, causing all pavement layers to deflect as a consequence of the introduced stresses and the resulting strains. The amount of deflection in each layer will generally decrease with depth and radial distance and will vary depending on the layer properties. Beyond a certain depth and radial distance, the induced stresses become negligible and the materials are not affected by the applied load. Further, stronger pavements (i.c., those with good quality materials and thick layers) deflect less under a given load than do weaker pavements. Hence, pavement deflections are being used as indieators of pavement quality and as inputs (along with the layer thicknesses) to a mechanistic analysis routine to backcalculate the properties of the various pavement layers. Pavement deflection can be measured by using nondestructive deflection tests (NDT). An NDT consists of applying a known force to a pavement surface and measuring its response (deflection). In some test methods (e.g., the Benkelman beam), the pavement deflection (or more precisely pavement rebound) is typically measured at a point loeated between the two tires of the back axle of a single axle truck. In other methods (e. g. , dynaflect, falling weight deflectometer (FWD), road rater), the pavement deflection profile (deflection at several points located under, and at various radial distances away from the center of the loaded area) is typically 6 measured. Analysis of the measured pavement deflection provides a quantitative basis for evaluation of the pavement structural conditions at any time during its service life. In addition, important information regarding rehabilitation and maintenance requirements can be inferred from the deflection profile (deflection basin). Because they are nondestructive in nature, NDT are easily and quickly performed, the tests cause minimal hindrance to the normal flow of traffic, and they are less hazardous and more economical to perform. In addition, the measured deflections are representative of the actual pavement response to the applied load. As stated earlier, pavement surface deflection has traditionally been used as an indieator of the structural capacity of pavements. One of the earliest uses of pavement deflection was that made in California in 1938 and reported by Haveem (1938). He concluded that flexible pavements would have satisfactory performance if they deflect less than 20 mils under a 15000 lb axle load. The WASHO Road Test (conducted in Huba Valley on flexible pavements) results showed that for a satisfactory pavement performance, pavement deflections should be limited to 30 to 40 mils for pavements located in cold and warm regions, respectively (WASHO, 1954; WASHO, 1955). Presently, NDT data are being used in conjunction with the pavement distress survey for evaluation, rehabilitation, and pavement management purposes. A proper analysis of the NDT data can provide information regarding: l. The need of a particular pavement section for an overlay and perhaps, the required thickness of the overlay. The degree of variability of the materials along the roadway. Potential locations of voids beneath the surface layer. The load transfer efficiency across joints in concrete pavements. The elastic properties of the various pavement layersand the roadbed soil. am???" The ability of the pavement structure to support traffic loads at the posted speed limits. 7 7. The effect of seasonal variations on the load carrying capacity of the pavements. 8. The need for, and perhaps the type of rehabilitation activities. 9. The in situ stress sensitivity of the paving materials. 10. The effectiveness and benefits of various rehabilitation techniques. The mechanistic analysis of pavement deflections to infer the structural properties (moduli and Poisson’s ratios) of the various layers is referred to as "backcalculation of layer moduli". The task of backcalculation of layer moduli, however, is a difficult one. This difficulty is related to several reasons including: 1. The variability of the pavement materials along a given stretch of roadway. 2. The changing characteristics of the pavement materials due to seasonal changes, time, and temperature. 3. The nonlinear behavior of the paving materials. 4. The lack of accurate information concerning layer thicknesses and depth to stiff layer (e.g., bedrock). 5. The existence (in some cases) of a very thin layer (e.g., a one-inch debonding layer between an original pavement and an overlay). Nevertheless, backcalclation of layer moduli techniques have been developed and have been successfully applied to both flexible and rigid pavements. Most of these techniques are based on elastic layered concepts and are directed at the incorporation of NDT data collected on flexible pavements and at mid-slabs of rigid pavements. Elastic layered analysis does not adequately model discontinuities such as joints and cracks in rigid pavements. The NDT equipment used to measure pavement deflections differ in the methods used in applying loads to pavement and in the number and location of 8 sensors needed for measuring the pavement response. The various types of NDT equipment can generally be divided into five categories as presented in the next section. 2.2 NONDESTRUCTIVE DEFLECTION TESTING DEVICES NDT devices can be categorized as follows: Static Deflection Equipment - Static deflection equipment measure pavement deflections or rebound due to the application of a gradually increasing or decreasing load. This type of equipment includes the Benkelman Beam (Moore, et al., 1978; Asphalt Institute, 1977; Epps, et al., 1986), Plate bearing test (Moore, et al., 1978; Nazarian, et al.,. 1989), Dehlen Curvature Meter (Guozheng, 1982), Pavement Deflection Logging Machine (Keneddy, et al., 1978), and C.E.B.T.P. Curviameter (Paquet, 1978). Several technical problems are associated with this type of equipment including: a) ' it is time consuming and laborious. b) The test requires closing the pavement section to traffic. c) Deflection can be measured only at one point (special arrangements need to be made to measure the deflection profile). (1) The test presents hazardous conditions to both the traveling public and the test operators. Automated Deflection Equipment - Automated deflection equipment delivers a gradually applied load to the pavement structure in an automated mechanism. This type of equipment includes the La Croix Deflectograph (Hoffman, et al., 1982; Keneddy, 1978) and the California Travelling Deflectometer (Roberts, 1977). The technical problems associated with this type of equipment are 9 similar to those associated with the static deflection equipment. Steady-State Dynamic Deflection Equipment - Steady-state dynamic deflection equipment (also called vibrators) produce a sinusoidal vibration in the pavement with a dynamic force generator (typically rotating eccentric load). The most popular devices include the Dynaflect, the Road Rater, the Cox Device, the Waterways Experiment Station (WES) Heavy Vibrator, and the Federal Highway Administration (FHWA) Thumper (Scrivner, et al., 1969; Smith, et al., 1984, Moore, et al., 1978). Each of these devices has some limitations and advantages that are addressed elsewhere (Bush, 1980; May, 1981). Impulse Deflection Equipment - Impulse deflection equipment delivers a transient force impulse to the pavement surface by means of dropping a weight. The most popular devices include the Dynatest Falling Weight Deflectometer (FWD), the KUAB Falling Weight Deflectometer, and the Phoenix Falling weight Deflectometer (Nazarian, et al., 1989; Hoffman, et al., 1981; Bohn, et al., 1972; Crovetti, et al., 1989; Claessan, et al., 1976). Recently, FWD devices have become popular and they are being used by an increasing number of State Highway Agencies (SHA). Wave Propagation Equipment/Method - The wave propagation method (also called surface wave testing) has been used with some success to backcalculate the layer moduli of pavements. The method is not very widely used because of the complex data analysis procedures associated with it (Thomas, 1977). A study carried out by Lytton et a1. (1990) lists in detail the characteristics, 10 operational costs, and data analysis techniques associated with different types of NDT equipment. The NDT equipment have been rated after assigning weights to various important factors associated with the use of the equipment. The study concluded that the FWD is the best equipment available for simulating the actual traffic loading and further ranked the Dynatest Model 8000 as the best FWD equipment available. Since the backcalculation technique presented in this thesis is based on NDT, only a brief summary of the wave propagation method is presented in the next section. The concepts regarding deflection testing and backcalculation of layer moduli, on the other hand, are presented in greater detail in subsequent sections. 2.3 SURFACE WAVE TESTING The surface wave testing technique involves the measurement of the velocity and length of the surface waves propagating away from the point of application of an impact load (Nazarian, et al., 1983; Nazarian, et al., 1984; Robert, et al., 1986). This technique was pioneered by the German Society of Soil Mechanics in the late 1930’s (Bernhard, 1939). The wave propagation theory is based upon the fact that wave velocities in a homogeneous and isotropic half space subjected to external impact load can be expressed by the following equation: v = . rm (24> where E = modulus of elasticity; p = mass density; and a = a coefficient that is a function of Poisson’s ratio of the medium. Upon excitation by an impact or a vibratory load, three types of waves are generally transmitted through and along the pavement. The three wave types and the percent of a“? - r... r3) I1." 124' in" I!" 1 1 the applied energy that is dissipated by each type are tabulated below (Miller, et al. , 1955). Have type Energy dissipation (X of applied energy) Culpression (P) 7 Shear (S) 26 Rayleigh (R) 67 For a multi-laycr pavement structure, analysis of these waves is a complex proposition. Extensive mathematical analysis of these waves can be found elsewhere (Thomas, 1977). The remaining part of this section summarizes two techniques that can be used to induce the three types of waves in a pavement structure and the advantages and disadvantages of the surface wave method. 2.3.1 Impulse Load Testing . In this technique, an impulse (impact) load is applied at a point (called the source) on the surface of the pavement section. Any impact device (e.g., sledge hammer or impact hammer) can be used for this purpose. Upon impact, the three types of waves will propagate away from the source either along the pavement surface or with depth. At an interface between two pavement layers, the wave fronts undergo both, reflection and refraction. During the test, the time of arrival of several wave fronts at different points along the pavement surface is recorded in order to measure the travel paths of these waves through different layers and to deduce their velocities in each layer. This method is not applieable when high-velocity stiff layers, as is the ease for flexible pavements, are encountered at the top and the layers grow progressively weaker with depth (Moore, et al., 1978). As such the method cannot be 12 applied to backcalculate the layer moduli of pavements. 2.3.2 Steady-State Vibration In the steady-state vibration technique, a vibratory source is placed at an arbitrary point on the pavement surface and a vibrating load (typically sinusoidal) is applied at the source. In this technique, the phase lag (the time delay which occurs between the motion of the pavement and that of the source as the waves travel away from the source) is measured. High frequency waves have shorter wave lengths and hence can penetrate only the surface layer. On the other hand, if low frequency waves are used they will have lengths several times that of the pavement thickness and their speed will be determined by the properties of the roadbed soils. Intermediate wave lengths can be used and their speeds can be related to the elastic properties and thicknesses of the underlaying layers (Jones, 1960). 2.3.3 Advantages and Limitations The surface wave test method has several advantages over other NDT methods. These include: 1. Layer thicknesses of the various pavement layers can be calculated and as such no assumptions or approximations are required (Nazarian, et al., 1989). Therefore, the method is more pertinent when the dimensions of the structure is not known (Robert, 1986). 2. The depth to bedrock as well as the bedrock modulus can be accurately calculated. These two factors represent a major source of error for the deflection based backcalculation technique (l-Ieukelom, et al., 1962). 3. The elastic modulus of a thin or a thick asphalt concrete (AC) layers can be estimated. Variations of the modulus within, any paving layer can also be estimated. The method has the potential of being fully automated at a later l -.4 5'" 13 time (Hiltunen, et al., 1989). 4. The test method and equipment are simple to operate, but sophisticated data analysis is required (Robert, et al., 1986). The test result pertain to a wide area and not necessarily to the local pavement properties in the vicinity of the applied load (Watkins, et al., 1974). The major disadvantage of this method is that the testing and data reduction cannot be performed rapidly. It takes about 20 minutes to perform one test whereas a deflection test can be performed in less than 2 minutes (Wang, et al., 1989). Also the moduli obtained are for low strain levels which may not be an accurate estimate of the moduli under actual traffic loading (large load may cause the pavement materials to exhibit stress-dependent behavior). At best, the method is presently suited for project level surveys only (Nazarian, et al., 1989). Interpretation of the data is an extremely complex process which can only be undertaken by experts. The test results are not necessarily unique and the method works only for a few special structures (Thomas, 1977). 2.4 NONDESTRUCTIVE DEFLECTION TESTING Applying a known force to measure the deflection response of a pavement structure is the essence of NDT. The earliest device used in the United States (U .S.) for measuring pavement deflection is the General Electric Travel Gage in 1938 (Moore, et al., 1978). The tests showed that the pavement deflection can be measured to a depth of 21 feet and that the main contribution to the total deflection comes from the upper 3 feet of the structural section (Haveem, 1938). During the WASHO Road Test (1954; 1955) an improved version of the General Electric Travel Gage, incorporating a Linear Variable Differential Transformer (LVDT), was used. In 1952, A.C. Benkelman developed a simple and 14 easy to use instrument for measuring pavement deflections called the Benkelman Beam. The Benkelman Beam is still widely used in many countries across the world for measuring pavement deflections. Recent years have witnessed the development of equipment with better capabilities than the Benkelman Beam and as a result new NDT methods have been developed. Currently used techniques can be placed into one of the following categories based on the method by which the deflections are induced: 1. static force-deflection 2. dynamic steady-state vibrations 3. dynamic impulse force-deflection 2.4.1 Static Force-Deflection In this procedure, the response of a pavement structure to gradually applied or gradually removed loads is measured. The force may be applied by a slow moving vehicle of known weight or through a rigid plate of specified diameter that is part of a stationary loading frame. Static or quasi-dynamic measurements of either rebound or loading deflection are made by making the load vehicle pass a point located on the pavement surface at a creep speed. Deflection and rebound deflection testing procedures have been published by AASHTO (1982) and the Asphalt Institute (1977), respectively. 2.4.1-l W The basic advantage of the static deflection method can be attributed to its simplicity, cheaper equipment, and low maintenance costs. The disadvantages are: 1. It is difficult to obtain an immovable reference point for making deflection measurements. 2. All the devices measure only a single deflection making it difficult to obtain valuable information regarding the shape and size of the deflection basin. 15 Moreover, no information on the critical strain in the upper layer is obtained. 3. The automated beam equipment is further handicapped by the fact that it is difficult to test a specific point on the pavement. Further if the deflection basin is large, the reference point may be located within the basin itself. 4. The method is suitable only for use with static analysis and dynamic effects cannot be analyzed. 2.4.2 Dynamic Steady-State Vibrator Essentially all steady-state vibrator equipment induce a steady state sinusoidal vibration in the pavement using a dynamic force generator. The dynamic force is superimposed on the static force exerted by the weight of the force generator (Figure 2.1). The dynamic load causes the pavement system to vibrate at the same frequency as the load. The deflection response of the pavement is usually measured with inertial sensors. Velocity sensors (called geophones) are commonly used, although some equipment make use of accelerometers as well. Many of the devices can vary both the amplitude and the frequency of the excitation (Moore, et al., 1978). 2.4.2.1 - mi R f v Under static loads, pavement deflection is normally proportional to the applied forces, and substantial recovery is obtained when the load is removed. Dynamic response is no different in that, at any specific driving frequency, the amplitude of the dynamic deflection is approximately proportional to the amplitude of the applied force (Moore, et al., 1978). Green and Hall (1974) obtained results using a 16 kip vibrator at three different driving frequencies (Figure 2.2). The test results showed that the deflection is almost proportional to the loads at 15 Hz and 40 Hz and somewhat non linear at 10 Hz. The overall rigidity of road construction, S, defined by Van der Poel (1951) 16 .88. :3 Lo .2025 850:...» 8.8 25m 383 Masai? a .6 59:0 .33»... ._.~ 2:3... 11.92.... mecca a 0.25 - 1 hzm§w>a 17 .32 ..a 6.8.9 a: 9. as .2 .2 a segues... 8.. 8:8 eeeecuvefi 3.5. .2 been 2:. 493.. v. u. a. = o a .w 00 “We \ X 38.0 \ «1 av . urn. .. \ \ \\‘\ uIO-l 1 .¢Oh(II.> $3.10. CI. 85.; 0‘8.(PI° UIU’ (P‘O «85.08 18 as the amplitude of the dynamic force required to produce a unit amplitude in the deflection of the pavement surface. He pointed out that S is not constant, it depends upon the driving frequency. The Kelvin model has been used to represent the pavement response to. dynamic loading as shown in Figure 2.3 (Lorenz, et al., 1953; Van der Poel, 1953). The equation of the motion and significance of different parameters involved in the model can be found elsewhere (Baladi, 1976; Baladi, 1979; Taylor, 1978; Thompson, 1972; Heukelom, et al., 1960; Heukelom, 1961; Szendirei, et al., 1970). 2.4.2.2 v n n Lirn i The advantages of NDT are: l. Accurate deflection basin measurements can be made with respect to an inertial reference. 2. Steady state dynamic deflection devices correlate well with the static deflection measurements. Many agencies make use of these correlations for pavement evaluation. The disadvantages are: 1. These types of measurements represent the stiffness of an entire pavement structure. The separation of the effects of all the pavement components with measurement of the deflection basin has not yet been accomplished (Nazarian, et al., 1989). 2. The commercially available machines operate at light loads and hence the pavement is not stressed to traffic loads. As a result, the effect of any non- linearity in the paving materials is neglected (Fell, et al., 1972). 3. The steady state deflections are observed to be greater in magnitude than rebound deflections for Bankelman Beam (Hoffman, et al., 1981), while l9 Force generated Fer-ea exerted on pave-en: (1') Pennant S '- ~- Effective Haas (K) helping (C) nut! Supper: '""i"?;’%§‘/71’2’47?“ Z/.”////////// // 7W / ’/ 144/w Figure 2.3. Mass-spring-dashpot representation of a pavement structure subjected to a forced dynamic vibration (Lorenz, et al., 1953). 20 the deflections under moving vehicle are found to be smaller than those for equivalent static loads (Lister, 1967). 2.4.3 Impact Loading Impact loading devices deliver an impulse force to the pavement surface and measure the transient response. Force impulses are normally generated by dropping a known weight from a known height on a plate placed on the pavement surface. Inertial motion sensors are normally used to record the pavement response. In the U.S. , the Cornell Aeronautical Laboratory (CAL) was the first agency to use a trailer mounted force generator for impulse loading (Moore, et al., 1978). The pavement properties can be investigated by using the classical Kelvin single degree of freedom (SDOF) system. Fourier transform of an instantaneous impulse response deflection can provide complete information regarding the steady- state frequency response. Practically, however, it is impossible to generate an instantaneous impulse. Based on their test results, Szendrei and Freeme (1970) and Moore et al. (1978) concluded that a load pulse duration must be less than 1 msec to be considered instantaneous. Longer force impulses do not contain all the steady-state frequency response. Analytical treatment of the instantaneous force impulse of the classical model can be found in the literature (Richart, 1970; Tayabji, et al., 1976; Hansen, et al., 1956). 2-4-3-1 W The advantages of the impact load test include: 1. The loading most closely resembles actual traffic loading (Hoffman, et al., 1982). Therefore the shape of the deflection basin and hence the developed strains closely reflect those due to actual traffic load. 2. The actual duration of the test is only a few minutes and the measured data 2.5 21 gives sufficient information regarding the deflection basin to investigate the layer properties (Moore, et al., 1978). Results can easily be correlated with those of the static load test. The test equipment is simple to operate and can be maintained at reasonable costs. The disadvantages are: Complex analysis is required to model the response because of the difficulty of producing an instantaneous impulse (Taylor, 1978). Analysis of longer force impulses are even more complex. It is a problem to obtain the response in the low frequency range because of the low output characteristics of the motion sensors. DEFLECTION RESPONSE OF PAVEMENTS Based on review of the literature, certain concepts regarding the deflection response of the pavements can be expressed (Moore et al., 1978; Taylor, 1978): l. A maximum tolerable deflection level can be assigned to each pavement structure. This level is typically a function of the layer thicknesses and properties and the traffic load and volume. Overlaying a pavement will reduce its deflections. The deflection history of a well designed pavement can be traced through three phases (Figure 2.4): a. 111mm: Just after construction, the pavement undergoes consolidation and the deflections show a slight decrease. b. W: The deflections remain constant or increase slightly. c.«Eajlu1e_phase: The deflections increase rapidly. 22 Dellactlon —e —-"""""""1 Functional Phase Initial Phase Failure Phase Time ——> Figure 2.4. Well-designed pavement deflection history curve (Moore, et a1. , 1978). 23 4. The deflections of a flexible pavement increase with the increase in temperature of the bituminous surface depending upon the thickness of the bituminous layer. 5. The deflection history of the pavement varies throughout the year depending upon the environmental factors (Izada, 1966). A typical annual deflection history of a pavement subjected to frost and spring-thaw actions can be divided into four periods (Figure 2.5): a. A deep frost period when the pavement is frozen. b. A spring-thaw period during which the pavement deflections rise rapidly. c. A rapid strength recovery period during which water from melting frost starts draining and evaporating from the pavement. d. A relatively dry period during which the pavement deflections level off. 2.6 INTERPRETATION OF DEFLECTION DATA Interpretation of NDT deflection data is a complex and difficult task. The techniques used to extract useful information from deflection data can be divided into three basic groups 1. Empirical Analysis 2. Rational Analysis 3. Mechanistic Analysis 2.6.] Empirical Analysis In the past most pavement design procedures were empirical in nature. An empirical approach relies upon the results of past experiments and experience. Generally, deflections are directly related to the pavement conditions and other 24 .Gco_ Jean: 338:2. .5538 :o :88“ me 8:25:— .n.m 23E m2; V A V xllellvA :5“— 032.... .253 TthEzm Bach mctam floor... .853 SNOIlOE‘HEG >._o>ooo._ £9.25 32m we noted _ .AIIIIV.“ V .30. £9.23... «mo: ammo . _ o 0.5 >..o>ooo._ £9.23 a o . n: he noted Eon. __o veron— 0t" 1 25 variables such as traffic, pavement type, and environment factors. Results of a number of experiments are used to obtain a relationship between the variables and the outcomes. The relationship is typically not supported by theory. Statistics rather than the phenomena shaping the results are given more importance. The Utah Department of Transportation (UDOT) procedure (Molenaar, et al., 1982) is one such example. The maximum deflection (DMD), the Surface Curvature Index (SCI), and the Base Curvature Index (BCI) are compared to the acceptable values of these parameters which are inferred from a long term pavement performance (LTPP) (Figure 2.6; Table 2.1). The California DOT (1973), Asphalt Institute (1974), Oklahoma DOT (AASHTO, 1972), Louisiana DOT (Kinchen, et al., 1977), and Texas DOT (Brown, et al. , 1970) methods are other such examples where the pavement deflections have been utilized directly to infer information regarding the pavement condition and its capability to carry projected future traffic. 2.6.2 Rational Methods Rational methods utilize basin properties such as spreadability or representative structural properties to describe the pavement strength (Lytton, et a1. , 1990). The representative structural properties of a pavement are normally taken as the effective thickness of the pavement (McComb, et al., 1974; Kinchen, et al., 1980), effective thickness of the asphalt concrete and base courses (Vaswani, 1971), or the effective modulus of the pavement (Asphalt Institute, 1977; Lytton, et al., 1990). 2.6.3 Mechanlstlc Analysis A mechanistic approach refers to the calculation of induced stresses and strains to determine the response of the pavement structure to applied load. In order to calculate the response of the pavement structure, certain fundamental material properties along with the layer thicknesses must be known. The results obtained can ASE :E .o gaze—es: 8:38:— =wfioe .365 :5: Sec .393. .8325” .5826; 832a 2 303888 Ems eouoocoo Co om: .c.~ 0.59". O... a<20m c0 25:. mama-m.“ 25:. cwbkwmo 9—0 GZOEm wo<¢om=m o2< hzw2m>> mo§ go 93 .2265: mSPerm 26 go weapon»: .55 use .2265: Emzm>§ James moéomao vim; Emzm>> mosh—03m 02¢ Emu-2mg}. :lO NOLUGNOO (SIP-‘0 MVAHI'IO (311m) XBClNl (emu) MOUSE-1:30 (0W0) WWXVW asmvm aovsms mums narrative .A-A.3- ..-: are .KU.:~::PCV Ere-re.:=.-=q~ neat-n: tansenvre~r~h¢v .hae sale-apneeee-env. .~ .hr. flasks-Ah. 27 EEO—0 flu ”8:3 sausage .8 25% 88m age—eon S .D - one 75. I an 38m 55938 535 Susana u a 223 .3 n 5 x85 wagon 9E canto 223 .83 .10 one .338 333. can . 38 nan sans unmade can 3 Box 2on— .. Son u Um x35 unease 8am DE 85 can u a .33“ ROM .5 as ham fl .— 83m 883.8 page ..o .8 £8 53 ”ago 083m . 5 \ no. é a a QR 62558... a seesaw some 9E +eoae~+eQé~+=eu< 32 as an 88% E .o 3.33% banana—35 «a 2.550 .8 33 BE .9358 Edam 58.38 8 8535832 82.. Esau: 38.8 gag .292 :3 .o .mocogezv 28082:. 539 Snooze—o mo S885 Ado—nap. 28 thus be explained by theory and changes in response due to changes in any variables can be predicted more rationally. The advantages of such an approach include (Mahoney, et al., 1991): l. The accommodation of changing loads. 2. The ability to account for changes in materials and environmental conditions. 3. Improvement in the reliability of performance prediction models. 4. Better assessment of the performance of various paving materials. The aim of backcalculation methods based on mechanistic analysis, is to backcalculate the layer moduli of different paving layers. Which in turn are used to calculate stresses and strains induced in the pavement structure due to a given load. Mechanistic properties can be used for both the design and evaluation of pavement structures. The mechanistic design procedures use laboratory determined material properties to calculate a layered system response to an applied load. The mechanistic evaluation procedures, on the other hand, uses the measured pavement response under a known load to backcalculate the layer properties (Robert, et a1. , 1986). The two procedures are illustrated in Figure 2.7. In general mechanistic-based backcalculation procedures require computer based solution. Some of the better known mechanistic models are discussed in detail in Section 2.8. 2.7 SPECIAL PROPERTIES OF PAVING MATERIALS Stress sensitivity of unbound materials and temperature dependency of asphalt concrete have considerable affects on the backcalculated layer moduli. These two properties of paving materials are discussed next. 29 5222:; 325%; go £335 82%” ”:22, 2.25350 gum oSmE eowuo~vcmuuucugu . mmeonmum «5:253 I swam? pone: vacuum; a go cause =m.m= moansou .m "meoFuuopmme «unmeam «cameo: .N :32 5.. 33¢ .— ueosm>aa a mo mcvumwk m_mum-—P:u mmeonmom covunuvcuuuocngu 333nm ' Emcee—.3 c.~gum upnueo>euo¢ mmucum coauw>oo u m "even; use so ummoa gm: moansou .m “coauuseomuu pavxo nuance: .N wane. _o.xe a_aq< .~ all) Tlllllr al— cueaum m a mo cwumoh eoaueonoa 30 2.7.1 Material Non-Linearity The response of common highway materials to traffic loading typically includes elastic, viscoelastic, and plastic components. During the initial cycles of stress, at slow loading rates or high stress levels, the viscous and plastic components may be dominant (Dehken, 1978). The stress-strain response of different highway materials is different under changing stress conditions depending upon their properties and is discussed separately for each material type. 2.7.1.1 W The resilient modulus of sands andgravel is reported to increase with confining pressure. The magnitude of repeated deviator stress, unless high enough to induce shear failure, has no effect on the resilient modulus (Dehken, 1978). Biarez (1962) suggested a relation for the resilient modulus (MR) after performing tests on uniform sand in a triaxial apparatus: MR 3 K1 (6)9 (2.2) where K, and k, are material constants, and 0 is the sum of the principal stresses (identical to the stress invariant 1,). If 0 is increased in a manner that (a, - 0,) are constant essentially there will be no change in the modulus but the model fails to address this condition. Trollope, et al. (1963) observed that the rebound modulus increased with confining pressure and also confirmed earlier observation that as long as failure conditions are not approached the modulus was not affected by the axial stress. Morgan (1966) performed repeated load triaxial tests on two sands and observed marked increase in resilient modulus with increasing confining pressure and a slight decrease with increasing deviator stress. He also stated that the resilient Poisson’s 31 ratio remained unaffected by both the deviator stress and the confining pressure. Mitray and Seed, et a1. (1964; 1969) found the M, of sand subjected to triaxial repetitive load tests can be expressed as: M, = [(3.039 (2.3) where a, is the confining pressure and K3 and K4 are material constants. Equations 2.2 and 2.3 were also found to apply to gravel with changed values of exponents K2 and K,. Jones (1960) observed stress-dependency of the modulus of sand subjected to dynamic (vibratory) tests. Hardin and Black (1966) found that the shear modulus, G, could be represented by: G = KS (6)“ (2.4) where 0 is the sum of normal stresses. The exponent K,5 was 0.6 for 0 < 42 psi. and 0.5 for greater values. Stress sensitivity of cohesionless roadbed soils of Michigan was confirmed by Baker (1978). 