BACKCALCULATION OF ASPHALT CONCRETE COMPLEX MODULUS CURVE BY LAYERED VISCOELASTIC SOLUTION By Ligang Lei A DISSERTATION Submitted to Michigan State University In partial fulfillment of the requirements For the degree of DOCTOR OF PHILOSOPHY Civil Engineering 2011 ABSTRACT BACKCALCULATION OF ASPHALT CONCRETE COMPLEX MODULUS CURVE BY LAYERED VISCOELASTIC SOLUTION By Ligang Lei In-situ evaluation of material properties is very important for estimating the structural adequacy of a pavement section under traffic loading. The fundamental material property for the AC layer in a flexible pavement is the complex modulus |E*| or the equivalent relaxation modulus E(t). In-situ |E*| can be used as a quality control tool as well as in estimation of remaining service life of existing pavements, and it is a critical input to the new Mechanism-Empirical Pavement Design Guide (M-E PDG). In this research, a method is presented to backcalculate |E*| or E(t) for the AC layer, in addition to base and subgrade layer moduli, using Falling Weight Deflectometer (FWD) surface deflection time histories. First, the deflection time histories from FWD test data were separated as the dynamic response part, and the viscoelastic response part: The time delay of each sensor is related to wave propagation through the pavement system while the shifted deflection time history without time delay can be considered as the viscoelastic response of the pavement. Second, the time delay of each sensor was used to estimate the elastic modulus of the subgrade layer, based on the wave propagation theory. Third, a forward layered viscoelastic solution was developed based on Schapery’s ‘quasi-elastic’ approximation. Finally, backcalculation was numerically done by Newton’s method, and followed by the MATLAB internal function ‘fminsearch’, to match the predicted deflection time histories from estimated modulus of each layer against the shifted measurement from the FWD test. The backcalculation results show very good agreement with the actual modulus values for both numerical examples and field FWD test data. The error in the base and subgrade moduli is generally less than 3%. The AC layer relaxation and complex modulus curves match the actual functions within the ranges of 0.0001 to 0.1 sec, and 10 to 10,000 Hz, respectively. The sensitivity analysis shows that errors in the deflection time-histories are the most significant factor affecting the backcalculated AC relaxation and complex modulus curves. Also, the layer thicknesses should be as accurate as possible. TO MY PARENTS, FAMILY AND FRIENDS iv ACKNOWLEDGEMENT The author would like to express his great appreciation to his major advisor Dr. Karim Chatti, Professor of Civil Engineering at Michigan State University, for his continuous academic and financial support throughout the author’s study at Michigan State University and while preparing this dissertation. The author would like to thank his co-advisor, Dr. Emin Kutay, Assistant Professor of Civil Engineering, for his valuable suggestions and kind help. This dissertation could not have been completed without their help. The author also appreciated help from other committee members, Dr. Gilbert Baladi, Professor of Civil Engineering, and Dr. Tom Pence, Professor of Mechanical Engineering, for their continuous advice and support. Thanks are also due to the Federal Highway Administration (FHWA) for financially sponsoring this research, and to the author’s graduate student colleagues at the Department of Civil and Environmental Engineering at MSU. Last but not least, the author would like to thank his wife, Huajing Yang, for her continuous and unselfish support and company throughout his stay at Michigan State University. v TABLE OF CONTENTS LIST OF TABLES......................................................................................................................... ix LIST OF FIGURES..................................................................................................................... xii Chapter 1. Introduction ................................................................................................................. 1 Chapter 2. Literature Review and Background ............................................................................ 5 2.1 Backcalculation Approach ................................................................................................. 5 2.1.1 Static Backcalculation Methods..................................................................................... 6 2.1.2 Dynamic Backcalculation Methods ............................................................................... 7 2.2 Primary Response Behavior of Asphalt Concrete............................................................ 8 2.3 Interconversion between E(t) and |E*| for AC Mixture ................................................ 13 2.4 Interconversion between E(t) and D(t) for AC Mixture ................................................ 15 2.5 Motivation for the Study .................................................................................................. 19 Chapter 3. Estimating the Subgrade Elastic Modulus using Wave Propagation Method ........ 22 3.1 Introduction....................................................................................................................... 22 3.2 Proposed Method to Estimate k-value ............................................................................ 23 3.2.1 Theoretical Formulation for Spectral Element Method ............................................... 24 3.2.2 Dominant Frequency of the FWD Test ........................................................................ 31 3.2.3 Numerical Example to Calculate k-value .................................................................... 32 3.2.4 Sensitivity Analysis for k-value ................................................................................... 33 3.3 Proposed Procedure to Estimate Subgrade Elastic Modulus ....................................... 38 3.4 Verification of the Proposed Method............................................................................... 42 3.4.1 SAPSI........................................................................................................................... 42 3.4.2 LAMDA....................................................................................................................... 43 3.4.3 Case of Subgrade without Ground Water Table........................................................... 44 3.4.4 Case of Subgrade with GWT ....................................................................................... 46 3.5 Validating the Proposed Method Using Case Studies from LTPP................................ 49 3.5.1 Example of FWD Test at LTPP 04-1036 Station in 1998 ............................................ 49 3.5.2 Example of FWD Test at LTPP 32-0101 Station in 1996 ............................................ 52 3.5.3 Example of the FWD Test at LTPP Station 06-0565 in 1999 ...................................... 56 3.5.4 Discussion of Case Studies Obtained from LTPP........................................................ 60 3.5.4.1 Effect of Confinement....................................................................................... 60 vi 3.5.4.2 Effect of Sampling Intervals ............................................................................. 63 3.6 Numerical Example of Bedrock under Subgrade layer ................................................ 64 3.7 Summary............................................................................................................................ 67 Chapter 4. Layered Viscoelastic Forward Solution .................................................................... 68 4.1 Introduction....................................................................................................................... 68 4.2 Layered Viscoelastic Forward Solution Algorithm........................................................ 68 4.2.1 Theoretical Background............................................................................................... 68 4.2.2 Algorithm Steps ........................................................................................................... 71 4.3 Verification of the Layered Viscoelastic Forward Solution........................................... 76 4.3.1 Verification of Quasi-Elastic Approximation............................................................... 76 4.3.2 Comparison between the Layered VE Solution and Dynamic Solutions .................... 78 4.3.3 Verification Examples using Pavement Structures from the LTPP Database .............. 83 4.3.4 A Numerical Example with the Bedrock ..................................................................... 89 4.4 Time Efficiency for Layered VE Forward Solution....................................................... 91 4.5 Summary............................................................................................................................ 