2.7.1.2 iv i Resilient modulus of cohesive soils is reported to decrease with increasing deviator stress and is little affected by the transverse stresses (George, 1969). Seed, et a1. (1972) reported results of repeated load triaxial tests on silty clay. The resilient modulus was found to decrease rapidly with increase in the deviator stress up to 15 or 20 psi, any further increase in deviator stress resulted in an increase of the modulus. The variation in modulus between deviator stress values of 3 and 15 psi was found to be about 400 %. _ Kallas and Rilley (1977) found that in repeated compression tests on silty clay, the values of MR increase slightly with increasing confining pressure and decrease more markedly with increasing deviator stress. Sparrow and Tory (1966) performed H 32 in-situ tests on medium plastic clay and reported an increase in modulus with increase in depth and offset radius. The secant modulus was found to decrease with increase of the major principal stresses, the effect being more pronounced at lower stresses. These results were confirmed by Brown and Pell (1967). The most commonly used model representing the stress dependency of fine grained soils is reported by Young and Baladi (1977) as follows: Mn = Klarogc’ (2'5) where on = deviator stress; and K, , K, = material constants. Some mechanistic pavement design procedures use non-linear models. For example, MICI—IPAVE and ILLIPAVE computer programs use a bilinear model for cohesive soils. The relationship is expressed as: M = ‘2 + kslkr - (01 - 03)] for k1> ((11-03) (2.6) ' k2 + 4““: ' 0,) ' k1] for ki‘ (or-03) where (a, - a3) = deviator stress; and k,, k2, k,, and k, = material constants. The relationship is illustrated in Figure 2.8. 2.7.1.3 W Fossberg (1969) conducted tests on highly plastic clay stabilized with lime and concluded that the resilient modulus increased with increasing confining pressure and decreased with increasing deviator stress. Mitchel, et al. (1977) observed a decrease in MR of cement stabilized sand with increase in deviator stress. Wang (1968) found that the resilient modulus of a silty clay stabilized with cement increase with increasing confining pressure and decreased with increasing deviator stress. Both researchers concluded that cement stabilized materials behave essentially linearly 33 The Resilient Modulus (pal) ,.._ 1:1 The Deviacor stress («71 - 03. ps1) Figure 2.8. Typical variation of resilient modulus with repeated stress for cohesive roadbed soil (Ming-Shan, 1989). ll in inns Chat T3116“ J Bot. 34 under repeated flexure, but non-linearly under compression. 2.7-1.4 mm Asphalt cement exhibits a linear viscoelastic response, but asphalt aggregate mixtures were found to display non-linear viscoelastic behavior, or even behavior which was not viscoelastic (Krokosky, et al., 1963). Based on uniaxial creep tests, Monismith, et al. (1966) concluded that for uniaxial strain of less than 0.1%, asphalt aggregate mixes behavior is linear. Terrel (1967) observed that the MR of asphalt mixtures increases slightly with increase in lateral pressure and it decreases as the deviator stress was increased. Chatti (1987) and Baladi (1988) also found similar trends in another study. Through repetitive triaxial confining tests Trollope, et al. (1962) found that the rebound modulus increased almost linearly with increase in the confining pressure. 2.7-1.5 W Sparrow, et al. (1966) performed plate bearing tests on a roadbed soil consisted of a homogeneous silty clay of medium plasticity in a test pit. Tire test results showed stress-softening material non-linearity. This observation was also confirmed by others through independent tests on silty clay (Mitray, 1964; Wang, 1968; Terrel, 1967). The most marked non-linear effects were observed at pressures below 3 to 6 psi, at higher pressures a linear behavior was observed. Mitray (1964) observed stress hardening behavior for pavement structures consisting of a gravel base over a highly plastic roadbed soil. The response of the roadbed soil was observed to be of the stress softening type but the overall results were controlled by the over riding stress hardening behavior of the base material. Shifley (1967) performed similar test on an actual pavement structure. During its construction, different paving layers were subjected to plate bearing tests. The clay 35 roadbed was found to be of the highly stress softening type, but an addition of 11 in. thick crushed base aggregates changed the behavior to a stress hardening type, and the subsequent addition of a 2.4 in. thick asphalt concrete (AC) layer rendered the pavement response almost linear. The pavement when completed with a further addition of 4.8 in. AC layer showed markedly stress softening behavior. The test results at the AASHO road test (1962) confirmed the pattern of non- linearity described above under actual traffic loading. The non-linearity of the measured pavement deflections was found to be much less pronounced than those of the constituent materials in the laboratory. This phenomenon can be explained by the fact that in a layered system consisting of stress-hardening and stress-softening materials, the opposite effects counter each other, reducing the overall effect of material non-linearity. 2.7.2 Temperature Dependency Asphalt cement and asphalt-aggregate mixes are known to behave elastically at low temperatures, whereas the behavior tends to be viscoelastic at higher temperatures. Temperature dependency of the asphalt cement behavior was related to its stiffness characteristics (Heukelom, 1969; Mcleod, 1969). Van der Poel (1954) devised the Shell nomograph for estimating the asphalt layer stiffness with changing temperatures. Heukelom (1969) modified the Shell method to make it applicable to North American asphalts. Further, modification was suggested by McLeod (1969), who proposed the Penetration-Viscosity-Number (PVN) as a means of characterizing the asphalt. As a result of a study conducted at the state of Washington (Bubusait, et al., 1974), an empirical relationship was suggested for temperature adjustment of the asphalt cement modulus. The results of the study also indicated that the sensitivity of the backcalculated asphalt layer modulus to temperature depends on the condition of the pavement. New pavements being affected more than distressed ones. .‘EF J -.. -vn . 9 . " 36 The problem of temperature measurement and application of temperature correction to the asphalt layer modulus is discussed in more detail in section 2.9.1. 2.8 MECHANISTIC ANALYSIS MODELS The load-carrying capacity of a flexible pavement is enhanced by the load- distribution characteristics of its layered system. The system consists of various paving layers with the highest quality material placed at the top (Yoder, et al., 1975). The load distribution over the roadbed soil is achieved by building up thick layers of paving materials. Early calculations of stresses in flexible pavements were based on linear elastic theory. Since then efforts have been made to improve the basic models to incorporate non-linear and material damping effects under traffic loads. Although more complicated techniques to model the pavement response are now available, (e. g., dynamic, viscoelastic) layered elastic analysis is still widely used because of its simplicity and ease with which the required input data can be acquired in practice. Elastic layer analysis and some other models which have traditionally been used are reviewed in this section. 2.8.1 Layered Elastic Model Multi-layered elastic theory has been extensively used to model the stresses and strains in flexible pavements. The basic multi-layered system as pictured by Yoder, et al. (1975) is shown in Figure 2.9. The analytical solution based on elastic theory has several inherent assumptions including: 1. The material properties of each layer are homogeneous and isotropic. 2. Each layer has finite thickness except the roadbed soil, and all are infinitely wide in the lateral directions. 3. Full friction between the paving layers is developed at each interface . 37 hr.£r.ur ' .MQl as.» ’3‘ 8‘ . W2 Ila-Esau imbue-1 h--.£:ua Figure 2.9. Multi-layered elastic system (Yoder, et al., 1975). 38 4. There are no shearing forces at the pavement surface. Early calculations of stresses and strains in flexible pavements were based on Boussinesq’s equations originally developed for a homogeneous, isotropic, and elastic half-space subjected to a point load. The vertical stress at any point below the earth’s surface due to a point load at the surface is given by Boussinesq’s formula as a = k I— (2.7) 22 Z k = —3— ———1— (2.8) 2" [1+(r/z)2]”z where r = radial distance from the point load; and = depth. According to the above equation the vertical stress is independent of the properties of the medium and depends only on the vertical depth and radial distance from the load. By treating the whole pavement as a homogeneous and isotropic half- space, it is assumed that the contribution of pavement layers above the subgrade towards the total surface deflection is negligible (Yoder, et al., 1975). Burmister (1943; 1958) developed a solution for a two-layered elastic system. The materials in both layers are assumed to be homogeneous, isotropic and elastic. The surface layer has finite depth and is assumed to be infinite in the lateral direction, whereas the lower layer is infinite in both the lateral and vertical directions. Both layers are assumed to have a full contact and the surface layer is free of shear and normal stresses outside the loaded area. This model takes into account the properties of the materials above the subgrade. It also accounts for a uniformly distributed circular load which is a better representation of the wheel load than a point load. Further, the high stiffness of the surface layer has a pronounced effect on the vertical w 2.8 1‘ 39 stresses and strains. In contrast to Boussinesq’s solution, the stress gradients obtained by two layer theory are appreciably different from those obtained using a homogeneous half-space. Burmister and other researchers (Acum, et al., 1951; Jones, 1962; Peattie, 1962) expanded the solutions to three layer systems and presented the radial and vertical stresses in tabular and graphical forms. The advent of microcomputers made it possible to extend the linear elastic layer analysis to systems with more than three layers. Many computer programs are now available which can handle up to ten paving layers. Mathematical derivations and comprehensive explanation of concepts and assumptions pertaining to the elastic layer theory can be found elsewhere (Higdon, 1967; Westergaard, 1964; Timoshenko, 1987). Despite the availability of more advanced and complicated analysis models, the elastic layer solution is still widely employed for pavement analysis because of its practicality and simplicity. The basic advantages of elastic layer theory are: 1. It satisfies the laws of mechanics and is thus capable of making consistent calculations. 2. It is a relatively simple method and the calculation effort and capabilities required of computers are small. 3. Iterative variations of elastic layer theory that can approximately account for the non-linear variation of material properties in the vertical direction have also been developed, but for highly stress sensitive materials these may be inadequate. 4- Only two parameters, resilient modulus E and Poisson’s ratio ,1, are required in elastic layer theory. The results are not too sensitive to the value of 40 Poisson’s ratios and reasonable values are assumed in most cases. This leaves E as the only required material property, along with the layer thicknesses, as an input to perform the analysis. The limitations and inconsistencies related to the elastic layer assumptions are: 1. The behavior of the paving materials is not purely elastic. It has plastic, and visco-elastic components. 2. The stress-strain relationships for the materials are not linear for the range of stress levels encountered in pavements. Linear elastic theory may be inadequate to model highly stress sensitive materials when encountered. 3. Most paving materials are particulate in nature, hence, they are neither homogenous nor isotropic. 4. The stress-strain characteristics of most materials vary over time in all three dimensions. 5. Boundary conditions are quite complex and different from those assumed by elastic layer theory. 6. Actual traffic and the FWD apply dynamic loads to the pavement, elastic layer analysis is typically used with static loading. 2.8.2 Hogg’s Model Hogg (1938; 1944) modeled the pavement as a thin plate resting on an elastic subgrade. This model assumes that the vertical stresses. within the pavement structure are small and ean be neglected. The model has yielded good results for the backcalculation of roadbed modulus values. A major advantage of this model is that the roadbed modulus can be estimated without a prior knowledge of the characteristics of the pavement layers (Wiseman, et al., 1985). The model was used to evaluate three layer pavements using deflection data obtained by the La Crox-L.C.P.C. mu 41 Deflectograph. The results were satisfactory (I-oninck, 1982). Details about the model have been presented elsewhere (Wiseman, 1975; Wiseman, 1983). 2.8.3 Equivalent Thickness Model All equivalent layer models developed for estimating deflections of multilayered pavements share Odemark’s assumption (Odemark, 1949). Odemark’s assumption is used to convert a multilayered elastic system to a single layer elastic model. He suggested that deflections of multilayered pavement with moduli E., and layer thicknesses hi, can be approximated by a single layer thickness, H, and a single modulus E,, if the thickness H is calculated as; 2.9 H = 21‘ Ch, (E, / 5y” ( ) f i? 3 :2 II the equivalent thickness; hi = actual thicloiess of the ith layer; Ei = the elastic modulus of the ith layer; 1?.0 = the modulus of a single layer to which the multilayered system has been converted; and C = constant. The equivalent thickness method has limitations in its application and for certain types of pavements is known to give erroneous results (Kuo, 1979; Lytton, et al., 1979; Hung, et al., 1982). The method has the advantage of simplicity and speed of computation. Ulliditz (1978) used the equivalent layer model successfully for pavements with all linear elastic materials and also for pavements with non-linear roadbed soils. Lytton (1989) also used the model with some modifications for the analysis of flexible pavements. 42 2.8.4 Finite Element Method This is the only method which theoretically can adjust the stiffness of each element according to its own stress state. The method recognizes that for a nonlinear material the modulus is not characteristic of the whole layer but instead pertains to a point within that layer. MICHPAVE (Ming-Shen, 1989) is one such program which uses finite element method for linear and non-linear elastic analysis of layered system. In this program the limitation of modeling an infinite subgrade by deep fixed boundary has successfully been overcome by the incorporation of a flexible boundary concept. There are no known backcalculation programs directly based on finite element analysis because of computational time required. Instead, the data generated from finite element programs has been used to generate regression equations in order to account for the non-linearity of the paving materials (Hoffman, et al., 1982). Such an approach requires lesser time for backcalculation but all the limitations of regression type analysis of deflection data are applicable. 2.8.5 Dynamic Analysis Static analysis are generally used to backcalculate the layer moduli of pavements regardless of the load application mode. A load applied dynamically is not equivalent to the static loading and so should not be the stress and strain fields induced by each loading mode (Wiseman, et al., 1972; Stolle, et al., 1989). However, conflicting views regarding the magnitude of the error induced by the inertial response of pavements to dynamic loading can be found in the literature. Some researchers have reported significant error (Davies, et al., 1985), while others found the error to be insignificant (Roesset, et al., 1985). Davies and Mamlouk (1985) argued that the single-degree of freedom (SDF) models employed by most researchers are inadequate. They stressed the need to use 43 elastodynamic solution for the analysis of pavement deflection data, which can account for loading from multi-directions. In elastodynamics, Helmholtz equation for steady-state harmonic motion is used (Erigen, et al., 1975): C,2 grade(dive u) - Cf curl(curl u) + p20.)2 a = 0 (2.10) where c, and c, = the pressure and shear wave velocities, respectively; u = displacement vector; and a: = the circular frequency of the excitation. The displacement vector u can be expressed in the form: u(t) = u ‘e“’" (2'11) where u' = complex amplitudes of the displacement vector; t = time; and i = unit imaginary number. The wave velocities are related to the stiffness and mass density of the material by: r c = [ E(l-u) J3 (2.12) ’ (1+n)(l-2u)p 1 c = [ E J 2 (2.13) ’ 2(l+u)P where E, u and p are Young’s Modulus , Poisson’s Ratio, and mass density, respectively. A closed form solution for equation 2.12 is available only for a point load excitation on a homogeneous half-space. Numerical solutions, must be obtained for a multi-layered system. The usual assumptions of linear elastic material and isotropy are invoked. The soil and pavement layers are assumed to be unbounded laterally, bedrock is assumed at a finite depth and full bonding is assumed at all layer interfaces (Davies, et al., 1985). 44 Using the numerical solution technique presented by Kausel and Peek (1982) the in—phase and out-of-phase displacements at any location throughout the pavement can be obtained (Sebally, et al., 1986). Davies, et al. (1985), however, have stressed the complexity and difficulties associated with the analysis despite making a number of simplifying assumptions. 2.8.5.1 i D mi n l Loads applied by the FWD are transient in nature and not harmonic. The ‘ Fourier transform is used to represent the transient load by the sum of the harmonic load over different frequencies and amplitudes (Roesset, et al., 1985). Sebaaly et al. (1986) assumed a periodic loading impulse with period T, which they divided into a loading pulse-width, t,, and a rest period, TR, (Figure 2.10). The loading pulse width is a function of the loading device and pavement system properties, with typical values ranging between 25 and 60 msec for most FWD devices. The rest period Ta is chosen to be large enough such that the pavement fully recovers from deformation and hence the response of every drop is independent of the earlier one. The Fourier coefficients for the load impulse expansion are analyzed and then the phase lag, frequency, and amplitude of each harmonic response component are obtained. The harmonic responses are summed in the time domain to obtain the complete response due to the impulse. Similar solution was sought by Roesset and Young (1985). 2.8.5.2 gamma} 9n 9! Static and Dynamic Analyses 1. Static analysis of dynamically loaded pavements result in a significant error if the frequency of the applied load is approximately equal to the resonant frequency of the pavement system, or if the resonant frequency is so high that the inertial forces become dominant (Davies, et al., 1985). 45 [DAD TIME Figure 2.10. Assumed periodicity of FWD impulses (Sebaaly, et al., 1986). 2.8.6 46 Static interpretation of the deflections measured by an FWD test is reasonable when the depth of the bed rock is more than 60 feet. The resulting error increases when the subbase material is not homogeneous and/or when its stiffness increases with depth (Roesset, et al., 1985). Dynamic effects are less important for FWD loading because its load covers a wide band of frequencies. The results obtained by Roesset, et al. (1985) using dynamic and static analysis of FWD deflections showed that the difference in backcalculated moduli using two methods was small. In an independent study conducted on three different pavement structures in the United Kingdom, Tam, et al. ( 1989) concluded that the differences in the results of backcalculated moduli, for static and dynamic analysis using FWD deflection basins, were insignificant. Stolle and Hein ( 1989) observed from the results obtained by Sebaaly, et al. (1986) that better agreement exists between the deflection basins measured by FWD and that predicted by static analysis, than between that predicted by dynamic analysis (Figure 2.11). They also pointed out from a previous study (McCullough, et al., 1982) that while the accuracy of the measured deflection values is important to evaluate the roadbed modulus, the shape of the deflection basin is more important to accurately evaluate the pavement layer moduli. ' Nonlinear Elasticity Stress-strain curve for many paving materials are nonlinear. One simple method of dealing with such materials is to replace the elastic constants in the linear stress-strain relations with tangent moduli dependent upon stress or strain. The elastic constants can be obtained by using piece wise linear models. Such an approach is called a Cauchy elastic formulation (Chen, et al., 1985). 47 sunrace onseuceueur RADIAL DISTANCE Figure 2.11 Schematic showing measured and predicted surface deflection basins (Stolle, et al., 1989) 48 The Cauchy type of elastic models may generate energy for certain types of loading-unloading cycles. Hyperelastic models do not suffer from this drawback. Both of these models suffer from the disadvantage that they are independent of the stress or strain path, which is not true for soils in general. A more realistic and rational description is provided by the hypoelastic formulation in which the incremental stress and strain tensors are linearly related through variable material response moduli that are functions of the current state of stress or strain. Details of the three formulations and their respective characteristics can be found elsewhere (Chen, et al., 1985). Here, only the second order stress-strain relationships are reproduced (Uzan, 1993). The constitutive relationship for the Cauchy model is of the form a” = (C,I,+Czl,2+C,lz)bv + (C‘+C,I,)eu + Cseueu (2°14) while for the hyperelastic model it is o” = (2C111+3C2112+C312)50 + (c,+c,l,)eu + cseueu' (2.15) In the hypoelastic model, the stress rate is expressed in terms of the stresses and strain rates by: 60 == Coéuby + Clé,j + Czouéubu + 030.3.” + gave", + C,(oméw+éhow) + Csoflémby. (2-16) where ”v , 5,,- = components of stress and strain tensor; 0", , a", = components of stress and strain rate tensor; 1,, I2 = stress invariants; and C0, to C, = material parameters. 49 These formulations, while having a strong mechanistic basis, are not widely used beeause the material constants have little or no physical interpretation (Uzan, 1993). Uzan (1985) presented a simplified and general model for the secant resilient modulus of paving materials as: MR = klp‘ [3]“ [Ear (2.17) P. Pa Where p, = atmospheric pressure; rm, = octahedral shear stress; 0 = bulk stress; and kI , k2 , and k3 = material constants. This model is a simplified form of the non-linear Cauchy model, and though appealing, violates the laws of thermodynamics. In a recent paper Uzan (1993) presented a modification of this model where two more material constants have been added to account for the non-liner behavior of Poisson’s ratio and are derived by imposing the path independence of the strain energy density function. The five parameters for each paving layer cannot be directly backcalculated from the set of deflection data measured by FWD equipment. Hence, Uzan suggests that only k, be backcalculated, while k2 - k, be estimated from laboratory testing or existing data banks. This partially defeats the purpose of backcalculation, making non- linear models unappealing at present. 2.8.7 Viscoelastic Model Asphalt displays viscoelastic behavior, and since FWD loadings are essentially dynamic, some researchers have used viscoelastic models to obtain a more accurate representation. In general, even granular and cohesive material can be modeled as being viscoelastic. Viscoelastic material can be easily modeled by using a complex- 50 valued modulus (Wolf, 1985) as: E*(c->) = 5’0») + iE”(w)=E’(w)[l+i25] (2.18) where E', E', E" = complex-valued modulus and its real and imaginary parts; to = circular frequency; and = damping ratio. Granular materials are normally assumed to have a constant damping ratio (Uzan, 1993). Nonlinear viscoelastic models are used more and more for dynamic backcalculation but are relatively complex and hence have not been used widely in practice. The increased number of parameters makes it difficult to backcalculate all of them solely from FWD measurements. 2.8.8 Pavement Material Type and Choice of Analysis Model Differences in opinion regarding the degree of non-linearity of pavement materials can be found throughout the literature. Uzan (1993) recommended the use of complex moduli to model the viscoelastic behavior of pavements and recommended that data of the last few deflection sensors not be used in the backcalculation if the complex moduli tended to increase away from the applied load. He further emphasized the importance of using linear dynamic analysis when the bedrock is present at a shallow depth. For deep bedrock Uzan recommends the use of the static non-linear backcalculation as the state-of-the—art improves. Presently the use of non- linear or viscoelastic models require some of the material properties to be inferred from existing data banks or from laboratory tests. As a result, they have not found much use in practice, and backcalculation based on simple linear elastic models are still the most popular. 51 2.9 BACKCALCULATION METHODS Existing backcalculation routines can be classified into three major groups depending on the techniques used to reach the solution. These three techniques may have any of the forward analysis methods, discussed earlier, embedded in them. The first group is based on iteration techniques, which repeatedly use a forward analysis method within an iterative process. The layer moduli are repeatedly adjusted until a suitable match between the calculated and measured deflection basins is obtained. The second group, is based on searching a database of deflection basins. A forward calculating scheme is used to generate a data base which is then searched to find a best'match for the observed deflection basin. The third group is based on the use of regression equations fitted to a database of deflection basins generated by a forward calculation scheme. Some of the known backcalculation computer programs and their characteristics (adopted from Mahoney, et al., 1991) are presented in Table 2.2. 2.9.1 Iterative Methods The ultimate objective of most backcalculation methods is to find a set of moduli such that the calculated deflection basin match the measured one within a specified tolerance. This is usually achieved by minimizing an objective function which is commonly defined as the weighted sum of squares of the differences between calculated and measured surface deflections (Uzan, et al., 1989) i.e., minimize f = 2:1 “ltmlfwiy (2.19) where win = the measured deflection at sensor j; wjc = the calculated deflection at sensor j; and a,- = a weighing factor for sensor j. The flow chart (Lytton, 1989) presented in Figure 2.12, illustrates this process. The main steps of the iteration process are: 52 a as , : - - e82, _ 2.3. m u near...— Base 8> mice—ME 29586 83.2.32 eozm 2 has... 83m 528%: M .8388 38 2 3.2.3 e233. .2 m>F _ Step I. Step 2. Step 3. Step 4. Step 5. j 55 . Surface deflections at known distances away from the applied load are measured. Layer thicknesses, load application characteristics, and Poisson’s ratios for each layer are required to be input by the user. In almost all programs constant values of Poisson’s ratios are used. In order to start the forward calculation process, approximate layer modulus values (seed moduli) are required as input. Seed moduli are sometimes generated by the program using measured deflections and regression equations, or else they must be specified by the user. Some programs use a database approach at this stage to obtain seed moduli. The data specified in step 2 and the latest set of layer moduli are used to calculate surface deflections at the same radial offsets at which the deflections were measured. An error check is performed to assess if the measured and calculated surface deflections are within a specified tolerance limit. Different techniques are used at this stage to adjust the set of layer moduli so that the new set of moduli reduces the error quantified by the objective function. The method by which the moduli are adjusted is the main differentiating factor between most iterative procedure based programs. Steps 4 and 5 are repeated until the value of the objective function is sufficiently small or the adjustments to the layer moduli are very small. One of the problems faced with this approach is that the multi-dimensional surface represented by the objective function may have many local minima. As a “33““ the program may converge to different solutions for a different set of seed 56 moduli. Some programs overcome this problem by automated assignment of the seed moduli. Another problem is that the convergence can be very slow, requiring numerous calls to a mechanistic analysis program. ‘ An example of an iterative program is EVERCALC (Sivaneswaran, et al., 1991), which uses an efficient and general minimization method (Levenberg— Marquardt algorithm). The program seeks to minimize an objective function formed as the sum of squared relative difference between the calculated and measured surface deflections. EVERCALC is a robust, efficient, and accurate program, and uses the CHEVRON computer program for forward calculations. The "---DEF" series of programs use an assumed linear variation in _ logarithmic space to revise the layer moduli after each iteration. These programs employ a gradient search technique and the correct set of moduli is searched in an iterative manner. The CHEVDEF program (Bush, 1980) is one such example in which the CHEVRON program (Michelow, 1963) is used for forward calculations. A set of seed moduli are required to be input by the user in this program to start the iteration process. The simplified description of the process to find new layer modulus from an initial guess for one layer and one deflection is shown in Figure 2.13. For multiple deflections and layers a set of equations defining the slope and intercept for each deflection and each unknown layer modulus is developed as follows (Van Cauwelaert, et al. , 1989): log (deflection) = A}: + Sfi (logEi) (2.20) where A5, = intercepts; Sis = slope; j = 1,2,...ND (ND=No. of deflections); and i = 1,2,...NL (NL=N0. of layers with unknown moduli). 57 mm FROM LAYERED can»: I g I .- I a _, I I e I I g I I maceration-“suns: .. g 2 I - I I i I | l 3 l I I w I | I I I I I I l I I I ' I I I l V l l emu emu em LOG mom Figure 2.13. Basic process for matching deflection basins (Van Cauwelaert, 1989)- 58 p The linearization of the model in logarithmic space simplifies the search for new set of moduli. However the results obtained by these programs are highly dependent on the initial seed moduli. Some programs such as ELMOD (Ullidtz, et al., 1985) use an equivalent layer thickness concept along with Bousinesq’s equation in an iterative program to backcalculate the layer moduli. There are some other programs which are based on the elastic layer concept but which first estimate the subgrade modulus based on the deflections measured by the outermost senors. The subgrade modulus is then fixed and intermediate layer moduli are estimated using the middle sensor deflections and finally the AC modulus is inferred from the set of inner-most sensor deflections. The error in the estimated moduli of lower layers thus contribute to the errors in the moduli of the upper layers. 2.9.2 Database Approach In this method, a forward calculation program is used to generate a data base of deflection basins for different combinations of layer moduli, and specified layer thicknesses, material properties, pavement types, and loading conditions. The measured deflection basin is compared with the deflection basins in the database using a search algorithm, and a set of moduli are interpolated from the layer moduli which produced the closest calculated deflection basins in the database. The MODULUS backcalculation program (Uzan, 1985) is one such example which uses databases generated by WESLEA (Van Cauwelaert, 1989) program. The number of basins required to obtain a suitable database depends upon the number of layers and the expected moduli ranges provided by the user. Wide ranges require a greater number of basins to be generated than narrow ones. The generated deflection basins are then searched using the Hookes-Jeeves algorithm and a three-point Lagrangian method is used to interpolate the moduli. The program seeks to obtain a 59 _ set of moduli which will minimize an objective function defined as the relative sum of squared differences between the measured and calculated surface deflections. The program is known to converge always, although the chances of converging to a local minimum cannot be ruled out (Scullion, et al., 1990). The program performs a convexity test to determine the likelihood of having converged to a local minimum and the user is warned if this test is not satisfied. . Backcalculation based on a database search is especially suited when a large number of pavements with similar configuration are to be tested in continuation. For these situan'ons the data bank once generated can be used repeatedly to backcalculate moduli for all similar pavements, and the time required to generate the database can be minimized. This technique can be used with database generated from any linear or nonlinear program (Chua, et al., 1984). The results obtained are modemme accurate, but the accuracy of the results is sensitive to the expertise of the user and his or her knowledge of pavement materials. The COMPDEF (Chua, 1989) program also uses a database approach to backcalculate the layer moduli. The program uses a precalculated database of composite pavement deflection basins stored in a matrix, which is searched by an interpolation technique to find the layer moduli. 2.9.3 Statistical Analysis This method is similar to the database technique, the only difference being in how the database is used. The database is created by using any forward calculation routines, and then statistical analysis is performed to generate regression equations. These equations take the deflections as independent variables and attempt to predict the values of the layer moduli. Pavements of different configuration can be grouped separately to yield different equations for more accurate predictions. Different prediction equations are required for each pavement layer and pavements with a 60 different number of layers have to be treated separately. The LOADRATE program (Chua, 1989) belongs to this category and uses regression equations generated from a database obtained by using the ILLIPAVE non— linear finite element program (1982). This technique is best suited for agencies which deal with a limited and known type and configuration of pavements. Data bank generation to include all the expected combinations of pavement layers in the initial stages can offset this disadvantage to a large extent. Proper statistical interpretation of data can give reasonably accurate results. Once the regression equations are obtained this technique is simple, and extremely quick. The results on the other hand vary in accuracy depending on how well the database which was used to generate the statistical equations represents the pavement being analyzed. 2.9.4 Conversion of Backealculated Layer Moduli to Standard Conditions The modulus of the asphalt concrete surface course is significantly affected by the temperature and frequency of loading. On the other hand, the base, subbase and roadbed soil moduli are more affected by confining pressure, moisture, and stress levels. For a better interpretation of the modulus values, the backcalculated moduli should be converted to some standard conditions (moisture and temperature). The conversion to standard conditions is referred to as corrections to backcalculated moduli. 2.9.4.1 n n i The temperature correction procedure suggested by the Asphalt Institute is most commonly used for pavements with more than 2 in. thick AC layer (1977). The mean temperature of the pavement must be calculated for the same time when the pavement deflections are measured. This procedure requires exhaustive data 61 including: . l. The maximum and minimum temperatures for five days prior to the day the test is performed. 2. The pavement surface temperature at the time of NDT is conducted. 3. The frequency of loading or in case of FWD loading the time duration of the load impulse . 4. The percent asphalt cement content by weight. The chart in Figure 2.14 is used to determine the temperature at the top, middle, and bottom of the asphalt layer. The data in steps 1, 2, and 3 described above are required to use the figure. The mean of the three temperatures is considered as the mean temperature of the pavement. Southgate (1968) presented a slightly different procedure than the one described above to find the mean temperature of pavements having an asphaltic concrete layer thickness of less than or equal to 2 in. This procedure stresses that for thin asphaltic layer pavements, the hour of the day and the amount of heat absorption is more important than the maximum and minimum temperatures used in the Asphalt Institute method. Figure 2.15 is used to find the temperature on the underside of a thin asphaltic concrete layer. This temperature and that of the surface of the layer at the time of testing are then averaged. The frequency of loading is also required to use the Asphalt Institute equation (1982). The frequency in the case of a cyclic loading device, such as the Dynaflect or Road Rater, is the actual frequency of loading. In the case of a FWD device the frequency is obtained as: 62 do 'Hldafl .LV 3 UfllVBBd W3 J. 6N cc cm oo oo— ch ov— om. Akb— éaems “Rims 8.25852 .5833 38:85 .SN oSwE Ohwmab(& 2... o: 2: 2.. 8 3 on on 3 an on 2 o 2.. _ T _ _ . _ _ . _ . _ \ l 9...... I e .6. m. 1 a. 5:53 .5. S I on 5:58 I 3.. N: 5:58 on .6. 1 EEE: I 3 2.. u. 5&3 I o... 3.. : 55mm hzwzmza z. :Emo I on _ _ . — _ . . — . — — _ . I Ch 8“ 3a S“ can 2: 2: a: on. 2: a. an 3 on a uhwcahkzwazmh m2 zm* ‘ 8.93:2: + ecu w¥0m .X. N¥0m * 9.93:9... + l T 5.043102: IT l l 1- O (as) snlnpow paqpeou "I 10:13 4.4.1.1 112 Eff fA Thikn Errr anemen wihDiff nA mm The effect of incorrectly specified AC thicknesses on the layer moduli of pavements with different AC thickness was studied using three typical three-layer flexible pavements listed in Table 4.1. The AC thickness was varied between j; 10 % for all three programs. The results obtained from the programs are presented in Table 4.9. Examination of the results of the three programs indicate: 1. lnaccuracies in the AC thickness affect the AC modulus of thin AC layer pavements the most and thick pavements the least. The base modulus was affected the most for thick pavements and the least for thin ones. This observation was further investigated by Studying the change in the vertical stresses at the top of the base layer with inaccuracies in the AC layer thickness. The study reveals that owing to an inaccuracy of —10% in the AC thickness of the thin pavement, the increase in the vertical stress at the top of the base layer is about 6% , compared to about 20% for the thick pavement as shown in Figure 4.4. For EVERCALC, it can be noted that for the thick pavement, even for cases in which the error in the AC thickness is negative, the predicted AC modulus is always smaller than the actual. This is because of the relative inaccuracy of the older version of CHEVRON (used in EVERCALC 3.0), which becomes more significant for stiffer pavements. This discrepancy is highlighted in Section 4.10 where the results of different elastic layer programs are compared. 113 n.3- . o... co; ....««..- ....- ho. ......- o.o.- nos--1111! “ .... .... o... ...««. ....- o..- o...- «.o«- o.««. 0.. ... _ o... .... .... osm- «.«.- «.«. 3. o....- m...«. m2 _ r n. o... o... «.3. ....- ...... «.«- .....- «.o. no: m o... o... o... ...“. «...- ...o. .....- «....- ...«.- on n u w «.o .... o... 5..- oo. ...- ..v 3. o...- ms. _ o... o... o... of. o...- o.« «.«. o... o.«. no: _ ..o. o... o... o...« R.« «.... ..«.- ...? «... om .. H o... o... o... ...? o... o... «... o... .....- m2 o... o... o... «... .... .... o... «.... ...... no: 2v o... o... «...« .... ««.. an. o.« o.«. 0.. m- ..o. o... ..o. o..« o... ... «... o... «.n. m: o... ‘ o... o... as «.«. o... no. ...... ...... me: ..o. o... ..o. ...R . an. ... c..- ...: ...... um o.- «.o. a... «.o- 3.. an. «.« o... ...... o.«« ms. £88.33... 0... 2.96:... ..o.. «35$... 9.. 358% o... ... 88.0 e. 2... 2.62: .8»... 3.2.5.832... o5 ... Ecum— dé 03am. 114 .=e..8....oo..m «85.0.... U< 89.8... o. 0.... .93. 3.... o... ..o ..o. o... ... 9.2.... .85.? o... E 39...... ...... 2:9”. 3... «35.2....- o< :. .o...m c w o o e N c N. v- 0.. o.- c..- . . . . . . . _ _ ON- IO..- . .... m e u. u a _ m. 0 law W E .... - S s .3. w 32:26.. ..o...... um 32:25.. 25.0.! +- EoEoZ... «......- .o. an 115 4.4.1.2 E1131 of AC Thickness Eggpr 9n Eavgmgngs with Different AQ Stiffngs The medium thickness pavement listed in Table 4.1 was used to study the effect of error in the AC thickness for pavements having different AC stiffness. The AC moduli of 300, 500, and 800 ksi was used to simulate soft, medium and stiff pavements. The resulting errors in the different layer moduli due to the error in the AC thickness are presented in Table 4.10. As it was expected, for the same inaccuracy in the AC layer thickness, the error in the backcalculated AC layer modulus is higher for stiffer pavements for all three programs. The MICHBACK and EVERCALC programs have similar trends for the base modulus also (i.e., having greater error for stiffer pavements), but the MODULUS program gave erratic results not indicating any trends. The roadbed modulus remained insensitive to inaccuracies in the AC thickness for all three programs as observed earlier. 4.4.2 Base Thickness The percent errors in the backcalculated layer moduli due to an incorrectly specified base layer thickness are presented in Table 4.11, and Figures 4.5 through 4.7. The observations from these results for all three programs are: 1. The roadbed modulus is relatively insensitive to errors in the base thickness. 2. Both the AC and the base layer moduli are significantly affected by inaccuracies in the base layer thickness. With the base layer being affected the most. 3. The base layer modulus shows stiffening effect due to negative errors in the base layer thickness and vice versa- 4. The AC modulus was found to be positively related to the errors in the base layer thickness (i.e., negative errors in the specified thickness produce 116 ”.3- o... «.«.- «.«.- «.«.- «.«.- «....- ..o .... o... «.«. «.... «.«- ««.«- «.8. «.«.- om ... «... .... .... .....- «.«.- «.«.- ...- ...:- «....- m: o... o... o... «.«- 2.. «.«- .....- ...... «.«- no: o... o... a... «..- «.7 «.«- «.«.- «....- oé- om « .... .... o... «.... «o. ..o. «.«- o...- o.«- ms. o... o... o... ...... o...- o.« «...- o... «.«. no: a... o... o... o... ««.« .... «.«. ...... «.«. om o o... o... o... «.o. o... a... «... o... «... m... o... o... c... an ..« «o «... «.« ...o no: o... o... .... «.o. ..o 8s o...- o.« o.« 8. «- o... o... o... «o oo «.« «.... om «... ms. o... o... o... o.« «.«. ...... ...... «.«. .... no: «3.69: U< Ego-:3 .8 «85.0.... Ga. 358% o... ... «.83 o. one =ng .93. 33.83833 0... ... «atm— d. .v 033. 117 — .om c.0- n..- ...... n3- 2. ......- ..3.. can- m. 3H ...: nli 1! 3 # o... 2.. ...... o... ...-N- «.3- n...-- .... an. ...: n... ...... o" _ _ o... no. ...... o... n...- ...m.- ...:- 2 a... an o... o...- o. w o... ..o. n...- c... as. ...... ....- 2 ....- o... ..n ...... ... H o... o... o... c... o...- 3 ...o o... o... ...? o... a... o , a... a... ...... c... 2.. a... I. m. we. ....- ..n. ...... n. . ‘ o... I. e... a... an ....N E." w. a...“ 2.- .....- c.... 2. _ _ a. ...o «... o... ...S «.3 ...n n. 2."- ..fi. ...:- o.... 8. * h E .... ...n a... ...:N no: ...... a. ......- 37 ..2. ...... 8- . .liwitf ..l al-- I --.-.i l 1-. i31-;-.-!-!il-...,!!! l- f- 1i--l: I 3-14 no: R. a: ...: no: 0.... ...z ...: no: on a: ...: 3. v E” M m 3‘8... u< . .3052... 3.3 o... 5 Echo 2.8.0.. o. 0.... 22.0.... ..ob: 3.33—8.33 05 5 Echo Eco-.2. 2F ._ _ .v 2an 118 dosaoEoonm 30.5.3... 8.... 82.80... o. 2... 3.3.0... U< 3.23—«8.03 o... ... «go-cm .n... «Sm... GS woos-0...... omum :. Etm av cm cw O .- o o .... ON. an- Ge... _ _ . . _ _ _ Om.- m:..:oo_2 .... o..m * $65.20.... + cc. 3 56310.5...- - w on- m. V 0 w m. x -, I - I o m n.- . . s .\ W 4\ I am cc 119 .=0..8.-..00% 305—0.... 0...... 80.80:. c. 0.... 3.2.0... 0...... 00.230.08.03 0... ... ”...-...... .0... 0...»... 3... 0005.02... 003 c. ..o..m cc cn ON or c c .... ON- en.- ce- _ (J . _ _ . _ OOP- I Om- fl"! . .4 O m. , m w.- 8 B S 8 W O D. n .... 0.2055 x. ( Nxom * Mm 9.93on + -n r¥0m IT ¥0m 1.. .. om P 8.93:2: .... - . 8N 132 55.8583 5%.. .23. ......m 80:85 2 0.... 333:. 35.8. 3.53.8523 o... 5 385 ._ _ .9. 2:9... ...... £63 .26.. ....m ... .o..m 2. on em 3 o 3- an. an- ac. . . . . . . 4 cm- m:.:.oo.2 2. o..m + .... c..- 20320:... ... _. on. l l O ‘1‘ x \ \.\ (96) snlnpow paqpeou "I 10113 I I I I O O O O N v- 1'- O 0') 133 with each other, and the over-estimation of one results in the under—estimation of the other. 4. For all three programs the roadbed modulus was found to be very sensitive to error in the stiff layer depth. Therefore, the base modulus always showed an opposite trend than that of the roadbed compensating for the sensitivity of the adjacent layer. Thus, an over-estimation of the roadbed thickness resulted in the base modulus being under-predicted and the AC modulus being over- predicted, respectively. 4.5.2.2 ’ff 1 M I W Generally, the stiff layer modulus is more difficult to predict than the depth. This is especially true when the stiff layer effect is being observed not due to the presence of a hard layer but because of the soil overburden or stress stiffening characteristics of the roadbed soil. In the latter case the roadbed modulus gradually increases with depth, making the prediction of its modulus difficult. The effect of the modulus of the stiff layer on the backcalculated layer moduli are illustrated in this section. The three layer medium thick pavement (Table 4.1) was used to study this effect, and the depth to stiff layer was varied between 36 and 240- inch. 1n the first case the correct depth to the stiff layer (i.e. the same as that used to generate the deflection data using CHEVRONX) was used. The data was generated using a value of 1000 ksi for the modulus of the stiff layer, while for the purposes of backcalculation, the modulus was varied between 500 and 7000 ksi (error range of - 600% to 50%). The results presented in Table 4.16 indicate that: 1. Within the wide range of error in the stiff layer modulus, the observed error in the backcalculated layer moduli was reasonably small for the shallow stiff layer and almost insignificant for the medium and deep stiff layers. 134 Table 4.16. The effect of stiff layer modulus on the backcalculated layer moduli (stiff layer depth fixed). Stiff layer Stiff layer Error in stiff layer Error in backcalculated moduli (%) ll 1h modulus i. modulus % RMS dc” 0“ ) ( ) AC Base Roadbed 7000 -600 3.7 -5.3 8.3 4.9 Shallow (36 in.) 5000 400 3.5 4.9 7.7 4.6 1000 0 0.03 0.0 0.0 0.0 500 50 4.5 6.8 ~10.1 6.5 7000 -600 0.1 -3.2 3.8 -0.8 Medium (144 in.) 5000 400 0.1 -3.0 3.6 -0.7 1000 0 0.0 ~2.7 0.7 -0.1 500 50 0.1 2.4 -2.8 0.7 7000 -600 0.1 -1.8 2.0 -0.4 (240 in.) 5000 400 0.1 -1.7 1.9 -0.3 1000 0 0.1 -0.6 0.6 0.0 500 50 0.03 0.7 -1.0 0.4 .— E= 2. 135 The RMS error decreases as the stiff layer depth increased for the same degree of error in the stiff layer modulus. The example above was repeated for all cases, but this time instead of providing the actual stiff layer depth and fixing it MICHBACK was allowed to find the stiff layer depth iteratively along with the layer moduli. The stiff layer modulus was deliberately varied between the above mentioned limits, and the seed value for the stiff layer depth was provided by the regression equations. The results presented in Table 4.17 indicate: 1. The MICHBACK program compensates for higher values of the stiff layer modulus by reducing its depth keeping the stiffness of the other layers almost the same. In the case of an incorrectly specified stiff layer depth, the backcalculated moduli have a slightly larger error than that when the true depth is specified. This observation is more pertinent to the shallow stiff layer. The reason for this is that as the stiff layer depth decreases, the backcalculated results become more sensitive to the stiff layer pr0perties. At shallow depths, small errors in the backcalculated thickness coupled with the error in the stiff layer modulus can produce moderate errors in the backcalculated layer moduli. The above exercise indicates that the stiff layer modulus has a negligible effect on the backcalculated layer properties, especially when the stiff layer depth is not very shallow. It is further emphasized that the error range for the stiff layer modulus tested in the above exercise was extremely wide. Hence, making a reasonable guess regarding the stiff layer modulus will not have a significant effect on the backcalculated layer moduli. Also interaction between the stiff layer depth and modulus will only make the iterative process more complicated and prone to 136 Table 4.17. Effect of stiff layer modulus on the backcalculated layer moduli and stiff layer depth. Stiff layer Stiff layer Error in RMS Stiff la er depth modulus(ksi) modulus (%) dept 88'133’ 1‘1” ‘35 (2983.) 137 enon 5. Based on the above observations, no effort has been made in the MICHBACK program to include the stiff layer modulus as an unknown in the backcalculation process. 4.6 CONVERGENCE CHARACTERISTICS Newton’s method is a rapidly convergent and accurate optimization technique. The speed of convergence, in MICHBACK, has been enhanced by the logarithmic transformation of the gradient matrix as explained in Chapter 3. The convergence characteristics have been tested in this section using the deflection data generated by CHEVRON X. The MODULUS and EVERCALC programs being used for comparison are limited to a maximum of four pavement layers, and therefore no comparison could be made for the five layer example. The EVERCALC program uses the original version of the CHEVRON program for forward calculations, which being less accurate affects the backcalculated results. In order to make the comparison fair for EVERCALC program, the backcalculation was conducted by using theoretical deflection basins generated by both CHEVRONX and CHEVRON. The results where, CHEVRON generated data was used in the backcalculation, are denoted by EC-ALT. For all examples, surface deflections were rounded to the nearest hundredth of a mil. An improved accuracy was obtained for MICHBACK when the surface deflections were input to a greater precision, especially for the composite pavements. The other two programs do not allow the surface deflections to be input to a precision greater than hundredth of a mil. Although, such a precision is unrealizable in the field, this observation with MICHBACK indicates the sensitivity of the backcalculated results for composite pavements to even small changes in the measured deflections. 138 4.6.1 Three—Layer Flexible Pavements The properties of the three layer flexible pavements used in the analysis are listed in Table 4.1. The backcalculated results along with the maximum error in the moduli and the RMS error specified by Equation 3.23 are given in Table 4.18. The MICHBACK program yields accurate results for all three layers. MODULUS. on the other hand, has comparatively larger error for the base modulus, 4% being the largest. For other layers, the errors are smaller. The EVERCALC program has the largest error of the three programs, mainly because the CHEVRON program is used in the backcalculation algorithm. Also it can be seen that EVERCALC progressively calculates poorer results as the pavement becomes stiffer. This indicates that the difference between the modified and older versions of the CHEVRON program increase for the stiffer pavements. When the deflections generated by the old version of the CHEVRON program is analyzed EVERCALC (EC-ALT) also yields excellent results for three layer pavements (similar to MICHBAC K). 4.6.2 F our-Layer Flexible Pavement . The actual properties of the pavement are given in Table 4.2. The backcalculated results (Table 4.19) indicate that as the number of layers in the pavement increases MICHBACK clearly produces better results than the other two programs. EC-ALT results are comparable to those of MICHBACK but the largest error is more than 4% compared to less than 1% for MICHBACK. For pavements with more than three layers the MODULUS program yields a poorer result at least for one of the layers. MICHBACK produced consistently accurate results for all other four layerpavements tested as well. 139 Table 4.18. Comparison of the results of three programs for a three layer pavement. Pavement Backcalculated modulus (ksi) Max. error RMS Program type AC Base Roadbed 1n mggull error (%) Thin 497.9 45.0 7.5 0.42 .02 MICHBACK Medium 499.9 45.0 7.5 0.02 .03 Thick 501.2 44.6 7.5 0.84 .01 Thin 485.4 45.9 7.5 2.92 .37 MODULUS Medium 503.1 44.6 7.5 0.89 .11 Thick 485.4 46.8 7.5 4.00 .14 Thin 503.6 44.9 7.5 0.73 .02 EVERCALC Medium 477.7 46.0 7.5 4.45 .06 Thick 439.6 58.0 7.5 28.87 .13 Thin 500.2 44.9 7.5 0.11 , .02 EGALT Medium 502.9 44.8 7.5 0.57 .02 ’ Thick 500.5 45.8 7.5 0.17 .15 Table 4.19. Comparison of the results of three programs for a four layer pavement. Backcalculated modulus (psi) Max. error RMS error in Program in moduli deflections AC Base Subase Roadbed (%) (g) MICHBACK 500.1 45.1 14.9 7.5 0.6 .01 MODULUS 544.9 36.0 22.3 7.6 48.7 .16 EVERCALC 476.2 46.3 14.7 7.5 4.8 .09 EC-ALT 495.0 46.3 14.3 7.5 4.6 .06 140 4.6.3 Four-Layer Composite Pavements The properties of the two composite pavements analyzed are listed in Table 4.3. One composite pavement (composite 1) had a separation layer between the AC overlay and the PCC slab, whereas the other pavement (composite 2) had no such layer. The results shown in (Tables 4.20 and 4.21) indicate that while MICHBACK converges reasonably well for both composite pavements, the other two programs yielded considerable errors especially for the composite 2. It has been pointed by various researchers that for composite pavements, the modulus of the layer immediately under the slab is the most difficult to predict unless the lower layer is the roadbed soil. For this pavement the maximum error for MICHBACK is 8% compared to about 61% for MODULUS and 131% for EVERCALC (using the data generated by old CHEVRON (EC-ALT». This indicates that MODULUS and EVERCALC do not produce accurate results for composite pavements. MICHBACK has also shown some problems in predicting the modulus of the layer immediately under the slab. However the magnitude of the error is comparatively smaller. For all other composite pavement examples, the MICHBACK produced better results than both EVERCALC and MODULUS programs. 4.6.4 Three-Layer Pavements Over a Stiff layer The medium thickness three layer flexible pavement (Table 4.1) was underlain by a stiff layer at two different depths, 36 and 240 inch. All other parameters were the same as for the three layer flexible medium thickness pavement analyzed without the stiff layer. The results are presented in Table 4.22. It can be seen that the results of MICHBACK and EVERCALC (EC-ALT) are comparable and better than those of MODULUS. 141 Table 4.20. Comparison of the backcalculated results for a composite pavement section. Backcalculated modulus (ksi) Max. error RMS error Program in moduli (%) (%) AC Slab Base Roadbed MICHBACK 499.4 4516.3 23.0 7.5 8.0 0.01 MODULUS 527.7 4471.1 9.8 7.6 60.8 0.07 EVERCALC 1582.2 2297.1 13.2 7.5 216.4 1.53 EC-ALT 494.7 4217.1 57.8 7.4 131.1 0.06 Table 4.21. Comparison of the backcalculated results for a composite pavement section consisting of a granular separation layer. Backcalculated modulus (psi) Max. error RMS error Program in moduli (%) (95) AC Base PCC Slab Roadbed MICHBACK 499.4 24.8 4472.9 7.5 0.6 0.03 MODULUS 492.1 25.4 4402.5 7.5 2.2 0.08 EVERCALC 622.7 27.7 3959.7 7.5 24.5 0.75 EC-ALT 491.5 25.5 4412.6 7.5 2.1 0.11 142 Table 4.22. Comparison of the results of three programs for a three layer pavement over stiff layer. Stiff layer Backcalculated modulus (psi) Max. error RMS error in Program location in moduli deflections AC Base Roadbed (%) (95) Deep 501.7 44.9 7.5 0.34 0.04 MICHBACK Shallow 499.8 44.9 7.5 0.19 0.09 Deep 502.0 44.8 7.5 0.44 0.19 MODULUS Shallow 508.9 43.9 7.7 3.55 0.07 Deep 796.8 31.2 7.5 59.36 1.15 EVERCALC Shallow 598.0 40.3 7.6 19.60 0.60 Deep 498.3 45.2 7.5 0.41 0.02 EC-ALT Shallow 501.4 44. 8 7.5 0.34 0.04 143 4.6.5 F ive-Layer Flexible Pavement For the five layer pavement, the results of MICHBACK are presented in Table 4.23. The other two programs cannot analyze a five layer pavement. The five layer pavement configuration used for the analysis is specified in Table 4.4. It can be seen that the maximum error produced by MICHBACK for the modulus of any layer is less than 1%. It indicates that, unlike the other programs, there is no decrease in the accuracy of the backcalculated results with the increase in the number of pavement layers for MICHBACK. 4.6.6 Performance Comparison Performance comparison has mostly been restricted to the examples presented in this section only and to MICHBACK and EVERCALC programs because MODULUS is not an iterative program. For MODULUS, the number of deflection bowls generated depends upon the number of layers in the pavement as well as on the range of the moduli provided by the analyst. The range in turn affects the backcalculated results and a closer range was provided to keep the convergence performance of the program compatible. The MODULUS program has, therefore, not been included in the performance comparison. The results of the comparison are provided in Table 4.24 and Figure 4.12. The number of calls for EVERCALC are based on deflections generated by the old CHEVRON (the number of calls for data generated .by CHEVRONX were a little higher). The designation MICHBACK(N) refers to the use of arithmetic scale, MICHBACK(M) represents the results for the modified Newton method and MICHBACK(L) refers to the use of the logarithmic scale together with the modified Newton method, respectively. ‘ The results indicate that after logarithmic transformation, the performance of MICHBACK is somewhat better than that of EVERCALC. The effect of logarithmic 144 Table 4.23. The MICHBACK backcalculation results for a five layer pavement. Backcalculated modulus (psi) Max. error RMS error in in moduli deflections Subbase I Roadbed (%) (%) AC I Treated base Base 0.02 497.0 I 100.4 45.1 14.9 I 7.5 -0.96 £ 145 Table 4. 24. Cgm gall-{son of the performance of MICHBACK and EVERCALC P 8 Program Number of times ogretion was performed ,2 CHEVRON called __ _ _W MICHBACK(N) 22 5 5 Three layer (min) MICHBACK(M) 19 3 8 MICHBACK(L) 14 2 6 EVERCALC 17 4 4 MICHBACK(N) 26 6 6 Three layer (Medium) MICHBACKOVI) 24 4 10 MICHBACK(L) 13 2 5 EVERCALC l7 4 4 MICHBACK(N) 25 5 5 1a er midi) MICHBACK(M) 20 3 5 ' MICHBACK(L) 14 2 6 EVERCALC l7 4 4 MICHBACKm) 37 7 7 Four layer MICHBACKM) 23 3 9 MICHBACK(L 16 2 6 EVERCALC 21 4 4 . MICHBACKm) 32 6 6 W” MICHBACK(M) 28 4 10 MICHBACKG.) 14 2 4 EVERCALC 21 4 . MICHBACK(N) 32 6 6 (with grantlzlar MICHBACKGQ 28 4 10 mm 1"“) MICHBACKG.) 21 3 7 JWRCALC 21 4 4 MICHBACK(N) 32 5 5 Five Layer MICHBACK(M) 26 3 9 MICHBACK(LL 18 2 6 MICHBACKm) 36 7 7 (Deep Stiff uhyer) MICHABCK(M) 24 4 10 MICHBACK(L) 12 2 5 EVEALCRC ,_ __ _ 17 4 146 3 E Emir. .. 1.... *iixitvirn a 5 4 'o F m i i ’5 7////////2 2642/2/40?” ////////////////////// , ,g/ ’/’///// ”/////////// < t?) O Li) O 1}: t': 115 o 8 ('3 (‘0 N N 1- v- aEDWIN 9090 Pavement Type I- MICHBACK(N) - MICHBACK(MN) m MICHBACKLOG [mm EVERCALC Figure 4.12. Comparison of the performance of MICHBACK and EVERCALC programs. 147 28888. .886... 3.9.3.. ...... 29.8.0.2 .6 858.82. 2.. .e .8828 .N. ... as»... 0.205%. mm... 60.20319... % 22.20%... o... I 2.2085... / .93.. oz...— onE. 39.8.66 263500 .26.. Sen. 9pr sulao 148 transformation becomes more significant as the nature of the problem becomes more complicated. 