93 Chapter 5. Backcalculation Algorithm........................................................................................ 94 5.1 Introduction....................................................................................................................... 94 5.2 Variable Identification ...................................................................................................... 94 5.3 Variable Boundaries.......................................................................................................... 95 5.3.1 Boundary Limits for |E*|.............................................................................................. 95 5.3.2 Boundary Limits for E(t) ............................................................................................. 97 5.4 Inverse Solution Theory ................................................................................................... 99 5.4.1 Basic Theory Background............................................................................................ 99 5.4.2 SVD Method .............................................................................................................. 102 5.4.3 Truncating Singular Values ........................................................................................ 103 5.5 VE backcalculation of E(t) ............................................................................................. 104 5.6 VE backcalculation of |E*| ............................................................................................. 107 5.6.1 Forward calculation of E(t) from |E*| and φ .............................................................. 107 5.6.2 Calculation of E(t) and φ from |E*| by Iteration ........................................................ 108 5.6.3 Calculation of E(t) from |E*(ω)| by Iteration φ(ω) .....................................................112 5.6.4 Flowchart of VE backcalculation of |E*| ....................................................................114 5.7 One Numerical Example of Backcalculation Result.....................................................116 5.7.1 Result of VE backcalculation of E(t) ..........................................................................116 5.7.2 Result of VE backcalculation of |E*| ..........................................................................118 5.8 Summary.......................................................................................................................... 120 Chapter 6. Verification and Sensitivity Analysis ....................................................................... 121 vii 6.1 Introduction..................................................................................................................... 121 6.2 Verification of the VE Backcalculation Algorithm ...................................................... 121 6.2.1 VE Backcalculation using Time Histories from VE Solution.................................... 123 6.2.1.1 Numerical Example 1 ..................................................................................... 124 6.2.1.2 Numerical Example 2 ..................................................................................... 129 6.2.1.3 Numerical Example 3 ..................................................................................... 133 6.2.1.4 Numerical Example 4 ..................................................................................... 137 6.2.1.5 Discussion ....................................................................................................... 141 6.2.2 VE Backcalculation using Time Histories from Dynamic Solutions......................... 141 6.2.2.1 Example 5: Michigan Site............................................................................... 141 6.2.2.2 Example 6: Texas Site..................................................................................... 145 6.2.2.3 Example 7: Florence Site ................................................................................ 150 6.2.2.4 Example 8: Kansas Site .................................................................................. 154 6.2.2.5 Discussion ....................................................................................................... 158 6.2.3 VE Backcalculation using Field FWD Test Data....................................................... 158 6.2.3.1 FWD Test Conducted at L9S4 ........................................................................ 159 6.2.3.2 FWD Test Conducted at L10S4 ...................................................................... 161 6.2.3.3 FWD Test Conducted at L11S4 ...................................................................... 163 6.2.3.4 Discussion ....................................................................................................... 165 6.3 Sensitivity Analysis for the AC Layer ........................................................................... 165 6.3.1 Sensitivity Analysis for Deflection Measurement of the First Sensor....................... 168 6.3.2 Sensitivity Analysis for Time Shift Error of the First Sensor .................................... 172 6.3.3 Sensitivity Analysis for Post-Peak Time Offset of the First Sensor .......................... 177 6.3.4 Sensitivity Analysis for Layer Thickness .................................................................. 181 6.3.5 Discussion of Sensitivity Analyses ............................................................................ 186 6.4 Summary.......................................................................................................................... 186 Chapter 7. Conclusions and Recommendations ....................................................................... 188 7.1 Summary.......................................................................................................................... 188 7.2 Conclusions...................................................................................................................... 189 7.3 Recommendations ........................................................................................................... 190 BIBLIOGRAPHY....................................................................................................................... 192 viii LIST OF TABLES Table 1. Prony series coefficients for E(t) and D(t) for a typical AC mixture ...................... 18 Table 2: The equivalent frequency for different load duration times.................................... 32 Table 3: The average k-value for different cases of Poisson’s ratio ..................................... 37 Table 4: Basic information of a three-layer pavement structure without GWT ................... 44 Table 5: Time delay of each sensor simulated by SAPSI ..................................................... 44 Table 6: Basic information for a three-layer pavement structure with GWT ....................... 46 Table 7: Time delay of each sensor by LAMDA .................................................................. 47 Table 8: Time delay and peak deflection of each sensor due to the FWD test at station 04-1036 in 1998 ............................................................................................................ 50 Table 9: Structural information and MODCOMP5 backcalculated modulus at LTPP station 04-1036 in 1998 ............................................................................................................ 50 Table 10: Time delay and maximum deflection of sensors during the FWD test at LTPP station 32-0101 in 1996 ................................................................................................ 53 Table 11: Basic information and MODCOMP5 backcalculated modulus at LTPP station 32-0101 in 1996 (Trial: I) ............................................................................................. 54 Table 12: Basic information and MODCOMP5 backcalculated modulus at LTPP station 32-0101 in 1996 (Trial: II) ............................................................................................ 54 Table 13: Elastic modulus of the subgrade layer at LTPP station 32-0101 in 1996 ............. 55 Table 14: Time delay and maximum deflection of the sensors due to the FWD test at LTPP station 06-0565 in 1999 ................................................................................................ 57 Table 15: Structural information and MODCOMP5 backcalculated modulus at LTPP station 06-0565 in 1999 (Trial: I) ............................................................................................. 58 Table 16: Structural information and MODCOMP5 backcalculated modulus at LTPP station 06-0565 in 1999 (Trial: II) ............................................................................................ 58 Table 17: Elastic modulus of the subgrade layer at station 06-0565 in 1999 ....................... 59 Table 18: Results of data regression of resilient modulus on stress parameters using Uzan ix and Cornell Models....................................................................................................... 61 Table 19: The stress status of the subgrade layer in the FWD test for different K0 values... 62 Table 20: The range of elastic moduli predicted by nonlinear models, measured in the lab, estimated by the proposed method, and backcalculated by MODCOMP5 for the case studies (MPa) ................................................................................................................ 63 Table 21: Arrival time for S-wave and R-wave for each sensor due to the FWD test.......... 64 Table 22: Material properties of a three-layer pavement structure with bedrock ................. 65 Table 23: Time delay of each sensor by SPASI simulation with bedrock ............................ 66 Table 24: Material properties of the pavement structure used in the verification example .. 76 Table 25: Properties of a pavement used in comparison of dynamic and VE solutions ....... 80 Table 26: Several examples of pavement structure in the SPS-1 project ............................. 83 Table 27: Layer information for each pavement structure.................................................... 84 Table 28: The computation times for different numbers of discrete time steps.................... 92 Table 29: The |E*| parameters for different AC mixtures ..................................................... 96 Table 30: The average and boundaries for the four parameters of |E*|................................. 97 Table 31: The E(t) parameters for different AC mixtures ..................................................... 98 Table 32: The mean and boundaries for the four parameters of E(t) .................................... 99 Table 33: Basic information of a three-layer pavement structure and backcalculated result ......................................................................................................................................116 Table 34:Basic information of pavement structure and backcalculated result for numerical example 1 .................................................................................................................... 124 Table 35:Basic information of pavement structure and backcalculated result for numerical example 2 .................................................................................................................... 129 Table 36:Basic information of pavement structure and backcalculated result for numerical example 3 .................................................................................................................... 133 Table 37:Basic information of pavement structure and backcalculated result for numerical example 4 .................................................................................................................... 137 Table 38:Basic information of Michigan pavement structure and backcalculated results142 x Table 39:Basic information of Texas pavement structure and backcalculated results..... 146 Table 40: Basic information of a Florence pavement structure and backcalculated result. 150 Table 41: Basic information of a Kansas pavement structure and backcalculated result ... 154 Table 42: Basic information of L9S4 pavement structure and backcalculated result......... 159 Table 43: Basic information of L10S4 pavement structure and backcalculated result....... 161 Table 44: Basic information of L11S4 pavement structure and backcalculated result ....... 163 Table 45: Layer information of each pavement structure for sensitivity analysis .............. 166 xi LIST OF FIGURES Figure 1: Organization of this dissertation.............................................................................. 3 Figure 2: Sinusoidal Stress and Strain in cyclic loading......................................................... 9 Figure 3: Mastercurve |E*| of one asphalt concrete mixture (PG64-28) .............................. 10 Figure 4: Phase angle φ(ω) of one asphalt concrete mixture (PG64-28)...............................11 Figure 5: Physical meaning of each parameter on the sigmoidal function ........................... 12 Figure 6: Relaxation modulus |E(t)| of one asphalt concrete mixture (PG64-28)................. 13 Figure 7: The generalized Maxwell model ........................................................................... 14 Figure 8: E(t) and D(t) from interconversion for a typical AC mixture................................ 19 Figure 9: Wave system from surface point source in ideal medium ..................................... 30 Figure 10: Phase angle of each point at the surface for a specific example (t =0 s)............. 32 Figure 11: Sensitivity analysis for k-value (ν= 0.20)............................................................ 34 Figure 12: Sensitivity analysis for k-value (ν= 0.25)............................................................ 34 Figure 13: Sensitivity analysis for k-value (ν= 0.30)............................................................ 35 Figure 14: Sensitivity analysis for k-value (ν= 0.35)............................................................ 35 Figure 15: Sensitivity analysis for k-value (ν= 0.40)............................................................ 36 Figure 16: Sensitivity analysis for k-value (ν= 0.45)............................................................ 36 Figure 17: Sensitivity analysis for k-value (ν= 0.49)............................................................ 37 Figure 18: Example of FWD load and deflection time histories at LTPP station 04-1036 in 1998............................................................................................................................... 39 Figure 19: Time delay vs. location for the example of FWD test at LTPP station 04-1036 in 1998............................................................................................................................... 40 Figure 20: Flow chart to estimate subgrade elastic modulus using k-value ......................... 41 Figure 21: Time delay vs. location for a numerical example of FWD test (without GWT). 45 Figure 22: LAMDA simulated time-history due to the FWD test (with GWT) ................... 47 xii Figure 23: LAMDA simulated time-delay for the FWD test (with GWT) ........................... 48 Figure 24: Deflection basin from FWD test and MODCOMP5 Simulation at LTPP station 04-1036 in 1998 ............................................................................................................ 51 Figure 25: Example of FWD test results at LTPP station 32-0101 in 1996.......................... 53 Figure 26: Deflection basin due to the FWD test and two cases of MODCOMP5 simulation at LTPP station 32-0101 in 1996................................................................................... 55 Figure 27: Example of the FWD test history at LTPP station 06-0565 in 1999 ................... 57 Figure 28: Deflection basin during the FWD test and two cases of MODCOMP5 simulation at LTPP station 06-0565 in 1999................................................................................... 59 Figure 29: SAPSI simulated time-history of FWD test with bedrock underneath ............... 65 Figure 30: Time delay vs. location for a numerical example of FWD test with bedrock..... 66 Figure 31: Typical geometry of a pavement structure .......................................................... 70 Figure 32: Discretization of stress history ............................................................................ 72 Figure 33: Discretization of the relaxation modulus mastercurve ........................................ 72 Figure 34: Deflections calculated for points at different distances from the centerline of the circular load at the surface ............................................................................................ 73 Figure 35: Illustration of the dσ (τ j ) in Eq. (32b) for each time step τ j .................... 74 Figure 36: Example of a pavement structure used for VE forward calculation.................... 75 Figure 37: Examples of computed VE surface deflections at different radial distances from the centerline of the load............................................................................................... 75 Figure 38: Implemented and actual E(t) for one AC mixture ............................................... 77 Figure 39: Implemented and actual E(t) for one AC mixture ............................................... 77 Figure 40: FWD test simulation by layered VE solution for a half-space AC layer ............ 79 Figure 41: Sensor deflection time histories predicted by SAPSI.......................................... 80 Figure 42: Sensor deflection time histories predicted by LAMDA...................................... 81 Figure 43: Comparison of deflection time histories from VE and SAPSI solutions ............ 82 Figure 44: Comparison of deflection time histories from VE and LAMDA solutions......... 82 xiii Figure 45: Time-shifted dynamic solution and VE solution for case 116............................. 