4.7 UNIQUENESS OF THE BACKCALCULATED RESULTS Many backcalculation programs suffer from the disadvantage that the backcalculated results are highly dependent on the md modulus values provided by the user. The farther the guess is from the true values, the higher are the chances of converging to a wrong solution. This is especially true for the methods which seek the minimization of an objective function where the chances of converging to a local minimum are higher. The convergence of Newton’s method in general is also not global, but is problem dependent. For many complicated problems, the solutions are reported to be governed by the starting values. However, backcalculation of layer properties from FWD deflection data appears to be a well behaved problem, especially for flexible pavements, and the results obtained using Newton’s method seem to be independent of the starting values. Many researchers (Sivaneswaran, 1991) have also pointed out, though not with absolute certainty, that the criterion function constructed by minimizing the squared difference between measured and calculated deflections is convex in shape and hence will have a unique minimum. For the flexible pavement examples, it was observed that the results obtained by MICHBACK are independent of the seed values. The results for three, four, and five layer pavements are presented in Tables 4.25 through 4.27, respectively. The deflection data was generated by using the pavement cross sections introduced earlier but the layer moduli were changed as shown in the respective tables. The seed moduli were chosen randomly but far from the true values to test the capability of the program to converge to the correct solution even for unreasonable seed values. It can be seen from these tables that the results obtained by MICHBACK are not affected by the seed moduli at all. The only difference is in the number of 149 Table 4.25. Uniqueness of the MICHBACK solution for a three layer flexible pavement. Ex. Actual modulus (ksi) Seed modulus (ksi) Bbackcalculated modulus (ksi) II El E2 E3 1000 500 100 499.2 75.10 15.0 1 1 1 499.2 75.10 15.0 2000 100 100 801.3 44.91 7.5 3 800 45 7.5 l 1 1 801.3 44.91 7.5 II j——————+_‘* Table 4.26. Uniqueness of the MICHBACK solution for a four layer flexible pavement. Ex. Actual modulus Seed modulus Backcalculated (ksi) (ksi) modulus (ksi) E2 E3 E4 El E2 E3 E4 1 El E2 E3 2000 100 100 100 500.1 45.1 14.9 45 15 7.5 1 1 1 1 500.1 55.1 14.9 2000 500 50 50 803.0 74.6 25.3 15.0 75 25 15 1 l l 1 803.1 74.6 25.3 15.0 Table 4.27. Uniqueness of the MICHBACK solution for a five layer flexible pavement. II Seed Modulus (ksi) Backcalculated Modulus (ksi) El E2 E3 E4 E5 E1 E2 E3 E4 E5 1 l 1 1 1 511.0 I 96.1 48.1 ' 13.3 7.52 100 1000 15 5 1.5 510.8 I 96.2 48.0 13.3 7.52 150 iterations required to meet the given convergence criterion. The upper and lower limits for all the layer moduli were set at 10,000,000 and 1 psi respectively. This capability has been provided in the MICHBACK program so that the analyst can specify bounds for the backcalculated moduli. This further ensures that the solution should remain within the expected or even known moduli ranges for the pavement materials. In the case of flexible pavements this capability was not invoked. For the composite pavements convergence to the correct results from excessively erroneous seed values can only be achieved by intelligent use of the bounding values. However, even for composite pavements, convergence from reasonable seed values (expected even from a novice) is not a problem. Unreasonable seed values were used only to check the robustness of the program. This problem has been partially addressed in the program by automatically setting a lower bound of 1 million on the backcalculated results whenever the analyst recognizes the pavement to be composite and not built on a rubbled slab. No upper bound is required to be set. The pre-specified lower limit assures convergence to the correct results even from unrealistic seed values. 4.8 EFFECT OF INACCURACIES IN DEFLECTIONS AT SIMULATED SENSOR LOCATIONS ON BACKCALCULATED RESULTS The accuracy of each sensor of FWD is about :1; 2% of the sensor’s range. Hence, similar range in the accuracy of the backcalculated results should also be expected. Also it is a general belief that since the deflections at the outer sensors are comparatively smaller, inaccuracies at these sensors have a larger contribution towards the overall error especially for the lower layers. In this section, the effect of inaccuracies in deflections at different sensor locations on the backcalculated layer moduli are examined. It was feared that logarithmic transformation may make the backcalculated results more sensitive to the inaccuracies in the measured deflections. 151 Therefore, a comparison has also been made between the sensitivity of the backcalculated results to the surface deflection for both arithmetic and logarithmic scales. The thin, medium, and thick three layer flexible pavements (Table 4.1) were used to study the sensitivity of the backcalculated results to the surface deflections. The medium three layer pavement was used to examine and compare the sensitivity of the backcalculated results when working in logarithmic scale. The surface deflections were generated using the CHEVRONX program. An error of i 2% was deliberately introduced in each sensor deflection individually. The results for the arithmetic and logarithmic scales are presented in Tables 4.28 and 4.29, respectively. The results indicate: 1. lnaccuracies in the deflections of these locations close to the load (for fixed percent error) induce larger error in the backcalculated layer moduli than the other locations. 2. For the same absolute magnitude of the error, negative errors induce greater errors in the backcalculated results than the positive ones.- 3. For almost all locations, the AC modulus is the most affected followed by the base modulus. The roadbed modulus is least affected by inaccuracies in the surface deflections at any locations. Although errors in the last few locations do affect the roadbed modulus slightly more than errors in the others, the maximum error in the roadbed modulus remains well below 1.5%. 4. The AC modulus of thin layers is affected the most by deflection inaccuracies. This is because pavements with a thick AC layer can compensate for the erroneous deflections by smaller changes in the AC modulus, whereas to redress the same amount of error, the AC modulus must undergo a bigger change for pavements with a thin AC layer. 5. The base modulus of a pavement with a thick AC layer is affected more than 152 N.. n. .... .....- 3.. ...? .... .... ... N- N..- n- ..c. ...m. ..4 m... N6. N.N- ...? N+ .. N.. _ N... 8...- Ne o.N- ... a... o... N- N..- a.- ...? a... ...N N.N m..- ...... o8. N+ o o... e- e... N... N.N. ...n- N.... N...- ..N. N. n? 2.. e...- no. N.N N... ...v ...o N.- N+ m .2 ... 2... ...o. a... an- o...- os- no. N- N... ...- S.- n.e.- o6. an as I. ...n- N+ .. 24 a- 8... SN ...... N.N a...- o.«.. ......- N- S. ... o...- o..N- 3.. an- e... N... N...N T. n ...... e.- N.. ...: ...: N.NN 5.. .... .- .8. N- w... c .... ea.- n.N.- n.2- e... ...... 3... N+ N .... m. 2.. :2. N2..- N.n- SN ...aN ...:- N. 8.... ... N... 28 ...... an ......- 3.N- 4.9.. N+ . .85 .82 :2... 8E .82 :2... ......F .82 ...E. .5 .8882 34.. 0... 82.8.86 .62 5 ..otm e538... $1.582 8.6.8.6362. ... .oem. .28... 5.2.55.3 - 25.8%.. 5 888 o. 0.6 =ng .93. 8.53.858; o... 5 9.8.0 2.8.-em .w~.v 25¢. 153 6.. o._ T ...... N- N and. m... : mud. N+ $6 36- «N; N- o S...- emN m6- N..- mm... o6 ms- N- n NNd- 36. m.» N..- m... ...... N....... N-. v No...- 86- em... N+ c... a... o. .a- N- n 8... Cd- ood N+ 3...- , MN... 3..- N- N ac... N5..- mN.N N+ 2... a; _- mndN N- _ vod- 3... deN- N+ .... . 2x. 255...... 2*. «...—5...... AR. 9.3.6:. 35.3%.. ...2 35...... 5 ...:m 8.3 5 ...-cm U< 5 .....m 5 ......m 55.83 .58. 5655.»... - 85.8%.. 5 9.8.6 ... 0.... :55... .9»..— B.N_..5mo._omn on. 5 32.8 58.0.. 6N... 2an 154 that of a pavement with a thin AC layer. This is probably because the inaccuracies in the AC modulus for a pavement with a thick AC layer affects the overall strength of the pavement more than for a thin pavement. Therefore, in trying to adjust the overall stiffness the base modulus for thick AC pavement is affected more. Results of a similar test for the medium thick pavement performed after the logarithmic transformation are presented in Table 4.29. It can be observed that this change does not significantly affect the sensitivity of the backcalculated results to the surface deflections. Except for the last deflection location, for most other sensors the error in the backcalculated layer moduli is slightly less than those arising from the arithmetic scale. The above exercise was repeated by fixing the induced difference in the measured deflections at :1: 0.5 mils for all the sensors. The results are presented in ‘ Table 4.30. The errors in the backcalculated moduli are a little higher than in the case where the errors were induced as percentages of the measured deflections. This is because of the fact that the 1: 0.5 mil error is larger than the 2% error. 4.9 EFFECT OF POISSON ’S RATIO ON THE BACKCALCULATED LAYER MODULI Several studies were conducted to assess the effects of Poisson’s ratios of the various pavement layers on the calculated deflections. Pichumani (1972) concluded that Poisson’s ratio of only the roadbed soil has some appreciable effect on the surface deflections. Variations in the Poisson’s ratios of the other layers were found to have little effect on the surface deflections. Another study, conducted at the University of Utah (Hou, 1977), suggested that for three-layer pavements, variations in Poisson’s ratio of each layer including the roadbed soil from 0.25 to 0.45 have no 155 Table 4.30. Percentage error in backcalculated layer moduli due to i 0.5 mil error in deflections-log scale. Samar Error AC Base No. +0.5 -25.8 17.7 1 -0.5 32.7 -18.6 +0.5 19.5 -19.3 2 -0.5 -21.7 25.3 +0.5 12.4 -10.9 3 -0.5 -11.7 . 11.2 +0.5 9.9 -7.1 4 -0.5 ~8.3 5.9 +0.5 3.6 -0.2 5 -0.5 ~3.0 -0.5 +0.5 7.2 12.4 -12.2 24.9 -29.6 156 significant effect on the surface deflections. These and other similar findings have led to a general consensus that since Poisson’s ratios of the paving layers have little influence on the surface deflections, their effect on. the backcalculated layer moduli must also be negligible. No study appears to have investigated the direct effect of Poisson’s ratios on the backcalculated layer moduli. In this study, this issue was investigated and the results are presented in this section. The deflection basins used in the previous sections were generated by using constant Poisson’s ratios of 0.35, 0.4, 0.45, and 0.45 for the AC, base, and subbase layers and for the roadbed soil, respectively. To assess the effect of Poisson’s ratio on the backcalculated layer moduli, the value of Poisson’s ratio of one layer at a time was varied by i 0.05 from the true value. The results of this analysis for the medium thick flexible pavement of Table 4.1 an listed in Table 4.31. Results of similar tests for composit pavement (Table 4.3) are presented in Table 4.32. Examination of the results indicate that: 1. An error of i .05 in Poisson’s ratio of the AC layer introduces about a 4% error in the backcalculated modulus of the AC layer. The errors in the moduli of the other two layers are comparatively small and the roadbed soil modulus is the least affected. 2. As the stiffness of the AC layer increases so does the effect of Poisson’s ratio. 3. Poisson’s ratio of the base layer has some impact on its modulus and the least effect on the backcalculated results of the other layers. 4. An error of i 0.05 in Poisson’s ratio of the roadbed soil has the largest effects on the backcalculated results. It introduces a 10 % error in the AC modulus, a 9 % error in the base modulus, and a small error in the roadbed soil modulus. 5. For composite pavements, the backcalculated layer moduli appear to be very sensitive to errors in the value of Poisson’s ratio. For example, an error of 157 Table 4.31. Percent errors in the backcalculated layer moduli due to error in Poison’s ratio for flexible pavements. Layer Error in Error in backcalculated moduli (%) Poiszon’s ra 0 AC Base Roadbed Table 4.32. Percent errors in the backcalculated layer moduli due to error in Poison’s ratio for composite pavement. Layer Error in Porsson,s 158 300 % in the subbase layer modulus (the layer immediately beneath the PCC slab) was observed. The errors in the modulus values of the other layers were less than 20 %. 6. The effects of Poisson’s ratios on the calculated deflection basins is negligible. The above observations do not negate the past findings, they only emphasize that small changes in the deflection basins can significantly affect the values of the backcalculated layer moduli. I It should be noted that the results presented above are pertinent to the MICHBACK program (which uses the CHEVRONX as the forward analysis program). The sensitivity of the various backcalculation techniques to Poisson’s ratios or to small changes in the deflection basins may vary. One point can be made here is that, given the present state of the equipment (deflections at only seven sensor locations are measured), and the number of unknowns that need to be estimated, the inclusion of Poisson’s ratios of the various layers in the pool of unknown to be calculated will only make the backcalculation more complicated and prone to higher errors. Furthermore, the intention of the analysis presented above is to warn the analyst that reasonable ranges of Poisson’s ratios of the different paving materials must be known and must be used cautiously. 4.10 COMPARISON OF DEFLECTION OUTPUT OF DIFFERENT ELASTIC LAYER PROGRAMS The advent of fast micro-computers, made automated backcalculation of layer moduli possible. Most backcalculation programs make use of a multi-layer elastic routine as a forward analysis programs in one way or another. Hence, the backcalculated results are not only affected by the backcalculation technique, but also by the precision of the forward analysis routine. Surface deflections are the common 159 output used from these routines in the backcalculation process. The accuracy of the multi-layer elastic routines is generally established by comparing the deflection results to those of an established and reputed one such as the "BISAR". MICHBACK initially used CHEVRON as the forward analysis program. At the onset of this study, the analysis of deflection data received from various agencies have indicated that differences between the deflections obtained from the CHEVRON program and those from the BISAR program exist. Consequently, a corrected version "CHEVRONX“ of the CHEVRON program was obtained from Dr. Lynne Irwin at Cornell University and it was embedded in the current version of MICHBACK. It should be noted that, originally, results of the CHEVRON program was validated by Lee, et. al. (1988). They concluded that the output of the CHEVRON program differs slightly from that of the BISAR program for flexible pavements and that the program should not be used for the analysis of stiff flexible and composite pavements. In this section surface deflections from four elastic layer programs, (CHEVRON . ELSYMS, WESLEA, and CHEVRONX) have been compared with those of BISAR. Once again, CHEVRONX is the enhanced version of the CHEVRON program and it is used in MICHBACK. The five pavement sections listed in Tables 4.1 through 4.4 are used in this comparison. Further, an AC modulus of 800 ksi (instead of 500 ksi) was used for the medium thick three-layer flexible pavement. The stiff layer depth for the pavements was set at 144-inch to represent the presence of bedrock. 4.10.1 Comparison of Deflection Output The deflections from all the five programs are presented in Table 4.33. The deflections were rounded to the nearest hundredth of a mils to represent the FWD readings. It can be seen that the outputs of BISAR, WESLEA and CHEVRONX are essentially the same. Table 4.34 provides a list of the percent differences in the 160 Table 4.33. Deflections from different elastic layer programs. Pavement Deflections (mils) Type Program d0 d1 d2 d3 d4 d5 d6 BISAR 24.400 20.900 18.600 15.600 13.100 9.500 5.530 Three CHEVRONX 24.370 20.870.18.570 15.580 13.120 9.498 5.534 Layer CHEVRON 24.260 20.870 18.570 15.580 13.120 9.498 5.534 Medium ELSYM5 24.260 20.870 18.570 15.580 13.120 9.497 5.534 WESLEA 24.369 20.872 18.570 15.578 13.123 9.498 5.533 BISAR 22.200 19.600 17.700 15.100 12.900 9.500 5.600 Three CHEVRONX 22.200 19.570 17.700 15.120 12.900 9.497 5.595 Layer CHEVRON 22.070 19.560 17.700 15.120 12.900 9.497 5.595 Medium ELSYM5 22.070 19.560 17.700 15.120 12.900 9.497 5.595 Stiff WESLEA 22.203 19.572 17.700 15.121 12.905 9.497 5.599 BISAR 16.600 14.900 13.900 12.500 11.200 8.980 5.810 Three CHEVRONX 16.650 14.900 13.910 12.530 11.230 8.977 5.813 Layer CHEVRON 16.470 14.790 13.920 12.530 11.230 8.977 5.813 Thick ELSYM5 16.470 14.790 13.920 12.530 11.230 8.977 5.813 WESLEA 16.649 14.899 13.913 12.530 11.234 8.978 5.813 BISAR 19.600 17.000 15.400 13.300 11.600 8.930 5.660 Four CHEVRONX 19.630 17.030 15.420 13.330 11.590 8.929 5.662 Layer CHEVRON 19.530 16.960 15.420 13.330 11.590 8.929 5.662 ELSYM5 19.530 16.960 15.420 13.330 11.590 8.929 5.662 WESLEA 19.634 17.032 15.421 13.331 11.593 8.929 5.658 BISAR 20.900 17.400 15.100 12.100 9.710 6.140 2.340 Three CHEVRONX 20.910 17.420 15.120 12.150 9.709 6.140 2.345 Layer CHEVRON 20.310 16.700 14.950 12.240 9.739 6.134 2.345 With ELSYM5 20.310 16.700 14.950 12.240 9.739 6.134 2.345 StiflLayer WESLEA 20.910 17.420 15.120 12.150 9.709 6.140 2.345 BISAR 9.480 8.800 8.600 8.270 7.890 7.080 5.520 Composite CHEVRONX 9.479 8.800 8.599 8.268 7.890 7.080 5.521 81111 CHEVRON 10.040 8.766 8.479 8.272 7.896 7.090 5.520 ELSYM5 10.040 8.766 8.479 8.272 7.896 7.090 5.520 WESLEA 9.479 8.800 8.599 8.268 7.890 7.080 5.521 Tab|e4.34. Perceritdlflereneeinfiiedeflectionaofdiflemntelasfielayerprograma. 161 Pavement Diflerance in Deflections (96) Type Program d0 d1 d2 d3 d4 d5 d6 BISAR 0.000 0.000 0.000 0.000 0.000 0.000 0.000 Three CHEVRONX 0.123 0.144 0.161 0.128 0.153 0.021 0.072 Layer CHEVRON 0.574 0.144 -0.161 0.128 0.153 0.021 0.072 Modlum ELSYM5 0.574 0.144 -0.161 0.128 0.158 0.032 0.072 WESLEA 0.128 0.133 0.182 0.140 0.173 0.026 0.060 BISAR 0.000 0.000 0.000 0.000 0.000 0.000 0.000 Three CHEVRONX 0.000 0.153 0.000 0.132 0.000 0.032 0.089 Layer CHEVRON 0.588 0.204 0.000 0.132 0.000 0.032 0.089 Medium ELSYM5 0.586 0.204 0.000 0.132 0.000 0.032 0.089 61111 WESLEA 0.013 0.141 0.000 0.139 0.036 0.028 0.010 BISAR 0.000 0.000 0.000 0.000 0.000 0.000 0.000 nine CHEVRONX 0.301 0.000 0.072 0.240 0.268 0.033 0.052 Layer CHEVRON 0.783 0.738 0.144 0.240 0.268 0.033 0.052 Thick ELSYM5 0.783 0.738 0.144 0.240 0.268 0.033 0.052 WESLEA 0.293 0.008 0.094 0.238 0299 0.023 0.049 BISAR 0.000 0.000 0.000 0.000 0.000 0.000 0.000 Four CHEVRONX 0.153 0.176 0.130 0226 0.086 0.011 0.035 Layer CHEVRON 0.357 0235 0.130 0.226 0.086 0.011 0.035 ELSYM5 0.357 0.235 0.130 0.226 0.086 0.011 0.035 WESLEA 0.172 0.191 0.135 0232 0.057 0.007 0.028 BISAR 0.000 0.000 0.000 0.000 0.000 0.000 0.000 111m CHEVRONX 0.048 0.115 0.132 0.413 0.010 0.000 0214 Layer CHEVRON .2823 4.023 0.993 1.157 0299 0.098 0.214 With ELSYM5 -2.823 4.023 0.993 1.157 0.299 0.098 0214 Stll'fLayer WES-LEA 0.048 0.115 0.132 0.413 0.010 0.000 0.214 BISAR 0.000 0.000 0.000 0.000 0.000 0.000 0.000 Composite-CHEVRONX 0.000 0.000 0.000 0.000 0.000 0.000 0.000 8011 [CHEVRON 5.907 0.386 -1407 0.024 0.076 0.141 0.000 ELSYM5 5.907 0.386 -1.407 0.024 0.076 0.141 0.000 [WESLEA 0.000 0.000 0.000 0.000 0.000 0.000 0.000 162 deflection at each sensor location relative to the BISAR calculated deflection. It can be seen that: 1. Except for rounding-off precision, there are no significant differences in the outputs of the CHEVRONX, WESLEA, and BISAR programs. 2. The deflection outputs of the CHEVRON and ELSYM5 programs are the same. However, the following observations can be made relative to the differences between these two outputs and that of the BISAR program: a) The differences are more pronounced at the first two sensor locations and negligible at the outer sensors. b) The differences increase as the stiffness of the pavement increases. For the medium thick pavements with a stiff layer, the differences at all sensor locations are higher than those for the same pavement without a stiff layer. The maximum difference of 4% is observed at the second sensor location. For flexible pavements, the CHEVRON deflections are generally lower than those of BISAR. c) For the composite pavements, significant differences were found between the CHEVRON and the BISAR deflections. 4.10.2 Comparison of Backcalculated Results The deflections from the CHEVRONX and CHEVRON programs were used in MICHBACK to backcalculate the layer moduli. The results are presented in Table 4.35. It can be seen that: l. The backcalculated results are appreciably different for the two sets of generated data. The difference increases as the overall stiffness of the pavement increases. 2. For the flexible pavement with a stiff layer, the maximum error is about 34%. 3. For the composite pavement, the results are quite erroneous, with a maximum 163 Table 4.35. Backcalculated results of MICHBACK for deflection data generated by different elastic layer programs. Pavement Error in backcalculated layer moduli (%) I ype AC Base Subbase Roadbed Three layer CHEVRON 5.39 3.05 - 0.05 ’ 0.05 medium CHEVRONX -.02 -0.07 - 0.0 0.03 Three layer CHEVRON 11.48 -16.1 - 0.2 0.15 thick CHEVRONX .25 -0.84 - 0.04 0.0] Three layer CHEVRON 5.74 -3.71 - -0.01 0.09 medium stiff CHEVRONX -.28 0.69 - -0.06 0.02 Three layer CHEVRON 34.05 42.01 - 0.2 1.04 with stiff layer (144 in.) CHEVRONX .43 044 - 0.04 0.03 CHEVRON 4.26 -O.48 -2.83 0.09 0.08 Four layer ‘ CHEVRONX .02 0.22 -.05 0.03 0.03 Composite CHEVRON 66.1 77. 13 -99.4 231.7 0.01 CHEVRONX -.14 0.34 -7.9 0.07 0. 89 164 error of about 231%. The scenario presented in this and in the previous section demonstrates that the results of the backcalculation are, in general, affected by the accuracy of the employed forward analysis program. Based on these results, the use of the CHEVRON and the ELSYM5 programs for the backcalculation of layer moduli of any pavement type is not recommended. 4.11 COMPARISON OF MICHBACK RESULTS WITH SHRP STUDY In the previous section, the backcalculated results of MICHBACK were compared with two leading programs and various performance aspects were highlighted. In this section, additional comparison of the MICHBACK results with three programs (MODCOMP, MODULUS, and WESDEF) is presented. In this comparison, the pavement cross sections and the deflection basins that reported by Rada, et al., 1992 and listed in Table 4.36 were used. The true values of the layer moduli and those backcalculated by using the three programs are listed in Table 4.37 (Rada, et al., 1992). Table 4.38 provides a list of the layer moduli of the first six pavement sections of Table 4.37 that were obtained by using the MICHBACK program. Table 4.39 provides a summary of the error in each layer modulus as well as the accumulated absolute error in the backcalculated results for all four programs. It should be noted that the errors the MODCOMP, MODULUS, and WESDEF programs were obtained from Rada et al. study (1992). The accumulated absolute errors of the four programs for the six pavement sections are shown in Figure 4.13. It can be seen that the MICHBACK results have consistently the lower cumulative error. Table 4.40 provides a list of minimum, maximum, average, and standard - deviations of the accumulated absolute errors in the moduli values of the six pavement sections for the four programs. It can be clearly seen that the results of MICHBACK are much more consistent than those of the other programs. This comparison of the 165 Table 4.36. Deflection and cross sectional data for nine test sections (after Rada, et al., 1992). hummus) ID Layer 5“ nua- (H3) r-0' r-l'r-12'r-18'r-24'r-36'r-dr 1 men-area 3 1 2 Granalarflaae 6 32.90 23.11 17.81 12.60 951 6.15 3.57 3 Sabin“ 1 AsphaltConaeae 6 2 2 6mm Base 12 30.10 2450 21.70 1850 15.90 12.20 7.73 3 SW 1 AsphatrCeaeme 3 2 Cementsmaase 6 8.97 7.95 756 7.07 6.57 5.61 4.1!) 3 81:09:68 1 PCCSIab 4 2 LineSuuBaae 6 8.94 8.41 8.13 7.48 6.91 5.80 4.02 3 SW 1 PCCSIab 6 5 2 SW 18.10 16.60 1550 13.70 11.!) 8.97 5.26 1 PC1281» 12 6 2 Cement Stab. Base 6 8.18 8.06 7.91 7.71 7.50 7.05 6.10 3 SW 1 AsphaltCoaa'eae 7 2 new 9 7.50 6.13 5.87 5.43 4.97 4.09 2.72 3 Snag-ace 1 AsphaltConcrei-e 5 2 mm 10 8 3 IJneSnaBase ' “65.485.275.01 4.73 4.14 3.08 4 800m r AsphaltConaete 4 2 POCSlab 9- 3 mm” ' 6.89 5.87 5.74 559 5.42 5.05 4.28 4 Saw-do mil-1m” WWI-5.91m 166 8..> 5.582 3.... . s... 85.2% .33... i E... .8... 85.358 .8... .. ...... .8... 88......» .88... t Bu 838318.233633102 .98§<3ecafleiae.saaeufia&wQ OE: hOh 3:505 MU 6' 2 Thick > 6' 4 Cross Section AC Overlay 2 Layer: 3 1 Course 2 3 layers 2 2 Couree 2 4 layers 2 3 or more 3 Courses AC Cour-ea (no Number of Overlays overlay). l Overlay 2 Lean than 3 I 2 or more 2 More than 3 1 Overlays AC (overlay) l lava-lay Age (years) Lesa than 3 2 . New Mix 2 Roadbed 8011 Sand 2 Old Mix 2 Clay 2 Pavcmcrn Age (yeare)L Lear than 3 190 score consists of the sum of the weights assigned to each variable. The pavement section with the highest score was given the highest priority. Based on the pavement score (priority), the 49 flexible and 15 composite pavement sections with the highest priorities were chosen and they are included in this study. Tables 6.2 and 6.3 provide lists of these flexible and composite pavement sections. Other information such as pavement type, route number, direction (north, south, east, or west bound), district, control section number, and the beginning and ending mile post of each pavement section are also listed in the tables. Figure 6.1 shows the distribution of the selected pavement sections across the State of Michigan. For each of the 49 flexible and 15 composite pavement sections, the pavement condition data (obtained from MDOT data bank) was examined. It should be noted that the MDOT data bank does not identify fatigue cracking as a separate distress eategory. Rather, it identifies the severity and extent of cracking which includes various types of cracks. Therefore, only the rut data in the MDOT data bank were considered. Each pavement section was divided into several lOO-feet long test sites. For some sections, the test sites were adjacent to each others while for some others, they were separate. During the Summer of 1991, four members of the Michigan State University (MSU) research team visited each test site. The purposes of the visit were to: 1. Verify the location reference point, pavement type, and general conditions of the sites. 2. Mark the test sites. 3. Inspect, measure, and record the extent and severity of rutting, fatigue cracking and other types of distress. 4. Identify those test sites to be cored by MDOT. 5. Mark locations for nondestructive deflection tests (NDT) within each test site. 191 Table 6.2 Flexible pavement sections selected for study. Section Route Dir. District Control Mile Post NO. I =11 N=1 Sectio FROM TO Remarks us=22 s=2 M =33 E=3 w=4 IMSUOlF 3355 1 3 40031 200 13.00 [fisuo2r 3350 3 3 23052 1.30 14.00 [rusuoer 3334 3 3 46041 1.50 11.20 [MSUO4F 3372 3 3 40023 0.00 3.00 [usuosr 3323 3 1 66022 3.00 14.40 [MSUOGF 3350 4 5 34021 4.50 3.00 Wsuwr: 22131 1 ' 5 54014 3.00 11.10 Cored [msuoer 33133 3 5 79011 15.70 19.90 IMSUOQF 3320 4 5 54022 0.00 0.50 [MSUlOF 3323 3 1 31021 4.60 9.50 Cored [fisunr 3344 3 5 41051 4.20 5.20 [MSU12F 1195 3 5 61152 1.20 5.40 @sumr 2227 1 3 13034 1.31 1.30 Cored [MSU14F 3377 1 2 75052 3.00 10.00 Cored 5150156 2227 2 4 20016 0.00 6.30 [MSU16F 22131 1 3 33031 2.50 3.00 wsum= 22131 1 5 54013 0.50 3.41 flaswer 2227 1 4 20016 0.00 5.30 “161.11% 1175 1 4 59013 0.00 5.00 cored Wsuzor 3365 1 5 59051 11.90 13.20 6490211: 3319 1 5 74032 0.00 10.00 lstuzzF 3332 4 5 52041 0.00 5.90 [MSU23F 3323 3 1 66023 6.60 11.00 flusuz41= 3399 1 3 33011 4.30 4.60 5450255 22131 2 3 33031 3.00 2.00 [MSUZGF 3357 3 6 25102 2.35 2.95 [MSU27F 1175 1 4 72061 7.00 13.40 [msuzer 1 175 1 4 15093 0.00 5.90 [MSU29F 3399 1 3 23092 2.20 7.30 cored [We-0305 1175 1 4 69013 5.00 7.00 [NTSU31F 3325 4 1 66051 0.00 4.00 [rTsust 22131 1 5 59012 11.00 11.10 192 Table 6.2 Flexible pavement sections selected for study (continued). Section Route Dir. District Control Mile Post Designation l=11 N=1 Section FROM TO Remarks 03-22 3:2 M =33 E=3 W=4 [MSUSSF 22131 2 5 54014 3.50 3.50 [Ms034r= 33129 1 2 17072 12.10 19.30 [MSU35F 1175 1 4 16091 0.00 1.50 [MSU36F 1175 2 4 72061 13.30 7.00 [Ms037F 1175 1 4 72061 19.00 23.66 [Ms033F 3365 1 3 57013 3.20 12.71 [Ms039F 222 3 2 21024 4.30 14.30 cored [MSU40F 1175 1 4 20015 4.10 9.10 [MSU41F 1175 1 4 20015 9.10 14.20 [MSU42F 1194 3 7 11015 13.00 15.00 [MSU43F 1194 3 7 11015 3.00 7.00 cored [M5044F 2223 1 4 71073 24.70 26.71 [M50451= 1175 1 4 20015 0.00 1.00 [MSU46F 3366 1 3 40031 0.00 1.22 [MSU47F 3351 3 3 13041 3.65 3.92 [MSU48F 3349 1 3 30011 11.00 17.00 [M5049F 1175 2 4 72051 15.30 17.50 193 Table 6.3 Composite pavement sections selected for study. Section Route Dir. District Control Mile Post Designation l =1 1 N=1 Section FROM TO Remarks US= 22 S=2 M =33 E=3 W=4 MSU01 C 1194 4 8 81104 5.60 7.90 cored MSUOZC 1194 4 8 81104 1.90 5.60 MSU030 3350 3 8 46081 1 .10 3.00 MSU04C 22127 1 8 30071 0.00 4.90 cored MSU050 2241 1 1 7023 8.48 14.01 cored MSU06C 22223 3 8 46061 2.80 3.60 MSU07C 22127 1 8 46011 4.50 5.18 MSU08C 3350 3 8 46081 8.30 13.20 cored MSUOQC 2241 1 1 7013 0.00 3.10 MSU1OC 3350 3 8 46081 3.00 8.30 MSU11C 3325 3 6 32012 12.00 19.90 cored MSU1ZC 3325 3 6 32012 19.90 27.90 MSU130 22127 1 8 30071 9.80 10.30 MSU14C 22127 1 8 30071 4.90 9.80 MSU15C 1 175 1 9 63173 9.25 9.50 194 Figure 6.1. Distribution of selected pavement sections across the State of Michigan. 195 All sites were visited again during the summers of 1992 and 1993. During each visit, the rut and fatigue cracking data were collected and the general conditions of the sites were recorded. The rut depth was measured by using a six feet straight-edge leveling rod and a graduated triangular wedge with an accuracy of 0.025-inch. The rut was measured in the outer wheel path over each marked core location and at several other locations along the test site. The fatigue crack condition was recorded in terms of severity and the percent of the lOO-feet long test site showing alligator cracking. Further, a total of 106 locations were designated for pavement coring. Each test site and NDT and core locations were given specific designation numbers. The coring method and the designation numbers are presented in the next section. 6.3 MARKING, CODING, CORING AND NDT Each test site was designated by a two-number system. One number designates the pavement section and the other designates the test site. For example, a test site designation of 29-1 indicates the first test site of pavement section number 29. It should be noted that, for all pavement sections the test site number increases from south to north and from west to east. Some test sections were selected for coring. The cores were located either in the outer wheel path, between the wheel paths, or in the inner wheel path of the traffic lane. In addition, some cores were located over an existing longitudinal crack. Each core location was designated using a seven digit number. Starting at the left most digit, the first two digits indicate pavement section number; the third digit indicates pavement type (1 for flexible and 2 for composite); the fourth digit designates the site number; the fifth and sixth digits indicate the distance of the core location from the beginning of the site; and the seventh digit indicates the core number within that site. For example, a core location designation of 2712402 indicates (left to right) that the c0re is obtained from section 27, flexible pavement, 196 test site number 2, at 40 feet from the beginning of the site, and the second core of the site. The cores were obtained by using a power auger equipped with a 6-inch coring bit. A hand auger was preferred over a power auger to obtain undisturbed samples from unbound bases, subbases and roadbed soils so that different layers could be easily identified. Inspite of this, for most pavement sections, separate and accurate identification of the base and subbase layer thicknesses was not possible because of the type of materials encountered (cohesionless and moist). Non-destructive deflection tests (NDT) using a falling weight deflectometer (FWD) were conducted at several locations along each test site. The NDT tests were divided into two categories as follows. 1. Regular tests - Each test site ( 100-feet long) was subjected to five FWD tests (a test is 3 drops). The tests were conducted at equal intervals of 20-feet starting at the beginning of the test site. 2. Additional tests - For cored test sites, additional FWD tests were conducted over the core locations. The NDT were performed during various seasons as follows: 1. Summer 1991 - All test sections at about 9000 lb load. Fall 1991 - All test sections at about 9000 lb load. Spring 1992 - All test sections at about 9000 lb load. PP.“ Summer 1992 - All test sections excluding core locations were tested at about 9000 lb load. For the cored pavement sections the FWD test were conducted at three load levels 4500, 9000, and 15000 lb. 5. Spring 1993 - Cored sections only at about 9000 lb load. Each FWD test and the resulting deflection files were designated by using an eight digit number. Starting at the left most digit, the first two digits indicate pavement section number; the third digit indicates pavement type (1 for flexible and 2 197 . for composite); the fourth digit designates the site number; the fifth and sixth digits indicate the distance of the test location from the beginning of the site; the seventh digit indicates the sequential drop number (drop number 1, 2 or 3); and the eighth (right most) digit indicates test location (0 for regular test in the outer wheel path, 1 for additional test in the outer wheel path, 2 for additional test at the center of the lane, 3 for additional test in the inner wheel path, 4 for a test at a joint in the outer wheel path, and 5 for a test at a joint in the inner wheel path). For example, an FWD test designation of 27124020 indicates (left to right) that the test is conducted on section 27, flexible pavement, test site number 2, at 40 feet from the beginning of the site, and the second drop of a regular test in the wheel path. 6.4 BACKCALCULATION or LAYER MODULI FOR SELECTED PAVEMENT SECTIONS As noted earlier, during the Summer of 1991, NDTs were conducted on all cored and uncored flexible and composite pavement sections. The measured deflection basins are tabulated in Appendix C and the pavement layer thicknesses are included in Appendix B of a research report (Mahmood, 1993). It should be noted that for the uncored pavement sections, the thicknesses of the pavement layers were obtained from the proper MDOT files. For the cored pavement sections, the layer thicknesses were measured from the cores. Since the layer thicknesses of the uncored pavement sections are not accurately known, the results of only typical cored flexible and composite pavement sections are presented and discussed in this section. Recall that each NDT test consisted of three drops. The target load for each drop was set at 9000 pounds. The actual load delivered to the pavement section and measured by the load cell of the FWD however, varied slightly from the target load. Therefore, the average of the three deflection basins and the average load of the three drops were used to perform the backcalculation of the layer moduli using the MICHBACK computer program. 198 For most flexible pavement sections there are no significant differences between the base and subbase layer materials (Chapter 4, Field Manual of Soil Engineering, MDOT). In addition, as stated earlier, the similarity in the texture of the base and subbase materials made it very difficult to identify the thickness of each layer (the materials were soaked by water during the coring). Because of this similarity and due to the lack of accurate base and subbase thickness data, the two layers were combined into a single layer and all flexible pavements were analyzed as three-layer systems (AC, base, and roadbed). All of the composite pavement sections included in the study are listed in the MDOT proposed cross-section file as a three layer system. The construction procedure, however, recommends the inclusion of a 3-inch base layer under the PCC slab (Chapter 4, Field Manual of Soil Engineering, MDOT). During coring, a base layer of various thicknesses was detected under some of the PCC slabs. However, these thicknesses could not be accurately measured during coring. Therefore, all the composite pavements are also analyzed as three layer systems. A second peculiarity of the composite pavement sections is that the history of the pavement was not available from the MDOT records. All sections were originally constructed as PCC pavements. Only after their deterioration were they overlaid by an AC layer. The conditions of the pavement or the treatment given to the pavement before the overlay (joint repair, crack seating, etc.), could not be verified. During the visual inspection, only the distresses that could be observed on the AC surface were recorded. The condition of the slab after the overlay has a significant influence on the backcalculated moduli. Hence, major factors that can be used for interpreting the deflection data and backcalculated moduli are missing from the pool of explanatory variables. Tables 6.4 and 6.5 provide a summary of the general conditions (fatigue cracking, rut depth, and other types of distress) and layout (section and site numbers, 199 Table 6.4. Summary of the layout and conditions of the cored flexible pavement sections. Pavement Pavement site Distance Fatigue Rut variation Other distress Deflection section No. between sites cracking (in.) type along the desi '00 miles % vement Transverse Adjacent sites 50-100 0.19-0.50 cracking Variable MSUO7F .. 4 5 Adjacent sites 0-15 0.19-0.50 Variable 6 1 High severity . MSUIOF Adjacent sites 0.05-0.42 transverse Unrform 2 cracking, l Adjacent sites 50-100 0.12-0.25 Lane shouler Moderate _3__- Variation MSU13F 3 ' 0.30 «4 .. 4 Adjacent sites 0.12025 Lane shouler Moderate 5 separation Variation 6 1 High severity transverse __2__ cracks, lane MSU14F 3 Ad' t ' 100 0 10-0 50 shoulder High acen srtes . . tron, J and bad local 4 “993° 5 conditions 6 1 . . Uniform to __L__ Adjacent sites 0.20-0.28 ' d variati MSUI9F 1.0 .- Uniform to Adjacent sites 0.20028 ' d variati 200 Table 6.4. Summary of the layout and conditions of the cored flexible pavement sections (continued). Pavement Pavement Distance Fatigue Rut Other Deflection section site No. between cracking variation distress along the designation sites (miles) (%) (in.) type pavement 1 Block and _ transverse High 2 Adjacent 100 0.06-0.12 cracking, vanation 3 srtes and stnppmg 4 ~ MSU29F 5 Block and . . transverse Hrgh 6 Adjacent 100 0.06-0. 12 craclnng, vanatron 7 srtes and stripping 8 1 Block and Adjacent 100 0.12-0.30 transverse High 2 srtes crackmg MSU35F 3 4 . Block and . Adjacent 100 0.12-0.20 transverse ngh 5 srtes cracking 6 1 Adjacent 0.25-0.35 Medium 2 srtes 0.30 '< - 4 Adjacent 0. 38-1 .0 Medium 5 srtes MSU43F 6 1.0 7 Adjacent 0. 12—0. 19 Medium 8 srtes Summary of the layout and condition of the cored composite 201 Table 6.5. pavement sections. Pavement Pavement site Distance Distress type Rut variation Deflection section No. between sites (in.) along the dedgnation (miles) pavement I _ 1 . . Edse cracking. high awed HES?! Adjacent srtes lane shoulder separation an 0.25-0.31 variation 2 transverse cracking 3 MSU01C 1.0 4 Edge cracking, high severi High Adjacent sites lane shoulder separation 0.25-0.31 variation 5 transverse cracking 6 1 Edge crackin , transverse Very high Adjacent sites cracks and ongitudinal 0.25-0.56 var-ration 2 cracks 3 MSUOSC 1.0 4 Edge crackin , transverse Very high Adjacent sites cracks and ongitudinal 0.25-0.56 var-ration 5 cracks 6 l 5 to 7 even] spaced hi Hi Adjacent sites severity transir'erse cracksshin 0.25-0.56 vana‘trhon 2 each site 3 MSU08C 2.0 4 5 to 7 even] spaced hr h Hi Adjacent sites severity transzerse crackg rn 0.25-0.56 variagtrhon 5 each site 6 1 High MSU04C 2 Adjacent sites Pavement in good condition variation 3 202 and distances between the test sites) of eight cored flexible and four cored composite pavement sections, respectively. Descriptive terms relative to the variability of the deflection data along the various pavement sites are also listed in the right-most column of the tables. It should be noted that the location of any FWD test (a test consists of three drops) is designated by a location/station number. Location or station number 1 is always located at the beginning of the first test site of each pavement section. For any given pavement section, regardless of whether the test sites are adjacent to each others or not, a continuous numbering system (e.g., l, 2, ..., 8) is used to designate the location/station numbers. In the next subsections, examples of the actual variations of the various deflection values along the sites are presented. Further, typical deflection basins measured at various location/station numbers along various pavement sites are also presented. 6.4.1 Flexible Pavement Section MSUO7F - Variable Deflections There are a total of six test sites in this section, the first three and the last three sites are grouped together with a distance of 1.5 miles between the two groups. The first three sites have low severity fatigue cracking extending over 50 to 100 percent over each site, while the extent of cracking along the last three sites is between 0 to 15 percent. The second group of sites is located on a fill section. The variations of the deflections (measured bythe seven FWD sensors) along the length of the entire pavement section is shown in Figure 6.2. The backcalculated layer moduli are presented in Table 6.6 and they are discussed below in detail as an example of typical results for pavements with a relatively high variation in the measured deflections. 2 Examination of the values of the backcalculated layer moduli listed in Table 6.6 indicates that: 1. The AC modulus varies from a low value of 711,248 psi at test location 203 .EmeE 8:8» .5862..— .8.. oan canoe—SQ .me 2:»...— uenadn noaueoon unea a: .... 3630688 Ta If_ 204 Table 6.6. Backcalculated moduli for pavement section MSU07F. VET 2‘ "_ ‘ 2' if i 2 E Cl T 2‘ ‘ if 'L Ti ‘1 FWD test Test location Backcalculated moduli (psi) RMS Error 1 P code No. AC R oadbed (%) + ; 07112011 3 711248 33538 30701 1.04 41 3 07122411 12 1106757 34646 29804 0.72 " 07127721 15 1131439 33336 36455 1.25 07128332 17 1320628 54731 34356 1.55 07131211 19 1028024 43254 32759 1.46 07137022 24 767833 21714 27864 0.36 g 07141211 26 909228 20197 27628 0.79 205 number 3 to a high value of 1,320,628 psi at test location number 17 (a factor of about 1.86). The base modulus varies from a low value of 20,197 psi at test location number 26 to a high value of 54,731 psi at test location number 17 (a factor of about 2.71). The roadde modulus varies from a low value of 27,628 psi at test location number 26 to a high value of 36,455 psi at test location number 15 (a factor of about 1.32). The maximum values of the root-mean—square of the errors between the seven calculated and measured deflection values is 1.55 %. In order to discuss the variations in the values of the backcalculated layer moduli provided in Table 6.6, the measured pavement deflections at sensors 1 (the inner-most sensor) and 7 (the outer-most sensor) and the AC thickness for test location number 3 were designated as datum values and the percent differences between the datum values and the corresponding values at other test locations were calculated and are depicted in Figure 6.3. Examination of the figure indicates that: l. The relatively high AC modulus values at test locations. 12, 15, and 19 can be related to lower measured deflections (sensor 1) and AC thicknesses at these locations. Thin AC layers should lead to higher deflections. The low measured deflection values at these locations relative to those at test location number 3, imply that the AC layer is stiffer and hence its modulus is higher. For test location number 17, the measured deflection at sensor 1 is lower than that at test location number 3, yet the AC thickness is almost the same (slightly higher). Again, the stiffer (higher modulus) of the AC is the main contributing factor to lower deflection. For test locations 24 and 26, the measured deflections at sensor 1 are higher 206 $8sz .528... .5892. wee—e - 93 _ .828 do essence—o 05 Pa .8052... U< 05 .2358 so»... Basso—8.63 ...o coma-53 .nd 2:93 cosmoon.mmp ow em m P t m F N P - aomcow 53.3.0.2 venomous 3.2.22 33 a ow.- P newcomg 3.02... 0<§ 322002 0(- cv- om- o no w cm a U 0 oe ....r m on on cow 207 ‘ than those at test location number 3 and yet the backcalculated layer moduli are higher. One may expect that higher deflection values imply softer materials and lower moduli. Although this is true, the thicknesses of the AC layer at these two test locations are substantially lower than the AC thickness at test location number 3. Lower AC thicknesses cause higher deflections. The above scenario implies, as expected, that the deflection of a pavement structure is a function of both the modulus and the thickness of the AC layer. Variations in any one variable lead to variation in the measured deflection. At test location number 24 and 26, the thickness of the AC is the main cause of higher deflections. The variation in the values of the roadbed modulus is well explained by the variation in the measured deflection at sensor 7 (located at a lateral distance of 60-inch from the center of the applied load). It can be seen from Figure 6.3 that a higher roadbed modulus corresponds to a lower‘deflection at sensor 7. For test location numbers 12, 15, 17, and 19, the values of the base modulus is higher than that at location number 3, while they are lower for test location numbers 24 and 26. In order to probe further into the variation of the base modulus, the percent differences between the deflections measured at all seven sensor at test location number 3 and those at locations 15, 17, 19, and 26 are shown in Figure 6.4. It can be seen that for test location number 17, the deflections measured by all sensors are appreciably lower than those measured at other test locations. In particular, the deflections of the middle sensors (sensors 2 to 5) are much lower, which indicate the presence of a stiffer (higher modulus) middle layer (in this case, the base layer). Inversely, for test location number 26, the deflections measured by all middle sensors are appreciably higher, which indicate softer base layers. 58:22 5:08 22:33 wee—a 88:8 :38 =a e8 «cocoocou mo SEES .ve 223m ..mQEDZ howcmm 208 N. m m v m N F _ _ _ . _ _ _ CV: I. ON: 1 1 x 1 x . x x l o m y / / n m 1 fl ” fl ” u" n / / / / m I n y y y e w H ” ” fl u ” y H H -8 o e / / / / g y y H ) y fl ” fl % V g y ( fl ” r / I ow V cm 5:83 H 2 8:80.. E t 5.88.. E m. 5:80.. I om 209 The above observations indicate that the variations in the backcalculated layer moduli along pavement section "MSUO7F” are very reasonable and consistent with the variations of the measured pavement deflections and AC thicknesses. 6.4.2 Flexible Pavement Section MSU19F - Uniform Deflections This pavement section consists of a total of eight test sites divided into two groups (each consists of four sites) which are separated by a one mile stretch of pavement. The pavement is in a good condition with no significant distress. The AC thickness within the section varies from 5.1 to 85-inch. The deflections measured by all seven sensor locations are shown in Figure 6.5 and the values of the backcalculated layer moduli are provided in Table 6.7. This pavement section is representative of those sections whose behavior is more or less uniform along the length of the pavement. Examination of the values of the backcalculated layer moduli listed in Table 6.7 indicates that: 1. The AC modulus varies from a low value of 215,731 psi at test location number 45 to a high value of 391,656 psi at test location number 10 (a factor of about 1.82). 2. The base modulus varies from a low value of 56,760 psi at test location number 45 to a high value of 82,908 psi at test loeation number 17 (a factor of about 1.46). 3. The roadbed modulus varies from a low value of 40006 psi at test location number 40 to a high value of 50,857 psi at test location number 15 (a factor of about 1.27). 4. The maximum values of the root-mean-square of the errors between the seven calculated and measured deflection values is 3.54%. 210 $2 3m: .8508 30838 e8 0565 couoocoa HOSE—HQ QOHHGOOH 0.003 .3 2am F1 an I § o v 8 E III I . H I] ~ ll r5 I‘ If} all {.l N la ONOMQ‘M m Ann :6 30300:!— 211 Table 6.7. Backcalculated moduli for pavement section MSU19F . _ _ _ — A 7 i 7+ ‘— ._ _ ___H_________ _i., . ~~fm_._ __‘ FWD test Test location Backcalculated moduli (psi) RMS Error 1 °°d° N°° AC Base Roadbed (%) ‘ .______—— _.________ , _ ,__1 ? 19113011 2.99 3 346848 71962 50190 ' 19118311 8 321499 68224 48739 3.41 ' 19121611 10 391656 69329 46408 2.03 [I 19128211 15 271118 77296 50857 3.02 n 19131111 17 257348 82538 50809 1.71 ; 19132312 19 253084 82908 50598 3.39 : 19137411 22 364790 76309 50215 3.19 . 19141811 25 258563 72029 43146 2.57 19148111 31 317965 76003 40839 3.22 19150911 33 263608 74919 41806 3.54 19158411 38 250851 77889 49048 2.45 19161211 40 227458 72986 40006 3.40 19169111 45 215731 56760 41228 3.39 j 212 As for the previous pavement section, the values of the backcalculated layer moduli provided in Table 6.7, the measured pavement deflections at sensors 1 (the inner-most sensor) and 7 (the outer-most sensor), and the AC thickness for test location number 3 were designated as datum values and the percent differences between the datum values and the corresponding values at other test locations were calculated and are depicted in Figure 6.6. Examination of the figure indicates that: l. The variations in the measured deflections are small and so are the variations in the values of the backcalculated layer moduli. For test location number 8, as it is expected, the deflection measured at sensor 1 and the AC layer thickness are higher which lead to a lower AC modulus. For test location number 10, the deflection measured at sensor 1 is high because of the AC thickness which is substantially lower than that at location 3 (a difference of about 10 percent). Hence, for this test location high deflection is caused mainly by a lower AC thickness rather than by a softer AC layer. An apparent discrepancy exists for test locations 15, 17, and 19. For example, at test location number 17, the AC thickness is almost the same as that at the datum (test location number 3) and the difference in the measured deflection at sensor 1 is moderately low, and yet an appreciable difference in the backcalculated AC modulus can be noted. This apparent discrepancy led to a further examination of the measured deflection records. It was noted that the NDT tests along this pavement section was conducted at noon time on a sunny summer day. Further, the temperature of the AC surface (which was recorded during the test) at test location number 17 was ten degrees higher than that recorded at test location number 3. Hence, the lower AC modulus is mainly due to the higher temperature of the AC. In order to confirm this, the percent differences between the deflections measured at all sensor and those measured at test location number 3 are plotted in Figure 6.7 for test location numbers 8, 213 .maSwZ 5:8» 2059.2. ace... h 98 _ .528 ..e cozoocov 2.. P... 685.35 U< o... ....69: 3.3. 33.3—88.8: .3 5.353 COZGOOA 30... mm 2 t 2 S m n 32.5 22.360: nonuuomg 3.3622 3.52 F 8338 $0.5. o<§ 2.3.3.2 o ll\||\ll\l’)ll‘\ll[|’lf\/I\1l I a iIIIIK/llllllllltlf o A J t IfllJllrb.‘ 881'. . . Ia! I’Jflfllfllll .‘ lllJIIvs ._ a he.“ 0:1.ch Exes—hon— .— m 6 eawmm NOAH—dam flOflHCOOH HIGH. 222 on A on «I. £03002!— 223 conditions along the section. Table 6.10 provides a list of the values of the backcalculated layer moduli. It should be noted that some existing backcalculation routines (such as MODULUS 4.0) recommend the use of a fixed AC modulus for thin pavements. In this study, a fixed AC modulus was not used. All values were backcalculated using the MICHBACK computer program. Nevertheless, the trends between the values of the backcalculated layer moduli and the measured pavement deflections and AC thicknesses are reasonable and consistent. 6.4.6 Flexible Pavement Section MSU29F - Variable Deflections In this section there are two groups, each group consisting of three test sites. The groups are separated by a distance of 3.2 miles. Difficulties in the analysis of the measured deflection basins for this section led to a further investigation of its conditions. It was noted that at least a portion of the outer pavement lane (the traffic lane) and the shoulder are located on an old portland cement concrete (PCC) pavement. The section has 100 percent medium to high severity cracking (mostly a combination of fatigue and block cracking) which had been sealed. The first three test sites are in an uphill fill section, whereas the last three sites are in a downgrade. A wide variation in the drainage conditions along the section was also observed. Some of the driveways to existing homes along the road were blocking the natural water drains. High variations in the deflections measured at all sensors were observed and are shown in Figure 6.12. The variations follow a specific trend. The deflections between test location numbers 1 and 28 are comparatively low and are interspread with approximately equally spaced high blips. The deflections between stations 28 and 40 are however, consistently high for the first few sensors and consistently low for the last ones. After station number 40, the deflections taper off to lower values. In 224 Table 6.10. Backcalculated moduli for pavement section MSUl4F. , FWD test Test location Backcalculated moduli (psi) RMS Error 1 °°d° N°‘ AC Base Roadbed (%) + .__ - _e _ ,, .. _ _ __-- ___-fl- __-- v _ ' 1817913 21620 . 14113013 14121713 18887 14137811 31124 14143711 29577 14148213 1976844 29702 14151513 1217635 24865 14155113 38 1568918 23277 20567 1222 14161813 1187707 L____ ___._.__,_, _ __. ___ ________ ___ 2658550 1748687 2073122 225 90 £3sz 8.88 .52—88.. .8 0.2.9:. canoe—.225 .N. .0 2:3". "Gan flOfiOIOOH HIGH. A a— at. pagan—hon 226 addition, the shapes of the measured deflection basins at the various stations are not consistent as depicted in Figure 6.13. The values of the backcalculated layer moduli are provided in Table 6.11. It can be seen that the values are also highly variable and they follow the trends of the deflections along the road. For example, the values of the layer moduli are consistently high for all locations where the measured deflections are comparatively low. Further, the wide variation in the drainage condition has caused compatible variation in the base and roadbed moduli. Nevertheless, the backcalculated moduli values should be viewed with caution because of the presence of the PCC slab and the high distress conditions prevalent in the section. 6.4.7 Flexible Pavement Section MSU35F - Variable Deflections This test section consists of six almost adjacent sites (the first three sites are only 0.1 mile away from the last three). The thickness of the AC measured from the cores varies from 5 to 7.3-inch. A combination of low severity fatigue and block cracking was observed along the entire test section. One to two high severity transverse cracks were also noted along each of the test sites. Figure 6.14 displays the high variations in the deflections measured at all sensors along the pavement section. Except at a few locations, there is a recurring pattern of moderate deflection fluctuation which indicates the presence of equally spaced cracks. However, the shape of individual deflection basins was found to be consistent when compared to that of section 29F. The values of the backcalculated layer moduli are listed in Table 6.12. The values are very much consistent with the variations in the measured deflections and AC thicknesses. 227 .25sz 8.282 .5528. 26. an Ea .2... .... .22 00 av ...... 88.25 .38.. mm .2: 22:62.86. .8. .22 2.8.. 52.8%.. .83»... .VN NF .2 ... 22.2.". AN 8.283 .9 .2. 8.283 .... mm 8:83 + .12 cozmoo. ... 3 8.268.. * or- (sum) suonoeueo Table 6.11. ! 29111711 228 Backcalculated modulus for pavement section MSU29F. 535056 30363 24682 FWD test Test location Backcalculated moduli (psi) RMS Error é 1 code No. AC Base Roa db ed (%) w 2 3. 29117721 471364 39576 29134511 773636 11197 29146911 719813 22708 29172711 216579 6424 29176321 283858 8025 ‘. '29186711 50 325096 27059 17819 L.91332 336393 229 .mmmbmz .5208 208025 he oEoE cocoocon .2 6 2=mE NOAH—=3 flOdu¢GOOfi HQOB a.“ NH nu— m .6 son—boas: 230 Table 6.12. Backcalculated modulus for pavement section MSU35F. 35112211 35114222 35118931 35123311 35128321 862525 35132611 528756 35136221 718640 35143911 ' 501497 35152311 463521 35159121 365961 35161411 472025 ? 35167331 529678 231 p ‘ 6.4.8 Flexible Pavement Section MSU43F - Variable Deflections This test section has a total of three groups, each consists of three test sites. The groups are separated by distances of 0.3 and 1 mile. The pavement section is characterized by very high rut depths, especially in the first three sites. Low severity cracking was also observed along the entire section. Figure 6.15 depicts the high variations in the measured deflection profiles along the pavement section. It can be seen that continuously reoccurring blips of high deflections were measured. The shape of individual deflection basins however, is mostly smooth and consistent. The backcalculated layer moduli listed in Table 6.13 have moderate variations in the base and roadbed moduli and high variation in the AC modulus. These variations are very much compatible with those of the measured deflections. For example, the deflections measured at test station number 29 are relatively higher than those at station 30. Consequently, the values of the backcalculate layer moduli at station number 29 are lower than those at station number 30. 6.4.9 Composite Pavement Section MSU01C - Variable Deflections The eight test sites of this section are located in two groups having five and three sites, and are located 1 mile apart. Beside edge cracking, the main distress observed along this section is in the form of high severity sealed transverse cracks (reflective cracking). This section has very heavy truck traffic. Figure 6.16 depicts the high variations of the measured deflection profile along the road. The towering blips of high deflections are indicative of the presence of high severity transverse cracks which are closely spaced in certain locations. The shapes of the measured deflection basins are not consistent and are not smooth as shown in Figure 6.17. The values of the backcalculated layer moduli are listed in Table 6.14. The 232 .5352 c288 .5802... .8 0E9... 5.82.09 .26 Baum... henna: nowuaooH unoa on 8 am 2. on on 3 a .....m.m.s.~.m.m.o.o.s.~...” lift/211.1. .11! .../12.51114. N 11.111113711711. 431.22 11 1. 415111.11 - n ”v m .6 Us a .m A u— 13 segue—«on 233 Table 6.13. Backcalculated moduli for pavement section MSU43F. FWD Test Test Location Backcalculated Results (psi) RMS Error C°de N°' AC Base Roadbed V (%) 43111711 2 580509 22886 36221 1.35 43121811 9 873327 13926 46995 2.26 43131911 16 890107 14774 44970 1.65 43141611 23 875335 17540 41522 3.06 43148313 29 597983 19410 42198 1.28 4314831 1 30 422440 24850 38947 1.47 43152311 33 713238 17787 36701 1.10 43161411 39 483639 20058 34977 2.12 43178511 50 589905 19907 39101 1.07 431781 12 51 485790 20296 39462 1.60 43188111 57 531447 21100 40511 1.26 43191411 60 564445 19098 39715 1.64 234 am 05:22 5:09. 25:82... .8 .6an .528ch .2 .c 2sz Hog“ QOflHCOOH u. '03 @h-sflLflV‘I'flNv-ID A .m— mzu segue—E 235 .05sz 5.83 .5898 3.. pm vs. .3 .v 28:82 .8. .a «£8: 8.80:2. 30...»... ...... .e 2:»... E. 8:965 _mfimm oo 9. mm. 1N N _. o t. 8:83 * mm 8:86.. + m- v co_.moo._ .... 0') (SIM) suouoeueo Table 6.14. '__._._ _ Backcalculated moduli for pavement section MSU01C. Backcalculated moduli (psi) _M" 1 RMS Error 1 g °°d° N°' AC Base Roadbed ‘91) r1284611 4 337073 5406975 16271 2.28 I 1 01272411 12 319451 2636877 18653 2.36 01261611 17 282635 4694486 23958 1.01 II 01264221 20 195181 1520576 22012 0.92 ; 01251611 25 302239 3191544 18924 1.39 1 01240411 31 762348 5040840 18107 0.95 1 01248931 35 1012610 524600 25662 1.12 I 01232611 42 725415 3293690 22761 1.33 3 01221711 47 610169 4900569 35123 1.57 1-01114 55 523795 _ 1‘142 19998 116. 237 variations of these values along the pavement section are consistent with those of the measured deflection profiles. For example, the measured deflections at station location number 47 are relatively lower than those at station 17. Consequently, the backcalculated layer moduli are higher at the former station than at the latter one. 6.4.10 Composite Pavement Section MSUOSC - Variable Deflections There are six test sites inthis section located in two groups of three sites each which are approximately 1.0 mile apart. Lane-shoulder separation and medium to high severity transverse cracks (reflective cracking) were observed along the entire section. The transverse cracks are equally spaced and are occasionally connected by longitudinal cracks. Figure 6.18 displays the extremely high variations in the measured deflection profile along the road. The equally spaced high deflection blips in the figure correspond to the equally spaced high severity transverse reflective cracking. Table 6.15 provides a list of the values of the backcalculated layer moduli. Variations in the moduli values correspond to the variations in the measured deflection basins. For example, the measured deflections at location number 28 are much higher than those at locations 8 and 17. Correspondingly, the values of the backcalculated moduli of the first station are lower than those of the latter two stations. Further, the MICHBACK computer program did not converge on any set of moduli values for station number 45. The reason for this is the irregular shape of the deflection basin. Figure 6.19 displays the deflection basins measured at four location numbers 45, 46 (located only three feet away from station 45), 43, and 28. It can be seen that the shape of the deflection basins vary substantially from one station to another. Such variations in the shape of the deflection basins and in the values of the Mk deflections are the consequences of the state of distress of the PCC slab. Therefore, for any distressed pavements, the backcalculated layer moduli (even with 238 mm .08sz 8.38 32.5%.. .8 Eco... 5:8..8 .w. 6 Saw... Hausa: nouunoou boob mmcm mN ON 1-4-1 INQUIV'O'JNv-ID A 1 A l I m A an 16 503.0023 239 Table 6.15. Backcalculated moduli for pavement section MSUOSC. 