85 Figure 46: Time-shifted dynamic solution and VE solution for case 117............................. 86 Figure 47: Time-shifted dynamic solution and VE solution for case 120 ............................ 87 Figure 48: Time-shifted dynamic solution and VE solution for case 123 ............................ 88 Figure 49: Time-shifted SAPSI and VE solution for shallow bedrock................................. 90 Figure 50: Sensitivity analysis for relative error of maximum deflection between VE solution and SAPSI with bedrock depth ....................................................................... 91 Figure 51: Random Function to avoid local minimum problem......................................... 105 Figure 52: Algorithm of backcalculation procedure of E(t)................................................ 106 Figure 53: Algorithm for Calculation of E(t) and φ from |E*| by iteration..........................111 Figure 54: Algorithm for Calculation of E(t) from |E*| by iteration φ ................................113 Figure 55: |E*| input and calculated after converting to E(t) for one AC mixture...............114 Figure 56: Algorithm of backcalculation procedure of |E*|.................................................115 Figure 57: Input and backcalculated time-histories of E(t) in the numerical FWD test ......117 Figure 58: Input and backcalculated E(t) for the numerical FWD test................................117 Figure 59: Input and backcalculated time-histories of |E*| in the numerical FWD test ......119 Figure 60: Input and backcalculated |E*| for the numerical FWD test ................................119 Figure 61: Available range in E(t) function from field FWD test backcalculation............. 122 Figure 62: Useful range in |E*| function from field FWD test backcalculation ................. 123 Figure 63: Input and simulated time-histories by backcalculation of E(t) in example 1.... 125 Figure 64:Input and backcalculated E(t) in example 1 .................................................... 125 Figure 65: Input and simulated time-histories by backcalculation of |E*| in example 1 .... 126 Figure 66: Input and backcalculated |E*| in example 1 ...................................................... 127 Figure 67: Input and backcalculated φ in example 1 .......................................................... 128 Figure 68: Input and simulated time-histories by backcalculation of E(t) in example 2.... 130 xiv Figure 69:Input and backcalculated of E(t) in example 2................................................ 131 Figure 70: Input and simulated time-histories by backcalculation of |E*| in example 2 .... 132 Figure 71: Input and backcalculated of |E*| in example 2 .................................................. 132 Figure 72: Input and simulated time-histories by backcalculation of E(t) in example 3.... 134 Figure 73:Input and backcalculated E(t) in example 3 .................................................... 135 Figure 74: Input and simulated time-histories by backcalculation of |E*| in example 3 .... 135 Figure 75: Input and backcalculated |E*| in example 3 ...................................................... 136 Figure 76: Input and simulated time-histories by backcalculation of E(t) in example 4.... 138 Figure 77:Input and backcalculated E(t) in example 4 .................................................... 138 Figure 78: Input and simulated time-histories by backcalculation of |E*| in example 4 .... 139 Figure 79: Input and backcalculated |E*| in example 4 ...................................................... 140 Figure 80: Input and simulated time-histories by backcalculation of E(t) in Michigan site143 Figure 81:Input and backcalculated E(t) in Michigan site............................................... 143 Figure 82: Input and simulated time-histories by backcalculation of |E*| in Michigan site144 Figure 83: Input and backcalculated |E*| in Michigan site ................................................. 145 Figure 84:Input and simulated time-histories by backcalculation of E(t) in Texas site... 147 Figure 85:Input and backcalculated E(t) in Texas site ..................................................... 147 Figure 86:Input and simulated time-histories by backcalculation of |E*| in Texas site ... 148 Figure 87:Input and backcalculated |E*| in Texas site ..................................................... 149 Figure 88: Input and backcalculation simulated time-histories using E(t) in Florence site 151 Figure 89: Input and backcalculated E(t) in Florence site .................................................. 151 Figure 90: Input and backcalculation simulated time-histories using |E*| in Florence site 152 Figure 91: Input and backcalculated |E*| in Florence site .................................................. 153 Figure 92: Input and simulated time-histories by backcalculation of E(t) in Kansas site .. 155 Figure 93: Input and backcalculated E(t) in Kansas site..................................................... 155 xv Figure 94: Input and simulated time-histories by backcalculation of |E*| in Kansas site .. 156 Figure 95: Input and backcalculated |E*| in Kansas site..................................................... 157 Figure 96: Input and simulated time-histories by backcalculation of E(t) in L9S4 site ..... 160 Figure 97: Input and backcalculated E(t) in L9S4 site ....................................................... 160 Figure 98: Input and simulated time-histories by backcalculation of E(t) in L10S4 site ... 162 Figure 99: Input and backcalculated E(t) in L10S4 site ..................................................... 162 Figure 100: Input and simulated time-histories by backcalculation of E(t) in L11S4 site . 164 Figure 101: Input and backcalculated E(t) in L11S4 site.................................................... 164 Figure 102: Error caused by time shift of entire pulse for one sensor................................ 167 Figure 103: Error caused by expansion/contraction of deflection pulse in the post-peak zone for one sensor.............................................................................................................. 167 Figure 104: Sensitivity analysis of |E*| for deflection measurement error (Case A).......... 168 Figure 105: Sensitivity analysis of E(t) for deflection measurement error (Case A).......... 169 Figure 106: Sensitivity analysis of |E*| for deflection measurement error (Case B).......... 169 Figure 107: Sensitivity analysis of E(t) for deflection measurement error (Case B).......... 170 Figure 108: Sensitivity analysis of |E*| for deflection measurement error (Case C).......... 170 Figure 109: Sensitivity analysis of E(t) for deflection measurement error (Case C).......... 171 Figure 110: Sensitivity analysis of |E*| for deflection measurement error (Case D).......... 171 Figure 111: Sensitivity analysis of E(t) for deflection measurement error (Case D).......... 172 Figure 112: Sensitivity analysis of |E*| for time shift error of sensor 1 (Case A) .............. 173 Figure 113: Sensitivity analysis of E(t) for time shift error of sensor 1 (Case A) .............. 173 Figure 114: Sensitivity analysis of |E*| for time shift error of sensor 1 (Case B) .............. 174 Figure 115: Sensitivity analysis of E(t) for time shift error of sensor 1 (Case B) .............. 174 Figure 116: Sensitivity analysis of |E*| for time shift error of sensor 1 (Case C) .............. 175 Figure 117: Sensitivity analysis of E(t) for time shift error of sensor 1 (Case C) .............. 175 Figure 118: Sensitivity analysis of |E*| for time shift error of sensor 1 (Case D) .............. 176 xvi Figure 119: Sensitivity analysis of E(t) for time shift error of sensor 1 (Case D) .............. 176 Figure 120: Sensitivity analysis of |E*| for post-peak time offset of sensor 1 (Case A)..... 177 Figure 121: Sensitivity analysis of E(t) for post-peak time offset of sensor 1 (Case A) .... 178 Figure 122: Sensitivity analysis of |E*| for post-peak time offset of sensor 1 (Case B)..... 178 Figure 123: Sensitivity analysis of E(t) for post-peak time offset of sensor 1 (Case B) .... 179 Figure 124: Sensitivity analysis of |E*| for post-peak time offset of sensor 1 (Case C)..... 179 Figure 125: Sensitivity analysis of (t) for post-peak time offset of sensor 1 (Case C)....... 180 Figure 126: Sensitivity analysis of |E*| for post-peak time offset of sensor 1 (Case D) .... 180 Figure 127: Sensitivity analysis of E(t) for post-peak time offset of sensor 1 (Case D) .... 181 Figure 128: Sensitivity analysis of |E*| for layer thickness (Case A) ................................. 182 Figure 129: Sensitivity analysis of E(t) for layer thickness (Case A)................................. 182 Figure 130: Sensitivity analysis of |E*| for layer thickness (Case B) ................................. 183 Figure 131: Sensitivity analysis of E(t) for layer thickness (Case B)................................. 183 Figure 132: Sensitivity analysis of |E*| for layer thickness (Case C) ................................. 