1FWD FWDtest Test location Backcalculated moduli (psi) ~__RMS Error 1 1 code N o Roadbed (95 ) h . 1 05218611 1134797 8519459 30247 1.24 ; 05229211 17 1292937 7494653 34832 1.11 1 05241411 28 1662644 4757409 33519 2.38 1 1 05258311 43 912615 7436017 20367 2.28 1 05261712 45 No Convergence 1 #012 46 _ 1056789 7104805 19389 _8___ _1 240 om wv 0m v N N? 8 8:83 * 8 8:83 8.4 8 8:83 + me 20:80. .... P I 0882 8:8... 882... 8. 8 Ea .9. .9. .8 8:86. .8. a 2.8.. 8:88.. 8...... .2... as»... E: 88.65 886: (SM) suouoeueo 241 small root-mean-square error) must be viewed cautiously. 6.4.11 Composite Pavement Section MSU08C - Variable Deflections In this section, there are two groups separated by a distance of 2 miles. Each group consists of three adjacent test sites. There are 5 to 7 transverse (reflective) cracks in each site. The pavement is constructed on a 4-feet high embankment. Figure 6.20 depicts the deflection profiles measured along the pavement section. It can be seen that the profile is characterized by fluctuation caused by the presence of reflective cracks. These variations in the measured deflections cause the similar variations in the backcalculated layer moduli provided in Table 6.16. Figure 6.21 depicts the irregularities of the deflection basins measured at various stations along the road. As for the previous composite pavement section‘s, these irregularities raise questions about whether the backcalculated moduli are reasonable. 6.4.12 Composite Pavement Section MSU04C - Variable Deflections The three adjacent test sites of this section are apparently in good condition apart from the observed medium severity transverse (reflective) cracks. The section is located on a 4 to 5-feet high embankment. The thickness of the AC overlay varies along the pavement from 2.3 to 2.9-inch. During coring, a buried AC pavement was found underneath what was thought to be the roadbed soil. Because of the limitations of the MDOT coring equipment, the thicknesses of the various layers of the buried pavement structure could not be determined beyond the upper 3-inch. Figure 6.22 shows the unique and moderate variations in the deflection profile measured along the road. It can be seen that the four lines representing the first four deflection sensors intersect and criss—cross each other frequently. Further, the prominent high deflection values in Figure 6.22 correspond to the presence of transverse cracks. .08sz .838» 22:98 8.. .oan 5:00:09 dud Bani henna: nowuuoon vans mm on mN ON mu SH m a 242 I d I I I u I u I u I u I d I cw 8 so a m« on 8 . OOF‘DWQ‘MNI-IQ A an 2.6 £0303: 243 Table 6.16. Backcalculated moduli for pavement section MSU08C. I‘" Test Ilqcziation Backcalculated moduli (psi) RMf%FSrror ' AC hBase Roa__bed __ Faun 1033037 2993202 11126 2.54 fl' . 03213022 1707413 2639131 20307 1.67 ; 03221011 705630 6592433 10362 3.53 i 03231511 15 1516300 2365745 13153 1.86 ' 03241011 21 390504 3311759 20363 1.31 03251311 27 437404 4267114 14644 2.03 03260911 33 1350763 2221734 13371 1.31 03261622 34 301636 4523369 22724 2.14 244 .08sz .5208 26893 he an Ea .m .h .N 22.82 as .a «53.. Stucco—c .539? .36 Bani :5 8:320 _mfimm om we om. VN N P o ”I N c0380.. X. A 8 8:86.. 36 m c0380.. 1T - T w- n c0283 1.. lTll_\ G . m... m. m... v. m S M 7....1. .. N- 245 03sz 528.». 2.0826.“ ..8 cache cecoocoa dud 2=wE Honda: nowuaoon uaoa mu 8 ma 3 m a me my. . of . no 1 .o -H ghw ..m J ..m .... .n nan «Iv coma-bozo: 246 For this section, the MICHBACK computer program did not converge on any set of layer moduli for any of the test locations. The reason for this is the extremely irregular shapes of the measured deflection basins. Figure 6.23 depicts the deflection basins measured at three different locations. Beside the presence of the buried pavement structure, no other explanation can be offered at this time regarding the irregularities of the deflection basins. 6.5 BACKCALCULATION OF LAYER MODULI AT DIFFERENT LOAD LEVELS During the summer of 1992, the FWD testing was expanded to include different load levels. All the non-cored pavement sections were tested at the standard target load of 9000 lbs. At each test location of the cored pavement sections, the NDT procedure was modified to include seven drops as follows: 1. The first drop 4500 lb (pavement seating). 2. Second drop 4500 lb (the deflection data was used for backcalculation). 3. Next three drops 9000 lb (the average deflection and the average load were used for backcalculation) 4. Seventh drop 16000 lb (the deflection data was used for backcalculation). To avoid unnecessary repetitions, only the results of backcalculating the layer moduli for a few typical test locations along two flexible and one composite pavements are presented and discussed in this section. Table 6.17 provides a list of the types of the roadbed soils (the information is obtained from the 1970 MDOT Field Manual of Soil Engineering) encountered in the four pavement sections. 247 03sz 5:08 3659a c8 E 98 .2 .2 .v 30:82 .8. 3 «£23 550:3 .833. .36 Esmfi :5 8:265 _memm 00 av om VN N _. o “I up 8:83 3.. 0.. c2300.. Jr . 10.. v c0380.. .... _ ml - G .11.... ..- m m. U 3 ml ) , m. h...” N... —.1 o 248 v ode _d -.,--.:.;.~c - - v.” keen 832.3 08sz - _ _ a .2 mm 8 See 3.886 2282 v _ N 2 mm an 860 cacao .3sz _ 8:80 8:602 05m > . . «_U :5 65m 3550 ~5qu _ ouncan .5383 22:93 _ A2303 .3 $85809 25 :8 Bowman .2269. 22:33 025 be 82380820 3&on .2 .0 035. 249 , 6.5.1 Flexible Pavement Section MSU19F - Variable Load Level A brief description of this pavement is included in Section 6.4.5. The test section is located in Otsego County and the roadbed soil is granular in nature. The backcalculated layer moduli for the three load levels are listed in Table 6.18. The layer moduli backcalculated at the standard load of 9000 pounds are referred to in the remaining parts of this discussion as the reference values. The percent differences between the reference moduli and those calculated at the other two load levels were ‘computed and are listed in Table 6.19. Examination of the layer moduli calculated at the various load levels and the percent differences between those calculated at the standard load level and those at the other two load levels indicates that: 1. Relative to the reference moduli of the AC and base layers, the modulus values calculated at the low load level are consistently low, whereas those calculated at the higher load level are consistently high. 2. The variations in the backcalculated roadbed soil moduli are within a tolerable range. First, the variations in the AC and base moduli backcalculated at the different load levels were investigated in relation to the measured deflection basins. Figures 6.24 through 6.27 depict four typical deflection basins measured at three load levels at test location numbers 19113011, 19118331, 19121611, and 19128221, respectively. The dotted lines in the figures represent the linearly extrapolated response of the pavement structure at the low (4648-pounds) and high (16080—pounds) load levels assuming that the pavement is a linear elastic system. That is, if the deflection basin measured at the standard target load level of 9000 pounds is used as a reference basin, and if the pavement response is purely linear with the load, then the measured basins at the low and high load levels should be exactly the same as those shown by the dotted lines. Ingreference to Figure 6.24, it can be seen that: 250 1 -__._ __ __-- _ _J 2: 8...: 2: Swan ~82 ....... o5“ 88v 98° Sam 88 SSE a 8+ 3. 3:“ 3;. a; 3:- was" 89. 8a 3. 95. SM «8% SH cannon :52 m a: a was. an? $3.8 88 :22: _ :a 3 3N9. ad. “was. «.8. $8: $3. a Ea S- 58 «.2 85 92 82mm 38. m $6 a N9: ”69. «mag 38 _ mm: _ a _ mam ed «8: ”.2- 28... as- was: ...ch , ”.3 1. $8.. 98 «8% ma 3:: 252 n 8a «8:. 5% 253 R8 .825 8.3. «3 SE. . 3 _- as ea. £83 32. E s: s: A v snap—yawn— mfl—flgz OOH—euka—mfi— mud—=8: ”oceukbmfi— 9:582 R 8 Sum 25. 8% a as 0... 9%: gowns cue =38. 3283308 _ : _ 95m $9me couoa Eon—gum he «was «8.53 3 =28... Esau—8x08 .3 .c 3.3. Ilil, 251 8.. mmd ou.~- 9.? an... 3...- flag- anew. came gamma. 3... T .....- 5d- mad- av.» 26 3.2 one... .w.~ Ed mad- :6- had. 8.8- 3.9- .32 «one 2322 mm..- 8... mod wvd w. .h 2.....- and. $9. an. E... mmd- wee- 3.9- 3.2- SNT v80. ammm 3mm. 3. 3.2- 36. N...— vw... 2.6 0nd. mnd. 36.. no. 8.. 8...- 8.». mm..- xx..- 3&7 once. the Zen. 2... Ned. nnd- 9.0 8.. ma... med 3.... 3.9. N. e n v m a . 3: one... 33 22.82 529. ... 22.00:“... ... 855:5 .05. e8: .8. DE .3sz 5.69. 2.2522. 8.. 2.93 558:2. 3.29.8.5 3.8:: v5. 3.3.8.: c852. «8:80....5 .96 05¢... 252 A:2:::ofiaaynszc "8. ans. 5.89.. .5502... 3.. minus 538:3 8:38. 53:: 3.035 E... 3.38:. ..o Satan—o0 .vué 25E om E. 8555 36mm ow on vm NP 0 as 9.. coco. 33* .. 23 33+ V F.- NF- 0 m. mlUI. m m. U S m M. fl 253 .2528. 2.8 .8. GE. m2 am: 5.88 .5833 8.. 2.3: 5.80:2. 8:5 .35. 3.896 can 3.3.8:. ..o coming—coo .mufi. oSwE E. 85.20 Emma 8 9. on em N. o ...-In cocoa-cc =05..- . :8. 33% ouua usu..+u 33 33 .u. In "F 2 (Sum) suonoeueg In I 254 .2838. 2.8 .8. QB... ”8.sz 5.89. .5802... .5. 2:93 538:2. 8:88. .8...— eroaxo ...... 3536:. ..o catwanU .86 23E E. 8555 _mfimm 8 .... 8 «N N. o Egan Cocoa-o: .35.... . ON- :8. 33% 8a.. .23.. ..T noov 3'04 :0. to ‘T é (sutu) suouoeuea lo I 255 .2838. 88 ...... ASE. "5.sz .8qu .5802... 8.. «£3. 8.80:»... 8.8%.... .85. 3.898 ...... 52:82: 888.8858“. .88 2.3.". ..c_..mocm.m_o 3:2... 8 .... 8 ..m m. o m _.- 5:0 cocoa-o: 32...... . «no: 3.0.. X .. on: 2.3+ . V _... 03' use. .0. N T m o T U 9 O H... O U S ) w. ...... A» ..... - - ._.| 256 1. For the lower load level, the measured deflections at the first five sensors are higher than the linearly extrapolated responses, whereas, the measured deflections at the outer two sensors are almost the same as those of the linearly extrapolated responses. 2. For the higher load level, the measured deflections at the first five sensors are lower than the linear responses, whereas, the measured deflections at the outer two sensors are almost the same as those of the linearly extrapolated responses. The above two observations imply that the increase in the pavement response is not linearly proportional to the load level. Discussion of this apparent non-linearity in the pavement response is presented in subsection 6.5.4. In the remaining part of this section, the trend between the backcalculated layer moduli and the measured deflection basins at the three load levels is discussed. Table 6.19 provides a list of the differences between the deflection responses measured at the three load levels and those calculated by using a linear analysis. Figure 6.28 and 6.29 show the percent differences between the deflection measured at sensor locations 1, 2, and 7 due to the low and high load levels, respectively, and the corresponding linearly extrapolated deflections. Further, the percent differences between the layer moduli backcalculated at the same load levels and the reference moduli (obtained from Table 6.19) are also shown in the figures. Examination of the data listed in Tables 6.18 and 6.19 and shown in Figures 6.28 and 6.29 indicates that (for ease of discussion, consider test location number 19113011): 1. The measured deflection basins at the 4648 pounds load level are higher than the corresponding linearly extrapolated responses (Table 6.19). Consequently, the backcalculated AC and base moduli are lower than the reference values computed at the 9277 pounds load level (Table 6.18). 40 r- 257 8:08 EoEo>~q wee—a 5 Ba .N ._ 88:8 we couooaoe u... can :25:— 5.?— cozmofl 02:. .32... 32 32 a "823: 333.883 .8 883$ .83 28$ 5wa 5.. Sb rm 5? . memzmw ramp—.2. on. o o o '7 9‘ $3 (%) ewe-name ON n 82.5 Q manna—2 nonumom a N .623 a 3.385. ommm a on F .omcom m mas—cos. O< I ov 258 ..26. 82 fie. .8 85m: 5:09. 2.08on— weefi h 98 .N ._ 88:9. .3 558:8 05 can :32: .9?— §m_=o_~oxomn ..o count—3 dud Ezmfi 8:83 Es“. memNPmF armrmrma memFPmF FFomFror 0v- 0N- E / / E G / - / / _ / mm fl fl fl fl V. / fl / om w fl . V. . / \w / W r at n 59.8 g 3.322 nonnmom I N .533 a 3.:qu mmwm a co F 83% m 3:622 0(- 10 Ml me 259 2. The measured deflection basins at the 16080 pounds load level are lower than the corresponding calculated linear responses (Table 6.19). Consequently, the backcalculated AC and base moduli are higher than the reference values computed at the 9277-pounds load level (Table 6.18). 3. For the outer-most sensors, the variations between the measured values and those calculated by linear extrapolation are insignificant (Table 6.19). Since the outer sensor measurements strongly influence the backcalculated roadbed modulus, the values estimated at the various load levels are almost constant (they vary within a very reasonable range). See Table 6.18. The three observations indicate that the results produced by the MICHBACK computer program are compatible with the measured deflection basins. That is, for all tes¢ location numbers, the values of the layer moduli backcalculated for any load level are “‘8 her or lower than the reference values if the measured deflection basins are, “emf-i Vely, shallower or deeper than the linearly extrapolated responses. This tend to support the contention that the variations in the results obtained by the MICHBACK computer program are accurate and reflect the variations of the meaSUI‘ed deflection basins. 5-5'2 Flexime Pavement Sections MSU35F - Variable Load Level 'I‘his test section is located in Cheboygan County and the roadbed soil is similar to that for section 19F (Table 6.17). The trends of the deflection basins "‘qu at the three load levels and the corresponding linearly extrapolated basins for three test locations are shown in Figures 6.30 through 6.32. Table 6.20 provides a “St of the layer moduli computed at the three load levels and the percent differences between the reference modulus values (calculated at the standard target load of 9000 pounds) and those determined at the other two load levels. For the three load levels, T ' . . able 6.21 provides a list of the differences between the measured deflection basins 260 .283 _mm 0.88 32 9?: "$252 8:08 .5833 be v.53: 558:3 8.5%2 58:: 3330 v5 338.: ..o eaten—=00 .86 Semi :5 8855 _aeam 8 ma 8 am «F o ON- ...-nu cocoa-cc 325.. . vnoo— too; % OONO fiscal—I choc too.- .O. . . m ..- 2 (sum) suouoeueo a. I 261 .ANNNE .mm 3.8 8.... 95m. "Emma—2 5.88 .eoEoZ... 8.. «£82 5.80:2. 8.888 .82.: 3.896 vs. 3.88:. .8 5859.50 .86 8:»:— E. 88.85 .98.. 8 ma 8 em 3 o oN- scum cocoa-cc 30.....- . Nmoo' vcoa% opna vocal—I an: ...-0.. .0. m T s2 (sum) suouoeueo ID I 262 .9 88.8 88 .8. 8.8. ".2392 82.08 .8838. .5. 888.. 538:3. 8:88.. .88. 8.88.... ...... 8.8.88 ..o 88.82.82. .96 8.8.”. E. 8855 _afiam 8 8 8 .2. N. o ON- ...-an 02.083. 32...... . 0000' .30.. * oouo t-OJIT :3 ago.— .0. m T 'o (sum) suouoeueo In I 263 :2. 3. 35“ 3.8 an: 2: 38% m8: *3 am? 33 ”8mm 88 2822 :3 31 Sam 3- «39. En. 32.x 2.9. m3 3. 3m: 3m RE 3 33mm «82 _ :._ Roam 88* $3: 28 «$22 a; 92 Sam 3. £3. «.2- :33 a: x: 2. Emma 98 £8 2: 88c. «88 8a EB “82 $2.? :8 8253 ”S 3 23a .. I: «an? ”.2- n83 28 $3 $3 33 s: 853:5 8382 8§£E . 2382 8:235 3382 arm as”. “Samoa 9.8 0< s: 38 9.5 __..85 3233.98 .26. 83 :3 HE E .mmmDmZ .8qu 82:33 8.. £32 30E 3 =59: 32338.08 .omd ozfl. ommmhvmi CCmugahv ht‘E‘sI-u.‘ .hCI‘ al.-"pie!“ :FH-QECIQ \ a‘ . . — 1 .U.CEF£>Q ...Lctt: it.“ tea-.552: car». 52...: «COCOLJ ..CXH .~N. . we Oxawuwu 264 cod 36 $6- 3.x- 3.9- vde 8.2. man— ooma Snag?“ v.3- 3.». :8- 3. _- mm: onfi ov. : Kev «fin mod- 3.? 2.5- 3.9 02:- 9.6—- amoe— Ema Nani—mm 8.x- 36- 3.»- S6. 8.. Ed 96 on; mmd ow. T mod- $3. 86. 3.2- on; T ”.33 mama 59: _nm end- mmfi- cad- «Ya- 36. ac; and 2.3. b o n v m a _ 35 360 33 25:83 820m 3 2280:2— E 8:80:qu :33 33 “non. OE .mmmDmE cocoon EoEo>a he 2:23 538:3 3288.5 .355— 93 338:. 5233 moo—Eats 4N6 05E. 265 and those calculated for a linear elastic system. It can be seen that the trends of the data in the figures and tables are the same as those for pavement section MSU19F presented in the previous section. 6.5.3 Composite Pavement Section MSU01C - Variable Load Level This pavement section is located in Lenawee county in the State of Michigan. The pavement exhibited medium to high severity transverse reflective cracks. The roadbed soil is cohesive in nature (see Table 6.17). The percent differences between the layer moduli backcalculated at the three load levels are listed in Table 6.22. Table 6.23 provides a list of the percent differences between the measured deflection basins and those computed by assuming a linear elastic system. Figures 6.33 through 6.35 depict the measured deflection basins and the linearly extrapolated responses (dotted lines) at the low and high load levels for test location numbers 01284611, 01272411, and 01261611, respectively. Besides the difficulties associated with the analysis of a composite pavements with deteriorated concrete slabs, the trends of the backCalculated results are similar to those presented above. 6.5.4 Discussion In this section, the discrepancies between the deflection basins measured at the three load levels and those calculated by assuming a linear elastic system (i.e, the apparent non-linearity in the pavement response) is discussed. These discrepancies can be attributed to several factors including: 1. Poisson’s Ratio - During the analyses of the deflection basins, Poisson’s ratios of the AC, base and roadbed soil were kept at constant values of 0.35, 0.4, and 0.45 respectively. Although, these are typical values of Poisson’s ratios, the actual ones may vary. For example, Baladi ( 1988) stated that Poisson’s 266 3- as.“ an- 583. a: «88m 39: gown 38$ Rag 38 :288 3 88a 3 85% «.2- 29.2 as. on- Rae we- 30%? 3H ”82m :22 85 33an 3me 98 5.38 92 5% ....Y 3.3% v.2- 353 E: o3- 8»: ”a «88mm 3 $88 as: "as 38% 85a 88 23.38 8.» RS . 3 83am 3. 53 83. $3 $3 $3 853:5 «2:82 353:5 3:602 coup—eta 2:252 v2.33. 9.3 U< Sc 38 as ___.85 32832.08 vi 83 .8. 9E 08sz 5:68 “5893 e8 «32 maotg 3 :32: “.3233on .36 058. 267 mm; cmd be; 3.6 ood- 86 mod. vacc— moum _GGNS mmd- 3.0- and- ohm- wvd- Ed- ca.~ _Nov cod 8+ 34” :3: med $6 mad- oas— nvmo Swarm—o 8.3- «m6. m3- cad- mmd- NEH no; «new Q..— ovd Ed 2“.— 8; R..— 84 $.8— 8? 2335 vwd- «~41 3.? 36- cvd- mnfi- vmfi- new». 5 o w v m m g So 350 33 20:83 “Scum “a 22.8500 5 8.5.355 654 was .85. E .05sz 5:09. EoEozE .8 2:32 528:8 Ban—88.5 35:: 23 3.38:. 5923 «cantata .36 2%.”. 268 .2833... 2.8 .8. 9E. 05sz .538... .5838 .8 “Eng :ouoocov 8:88.. 86:: 3830 9:. 3.88:. .8 car—mason. and 2:»?— E. 8:93 _mfimm 8 we 8 «N m. o m 7 whoop 3.04% m Fl «one “:0le .. ¢ v F- (smu) suouoeueo 269 .2 :38 08.. .8. 9?: 05sz .528... ...oEoZK 3.. ...—:83 558...»... 8:88. :85. 3.898 23 @2888 .8 car—EEO“. .36 RawE :5 8585 _mfimm 8 M... 8 ..u N. o ON- ...-an Cocoa-0c 32...... . oopo— v.0.- x ovao vocal—l «sow 0.04.0. . . m .... F2 (sutu) suonoeuea I0 I 270 . .3 GEN—o 0.60 .8. D35 08sz 8:08 .5838 .8 Emma 838:8 8:88. .85— 385 93 3.3.8:. .8 :88800 .36 2:3". 00 we :5 8385 _mfimm mm .... m. o 5:0 3:09.: =25- . Soup 33* vvwa 23+ h pov .30.. .0. 'o m . (sutu) suonoeuea If) I 271 ratio of the AC mix increases with increasing load level. He reported a range of laboratory measured Poisson’s ratio from 0.1 to 0.45. Likewise, in the AAMAS study, Vanquintus (1989) reported a range of the Poisson’s ratio of the AC from negative values to values of over 0.5. Bouldin, et al. (1993) reported a range of Poisson’s ratio from 0.1 to 7. Therefor, it appears that the Poisson’s ratio of the AC mix is a function of the magnitude of the applied load and the stiffness of the AC mix. In order to assess the impacts of Poisson’s ratio of the AC on the backcalculated layer moduli, two typical pavement sections (19113011 and 35114222) were analyzed by using Poisson’s ratios of 0.25 , 0.35, and 0.45 for the target load levels of 4500, 9000, and 16,000 pounds, respectively. Tables 6.24 and 6.25 provide lists of the values of the layer moduli backcalculated at the three load levels with a constant Poisson’s ratio of the AC of 0.35 , and the backcalculated values for the other two Poisson’s ratios. The percent differences between the reference moduli (calculated at the target load level of 9000 pounds using a Poisson’s ratio of 0.35) and those calculated for the other two load levels are also provided in the table. In reference to Table 6.24., it can be seen that a significant part of the apparent non-linearity at the higher load level is eliminated. For example, The percent difference between the reference modulus of the AC and that at the higher load level dropped from 9.9 percent to only 3.00 percent. Less significant improvements were found at the low load levels and in the modulus of the base layer. Since, the exact values of Poisson’s ratios are not known, part of the apparent non-linearity in the values of the backcalculated layer moduli can be attributed to variations in Poisson’s ratios. Shape Factor - Throughout this study, and because of the configuration of the steel plate of the FWD, a circular loaded area with uniform contact pressure 272 03E 9:838 382 on. a __eto «cocoon . . 83% $39 mud , 38... «mag SS 8 82;. Evan «Fawn one: 3th ENS—u mad __uoncaem 8mm U< v2.38— 88 U< cone—8M 8am 0< cue __ 89: 2.3 was. ”28%.. 3.553 :50. v8— vouaofifi 2: a _8 E :26... coma. noise—806nm m2 334 5:08 .5538 ..8 3:8 “Eamon 2.9—0&6 3 :26... ~93. ..e cognac—838m .vud 03$. 38:2 62 .8: 273 med bv. n wm.w— —N.N n: Omit Nvé—I nNd m..N- W.VN. o6 8.0 8.9 8.6 O.N_ owd- CNdT mmd . 0:8 9:838 .. . 8.8:.E 05 a: 5:0 “:85: vmwnm ngm _Nwm— m 9.6 . ~mbwm Swnv nnhmbm and 32 n mnmhn . fimawmm QSNm cwocv Ebo— m bu awn 2”an bm~wn~ nmd _ won—unom 3am U< ton—vac“ unam— U< won—KOM 03m U< __ ~32 28 an: PNWMWE 3.553 _26. 22 8.865 2: a a: 5 :39: :93. “.03—=03on ”5.59.: 5:09. “5826: 8.. mean.— a.:8m_em 380%: 3 :32: :9E ..o 5558—80—03 .36 2%.: 55:2 62 a8: 274 was assumed. In reality, the pressure distribution is a function of the relative rigidity of the steel plate and the AC layer, the macro-texture of the pavement surface, the magnitude of the applied load and the interface conditions between the plate and the pavement surface. According to Saint-Venant’s principle (Timoshenko and Goodier, 1987) the effect of changes in the pressure distribution is more significant in the zone near the loaded area while the response further away is sensitive to the total applied load. This is similar to the observations of the measured deflections data in which the deviation from the linear response due to varying load level is more pronounced for the first few deflection locations than the far ones. Hence, part of the apparent non- linear behavior may be attributable to the shape of the loaded area and the pressure distribution. System Non-Linearity - An earlier study conducted by Nazarian and Chai (1992) studied the effect of load induced non-linearity. The test results indicate that the AC modulus increases and the base and roadbed moduli decrease with increasing load levels. This observation is confirmed by the results presented earlier. The base modulus was observed to decrease with increasing load levels, which is opposite to the trends noticed for base material for the Michigan pavements that were analyzed. But one discrepancy between the method of data analysis is that for the above quoted study the AC modulus was fixed at average values after initial backcalculation. In the initial backcalculation, however, it seems that all the moduli were varied simultaneously. Backcalculation of the base layer modulus was then performed at fixed values of the AC modulus. No explanation has been offered for deviation from the standard procedure of backcalculating all the layer moduli simultaneously. The same study also concluded that the load induced non- linearity is more pronounced for KUAB (used by MDOT) than for any other 275 FWD equipment, although, the problem with vibratory devices is certainly more severe. Stress Dependency - Boker (1978) conducted study on Michigan cohesionless soils concluded that the resilient modulus of the soils tested is a function of the stress level, water content and dry density. In a similar study conducted by Goitom (1981), the above observation was confirmed for cohesive soils of Michigan as well. Chatti (1987) found that the resilient modulus of asphalt paving mixtures increase with an increase in the cyclic load. Tests conducted by Pronk (1989) with field data using the FWD showed that small changes in the roadbed modulus due to non-linear behavior can induce greater changes in the moduli of the other layers. However, he attributed confining pressure or the dead weight of the pavement as the main contributing factor towards the non-linear behavior rather than the changes in the applied loads. Hence, the observed trends can in part be attributed to the non-linearity induced by loading as well as material behavior. More comprehensive studies are certainly required to make more forceful claims. 0 Dynamic Behavior - Most existing backcalculation routines, including the MICHBACK computer program, are based on a linear layered elastic system subjected to static loading. The FWD delivers dynamic loads to the pavement structure. Hence, the magnitudes of the measured deflections at the various sensors are a function of the dynamic (not static) properties of the pavement . and loading systems. These include, that part of the mass of the pavement structure affected by the load (higher loads affect a larger mass whose properties are not constant with depth), the velocities of the compression, shear, and surface waves, the magnitude of the induced vibration in the system ' (which causes some energy losses), and the damping (viscous properties) of the various pavement layers. Hence, a part of the apparent non-linearity of the 276 backcalculated layer moduli may also be attributed to the dynamic behavior of the system. To this end, one may argue that the above enumerated factors should be included in any backcalculation routines. Although, from the academic viewpoint, this is a reasonable argument, in practice, it makes an insignificant impact on the accuracy of the backcalculated results. To illustrate, consider a pavement structure with AC, base and subbase thicknesses of 6-inch each, and a roadbed soil thickness of 60-inch (a total depth to the bedrock or to a stiff layer of 78-inch). These thicknesses are not accurately known. For example, the true thickness of the AC may vary from 5 to 6- inch at best. The depth to the stiff layer may vary by several inches or several feet. For most pavement structures, a variation of the AC thickness of l-incyh from the mean may cause significant variations in the calculated moduli of the AC, base, and subbase. These variations, in some cases, are much larger than those due to the apparent non-linearity of the system. Besides, it is not possible to include all the factors affecting pavement responses in a backcalculation routine, since the number of system unknowns cannot exceed the number of deflection sensors (seven in the case of the MDOT KUAB FWD). Therefore, it is more reasonable to use a linear elastic layered system whereby the thicknesses of some pavement layers are treated as unknowns (they possess a significant impact on the backcalculated results), than to include other factors which have lesser impact. Nevertheless, it is very important that accurate measurements of the layer thicknesses must be made before one can achieve accurate backcalculation results. Systems for making such measurements are currently under development and the most promising one is a system based on the radar technology. The results of the analysis of a variety of pavement sections and the associated discussions presented in this chapter provides enough evidence that the backcalculated 277 ‘ results of the MICHBACK program are consistent with the measured deflections and thickness variations observed in the field. The difficulty associated with the analysis of composite pavements has also been highlihgted. CHAPTER 7 TEMPERATURE EFFECTS ON THE BACKCALCULATED LAYER MODULI 7.1 GENERAL In this chapter, the effects of temperature variations on the deflections and backcalculated layer moduli for Michigan pavements are examined. As stated in Chapter 6, the pavement sections were tested during various seasons to observe the seasonal variation effects. Additional tests were designed to monitor the pavement response to temperature variations and they were performed on three pavement sections during the summer of 1993. Discussion and analyses of the test results along with recommendations are presented in this chapter. 7.2 TEMPERATURE EFFECTS As noted earlier, nondestructive deflection tests (NDTs) are typically conducted to evaluate the structural capacity and the state of deterioration of the pavement structure with time. Unfortunately, the measured pavement deflections are influenced by seasonal variations in moistureand temperature. The moisture variation influences the stiffness of the base and subbase layers and the roadbed soil. Temperature variation, on the other hand, affects the stiffness of the AC layer. Further, any change in the stiffness of any pavement layer causes changes in the responses of the other layers and of the roadbed soil. This implies that the measured deflection data are influenced by variations in both the AC temperature and the moisture level in the other layers and in the roadbed soil. In order to conduct a proper engineering evaluation of the pavement structure and to assess its rate of deterioration with time, the deflection data or the backcalculated moduli of the pavement layers must be corrected to a standard temperature and a standard moisture level. 278 279 Two temperature correction methods can be found: the American Association of State Highway and Transportation Official (AASHTO) method which calls for the correction of the measured pavement peak deflection (the deflection under the center of the load), and the Asphalt Institute (AI) method which corrects the backcalculated AC modulus to a standard temperature of 77 °F. Existing backcalculation routines, however, use the temperature correction method. Most routines use the AI temperature correction equation or some modified form of . the equation. The AASHTO method is mainly used in the deflection-based design of pavement overlays. Presently, there is no standard practice regarding seasonal corrections. Some State Highway Agencies (SHA) have collected seasonal deflection data over several years. The data were examined and deflection correction factors were deduced. In some cases the correction factor is constant for all pavements, while in others the value of the factor is a function of the pavement cross-section and the material properties. Since the weather elements (freeze-thaw cycles, and amount of rainfall) are not constant from one year to another, the development of accurate seasonal correction factors is a long time process that requires deflection data to be collected on various pavement sections over a long-term period. In this study, deflection data were collected over a two year period and three seasons. In addition, during the summer of 1993, NDTs were designed to assess the impact of the AC surface temperature on its backcalculated modulus values. Originally, eight pavement sections (6 flexible and 2 composite) were selected for this part of the study. Unfortunately, because of the weather (the summer was cooler than normal) and equipment limitation, only three flexible pavement sections were tested. For each section, the tests were commenced in the morning when the pavement temperature was around 60 °F, and continued throughout the day at half-hour time intervals. The tests were conducted at the same location which was previously marked by the MDOT FWD crew. 280 . Two of the three pavement sections (MSU13F and MSUl9F) have already been introduced in Chapter 6. The AC thicknesses for these two pavements are 4.8- and 6.2-inch, respectively. The third pavement was not included in the study plan but it was selected because of its near proximity to the MDOT testing facilities. This pavement has an AC thickness of 4.5-inch. The pavement is in a good condition with no apparent distresses and it will be referred to as MSU52F. The test results of pavement section MSU52F are discussed in more detail, it was tested under better conditions (sunny day with the temperature rising consistently) and it has the least amount of distress. 7.2.1 Flexible Pavement Section MSUSZF - Temperature Variation Figure 7.1 depicts the relationship between the measured pavement surface temperature and the peak pavement deflection (deflection under the center of the load). The figure indicates that: l. The peak pavement deflection increases almost linearly with increasing pavement temperatures. A 2. For a constant temperature, the pavement deflection measured during the heating cycle is different from that measured during the cooling cycle. Therefore, based on the second observation, the FWD test results measured during the cooling cycle of the pavement are ignored. Table 7.1 provides a list of the deflections recorded at different sensors and the corresponding pavement temperatures. The deflections measured at the lowest temperature (78 °F) were taken as reference values and the percent; variation in the deflections at other temperatures were calculated and are listed in Table 7.2, and shown in Figure 7.2. Examination of Figure 7.2 indicates that: 281 13.00 12.50" 12.00‘ 11.50- Deflections (mils) N.N-r 10.50" 10.00 75 Figure 7.1. l l I I 80 3'5 so 9515015560115150125 Pavement Temperature (F) Variation of maximum deflection with temperature observed for pavement section MSU52F. 282 Table7.1. Deflectioneveriationwlthtemperatureforpavementsection MSU52F. Werner! Measured deflections at dWererrt sensor locations (mils) ‘ . Temp. fl 1 2 3__ 4 5 __6 7 78 10.23 8.11 6.51 4.80 3.62 2.11 1.11 86 10.56 8.24 6.55 4.73 3.53 2.09 1.10 96 1 (1% 8.42 6.53 4.59 3.42 2.04 1 .09 101 1 1 .08 8.27 6.40 4.46 3.29 2.00 1 .08 109 1 1.40 8.32 5.39 4.35 3.22 1 .97 1 .09 1 14 1 1.80 8.43 6.37 4.31 3.19 1 .95 1.08 120 12.23 8.39 6.29 4.25 3.14 1.95 1.10 122 12.23 8.12 6.14 4.15 3.09 1.96 1.08 Teble72. Percertverlefionlndeflectionsfrornloweatempereturerecorded forpevemeraeectionMSUSZF. 3 Ila—II 283 .mmmDmZ 5:08 22:98: :8 2382.82 55 838%: :_ :esatg 5:088 E 920383 EmEm>mm NNP our w: mow For mm mm \. .855 S 0 89.3 _E m aomcow fl ... Comcom a m 8.8% a m .083 E — 5.8% I .2. 2:3... o..- O 2 (%) eouelemo O N 284 1. The pavement deflections measured at all sensor locations are affected by the pavement temperature. 2. The deflections of sensor 1 and 2 increase with increasing temperature, while the deflections for sensor 3, 4, 5, 6, and 7 decrease with increasing temperatures. These observations were expected, and are due to the reduction in the ability of the pavement to spread the applied load radially as the AC layer looses its stiffness at higher temperatures. The backcalculated layer moduli are listed in Table 7.3 and shown in Figures 7.3 and 7.4. It can be seen that the AC modulus decreases almost linearly with increasing temperatures. The rate of change of the AC modulus with temperature is about 10900 psi per °F. The base modulus also increases with increasing temperatures as shown in Figure 7.4. The roadbed modulus, on the other hand, is not affected by the AC temperature. The above observations indicate that the results obtained with the MICHBACK computer program are consistent and compatible with the measured deflection basins. One may argue that the modulus of the base layer (granular material) should not be affected by the AC temperature. However, there are two factors which may contribute towards the increase in base modulus: l. The state of stress in the base layer changes with changes in the AC modulus and if the base is stress sensitive, then this could cause the base modulus to change. 2. The Poisson’s ratios of the AC mix (whichwas assumed constant at all temperatures) increases as the stiffness of the AC decreases. This also affects the state of stress in the base layer (see Chapter 6). 285 Table 7.3. Backcalculated layer moduli for selected pavement sections. Pavement Pavement Backcalculated moduli section temp. (F) (psi) AC Base Roadbed 56 421429 68053 49339 57 442388 64732 47933 59 430099 64580 47642 MSU19F 80 390032 67622 46814 86 367670 69960 46359 89 339443 70747 47287 97 333389 68595 47038 100 302871 69506 46253 56 878214 54197 30674 57 871973 53457 30654 MSU13F 58 867489 53876 31080 64 857004 53156 30535 75 745520 55459 30882 87 617385 55191 30426 78 1210232 29806 30562 86 1 128475 30243 30834 - 96 1004659 31229 31246 MSU52F 101 916238 32908 31634 109 833275 34108 31653 114 807198 34108 31958 120 751909 35564 31638 122 730563 36920 31881 286 mum—.52 5:03 .5538 ..8 238888 55 2:265 U< noun—83863 Co 52953 .9» 032m A”: oaaflmaanoh E9529”. «8 tr «2 BF «8 3 mm 3 mm k _ _ a _ _ _ . a _ com 1 8m V 1 con 0 w 0 p m. .. 25 m \XJ 1 8 E .1 com; 287 (st) snlnpow .mmmDmZ 5:08 .5838 ..8 2882.83 .53 E62: eon—08.. v5 32 3.283833 he counts, .vfi Bsz A“: 929883 EmEm>mm NNP NZ. Nww NOF NOF Na Na .Nw Nm RN _ _ _ _ _. L _ _ P mN manna: nonuuom .+. «2.622 amen .... INN I O) N I ,— CO I ('3 (0 I I0 CO 18 288 Therefore an increase in the AC temperature causes a decrease in its modulus and indirectly causes the base modulus to increase. 7.2.2 Flexible Pavement Section MSU13F - Temperature Variation The NDTs were conducted on a cloudy day and under a light rainfall. The pavement deflections measured at different sensor locations are listed in Table 7.4. Once again the deflections measured at the lowest temperature (57 °F) were taken as reference values and the percent variation in the deflections at the other temperatures were calculated and are listed in Table 7.5 and shown in Figure 7.5. Examination of Figure 7.5 indicates that: 1. Sensor 1 deflection increases consistently with rising temperatures. 2. Sensor 6 consistently registered lower deflections with higher temperatures. 3. The deflections at all other sensors are erratic and do not display a consistent change with temperature. Comparison of the response of the previous pavement (MSU52F) and this pavement indicates that the way in which the two pavements spread the load is different and depends on the AC stiffness and on the temperature range. The thicknesses of the two pavement sections are'almost the same. However, MSU52F had a higher original stiffness, having been constructed to study the characteristics of Stone Mastic Asphalt (SMA) mixes. Hence, there is a difference in thebasic AC mix properties of the two pavements. The backcalculated layer moduli are presented in Table 7 .3 along with the results of the other pavement sections and the variation in backcalculated AC modulus with temperature is shown in Fig 7.6. The rate of change of the AC modulus with temperature is about 8415 psi per °F. The base modulus increased by a modest 289 Table 7.4. Deflection variation with temperature for pavement section MSU1 3F. Pavement Deflections at different sensor locations (mils) temp. (F) 1 2 3 4 5 6 7 56 8.87 6.85 5.53 4.12 3.18 2.09 1.28 57 8.92 6.93 5.58 4.14 3.19 2.09 1.29 58 8.85 6.87 5.49 4.09 3.14 2.06 1.27 64 8.92 6.95 5.54 4.13 3.18 2.08 1.28 75 8.97 6.84 5.41 3.99 3.08 2.04 1 .26 87 9.35 7.07 5.52 3.97 3.08 2.03 1 .29 Table 7.5. Percent variation in deflections from lowest temperature recorded for pavement section MSU13F. Pavement Variation at d'rflerent sensor locations (%) temp. (F) 1 2 3 4 5 6 7 57 0.56 1.12 1.03 0.57 0.31 -0.32 0.52 58 -0.30 0.34 -0.72 -0.65 -1.15 -1.59 -1.30 64 0.49 1.41 0.18 0.32 0.00 -0.64 0.00 75 1.09 -0.19 -2.17 -3.00 ~3.14 2.71 -1.56 87 5.37 3.26 -0.18 -3.56 -3.14 -2.87 0.78 290 .m2 322 5:09. 2.0838 8.. Bafionea 55 acouoococ E .5235, 2.3.8.“ Ev mSfiEQEmP EoEm>mm no mm em mm mm _ p _ _ _ .3 93E n .623 S m .893 E m 52.3 a v .023 Z m 523 a N 829."... E F 82% I 0.. (%) eoueJeulo 291 .2289. 22:26; 3.8.9. ..8 2282.82 55 3:62: U< 328—8062 .3 =ou~t~> 6.x. 8&5 Ev oSBEaEmh E0599... 55:22.22: 8 a 8 B on E 8 5 mm — _ _ _ _ _ _ _ _ _ _ _ /I//l/IL mamas. * um :52 +1 ”a Sm: + o CON oov com com coo. F com. w oov. _. (18>!) snlnpow ov 292 1.83% over the same temperature range. The roadbed modulus, on the other hand is not affected by the AC temperature. The backcalcualated results seem to be consistent with the observed deflections. 7 .2.3 Flexible Pavement Section MSU19F - Temperature Variation This section has a greater AC thickness than the previous two sections and has a lower distress than MSU13F. The range of temperature during which the FWD tests were conducted is 56 to 100 °F. The deflections recorded at different sensor locations and the test temperatures are listed in Table 7.6. Once more, the deflections measured at the lowest temperature (56 °F) were taken as reference values and the percent variation in the deflections measured at other temperatures were calculated and are listed in Table 7.7 and shown in Figure 7.7. It can be seen that: 1. The deflections of sensors 1, 2, 3, 6, and 7 increased with increasing temperatures. 2. The deflections for sensors 4 and 5 initially increased with increasing temperatures and then decreased. The backcalculated moduli are presented in Table 7.3. Variations in the AC modulus are illustrated in Figure 7.6 along with the results of the other two pavement sections. The rate of change of the AC modulus with temperature is about 2695 psi per °F. The base and roadbed moduli remain almost unaffected by the pavement temperature. Temperature, deflections, and the backcalculated moduli for the three pavement sections have a wide variation of trends and ranges. Although the data is limited, some of the trends observed can be summarized as follows: 1. The effect of the AC temperature on its modulus increases with decreasing AC Table 7.6. Deflection variation with temperature for pavement section 293 MSU19F. Pavement Deflections at different sensor locations (mils) temp. (F) 1 2 3 4 5 6 7 56 7.14 5.05 3.85 2.72 2.03 1.28 0.80 57 7.21 5.14 3.99 2.82 2.08 1.32 0.82 59 7.24 5.23 3.98 2.81 2.10 1.31 0.83 80 7.33 5.19 3.95 2.77 2.08 1.31 0.85 86 7.34 5.13 3.89 2.73 2.06 1.30 0.86 89 7.42 5.05 3.82 2.66 2.01 1.28 0.84 97 7.54 5.22 3.88 2.69 2.03 1.30 0.84 100 7.74 5.17 3.89 2.69 2.02 1.31 0.85 Table 7.7. Percent variation in deflections from lowest recorded temperature for pavement section MSU19F. Pavement Variation at different sensor locations (946) Temp. (F) 1 2 3 4 5 6 7 57 1.027 1.7162 3.8126 3.672 2.6314 3.394 1.6606 59 1.4006 3.6303 3.3795 3.0602 3.7826 2.3499 2.9054 80 2.6144 2.7723 2.6862 1.7137 2.4671 2.8715 5.8096 86 2.7545 1.5842 1.1264 0.3672 1.6446 2.0882 7.4689 89 3.9216 0.0659 -0.693 -2.203 -0.823 0.2608 4.1502 97 5.5556 3.3004 0.8665 -1.346 0 1.8274 4.1502 100 8.4034 2.3762 1.213 -1.346 -0.165 2.8715 5.395 294 .m2 DmE 5:08 .5826.— cou 238383 53 82.8%.. 5 .8239 Eocene A“: 929353 EmEo>mn .5. as»: oo P mm mm mm om mm hm ‘ N. newcom § o 82% E m 52.3 H e .523 Z m .823 a w 89% § F 583 I In (%) eoueJelllo O 1- mp 295 thickness, and decreasing pavement distress. 2. For the same AC temperature, the pavement behavior under a cooling cycle is different than that under a warming cycle. Unfortunately, the above observations cannot be stated with a good degree of confidence because of the limited data. Consequently, it is recommended that additional tests be conducted on various pavement sections. The results of these tests should enable MDOT to establish temperature and seasonal correction factors. The applicability of the Asphalt Institute (AI) temperature correction procedure for the AC modulus and the AASHTO method to impart temperature correction to the peak deflection (Do) to Michigan pavements are examined in the next subsection. 7.3 THE ASPHALT INSTTTUTE TEMPERATURE CORRECTION PROCEDURE Temperature correction to the AC modulus was performed using the AI Equation introduced in Chapter 2 and repeated here for convenience: 1 1 r,_ logEo=logE+Pm[-(Z); -W] +o.ooooos,/P:[(:,) (0'1 t " r -.m189fi[%-%1+.9317[i5;-;n1 where l = 0.17033; n = 0.02774; E = backcalculated uncorrected modulus; at, = test and reference temperatures in °F; fJL = loading and reference frequency in Hz; tale The lem lam 296 = percentage of asphalt cement in the mix; = the corrected modulus; ,0 = 1.3 + 0.49825 log (f0): , = 1.3 + 0.49825 log (t); and page = the percent aggregate Passing the ”0' 20° Sieve' For all test sites and measured temperatures, the following data were used in calculating the AI correction factors: 1. P at = 7 percent (the average percent frne content in the AC mixes), 2. f= fo = 25Hz; and 3. to = 77 °F. The calculated AI temperature correction factors are listed in Table 7.8. Table 7.8 also provides a list of the backcalculated AC modulus and the ratios of the backcalculated value at the reference temperature 77°F to that at any other temperature. Figure 7.8 depicts these calculated ratios and the computed AI correction factors plotted against the temperature. It can be seen that: l. The AC mixes encountered in the three pavement sections are less sensitive to temperature changes than predicted by the AI equation. 2. Thin pavement sections are more sensitive to the AC temperature than thick sections. The first observation could be related to the asphalt grades used in Michigan pavements (softer asphalt grades are typically used in colder regions). The second observation was expected because the average temperature of a thin pavement is higher than that for a thick pavement. That is, for a thick pavement section, more time is required before the temperature at the bottom of the AC starts to increase due 297 Table 7.8. Observed and Asphalt Institute AC modulus temperature correction factor. Pavement Pavement Backcalculated AC modulus The Al section temp (F) and ratios correction Modulus (psi) Ratio factor 56 421425 0.937 0.463 57 442388 0.893 0.503 59 430099 0.918 0.524 MSU19F 77 395000 - 1.000 1.000 80 390032 1.013 1.137 86 367670 1.074 1.443 89 339443 1.164 1.558 97 333389 1.185 2.386 100 302871 1.304 2.586 56 878214 0.816 0.532 57 871973 0.822 0.546 58 867489 0.826 0.562 64 857004 0.836 0.691 MSU13F 75 745520 0.961 0.950 77 716667 1.000 1.000 87 617385 1.161 1.406 77 1313886 1.000 1.000 78 1289605 1.019 1.028 86 1095356 1.200 1.292 96 879087 1.495 1.786 MSU52F 101 759447 1.730 2.070 109 631174 2.082 2.841 114 546395 2.405 3.430 120 443078 2.965 4.228 122 408521 3.216 4.504 Note: Ratio = Modulus @ measured temp./ Modulus @ 77 F 298 .3338 U< 8a. 8908 Swag—.8 9.3quan 23:»:— .=2%< 98 Stage Co eaten—Eco 9; Cocoa... #:6863844 9233th m v m N F _ . “_wmams. *1 mm 522 + ”6sz .... .2 2am 0 O V ('3 N 1- (pe/uesqo) 10109;; tueunsnlpv GJmBJedal ID 299 to increasing surface temperatures. 7.4 THE AASHTO TEMPERATURE CORRECTION PROCEDURE The 1990 AASHTO Guide for the design of pavement structures introduces the concept of composite pavement stiffness (Ev) to be determined from NDT data. The Direct Structural Capacity Method (explained in Appendix L of the guide) requires the value of E, to be determined along the length of the design project. The guide further recognizes the fact that the AC stiffness changes significantly with temperature. Since the peak deflections (D0) are the only information from the NDT data used to calculate E,, the guide calls for correcting the D, values only. The reference temperature suggested in the AASHTO method is 68 °F, which is consistent with the one used in part II of the guide to determine the effective structural number (SNdf) of the pavement. Further, the AASHTO Guide provides temperature correction charts. The charts were derived by using the Asphalt Institute equation along with an elastic layer analysis. Table 7.9 provides a list of the peak deflection data measured at all 3 sites at the various test temperatures and Figure 7.9 shows a plot of these. The ratios of the peak deflection at 68°F (the reference temperature) to that at any other temperatures are also listed in the table along with the corresponding AASHTO temperature correction factors. Figure 7.10 shows a comparison of these ratios and the AASHTO factors. It can be seen that: 1. The trend in the ratios of the measured deflection at the reference temperature to that at any other temperature is similar for all three pavements. 2. The three pavement sections are less sensitive to temperature variations than predicted by the AASHTO method. The above two observations are similar to those presented in the previous 11] 300 Table 7.9. Observed and AASHTO temperature correction factors for Do. Pavement Pavement Measure deflections and The AASHTO section temp. (F) ratios correction Deflections (mils) Ratios factor 56 7.14 1 .020 1 .094 57 7.21 1.010 1.099 59 7.24 1 .006 1 .074 68 7.28 1 .000 1 .000 80 7.33 0.993 0.906 MSU19F 86 7.34 0.992 0.857 89 7.42 0.981 0.846 97 7.54 0.966 0.798 100 7.74 0.941 0.797 56 8.87 1 .008 1 .081 57 8.92 1 .002 1 .078 58 8.85 1 .010 1 .059 MSU13F 64 8.92 1 .002 1 .028 68 8.94 1 .000 1 .000 75 8.97 I 0.997 0.933 87 9.35 0.956 0.962 68 9.82 1 .000 1 .000 78 10.23 0.960 0.860 86 10.56 ' 0.930 0.813 96 10.96 0.896 0.805 MSU52F 101 1 1 .08 0.886 0.784 109 1 1 .4 0.861 0.787 114 11.8 0.832 0.762 120 12.23 0.803 0.541 122 12.23 0.803 0.541 Note: Ratio = Deflections @ measured temp. I Deflections @ 68 F _— 14F 301 35.88 22:93 3.8.8 8.. 9382.83 5.3 26:85.3 3.89. .«o 5.593 d... E. 9393.99. EoEm>md 9:3". mm.. m... mow mm mm mm. mm mm _ fl . . . _ v mmmaws. .x. ”.mPDwS. 1T “5sz .... .. o W W... «\a\IIIo\I 1w w m +\T +III+ W. l 0.. m. U s M. l Nw kw 1.3rr 302 38.80%: x8: 8.. 8208 5.93:8 9392.88 Ohmm<< cc: 328:: we saunas—cu .o. .5. 9:9". m... 59.192. 5Com“. 3:02.633. 9399.th N; E F mo :8 3 8.: no. . _ . _ . . 4 mo m lodm m n 1.3m V . .m. ximom w [mom a I? m m ”amaze: 3% ”55229 «cm. 9 ”5592... m ‘9. ,_ section. 1 to that of is confirr 7.5 S' T correctic analysis. 1. I l 303 , section. This suggests that the trend of the deflection data with temperature is similar to that of the backcalculated AC modulus. Hence the accuracy of the backcalculation is confirmed once again. 7.5 SUMMARY AND RECOMMENDATIONS The data available to examine the validity of two suggested temperature correction methods was limited to three pavement sections only. Based on the analysis, the following conclusions are made: 1. The pavement behavior during the warming cycle is different than that during the cooling cycle. Hence, different temperature correction methods should be devised for the two cycles. 2. The three pavement sections included in this temperature correction study showed less sensitivity to temperature than that predicted by the AI and the AASHTO methods. 3. It appears that the temperature correction factors should include other pavement characteristics (such as AC thickness and stiffness, and asphalt grade) that are not part of the AI and the AASHTO methods. 4. The NDTs should be expanded to include more pavement sites than those included in this limited study. 8.1 mi thal infr stn the acti bal prc PFC prc the the CHAPTER 8 SUMIWARY, ACCOMPLISHNIENTS, CONCLUSIONS AND RECOMIVIENDATIONS 8.1 SUMMARY The continuous investment required to design, rebuild, rehabilitate, and maintain the nation’s transportation infrastructure presents a major and complex task that challenges the ingenuity of every highway administrator and engineer. The infrastructure problems cannot be solved by simply studying pavement conditions and taking corrective actions. Accurate and comprehensive evaluation of the pavement structural capacity and their rates of deterioration allow the decision makers to direct the investment where nwded and to properly select and schedule rehabilitation activities. The accurate evaluation of the pavement structural capacity requires a balanced, accurate, robust, comprehensive, and mechanistically-based computer program for the analysis of the nondestructive deflection test (NDT) data. The problem associated with such analyses is that existing state-of-the-art cOmputer programs are often not accurate and, for most programs, the accuracy of their solutions is dependent on the initial estimates specified by the user. The inaccuracy of the solution is mainly related to several important problems that affect the analyses of the NDT data. These problems include: 1. For most pavement structures, the deflection of the roadbed soil due to an applied load represents the bulk of the measured deflection data. Hence, an accurate estimate of the roadbed stiffness at the onset of the analysis can significantly increase the accuracy of the estimates of the stiffnesses of the other pavement layers. 304 305 The thickness of the roadbed 305 (or the depth to a stiff layer) is typically not known. Erroneous estimates of this depth causes substantial errors in the solutions. Thus, an accurate estimate of the stiff layer depth based on the mechanical behavior of the pavement system should be obtained as a part of the overall analyses. The pavement cross-section varies from one point to another. Variations in the thickness of the asphalt concrete layer causes significant errors in the solution. Consequently, a mechanistic routine whereby this thickness can be corrected should be a part of the analyses of the NDT data. These and other problems were extensively examined during the course of this study and solutions were obtained and implemented in a computer program named MICHBACK. The accomplishments of this study are summarized in the next section. 8.2 ACCOMPLISHMENTS Several major accomplishments have been achieved in this research study and they are implemented in the MICHBACK computer program. These include: 1. A new algorithm to accurately predict the roadbed modulus at the onset of the analysis (after only one call to a mechanistic analysis program) has been developed, tested, and implemented. An algorithm using the modified Newton method. to backcalculate layer moduli and layer thicknesses has been developed, tested, and implemented. The Newton’s method has been successfully extended to mechanistically calculate the stiff layer depth from deflection data. Hence, one of the major shortcomings of existing backcalculation programs has been eliminated. The sensitivity of the backcalculated layer moduli of flexible pavements to the user’s initial estimates (seed moduli) has been minimized. User friendly-features have been developed and implemented in MICHBACK. 8.3 306 These features are designed not only to facilitate the use of the program, but also to encourage user interaction in the backcalculation process which is felt to be essential for obtaining meaningful results. CONCLUSIONS Based on the analysis of the theoretical and field measured deflection data, on an exhaustive testing of the MICHBACK program, and on a comprehensive comparison of the results of MICHBACK with those of other programs, the following conclusions were drawn: 1. The new algorithm that has been developed to predict the stiffness of the roadbed soil at the onset of the analysis by using only one call of the mechanistic analysis produces very accurate results. The algorithm using the Newton’s method that has been developed to mechanistically predict the stiff layer depth is accurate. The MICHBACK computer program is capable of correcting an erroneous estimate of any one layer thickness and of accurately estimating the moduli of the pavement layers. Poisson’s ratios of the different pavement materials affect the accuracy of the backcalculated results. Hence, for each material, a range of Poisson’s ratio must be known and used with caution to achieve good results. The effect of inaccuracies in the deflections measured by those sensors located close to the loaded area on the backcalculated results is higher than those measured by the other sensors. The original CHEVRON and ELSYM5 computer programs produce deflection basins that can be significantly different than those produced by the respected BISAR program (which is considered a very accurate computer program). This discrepancy increases with an increase in the stiffness of the pavement. 10. 11. 307 Although only limited deflection and temperature data were available in this study, preliminary results indicate that the Asphalt Institute (AI) and the AASHTO temperature correction methods have very limited applicability to the measured deflection basins. Further, results of the limited analysis of the NDT data obtained from those flexible pavements included in the study showed that the asphalt concrete layer may be less sensitive to temperature variations than those advocated by the AI and the AASHTO methods. The measured deflection basins and the backcalculated results appear to be affected by the magnitude of the applied load. However, the exact cause of this effect could not be accurately determined or generalized for all pavement sections. Possible causes of system nonlinearities could be related to load magnitude, Poisson’s ratios, and dynamic effects. For any one pavement section, the variation of the backcalculated results obtained with MICHBACK from one station to another is compatible with the variation in the measured deflection basins. The analysis of composite pavements is a difficult task compounded by the unknown state of distress of the PCC slabs. Comparison of the backcalculated results obtained using MICHBACK with those obtained from other programs indicates: a) For pavements where a stiff layer is encountered, the results from MICHBACK are significantly better than those obtained with the other programs. b) For flexible pavements with a very deep stiff layer, the results from MICHBACK are similar to those obtained with the other programs. The increased accuracy of the results from MICHBACK becomes noticeable as the number of layers increase. c), For composite pavements, the results obtained with MICHBACK are 8.4 I I the follc l. 8.4 308 significantly better than those of the other programs. d) The performance of MICHBACK, measured in terms of the number of calls made to a mechanistic program, is slightly better than that of the EVERCALC 3.0 (which is known for rapid convergence). RECOMMENDATIONS FOR FUTURE RESEARCH Based on the results, accomplishments, and conclusions of this research study, the following recommendations are made: 1. The effect of Poisson’s ratios on the backcalculated results should be examined in more detail. The feasibility of estimating some of the values of Poisson’s ratios from the deflection data should be explored. The effect of the asphalt concrete temperature on its modulus and on the measured deflection data should be examined in more detail to establish more dependable temperature correction functions. The effects of seasonal variations on the backcalculated results and on the measured deflection basins should be undertaken as a part of a long term pavement performance (LTPP) study. The MICHBACK computer program uses a linear elastic multilayer analysis program (CHEVRONX). The extent and causes of the apparent nonlinearity observed in the measured deflection data should be further investigated. The user-friendly features of MICHBACK are similar to those of MICHPAVE (a linear and nonlinear finite element elastic layer program for the analysis and design of flexible pavements). The two programs enable the MDOT and other users to perform forward and backward analyses. However, the MDOT capability for the design of an overlay is very much limited to experience, a standard overlay thickness, and existing empirical procedures. It is strongly recommended mechanistic-based overlay design program be developed as an 309 integral part of MICHBACK. This development will have a major contribution to the capability of MDOT to perform mechanistic analysis, design, and evaluation of their pavement structures. List of References American Association of State Highway Officials, 'AASHO Interim Guide for Design of Pavement Structures," Washington, D. C., 1972. American Association of State Highwy and Transportation Officials, ”Standard Recommended Practice for Pavement Deflection Measurements, " '1‘256—77 Standard Specifications for Transportation Materials and Methods of Sampling and Testing, Part II, 1982. Acum, W. E. A., and Fox, L., “Computation of Load Stresses in a Three-Layer Elastic System,” Geotechnique, Vol 2, pp.293-300, 1951. Anderson, M. , ”Data Base Method for Backealculation of Pavement Layer Moduli, " Nondestructive Testing of Pavements and Backcalculation of Moduli, ASTM STP 1026, A. J. Bush 111, and G. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 201-216. Asphalt Institute, ”Research and Development of the Asphalt Institute’s Thickness Design Manual (MS-1) Ninth Edition," Research Report No. 82-2, The Asphalt Institute, August 1982. Austin Research Engineering, Inc. , 'Asphalt Concrete Overlays of Flexibile Pavements Vol. 1, Development of New Design Criteria, " FHWA Report No. FHWA-RD-75—75, August 1975. Baladi, G. Y. , ”Invariant Properties of Flexible Highway Pavements,” Ph.D Dissertation, Department of Civil Engineering, Purdue University, December 1976. Baladi, G. Y. , ”In-Service Performance of Flexible Highway Pavements, " International Air Transportation Conference, Vol. I, 1979. Baladi, G. Y. , ”Integrated Material and Structural Design Method for Flexible Pavements, " Vol. 1: Technical Report, Publication No. FHWA-RD-88-109, December 1988. Baladi, G. Y., 'Statistieal Model for Predicting Stiff Layer Depth and Layer Moduli for Different Paving Layers, " Unpublished Data, Civil and Environmental 310 311 Engineering, Michigan State University, East Lansing, 1993. Bernhard R. K., "Highway Investigation by Means of Induced Vibrations," Pennsylvania State Engineering Experiment Station Bulletin, No. 49, 1939. Biarez, J. , "Contriabutation a 1 ’Etude des Properietes Macaaniquuues des Sols et des Materiau Pulverulents," D.SC Thesis, University of Grenoble, 1962. "BISAR User’s Manual: Layered System Under Normal and Tangential Loads," Shell-Koninil-ijke/Shell Laboratorium, Amsterdam, The Netherlands, 1972. Bohn, A., Ullidtz P., Stubstad R., and Sorensen A., "Danish Experiment with the French Falling Weight Deflectometer, " Mugs, Third International Conference on the Structural Design of Asphalt Pavements, University Of Michigan, Ann Arbor, Vol 1, September 1972, pp. 1119-1128. Boker, T. N. , "Resilient Characteristics of Michigan Cohesionless Roadbed Soils in Correlation to the Soil Support Values," Ph.D. Dissertation, Department of Civil and Environmental Engineering, Michigan State University, East Lansing, Michigan, 1978. Bouldin, M. G., Rowe, G. M., Sousa, J. B., and Sharrock, M. J., "Mix Rheology - A Tool for Predicting the High Temperature Performance of Hot Mix Asphalt" Paper submitted for publication to W, June 1993. Brown, J. L. , "The Mechanistic Analysis of Pavement Deflections on Subgrades Varying in Stiffness with Depth," Draft Report No. 1159, Texas Transportation Institute, College Station, 1991. Brown, J. L., and Orellara, H. E., "Utilizing Deflection Measurements to Upgrade Pavement Structures," Report No. 191-IF, Texas State Department of Highways and Public Transportation, Austin, December, 1970. Brown, S.F., and Pell, P. S., "Subgrade Stresses and Deformations Under Dynamic Load," Journal of the Soil Mechanics and Foundation Division, ASCE, January 1967. Bubusait, A.A, Newcomb, D. E., and Mahoney, J. P., "Development and Implementation of Overlay Design Procedure, Interim Report No. 2: Asphalt Concrete Stiffness-Temperature Relationship and Pavement Distress Modeling, " Report No. WA—RD 66.2, Washington State Department of Transportation, Olympia, 1974. Burmister, D. M., "The Theory of Stresses and Displacements in Layered Systems and Applocation to the Design of Airport Runways," Mugs, Highway Research Board, Vol 23, 1943. 