184 Figure 133: Sensitivity analysis of E(t) for layer thickness (Case C)................................. 184 Figure 134: Sensitivity analysis of |E*| for layer thickness (Case D)................................. 185 Figure 135: Sensitivity analysis of E(t) for layer thickness (Case D)................................. 185 xvii Chapter 1. Introduction Asphalt concrete (AC) mixtures consist of asphalt binder and aggregates from crushed stone or from natural resources. They were first used for pavement in 1870 in Newark, New Jersey. According to the Federal Highway Administration (FHWA, 2001), 2.5 million miles of road are paved in the USA, and 94% are paved with AC, i.e., about 2.3 million miles of road are surface-covered by AC, or approximately 9 times the distance to the moon. About 300 million tons of AC mixtures are used in highway construction every year (Papagiannakis and Masad, 2008). Before the 1920s, only the thickness of pavement was considered in pavement design and analysis. It was believed the thicker the pavement, the longer the pavement would last. After many years of experience, three key factors are found to significantly affect pavement life: traffic or loading, environment and material properties. Traffic information is estimated by regional planning departments, while national weather stations can provide environmental conditions for a specific location. The remaining challenge for pavement engineering is the material properties of the pavement, especially the onsite materials used many years ago. In recent years, the falling weight deflectometer (FWD) test has been the typical method for evaluating the material properties of in-service pavement. In order to accurately evaluate pavement structure, several FWD tests with different load levels are usually performed at the same location. The structural condition, or layer modulus, is the key factor for determining pavement rehabilitation strategies. The material property is usually obtained through static or dynamic backcalculation methods, 1 in which layer moduli are determined by matching the deflection basin measured under a known load in the FWD test, with the deflection basin generated through a theoretical model of the pavement (Ji, 2005). Static backcalculation is mainly based on the layered linear elastic solution, while dynamic backcalculation modifies the linear elastic property into the damped elastic property for each physical layer. However, both models of the pavement structure are theoretically incorrect, because the primary response, i.e., small load-level response, of AC mixture is viscoelastic, rather than elastic. The mechanical behaviors of viscoelastic materials are quite different from those of the elastic or damped elastic material, for example, granular materials, aggregate base, or subgrade. The objective of this research is to develop a robust methodology to backcalculate the |E*| and E(t) mastercurve of the AC layer, and the elastic moduli of unbound layers. The dynamic effect due to the FWD test should be investigated first, and the pavement response is expected to be simplified as a viscoelastic response. Additionally, the elastic modulus of the subgrade layer can be estimated or identified from the dynamic response. After that, a pavement structure is simulated as viscoelastic system, and backcalculation is done in viscoelastic solution. A brief flowchart and the explanation of this dissertation are shown as following: 2 Chapter 3: Wave propagation and estimation Esubgrade AC AC Ε(t), ν1 Ε ,ν Base H=Inf 2 Subgrade Chapter 4: Layered viscoelastic forward solution 2 Chapter 5: Backcalculation of E(t)/ E* and Eunbound Εn, νn Chapter 6: Backcalculation results & sensitivity analysis Figure 1: Organization of this dissertation (For interpretation of the references to color in this and all other figures, the reader is referred to the electronic version of this dissertation.) Chapter 2 provides a review of relevant literature on the backcalculation approaches from field FWD tests, background on the material properties of the AC mixture as viscoelastic materials, and the interconversions between |E*|, E(t) and D(t) within the frame work of viscoelasticity. The purpose of the study is briefly given at the end of this chapter. Chapter 3 presents a method for calculation of Esubgrade based on wave propagation theory. The concept of equivalent frequency is introduced, and the application of Rayleigh wave in analysis of FWD test data is explained. As a validation, four cases of field tests from Long Term Pavement Performance (LTPP) database are demonstrated. The merits and limitations are given. 3 Chapter 4 introduces a layered viscoelastic forward solution. The response of viscoelastic multi-layered pavement system under FWD test is presented. The pavement is modeled as a system of the horizontal layers, and AC mixture is simulated as viscoelastic material, and the rest layers are isotropic and linear elastic. The forward solution is verified against the semi-analytical viscoelastic solution of a step surface load, and then checked against well-known dynamic solution SAPSI and LAMDA. Chapter 5 contains the details of the backcalculation algorithm, and some brief information on the backcalculation algorithm developed in MATLAB. A hypothetical FWD test deflection time histories, simulated by dynamic solution SAPSI, was used as the input parameters, and the modulus of each physical layer was backcalculated to verify the accuracy of the algorithm. Chapter 6 presents the validation of the viscoelastic backcalculation algorithm using numerical and field FWD test deflection time histories. Sensitivity analyses of various parameters on the backcalculation were conducted. The effects of inaccuracies in deflections and duration of deflection records were studied. Chapter 7 contains a summary of the findings, the impact of the research, and recommendations for future research. 4 Chapter 2. Literature Review and Background 2.1 Backcalculation Approach The elastic moduli of pavement layers are obtained from backcalculation via field FWD tests. To find the elastic modulus of each layer, backcalculation from FWD test data is usually carried out by matching the measured deflections under a known load with theoretical deflections generated by an analytical model of the pavement, usually by Newton’s method (Harichandran et al, 1994). Such procedures usually use error minimization techniques to minimize either the absolute or the square error, with or without weighting factors for sensors in the FWD test. At present, pavement moduli can be backcalculated from the FWD deflection basin, using the peak values of the deflection time-histories (static backcalculation) or using the FWD full time-history (dynamic backcalculation). However, deflection basins under static loads differ from those under dynamic or impulse loads because of geometric damping of pavement structure, and dynamic effects, such as inertia, material damping, and resonance. Dynamic analysis, or the time-history record, would therefore provide a more accurate estimation of the pavement moduli, at the cost of time. Additionally, the interpretation of data still remains problematic (Ji, 2005). This is due to the limitation associated with the mechanical models incorporated into the backcalculation procedures and the uniqueness of the inverse solutions. Nevertheless, during the past few decades, there has been a significant improvement in the area of pavement modeling and non-destructive test (NDT) techniques. In the following sub-sections, the static and dynamic 5 methods and backcalculation schemes are discussed. 2.1.1 Static Backcalculation Methods The simplest way to model the behavior of flexible pavements is based on Boussinesq’s (1885) solution that models a flexible pavement as a homogeneous, isotropic, and elastic half-space. Later, Burmister (1943) modified this method into a two-layer system. Acum and Fox (1951) advanced this model into a three-layer system. Newer methods, including CHEVRON (Warren and Dieckmann, 1963) and KENLAYER (Huang, 2004) were implemented to analyze the interface conditions between layers. Finite element methods, such as MICH-PAVE (Yeh, 1989) were also developed. The FWD test load is simulated as a static load with the maximum magnitude of the impulse. The most recent layered elastic model is CHEVLAY2 (Stubstad et al, 2007). Backcalculation programs typically perform on various forward computations to match the deflection basins from FWD test to the computed deflections. There are three major groups of static backcalculation methods, and each utilizes different techniques to reach the solution. The first group is based on iteration techniques, where the layer moduli are repeatedly adjusted until a suitable match between the calculated and measured deflection basins is obtained. The typical programs for this method are MODCOMP (Irwin, 1994), BISDEF (Bush, 1985), BOUSDEF (Roesset et al, 1995), and CHEVDEF (Bush and Alexander, 1985). The second group is based on searching a database of deflection basins. A forward calculation scheme is used to generate a database, which is then searched to find a best match for the observed deflection basin. One typical example is MODULUS (Uzan, 1994). The third group is based on the use of regression equations fitted to a database of deflection basins generated by a forward calculation scheme. The LOADRATE program (Chua and Lytton, 6 1985) belongs to this category and uses regression equations generated from a database obtained by using the ILLIPAVE (Raad and Figueroa, 1980) nonlinear finite element program. A more detailed review on static backcalculation methods can be found elsewhere (Mahmood, 1993). 