312 Board, Vol 23, 1943. Burmister, D. M., "Evaluation of Pavement Systems of WASHO Road Test by Layered Systems Method," Highway Research Board Bulletin 177, 1958. Bush, A. J. III, "Nondestructive Testing for Light Aircraft Pavements: Phase II, Development of Nondestructive Evaluation Methodology, " Report No. FAA-RD-80-9- II, Department of Transportation, Federal Aviation Administration, Washington, D.C., November, 1980. California Division of Highways, "Methods of Test to Determine Overlay Requirements by Pavement Deflection Measurements, " Test Method California 356, October 1973. Chatti, K. , "Characteristics of Asphalt Paving Mixtures Under Static and Cyclic Loading," MS, Thesis, Department of Civil and Envioranmental Engineering , Michigan State University, East Lansing, Michigan, 1978. Chen. W. F-. and Baladi. 6- Y..Ssfl.21astisibt_lhm_and_1mrtlsmsntati9n. Elsevier Science Publication Company Inc. , New York, 1985. Chou, Y. , "Development of an Expert System for Nondestructive Pavement Structural Evaluation, " PhD. Dissertation, Texas A&M University, College Station, Texas, 1989. Chua, K. M., and Lytton, R. L., "Load Rating of Light Pavement Structures," Transportation Research Record No 1043, Transportation Research Board, National Research Council, Washington, D.C., 1984. Chua, K. M., "Evaluation of Moduli Backcalculation Programs for Low—Volume Roads," NH".1!I a9 ‘ ' -_ . - -. STP 1026, A. J. Bush 111, and G. Y. Baladi, Eds., Ameriean Society for Testing and Materials, Philadelphia, 1989. Claessen, H. I. M., Valkeming, C. P., and Ditsmarsch R., "Pavement Evaluation with the Falling Weight Deflectometer, " Wings, Association of Asphalt Paving Technologists, Minneapolis, Minnesota, 1976. Classen, H. I. M., and Ditrnarsch R., "Pavement Evaluation and Overlay Design - The Shell Method," Wings, Fourth International Conference on the Structural Design of Asphalt Pavements, University of Michigan, Ann Arbor, Michigan, Vol 1, 1977. Crovetti, J. A., Shahin, M. Y., and Touma, B. 13., "Comparison of Two Falling Weight Deflectometer Devices, Dynatest 8000 and KUAB 2M-FWD, " Wye 313 mummmtmqmmn ASTM STP 1026. A. 1 Bush III, and G. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 59-69. Davies, G., T., and Mamlouk, S., M., "Theoretieal Response of Multilayer Pavement System to Dynamic Nondestructive Testing, " Transportation Research Record No. 1022, Transportation Research Board, National Research Council, Washington, DC. , 1985, pp. 29-36. Dehken, L. G. , "The Effect of Non-Linear Material Response on the Behavior of Pavements Subjected to Traffic Loads," PhD Dissertation, Department of Civil Engineering, University of California, Barkely, 1978. DenniSJ E...Jr andSchanabcl RB Numensalflethadafqulnmmim WNW Prentice-Hall Inc” Englewood Cliffs, New Jersey, 1983. Epps, J. A. , and Monismith, C. L., "Equipment for Obtaining Pavement Condition and Traffic Loading Data, " National Cooperative Highway Research Program (NCHRP) Report No. 126, Transportation Research Board, National Reasearch Council, Washington, D.C., September, 1986. En'ngcn. A” C., and Sahubi. 13.. 8.. WW, Vol. 2. Academic Press, New York, 1975. Fossberg. P. E. , "Some Deformation Characteristics of a Lime Stabilized Clay, " Highway Research Board, National Research Council, Washington, D.C., 1969. George, L. D. , "The Effect of Non-linear Material Response on The Behavior of Pavement Subjected to Traffic Loads," Ph.D. Thesis, Department of Civil Engineering, University of California, Barkeley, March 1969. Germann, F. P., and Lytton, R. L., "Temperature, Frequency and Load Level Correction Factors for Backcalculated Moduli Values," W W ASTMSTP 1026 A 1 391811111 andG Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 431-451. Goitom, T. , "Chracteristics of Michigan Cohesive Subgrade Soils Under Cyclic Loading, " Ph.D. Dissertation, Department of Civil and Enviommental Engineering , Michigan State University, East Lansing, Michigan, 1981. Green, J. L., and Hall, J. W., "Nondestructive Vibratory Testing of Airport Pavements, " National Technical Information Service Report No. FAA-RD-73-205-1, August 1974. 314 Guozheng, Y., "The Radius of Curvature and the Fatigue Design of Bituminous Pavement," W. Bearing Capacity of Roads and Airfields, The Norwegian Institute of Technology, Trondheim, Norway, 1982. Hansen H., and Chenea P., W, John Willey and Sons, Inc., New York, 1956. Hardin, B. O., and Black, W. L., "Sand Stiffness Under Various Triaxial Stresses," Journal of the Soil Mechanics and Foundation Division, ASCE, March 1966. Harichandran, R. S., Mahmood, T., Raab, A. R., and Baladi, G. Y., "A Modified Newton Algorithm for Backcalculation of Pavement Layer Properties, " Transportation Research Record No. 1384, Transportation Research Board, National Research Council, Washington, D.C., 1993. Harichandran, R. S., Mahmood, T., Raab, A. R., and Baladi, G. Y., "Backcalculation of Pavement layer Moduli, Thickness and Stiff Layer Depth Using a Modified Newton Method," Paper presented at the 2nd International Symposium on Nondestructive Testing of Pavements and Backcalculation of Moduli, Atlanta, Georgia, June 23-24, 1993. Haveem, F. N. , "Pavement Deflections and Fatigue Failures," Bulletin 114, Highway Research Board, Washington, D. C., 1938. Heukelom, W., and Foster, C. R., "Dynamic Testing of Pavements," Wings, American Society of Civil Engineers, Vol. 86, No. SMI, pp. 1-28, February, 1960. Heukelom, W. , "Analysis of Dynamic Deflection of Soils and Pavements," Geotechnique, Vol. 11, No. 3, pp. 224-243, September, 1961. Heukelom, W., and Klop, A. J. 6., "Dynamic Testing as a Means of Controlling Pavements During and After Construction, " International Conference on the Structural Design of Asphalt Pavements Proceedings, University of Michigan, Ann Arbor, August, 1962. Heukelom, W. , "A Bitumen Test Data Chart for Showing the Effect of Temperature on the Mechanical Behavior of Asphaltic Bitumens, " Journal of the Institue of Petroleum, November 1969. Higdon, A., Ohlsen, E. H., Stiles, W. B., and Weese, J. A., Meehaniguf Materials, Second Edition, John Wiley and Sons, Inc. New York, 1967. Highway Research Board, "The AASHO Road Test, Report 5, Pavement Research," Highway Research Board Special Report 61B, 1962. 315 Hiltunen, D. R., and Woods, R. D., "Influence of Source and Receiver Geometry on the Testing of Pavements by the Surface Wave Method," W W. ASTMSTP 1026 A J 3918th andG Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 138-154. Hoffman, M. S. , and Thompson, M. R., "Nondestructive Testing of Flexible Pavements Field Testing Program Summary," Civil Engineering Studies, Transportation Engineering Series No. 31, Dept. of Civil Engineering, University of Illinois at Urbana-Champaign, June 1981. Hoffman, M. S., Mario, S., and Thompson, M. R., "Backcalculating Nonlinear Resilient Moduli from Deflection Data," Transportation Research Record No. 852, Transportation Research Board, National Research Council, Washington, D.C., 1982. Hoffman, M. N ., and Thompson, M. R., "Comparative Study of Selected Nondestructive Testing Devices " Transportation Research Record 852, Transportation Researh Board, National Research Council, Washington, D.C., 1982. Hogg, A. H. A., "Equilibrium of a Thin Plate Symmetrically Loaded Resting on an Elastic Foundation of Infinite Depth," Philosophic Magzine, Vol. 25, No. 168, 1938. Hogg, A. H. A., "Equilibrium of a Thin Slab on an Elastic Foundation of Finite Depth," Philosophic Magzine, Vol. 35, No. 243, 1944. Hossain, M., and Scofield, L. A., "Correlation Between Backcalculated and Laboratory-Determined Asphalt Concrete Moduli, " Transportation Research Record No. 1377, Transportation Research Board, National Research Council, Washington, D.C., 1992, pp. 67-76. Hoyinck, W. T., Van Der Ban, R., and Gerritesn, W., "La Croix Overlay Design by Three-Layered Analysis, " Wings, Fifth International Conference on the Structural Design of Asphalt Pavements, Delft, The Netherlands, Vol 1, 1982, pp. 410-420. Hung, J. T., Briaud, J. L., and Lytton, R. L., "Layer Equivalency Factors and Deformation Characteristics of Flexible Pavements, " Research Report 284-3, Texas Transportation Institute, Texas A&M University, College Station, Texas, January, 1982. "ILLIPAVE User’s Manual, " Transportation Facilities Group, Department of Civil Engineering, University of Illinois at Urbana—Champaign, May, 1982. Irwin, L. H., "User’s Guide to MODCOMP 2," Report No. 83-8, Cornell University Loeal Roads Program, Cornell Univeristy, Ithaea, NY, November 1983. 316 Irwin, L. H., Yang, W. S., and Stubstad, R. N., "Deflection Reading Accuracy and Layer Thickness Accuracy in Backcalculation of Pavement Layer Moduli," u919‘9..9°'rv ' 9'f vmn 99:...9 _9'M.91999_ ,ASTMSTP 1026, A. J. Bush III, and G. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 229-244. Izada Nelson M., "Detecting Variations in Load-Carrying Capacity of Flexible Pavemenmts," NCHRP Report No. 21, 1966. Jones, A., "Tables of Stresses in ThreeLayer Elastic Systems," Highway Research Board Bulletin 342,1962. Jones R. , "Measurement and Interpretation of Surface Vibrations on Soil and Roads, " Highway Research Board Bulletin No. 277, 1960. Jones, R. , "Measurement of the Dynamic Elastic Properties of the Sand Subbase of the Experimental Road at Alconbury Hill, " Road Research Laboratory, Great Britan, Research Note RN/3831/RJ, August 1960. Kallas, B. F., and Riley, J. C., "Mechanical Properties of Asphalt Pavement Materials, " Mngs, Second International Conference on the Structural Design of Asphalt Pavements, University of Michigan, Ann Arbor, Michigan, Vol ‘1, 1977. Kausel, E., and Peek, R., "Dynamic Loads in the Interior of a layered Stratum: An Explicit Solution," Bulletin of the Seismologieal Society of America, Vol. 72, No. 5, Octobor, 1982. Kennedy, C. K. , "Pavement Deflection: Operating Procedures for Use in The United Kingdom, " TRRL Report 835, Transportation and Road Research Laboratory, United Kingdom, 1978. Keneddy, C. K., Ferve, P., and Clarks, C. S., "Pavement Deflection: Equipment for Measurement in the United Kingdom," TRRL Report 834, Transportation and Research Laboratory, United Kingdom, 1978. Kinchen, R. W., and Temple, W. H., "Asphalt Concrete Overlays of Rigid and Flexibile Pavements, " Report No. FHWA-LA-109, Federal Highway Adrrrinistration, Washington, D.C., September 1977. Kinchen, R. W.,and Temple, W. H., "Asphalt Concrete Overlays of Rigid and Flexible Pavements," Report No. FHWA/LA-80/l47, Louisiania Department of Transportation and Development, October, 1980. Krokosky, E. M., Tons, E., and Andrewa, R. D., "Rheological Properties of Asphalt-Aggregate Compositions," 12mm. Amelia!!! Society for Tooting and .lu-erI—d “Ant—i (3'11‘11r‘ >—-~—rmt-‘ “mt-:1:- 317 Materials, 1963. Kuo, S. S. , "Development of Base Layer Thickness Equivalency, " Research Report No. R-1119, Research Laboratory Section, Testing and Research Division, Michigan Department of Transportation, Michigan, Lansing, June 1979. Lister, N. W. , "Deflection Criteria for Flexible Pavements Under Moving Wheel Loads, " W, Second International Conference on the Structural Design of Asphalt Pavements, University of Michigan, Ann Arbor, 1967. Lorenz, H. , "Elasticity and Damping Effects of Oscillating Bodies on Soil," Symposium on Dynamic Testing of Soils, American Society for Testing Materials, Special Technical Publication No. 156, pp. 113-122, July 1953. Lytton R. L., Moore, W. M., and Mahoney, J. P., "Pavement Evaluation: Phase 1 Pavement Evaluation Equipment, " Report No. FHWA-RD-75-78, Federal Highway Administration, March 1975. Lytton, R. L. , and Michalak, C. H. , "Flexible Pavement Deflection Equation Using Elastic Moduli and Field Measurements, " Research Report 207—7F, Texas Transportation Institute, Texas A&M University, College Station, Texas, September, 1979. Lytton, R. L., Roberts, F. L., and Stoffels, S., "Determination of Asphaltic Concrete Pavement Structural Properties, " National Cooperative Highway Research Program Project No. 10-27, Transportation Research Board, National Reasearch Council, Washington, D.C., April, 1986. Lytton, R. L, "Backcalculation of Pavement layer Properties, "W Wampum ASTM STP 1026 A J BushIII. and G. Y. Baladi, Eds., American Society for Testing and Material, Philadelphia, 1989, pp. 1-38. Lytton, R. L., Germann, F., and Chou, Y. 1., "Determination of Asphalt Concrete Pavement Structural Properties by Nondestructive Testing, " Phase I and II, Final Report for NCHRP, Transportation Research Board, National Research Council, Washington, DC. February 1990. Mahmood, T. , "Nondestructive Deflection Testing Data for Michigan and Summary of Michback Routines, " Report No. MSU-ENGR—008-93, Department of Civil and Environmental Engineering, Michigan State University, East Lansing, November 1993. Mahoney, J. P., Winters, B. C., Jackcon, C., and Pierce, L. M., "Some Observations About Backcalculation and Use of a Stiff layer Condition, " R ID P h Cu Cu ‘5 318 Transportation Research Record No 1384, Transportation Research Board, National Research Council, Washington, D.C., 1993. Mahoney, J. P, Newcomb, D. E., Jackson, N. C. and Pierce, M. L, "Pavement Moduli Backcalculation, " Prepared for Highway Pavements Short Training Course, Participant’s Notes, Michigan State University, East Lansing, September 1991. Mamlouk, M. S. , "Dynamic Analysis of Multilayered Pavement Structures-Theory, Significance and Verification, " Wings, Sixth lntemational Conference on the Structural Design of Asphalt Pavements, University Of Michigan, Ann Arbor, Vol 1, September 1987, pp. 446-474. May, R. W., "Interpreting Dynamic Surface Deflections in the Granular Base Layer of the Pavement Structure," Federal Highway Admirristeraion, 1981. McComb, R. A., and Lavra, J .J ., "A Review of Structural Evaluation and Overlay Design for Highway Pavements," Pavement Rehabilitation: Proceedings of a Workshop, Report No. FHWA-RD-74-60, Federal Highway Administration, June, 1974. McCullough, B., F., and Taute, A., "Use of Deflection Measurements for Determining Pavement Material Properties, " Transportation Research Record No. 852, Transportation Research Board, National Research Council, Washington, D.C., 1982, pp. 8-14. McLeod, N.W. , "Reduction in Transverse Pavement Cracking by use of Softer Asphalt Cements," Special Report No. 101, Highway Research Board, Washington, D.C., 1969. Michelow, J ., "Analysis of Stresses and Displacements in an N—layered Elastic System Under a Load Uniforrnally Distributed in a Circular Area," California Research Corp. , Richmond, 1963. Miller, G.F., and Pursey, H., "On the Partition of Energy Between Elastic Waves in a Semi-Infinite Solid," Mg: of the Royal Society, Series A, vol. 233, 1955. Ming-Shan, Y., "Nonlinear Finite Element Analysis and Design of Flexibile Pavements, " Ph.D. Dissertation, Department of Civil and Enviommental Engineering, Michigan State University, East Lansing, Michigan, 1989. Mitchel, J. K., and Shen, C. K., "Soil-Cement Properties Determined by Repeated Loading in Relation to Bases for Flexible Pavements, " meanings, Second lntemational Conference on the Structural Design of Asphalt Pavements, University of Michigan, Ann Arbor, Michigan, Vol 1, 1977. 319 Mitray, F. G. , "Determination of the Modulus of Resilient Deformation of Untreated Base Course Materials," Ph. D. Thesis, University of California, Berkeley, 1964. Molenaar, A. A. A., and Van Gurp, C. A. P. M., "Structural Performance Model and Overlay Design for Asphalt Concrete Pavements, " Transportation Research Record No. 888, Highway Research Board, National Research Council, Washington, DC. , 1982. Monismith, C. L., Alexander, R. L., and Secor, K. E., " Rheological Behavior of Asphalt Concrete," Wings, Association of Asphalt Paving Technologists, Vol. 35, 1966. Moore, W. M., Chas R., Hanson, D. I, and Hall, W. J. Jr., "An Introduction to Nondestructive Structural Evaluation of Pavements, " Transportation Research Circular No 189, Transportation Research Board, National Research Council, Washington, D.C., January, 1978. Morgan, J. R., "The Response of Granular Materials to Repeated Loading," Wings, Third Conference Australian Road Research Board, part II, 1966. Nazarian, S., and Chai, Y. B., "Comparison of Theoretical and in Situ Behavior of a Flexible Pavement Section," Transportation Research Record No. 1377, Transportation Research Board, National Research Council, Washington, D.C., 1992, pp. 172-182. Nazrian, 8., Stoke 11, K. H., and Hudson, W. R., "Use of Spectral Analysis of Surface Waves Method for Determination of Moduli and Thicknesses of Pavement Systems," Transportation Research Record No. 990, Transportation Research Board, National Research Council, Washington, D.C., 1983, pp. 38-45. Nazrian, S., and Stoke 11, K. H. "Nondestructive Testing of Pavements Using Surface Waves, " Transportation Research Record No. 993, Transportation Research Board, National Research Council, Washington, D.C., 1984, pp. 67-79. Nazarian S. , and Stoke 11, K. H. , "Nondestructive Evaluation of Pavements by Surface Wave Method, " Nondestructive Testing of Pavements and Backcalculation of Moduli, ASTM STP 1026, American Society for Testing and Materials, Philadelphia, 1989, pp. 119-137. Odemark, N. , "Investigations as to the Elastic Properties of Soils Design of Pavements According to the Theory of Elasticity, " Staten Vaeginstitute, Stockholm, Sweden, 1949. Paquet, J. , "The CEBPT Curvimeter, A New Instrument for Measuring Highway Pavement Deflections," C.E.B.P.T., Paris, 1978. 320 "Pavement Rehabilitation: Proceedings of a Workshop, " Report No. FHWA-RD-74- 60, Transportation Research Board, Federal Highway Administration, Washington, D.C., June 1974. Peattie, K. R. , "Stress and Strain Factors for Three-layer Elastic Systems," Highway Research Board Bulletin 342, 1962. Pell, P. S., and Brown, S. F., "The Characteristics of Materials for the Design of Flexible Pavement Structure," Wings, Third lntemational Conference on the Structural Design of Asphalt Pavements, London, Vol. I, 1972. Pronk, A. C., "Stress Dependency of the (Sand) Subgrade Modulus by the Dead Weight of the Pavement Layers: Field Experiment with the Falling Weight Deflectometer, " Nondestructive Testing of Pavements and Backcalculation of Moduli, ASTM STP 1026, A. J. Bush III, and G. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. $40-$55. Pitchumani, R. , "Applications of Computer Codes to the Analysis of Flexible Pavements, " Wings, Third lntemational Conference on the Structural Design of Asphalt Pavements, University Of Michigan, Ann Arbor, Vol 1, September 1972, pp. 506-520. Rada, G. R., C. A. Richter, and P. J. Stephansons, "layer Moduli from Deflection Measurements: Sofiware Selection and Development of Strategic Highway Research Program’s Procedure for Flexible Pavements," Transportation Research Record No. 1377, Transportation Research Board, National Research Council, Washington, D.C., 1992, pp. 77-87. Richard, R. M., and Abbott, B.J., "Versatile Elastic-Plastic Stress Strain Formula," Technical Note, Journal of the Soil Mechanics and Foundation Division, ASCE, Vol. 101, No. EM4, August 1975. Richart F. 13.. Woods R. D., and Hall J. R., YihrationuLSoiluniEmmdations Prentice-Hall, Englewood Cliffs, N. J ., 1970. Robert. A. D. , and George, L. E. , "Nondestructive Pavement Testing by Wave Propagation: Advanced Method of Analysis and Parameter Management, " Transportation Research Record No. 1070, Transportation Research Board, National Research Council, Washington, D.C., 1986, pp. 53-68. Roberts, D. V. , "Evaluation of the Cox Deflection Devices, " FHWA-CA-TL-3150- 77-14, California Department of Transportation, Sacramento, California, 1977. Roesset, M., J ., and Ko-Young, S., "Dynamic Interpretation of Dynaflect and Falling Weight Deflectometer Tests, " Transportation Research Record No. 1022, 321 Transportation Research Board, National Research Council, Washington, D.C., 1985, pp. 7-16. Rohde, G. T., and Scullion, T., "MODULUS 4.0: Expansion and Validation of the MODULUS Backcalculation System," Research Report 1123-3, Texas Transportation Institute, Texas A&M University, College Station, 1990. Rwebangira, T., Hicks, R. G., and Truebe, M., "Sensitivity Analysis of Selected Backcalculation Procedures," Presented at the 1987 TRB Meeting. Scrivner, F. H., Poehl, R., Moore, W. M., and Phillips, M.B., "Detecting Seasonal Changes in Load-Carrying Capabilities of Flexibile Pavements, " National Cooperative Highway Research Program Report No. 76, 1969. Scullion, T., Uzan, J ., and Paredes, M., "MODULUS: A Microcomputer-Based Backcalculation System," Transportation Research Record No. 1260, Transportation Research Board, National Research Council, Washington, D.C., 1990, pp. 180-191. Sebaaly, B., B., Mamlouk, S., M., and Davies, 6., T., Dynamic Analysis of Falling Weight Deflectometer Data, " Transportation Research Record No. 1070, Transportation Research Board, National Research Council, Washington, D.C., 1986, pp. 63-68. Seed, H. B., Mitry, F. G., Monismith, C. L., and Chan, C. K., "Prediction of Flexible Pavement Deflections from Laboratory Repeated Load Tests, " National Cooperative Highway Research Program Report No. 35 , Highway Research Board, 1969. Seed, H. B., Chan, C. K., and Lee, C. E., "Resilient Charecteristics of Subgrade Soils and their Relation to Fatigue Failures in Asphalt Pavements," Wings, Fourth lntemational Conference on the Structural Design of Asphalt Pavements, University Of Michigan, Ann Arbor, Vol 1, September 1972, pp. 416-426. Shifley, L. H., "The Influence of Subgrade Characteristics on the Transient Deflections of Asphalt Concrete Pavements, " Ph. D. Thesis, Department of Civil Engineering, University of California, Berkeley, August 1967. Sivaneswaran, N., Steven, L. K., and Mahoney, J. P., "Advanced Backcalculation Using a Nonlinear Least Squares Optimization Technique, " Transportation Research Record No. 1293, Transportation Research Board, National Research Council, Washington, D. C., 1991, pp. 93-102. Smith, R. E., and Lytton, R. L., "Synthesis Study of Nondestructive Devices for Use in Overlay Thiclmess Design of Flexible Pavement, " Report No. FHWA/RD-83l097, Federal Highway Administration, Washington, D.C., April, 1984. 322 Southgate, H. F. , "An Evaluation of Temperature Distribution within Asphalt Pavements and its Relationship to Pavement Deflection, " Research Report No. HPP; 1(3), Part II, KYHPR-62-20, Kentucky Department of Highways, Lexington, April, 1968. Sparrow, R. W., and Tory, A. C., "Behavior of Soil Mass Under Dynamic Loading," Journal of the Soil Mechanics and Foundation Division, ASCE, May 1966. Stolle, D. ,and Hein, D., "Parameter Estimates of Pavement Structure layers and Uniqueness of Solution" 9 .. . . MQQHII. ASTM STP 1026, A. J. Bush 111, and G. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 312-322. Szendirei, M. E., and Freeme, C. R., "Road Response to Vibrating Tests," Mus, American Society of Civil Engineers, Vol. 96, No. SM6, pp. 2099- 2124, November, 1970. Tam, W. S., and Brown, S. F., "Back-Analyzed Elastic Stiffness: Comparison Between Different Evaluation Procedures" W W, ASTMSTP1026, A. J. Busth, andG. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 189-200. Tayabji, S. and Thompson, M. "Finite Element Analysis of a Railway Track Support System, " Report No. 76-257, Federal Railroad Administration Office of Research and Development, Washington D.C., July 1976. Taylor M. L. , "Characterization of Flexible Pavements by Nondestructive Testing, " Ph.D. Dissertation, Department of Civil Engineering, University of Illinois at Urbana-Champaign, 1978. Terrel, R. L., "Factors Influencing the Resilient Characteristics of Asphalt Treated Aggregates, " Ph. D. Dissertation, Department of Civil Engineering , University of California, Berkeley, August 1967. The Asphalt Institute, "Asphalt Overlays for Highways and Street Rehabilitation, " Manual Series 17, 1977. Thomas, Y. H., "Evaluation of Layered Material Properties from Measured Surface Deflections, " Ph.D. Dissertation , Department of Civil Engineering, University of Utah, March 1977. Thompson. W. T., Weapons. Prentice-Hall. Englewood Cliffs, N. 1., 1972. Timoshenko, S. P., and Goodier, J. N., W, Third Edition, . 323 McGraw-Hill Publishing Company, 1987. Trollope, D. H., Lee, I. K., and Morris, J., "Stresses and Deformations in Two- Layer Pavement Structures Under Slow Repeated Loading, " Wings, Australian Road Research Board, 1962. Uddin, W., Meyer, A. H., and Hudson, W. R., "Rigid Bottom Considerations for Nondestructive Evaluation of Pavements, " Transportation Research Record No. 1070, Transportation Research Board, National Research Council, Washington, DC. , 1986, pp. 21-29. Ulliditz, P. , "Computer Simulation of Pavement Performance," Report No. 18, The Institute of Roads, Transport, and Town Planning, The Technical University of Denmark, January, 1978. Ullidtz, P, . and Stubstad, R. N. , "Analytical-Empirical Pavement Evaluation Using the Falling Weight Deflectometer," Transportation Research Record No. 1022, Transportation Research Board, National Research Council, Washington, D.C., 1985, pp. 36-43. Uzan, J, "Granular Material Characterization," Transportation Research Record No. 1022, Transportation Research Board, National Research Council, Washington, D. C., 1985, pp. 52-59. Uzan, J ., "MODULUS User’s Guide," Appendix E of "Determination of Asphaltic Concrete Structural Properties by Nondestructive Testing," by Lytton et a1. , Final Report, Texas A&M University, Research Foundation Project RF 7026, Submitted to National Cooperative Highway Research Program, November, 1985 . Uzan, J., Lytton, R. L., and Germann,F. P., "General Procedure for Backwculating layer Moduli" W W, ASTMSTP 1026, A. J. BushIII, andG. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 217-228. Uzan, J. "Advanced Backcalculation Techniques, " Paper presented at the Second International Symposium on Nondestructive Testing of Pavements and Backcalculation of Moduli, Atlanta, Georgia, June 23-24, 1993. Van Cauwelaert, F. J., Alexander, D. R., White, T. D., and Barker, W. R., "Multilayer Elastic Program for Backcalculating layer Moduli in Pavement Evaluation," \919'. 9| 919‘ ‘. 91 9 it. 91911? .19 9.. .9."- .999 9 u 99 °, ASTM STP 1026, A. J. Bush III, and G. Y. Baladi, Eds., American Society for Testing and Materials, Philadelphia, 1989, pp. 171-188. Van der Poel, C. ,"Dynarnic Testing of Road Construction," Journal of Applied 324 Chemistry, Vol 1, Part 7, pp. 281-290, July 1951. Van der Poel, C. , "Vibration Research on Road Constructions," Symposium on Dynamic Testing of Soils, American Society for Testing Materials, Special Technical Publication No. 156, pp. 174-185, July 1953. Van der Poel, C. , "A General System of Describing the Visco-Elastic Properties of Bitumens and its Relation to Routine Test Data, " Journal of Applies Chemistry, Vol. 4, 1954. Vaswani, N. K. , "Method for Separately Evaluating Structural Performance of Subgrades and Overlaying Pavements," Highway Research Record No. 362, Highway Research Board, National Research Council, Washington, D.C., 1971. Von Quintus, H., and Kennedy, T. W., "Asphalt Aggregate Mixture Analysis System Properties Related to Pavement Performance," W, Vol. 58,1989. Wang, F., and Lytton, R. L., "System Identification Method for Backcalculating Pavement Layer Properties, " Transportation Research Record No. 1384, Transportation Research Board, National Research Council, Washington, D. C., 1993. Wang, M. C., "Stresses and Deflections in Cement Stabilized Soil Pavements," Ph.D. Dissertation, Department of Civil Engineering, University of California, Berkeley, September 1968. Wang, M. C, William, S., Amend, J. and Green, J. L., "USAF Wave Propagation NDTMCIhOd," \919‘.91919°‘ '.91' 9 it: 11-11.‘ 19 i.-...9r- 19.9119 099 , ASTMSTP 1026, A. J. Bush 111, and G. Y. Baladi, Eds., American Society for ' Testing and Materials, Philadelphia, 1989, pp. 155-168. "WASHO Road Test," Part 1, Special Report No. 18, Highway Research Board, Washington, D. C., 1954. "WASHO Road Test," Part 2, Special Report No. 22, Highway Research Board, Washington, D. C., 1955. Watkins, D. J ., Lysmer, and Monismith, C. L., "Nondestructive Pavement Evaluation by Wave Propagation Method, " Report No. TE-74-2, Office of Research Services, University of California, Berkeley, 1974. Westergaard, H. M..1heanLQf_Elasitsrty_and£lastim Dover Publication. Inc., New York, 1964. it. 325 Wiseman, G., Uzan, J., Hoffman, 1., and Livenh, M., "Simple Elastic Models for Pavement Evaluation Using Measured Deflection Bowls," Wings, Fourth lntemational Conference on the Structural Design of Asphalt Pavements, University Of Michigan, Ann Arbor, Vol 1, September 1972, pp. 416-426. Wiseman, G. , "The Deflection of an Elastic Plate on an Elastic Foundation," Emceedings, Fifth Asian Regional Conference on Soil Mechanics and Foundation Engineering, Banglore, India, Vol 1, 1975, pp. 341-346. Wiseman, G. , "Comparison of Methods of Determining Pavement Parameters from Deflection Bowl Measurements, " Wings, Seventh Asian Regional Conference on Soil Mechanics and Foundation Engineering, Haifa, Isreal, Vol 1, 1983, pp. 158- 165. Wiseman, G., Greenstein, J ., and Uzan, J. "Application of Simplified layered System to NDT Pavement Evaluation," Transportation Research Record No. 1022, Transportation Research Board, National Research Council, Washington, D.C., 1985, pp. 29-36. Wolf. 1. P., W Prentice-Hall. Ino- Englewood Cliffs, New Jersy, 1985. Yoder, E. J ., and M. W. Witczak, Wm, Second Edition, John Wiley and Sons Inc., New York, 1975. Young, M. A., Baladi, G. Y., "Repeated Load Triaxial Testing: State-of-the-Art," An interim Report Submitted to the Michigan State Department of State Highways and Transportation, Division of Engineering Research, Michigan State University, March, 1977. l .