2.1.2 Dynamic Backcalculation Methods Most dynamic backcalculation methods use dynamic damped-elastic finite layer or finite element models for their forward solutions. The finite layer solutions are based on Kausel’s formulation (Kausel and Peek, 1982) which subdivides the medium into discrete layers that have a linear displacement function in the vertical direction and satisfy the wave equation in the horizontal direction. The solution is based on the premise that if the sub-layer thickness is small relative to the wavelength of interest (around one tenth), it is possible to linearize the transcendental functions and reduce them to algebraic expressions. The typical example of this method is SAPSI (Chen, 1987). Al-Khoury et al (2001) developed an efficient forward solution, LAMDA, for the dynamic analysis of flexible pavements using the spectral element technique for the simulation of wave propagation in layered systems. The method is able to model each layer as one element without the need for subdivision into several sub-layers. However, the horizontal range (R) is required to simulate the vanishing of the wave toward the infinity. Endiran (1999) developed a non-linear dynamic model in DYNARK, a computer program accounting for non-linear properties in granular material as well as subgrade soil. The backcalculation methods for dynamics are based on either frequency or time domain solutions. In the frequency domain, the applied load and measured deflection time-histories are transformed into the frequency domain by using the Fast Fourier Transform (FFT). Backcalculation of layer parameters is done by matching the calculated steady-state (complex) 7 deflection basins with the frequency component of the measured sensor deflections at one or more frequencies. In time domain backcalculation, the measured deflection time histories are directly compared with the predicted results predicted by the forward program. The advantages of backcalculation in the frequency domain are the computational efficiency and theoretical validity. The disadvantage is that the results are very sensitive to truncation, sampling time interval, and rest period after the loading impulse. Truncation of deflection history of the sensors in the FWD tests is very common. The advantage of backcalculation in the time domain is that the matching of deflection history can be achieved for any desired time interval. The disadvantage is too many outputs in the forward calculation, so it is hard to converge for forward and backcalculated results, perhaps even impossible. A more detailed review of dynamic backcalculation can be found elsewhere (Ji, 2005). 2.2 Primary Response Behavior of Asphalt Concrete It is well known that asphalt concrete (AC) behaves as a linear elastic material under small strain conditions (Kim, 2008). A typical stress/strain behavior of AC in cyclic tests is shown in Figure 2. 8 Stress/strain under cyclic loading Normalized stress/strain 1 Normalized Stress Normalized Strain 0.5 0 -0.5 -1 time Figure 2: Sinusoidal Stress and Strain in cyclic loading The time lag between stress and strain history illustrates that AC mixtures are viscoelastic materials. Typically, the magnitude of dynamic complex modulus |E*| and the phase angle φ are measured at different frequencies and at different loading temperatures. After eliminating the effect of temperature by the time-temperature superposition principle, the mastercurve |E*| is obtained, which is the magnitudes of dynamic modulus E* as a function of reduced frequency at reference temperature. Experimental data show both |E*| and φ are only function of reduced frequency ω for a specific mixture. One example of |E*| mastercurve is shown in Figure 3, and the phase angle function is shown in Figure 4: 9 100000 |E*| (MPa) 10000 1000 E* Average 100 10 1.E-06 Sigmoid Fit, |E*| Mpa 1.E-04 1.E-02 1.E+00 1.E+02 1.E+04 Reduced Freq. (Hz) Figure 3: Mastercurve |E*| of one asphalt concrete mixture (PG64-28) (Kutay, 2008) 10 40.0 35.0 30.0 Phase Angle 25.0 20.0 y = 0.2186x3 - 0.53x2 - 6.1464x + 26.906 R2 = 0.9785 15.0 10.0 5.0 0.0 -6.0E+00 -5.0E+00 -4.0E+00 -3.0E+00 -2.0E+00 -1.0E+00 0.0E+00 1.0E+00 2.0E+00 3.0E+00 4.0E+00 Log fr Figure 4: Phase angle φ(ω) of one asphalt concrete mixture (PG64-28) (Kutay, 2008) By regression, the mastercurve can be mathematically expressed as a sigmoidal function with four parameters (a, b, c, d) obtained by data fitting. log( E * ) = a + b 1+ e −c −d *log ( wR ) (1) where: ωR is the reduced frequency at the reference temperature. Each parameter in the sigmoidal function has a special physical meaning: ‘a’ represents minimum modulus values at low reference frequency; ‘a + b’ indicates maximum modulus values at high reference frequency; ‘c’ shows the horizontal position of the turning point; and ‘d’ influences the steepness of the function (rate of change between minimum and maximum 11 modulus values) (Kim, 2008). They are illustrated in Figure 5. The physical meaning of each parameter can guide the backcalculation program during convergence. Log |E*| d (in c re a s e ) (a + b ) C (N e g ) C (P o s ) a Log R e d u c e d F re q u e n c y Figure 5: Physical meaning of each parameter on the sigmoidal function (Kim, 2008) Relaxation modulus E(t) is another fundamental material property that describes the viscoelastic behavior of AC. E(t) can be computed directly from |E*| mastercurve using an interconversion technique as described in the next subsection. It can also be characterized using a sigmoidal function with 4 parameters, as shown in Eq. (2). The relaxation modulus for the same mixture above (PG 64-28) at the same reference temperature is shown in Figure 6. It shows that the parameter group (a, b, c, d) determines the material properties of the AC mixture. log(Et ) = a + b 1+ ec + d *log (t R ) where: tR is the reduced time at the reference temperature. 12 1 (2) Relaxiation Modulus E(t) Relaxation Modulus (kPa) 1.0E+08 Experiment Simulated 1.0E+07 1.0E+06 1.0E+05 1.0E+04 1.0E+03 1.E-14 1.E-11 1.E-08 1.E-05 1.E-02 1.E+01 1.E+04 1.E+07 1.E+10 1.E+13 1.E+16 Reduced time (second) Figure 6: Relaxation modulus |E(t)| of one asphalt concrete mixture (PG64-28) (Kutay, 2008) 2.3 Interconversion between E(t) and |E*| for AC Mixture Mastercurve (|E*|), relaxation modulus E(t), and creep compliance D(t) are equivalent for viscoelastic materials, i.e., there is a unique relationship between |E*|, E(t) and D(t). The interconversion between |E*| and E(t) is discussed in this section, and the interconversion between E(t) and D(t) will be discussed in the following section. In viscoelasticity theory, E(t) and D(t) are calculated from experimental results of mastercurve |E*|, and are often expressed as the Prony series. It is important to note that a single polynomial model cannot be used for fitting the entire mastercurve, because the polynomial swing at low and high temperatures causes irrational modulus value predictions when 13 extrapolating outside the range of data (Kim, 2008). Typically, E(t) curve is represented using the generalized Maxwell model, shown in Figure 7. E o E 1 E 2 ?2 E n s ?1 ?n s Figure 7: The generalized Maxwell model E*(ω) is mathematically expressed as the storage modulus (E’ (ω)) and loss modulus (E” (ω)). The relationship between them is given in Eq. (3): 2 N ⎧ ( ωTi ) ⎪ E ' ( ω) = E * ( ω) *cos ( φ ) = E 0 + ∑ E i 2 1 + ( ωTi ) i =1 ⎪ ⎨ N ωTi ⎪E" ω = E * ω *sin φ = E ( ) ( ) ∑ i 2 ⎪ ( ) 1 + ( ωTi ) i =1 ⎩ where: Ti = (3) ηi Ei , also called relaxation time. If all the coefficients of Ei and Ti are known, E*(ω) can be directly calculated by Eq. (3). On the other hand, if E*(ω) and φ(ω) are given, then Ti will be assumed for each component of 14 -15 the general Maxwell model, typically ranging from 10 15 to 10 seconds, and a group of linear equations is set up to solve for Ei. Usually there are more equations than unknowns, and the technology of least square root of error (LSRE) may be used to calculate Ei. If the coefficients of Ei are found, the relaxation modulus is mathematically expressed as: N E (t ) = E0 + ∑ Ei e − t Ti (4) i =1 2.4 Interconversion between E(t) and D(t) for AC Mixture The mathematical relationship between E(t) and D(t) in the time domain is given by the following integral (Kim, 2008): t ∫ E (t − τ ) 0 dD(τ ) dτ ≡ 1 dτ (5) Solving the integral for interconversion can be done using a numerical approach. This requires that the integral be divided into a large number of time segments, which is consistent with the relaxation modulus from the generalized Maxwell model. This can be done by using the Prony series form of the E(t) and D(t) functions. Given {τ j {ρi , Ei (i = 1,2,..., m ) and E0 } or , D j ( j = 1,2,..., m ) and D0 } and the target time constants, the unknown set of constants can be determined through a system of linear algebraic equations. For example, creep compliance in its Prony series form, Dj (j=1,2,…,n), can be determined from the relaxation 15 modulus E(t) as follows (Kim, 2008): [A]{D} = {B} Or: (6) Akj D j = Bk (j=1,2,…,n; k=1,2,…,p) where: t ⎧ ⎛ − k ⎞ ⎪E0 ⎜1 − e τj ⎟ + ⎟ ⎪ ⎜ ⎝ ⎠ ⎪ ⎪ m ρ E ⎛ − tk − tτk ⎞ ⎪∑ i i ⎜ e ρi − e j ⎟ ( ρi ≠ τ j ) ⎜ ⎟ ⎪ i=1 ρi − τ j ⎝ ⎪ ⎠ Akj = ⎨ t − k ⎞ ⎪ ⎛ τ E0 ⎜1 − e j ⎟ + ⎪ ⎟ ⎪ ⎜ ⎝ ⎠ ⎪ t m ⎛ − ρk ⎞ ⎪ t k Ei ⎜ e i ⎟ ( ρi = τ j ) ⎪∑ ⎜ ⎟ i =1 τ j ⎝ ⎪ ⎠ ⎩ 16 (7) m Bk = 1 − E0 + ∑ Ei e i =1 − tk ρi m E0 + ∑ Ei i =1 The symbol tk (k=1,2,…,p) represents a discrete time corresponding to the upper limit of integration in Eq. (5). Once the model constants Dj, and τj are found, the function D(t) can be obtained in its Prony series form. Similarly, E(t) can be calculated from D(t) by solving for the unknown constants of the Prony series representation of E(t). An example of such calculation is given here for an AC mixture. The source and target Prony series coefficients of a specific AC mixture are shown in Table 1, and the corresponding fitted data are plotted in Figure 8. 17 Table 1. Prony series coefficients for E(t) and D(t) for a typical AC mixture Relaxation Modulus (MPa) E∞ ρi(s) 1.E-10 1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07 1.E+08 1.E+09 1.E+10 1.E+11 1.E+12 1.E+13 Creep Compliance (1/MPa) D0 9.8 Prony Coefficients Ei (MPa) 1.14E+02 6.81E+01 1.81E+02 3.67E+02 6.27E+02 1.02E+03 1.92E+03 2.92E+03 3.82E+03 3.76E+03 2.38E+03 1.28E+03 4.93E+02 1.66E+02 5.59E+01 2.08E+01 8.64E+00 4.11E+00 2.13E+00 9.98E-01 8.00E-01 9.92E-02 3.25E-01 6.99E-02 τi(S) 1.E-10 1.E-09 1.E-08 1.E-07 1.E-06 1.E-05 1.E-04 1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02 1.E+03 1.E+04 1.E+05 1.E+06 1.E+07 1.E+08 1.E+09 1.E+10 1.E+11 1.E+12 1.E+13 18 5.20E-08 Di (1/MPa) 3.87E-11 3.03E-10 5.85E-10 7.52E-10 1.97E-09 3.42E-09 6.75E-09 1.35E-08 2.82E-08 6.47E-08 1.48E-07 3.48E-07 1.03E-06 3.02E-06 8.43E-06 1.63E-05 2.05E-05 1.76E-05 1.27E-05 8.80E-06 4.90E-06 3.64E-06 1.27E-06 2.11E-06 Relaxation Modulus (MPa) E(t) (MPa) D(t) (MPa) 1.E+04 1.E-04 8.E-05 1.E+03 6.E-05 1.E+02 4.E-05 1.E+01 2.E-05 Creep Compliance (1/MPa) 1.E-04 1.E+05 1.E+00 0.E+00 1.E-15 1.E-11 1.E-07 1.E-03 1.E+01 1.E+05 1.E+09 1.E+13 1.E+17 Reference Time (s) (S) Figure 8: E(t) and D(t) from interconversion for a typical AC mixture 2.5 Motivation for the Study Current backcalculation methods, both static and dynamic methods, ignore the viscoelastic behavior of AC layer. Assuming AC as a linear elastic material is no longer a valid and appropriate assumption. The new Mechanistic Empirical Pavement Design Guide (ME-PDG) utilizes the |E*| mastercurve of AC layer. Therefore, there is a growing need for estimating entire |E*| mastercurve of AC. There are three major challenges for the VE backcalculation. First, the dynamic response of the pavement under field FWD test should be estimated correctly. Second, the elastic modulus of the base and subgrade layer should be as accurate as possible. Third, the AC material relaxation 19 modulus E(t), or the dynamic modulus mastercurve |E*| should be expressed as a function instead of one constant value. The pavement response under the FWD test is the response of a dynamic system after load impulse, and the phenomenon of wave propagation, as evidenced from the time delay of each sensor. Additionally, backcalculation algorithms are typically more sensitive to the elastic modulus of the base and subgrade layer than the AC layer. This indicates that the accuracy of the modulus of base and subgrade layer significantly affect the accuracy of the AC layer. Based on the wave propagation theory, the elastic modulus of the semi-infinite subgrade layer (Esubgrade) can be calculated from the time delay of each sensor, if the interface of different physical layer is insignificant for wave propagation. This forward calculated modulus can be used as the seed value of Esubgrade in the VE backcalculation program, and it accelerates the convergence of backcalculation. A detailed discussion will be given in Chapter 3. In AC mixture modeling, the generalized Maxwell model is often employed to illustrate the viscous property of the material. It is the constitutive equation in the time domain. In order to capture the material constitutive behavior, usually there are more than 30 groups of Maxwell model as a whole in the physical model. If the backcalculation is done directly to calibrate the generalized Maxwell model, there would be more than 30 variables in the model, and it is almost impossible for backcalculation. If the sigmoidal function is chosen, the backcalculation is used to calculate the parameters of the sigmoidal function (a, b, c, d) in either E(t) or |E*|, although the VE forward calculation is done in time domain. The detailed explanation of the VE forward solution is presented in Chapter 4, and the backcalculation algorithm is given in Chapter 5. The objective of this study is to propose a methodology that is robust in the backcalculation 20 of both AC viscoelastic modulus and the elastic modulus of base and subgrade layer, based on viscoelasticity theory. More specifically, the thickness, the Poisson’s Ratio and mass density of each layer are accurately given as input in backcalculation, only the four parameters (a, b, c, d) of the sigmodial function of E(t) or |E*|, and the elastic modulus of the base (Ebase) and subgrade (Esubgrade) layers need to be identified. 21 Chapter 3. Estimating the Subgrade Elastic Modulus using Wave Propagation Method 3.1 Introduction The Spectral Analysis of Surface Waves (SASW) method is a dynamic non-destructive test method for determining the shear wave velocity and shear modulus of soils in-situ. The test procedure was developed in the 1980s at the University of Texas, Austin, and included three steps: data acquisition, dispersion analysis, and inversion. The test is done at different frequencies, and the shear modulus is calculated by the frequency and the phase angle difference between sensors at different locations, through the dispersion curve, which is a plot of phase velocity vs. frequency. A detailed procedure of this test can be found in Ganji et al. (1998). A similar method to estimate the elastic modulus of the subgrade layer is proposed in this chapter. It is based on the spectral element method and the wave propagation theory. The steps of the proposed procedure are as follows: Assume reasonable material properties for the subgrade layer, including mass density and Poisson’s ratio; Obtain the wave propagation velocity, or Rayleigh wave velocity (VR), based on the location and the occurrence time of the peak deflection for each sensor in the FWD test; Obtain the k-value, or the ratio of Rayleigh wave velocity (VR) and shear wave velocity (Vs); 22 Calculate the shear modulus of the subgrade layer, using Vs and estimated mass density; Calculate the elastic modulus, based on shear modulus and Poisson’s ratio. The mass density is often estimated by the level of compaction during pavement construction, based on the engineer’s experience; while the Poisson’s ratio is often estimated by the type of the soil material, i.e., clay, sand, or bedrock. The VR is calculated by the distance between sensors divided by the difference of the occurrence time of the peak deflection for each sensor, from the time-history in the FWD test. The only challenge is to find the ratio between VR and Vs, or the k-value. For a plane wave expressed in Cartesian coordinates, the ratio (K) of the velocity of Rayleigh Wave (VR) to the velocity of Shear Wave (VS) is only a function of Poisson’s ratio. K is the root of the following equation (Richart et al., 1970): ( ) ( ) K 6 − 8K 4 + 24 − 16α 2 K 2 + 16 α 2 − 1 = 0 , where: α2 = (8) 1 − 2ν 2 − 2ν The FWD test generates axial-symmetric waves instead of plane wave, above result may not be valid, and a proposed method to estimate k-value is briefly discussed in the following section. 3.2 Proposed Method to Estimate k-value First, the theoretical formulation for the spectral element method is briefly discussed. Second, the dominant frequency in the FWD test is briefly introduced. Third, numerical examples are presented to calculate the k-value for specific subgrade site at specific frequencies. 23 Fourth, sensitivity analysis is presented for k-value for a range of elastic moduli, with different Poisson’s ratios of the subgrade layer, under cyclic loading of different frequencies. 3.2.1 Theoretical Formulation for Spectral Element Method There are two types of waves in an infinite, homogeneous, isotropic elastic medium: waves of dilatation and waves of distortion. In a half-space medium, a third wave, the Rayleigh wave, is found as a solution for the equations of motion. It corresponds to a wave whose motion is confined to a zone near the boundary of the half-space (Richart et al., 1970). In a cylindrical coordinate system, assuming a wave traveling in the radial direction, particle displacement will be independent of the angular direction. The equations of motion of an isotropic linear elastic material can be expressed in terms of the displacements by use of Navier’s equations (Al-Khoury, 2001), as: .. ( λ + μ ) ∇∇.u + μ∇ 2 u = ρu , (9) The vector u represents the displacement components of the material, and ρ is the mass density of the material. ∇ indicates a vector differential operator, ∇ . u is the divergence of u, and ∇ 2 . u is the Laplace operator of u. For axial symmetry, ∇ 2 ∂2 1 ∂ ∂2 + ∇ = 2+ r ∂r ∂z 2 ∂r 2 is expressed as: (10) where λ and μ are the Lame Constants expressed as: λ= νE , 1 + ν )(1 − 2ν ) ( 24 (11) μ=G= E , 2 (1 + ν ) (12) where E is Young’s modulus and ν is Poisson’s ratio. In the Helmholtz decomposition, the displacement vector u is expressed as the sum of the gradients of a scalar potential φ and the curl of a vector potential ψ as: u= ∇ φ+ ∇ × ψ (13) In view of the axial symmetry, the vector ψ only has a component in the θ-direction, which is ψ=ψ eθ . Denoting the displacement components in the r and z directions by u and w, respectively, the relations between the displacement components and the potentials are shown below (Yeh, 1989): ∂ϕ ∂ψ ⎧ u= − ⎪ ∂r ∂z ⎪ ⎨ ⎪ w = ∂ϕ + 1 ∂ ( rψ ) ⎪ ∂z r ∂r ⎩ (14) Because of Eq. (9) and (13), the above potentials satisfy the following axial symmetric wave equations: ⎧ ∂ 2 ϕ 1 ∂ϕ ∂ 2 ϕ 1 ∂ 2 ϕ + 2 = 2 2 ⎪ 2 + r ∂r ∂z cp ∂t ⎪ ∂r , ⎨ 2 2 2 ⎪ ∂ ψ + 1 ∂ψ + ∂ ψ − ψ = 1 ∂ ψ ⎪ ∂r 2 r ∂r ∂z 2 r 2 c2 ∂t 2 s ⎩ 25 (15) where the constants cp and cs are defined as: ⎧ ⎪cp = ⎪ ⎨ ⎪c = ⎪ s ⎩ λ + 2μ ρ μ ρ (16) Based on Eq. (15), there are two types of waves: compression wave (P) and Shear wave (S) respectively. Because of isotropy, these waves are uncoupled from each other. Considering the harmonic loading case with frequency ω, the potentials can be solved by the method of separation of variables with some arbitrary constant k, assuming the format of the potentials are (Maurice et al., 1957): ⎧ ϕ = S ( kr ) T ( z ) ⎪ , ⎨ ⎪ψ = S1 ( kr ) T1 ( z ) ⎩ (17) provided that the function S and T satisfy the equations: ⎧ d 2S 1 dS ⎛ ω2 ⎞ 2 + ⎜ 2 − υ ⎟S = 0 ⎪ 2+ ⎟ r dr ⎜ c p ⎪ dr ⎝ ⎠ , ⎨ ⎪ d 2T 2 2 ⎪ 2 +υ T =0 ⎩ dz 26 (18) υ = 2 where ω2 c 2 p 2 − k 2 = kα − k 2 υ' = with 2 Similar equations hold for S1, T1, equation of Eq. (18) has a particular solution: ω2 cs2 2 − k 2 = kβ − k 2 . The second T = e − ivz , which remains finite as z -> +∞ . The first equation of Eq. (18) is a form of Bessel’s equation, which has the Bessel function solution J0(kr), which remains a non-zero value at r = 0. Thus, two particular solutions are obtained: ⎧ ϕ = Ae − ivz J 0 ( kr ) ⎪ , ⎨ − iv 'z ⎪ψ = Be J1 ( kr ) ⎩ (19) A and B are constants determined by the boundary conditions. The solutions vanish as z—> ∞ and also vanish as r—> ∞, because of the property of the Bessel function J0(kr). The displacements can be expressed as: ( ( ) ) ⎧u = − Ake− ivz + iv' Be −iv 'z J1 (kr ) ⎨ w = − iAve−ivz + Bke−iv 'z J 0 (kr ) ⎩ The stress equations in cylindrical coordinates are: 27 (20) ⎧ ⎛ ∂u ∂w ⎞ τzr = μ ⎜ + ⎟ ⎪ ∂z ∂r ⎠ ⎪ ⎝ ⎨ ⎪σ = λ + 2μ ∂w + λ ∂ ( ru ) ) ⎪ zz ( ∂z r ∂r ⎩ (21) By replacing u and w, the following equation is obtained at the surface z = 0: ( ) 2 ⎧ τrz = ⎡ 2ikυAu + k β − 2k 2 uB⎤ J1 ( kr ) ⎪ ⎣ ⎦ ⎨ 2 2 ⎡ ⎤ ⎪σzz = ⎣ − k β − 2k uA − 2ikυ 'uB⎦ J 0 ( kr ) ⎩ ( ) (22) To make the problem discrete, some boundary conditions in the radial direction are introduced. At the radial boundary r = R (far away from the source), the amplification of the oscillation is considered to vanish. Only the vertical displacement w is important in the FWD test, so the horizontal component will be ignored in the future. These considerations can be implemented by the infinitely many positive roots αm of the J0 function, as km= αm/R (Al-Koury et al., 2001). For FWD test loading, if the maximum pressure is P, the appropriate boundary conditions are (Maurice et al., 1957): ⎧ τrz z =0 = 0 ⎪ ⎨ ⎪σzz z =0 = PJ 0 ( kr ) ⎩ By combining Eq. (22) and Eq. (23), the coefficients A and B can be solved. 28 (23) ( ) 2 ⎧ k β − 2k 2 P ⎪A = F(k) μ ⎪ , ⎨ ⎪ B = − 2ikυ ' P ⎪ F(k) μ ⎩ ( F (k ) = 2k − k β where: 2 ) 2 2 (24) − 4k 2υυ ' , which is called Rayleigh’s function (Maurice et al, 1957). The expression of the wave propagation includes Rayleigh’s function, so this wave is called Rayleigh wave (Xiang, 2009). Fig. 9 shows different kinds of wave propagation on the surface of an ideal medium after a load impulse. It illustrates the vibration magnitude by different wave modes, and there is a tiny vibration for compression wave (P) just after the load impulse. However, the vibration of shear wave (S) and Rayleigh wave (R) happen almost simultaneously, especially if the location is near to the loading point. The time difference of S wave and R wave is insignificant as the velocities of S wave and R wave are almost identical. Thus, the distance between two adjust peaks in vertical displacement is considered the wave length. However, wave length varies by location or propagation time, and the velocity is hard to identify, because the interaction of shear wave and Rayleigh wave affects the peak displacement. 29 t t Figure 9: Wave system from surface point source in ideal medium (a) Effect of different wave modes on horizontal displacement (b) Effect of different wave modes on vertical displacement 1 (c) The combined vibration for particle at Point ○ (Richart et al., 1970) Typically, the S wave or R wave velocity is around 150 m/s (500 ft/s) for subgrade layers, and the effective frequency in FWD test is less than 50 Hz, i.e., the S wave length is more than 3 m (10 ft), which is beyond the range of the sensor distribution in the FWD test. Thus, the occurrence time of peak deflection for each sensor is considered, noted as the time delay, rather than the deflection basin, and by the difference of the occurrence time, the velocity of R wave is identified. 30 3.2.2 Dominant Frequency of the FWD Test Eq. (24) indicates the Rayleigh wave propagation velocity is a function of frequency, while the shear wave velocity is independent to frequency. K-value should be dependent on the loading frequency. Thus, the equivalent frequency of the FWD test, or the dominate frequency range in the FWD test should be identified. The equivalent frequency of a loading history can be identified by: (a) the centroid of the area formed by Fourier Series; (b) the centriod of the area formed by the power spectral density (PSD) in the frequency domain; (c) the central frequency method from earthquake engineering, as defined below. In method (a), the coefficient of Fourier series is found first, and then the centroid of the coefficient corresponding to the frequency is obtained as the equivalent frequency. Method (b) is well-known as Parseval’s theorem or the principle of energy conservation in time and frequency domains. The power spectral density function (G (ω)) is calculated by coefficients of Fourier series first, and the centroid of the PSD is called the equivalent frequency. In method (c), the equivalent frequency is calculated as: ω= λ2 λ0 where λi =∫ ωN 0 w i G (ω )dω . For a typical FWD loading history, the loading duration is 35 ms with a haversine function. It is assumed that the rest time is 1 s with a sampling interval of 1ms, which only affects accuracy by a numerical error of 2 - 3 %. The equivalent frequency is calculated by these methods as: (a) f = 17.23 Hz. (b) f = 13.25Hz. (c) f = 16.38 Hz. Considering the actual loading is not a haversine function, and the loading time may differ from 35 ms, the equivalent frequencies for the three methods are listed in Table. 2. It shows the typical range of the dominant frequency in the FWD test is 5 - 25 Hz. 31 Table 2: The equivalent frequency for different load duration times Duration time (ms) 25 30 35 40 45 50 Method (a) 24.09 Hz 20.10 Hz 17.23 Hz 15.09 Hz 13.42 Hz 12.08 Hz Method (b) 18.56 Hz 15.47 Hz 13.25 Hz 11.60 Hz 10.31 Hz 9.28 Hz Method (c) 22.93 Hz 19.11 Hz 16.38 Hz 14.33 Hz 12.74 Hz 11.46 Hz 3.2.3 Numerical Example to Calculate k-value Eq. (24) indicates the VR under cyclic loading depends on the loading frequency, elastic modulus and Poisson’s ratio. One example of wave propagation is illustrated for the hypothetical profile which only includes the half-space subgrade layer. The cyclic loading frequency is 5 HZ, and the material properties are: elastic modulus = 68.95 MPa, and Poisson’s ratio = 0.45. The phase angle of each point at the surface is shown in Fig. 10 for t = 0 second. Phase angle at the surface of the subgrade layer at t=0 (E = 68.95 Mpa, v = 0.45, f = 5 Hz) Phase angle (o) 90 60 30 0 -30 -60 -90 0 10 20 30 Location (m) 40 50 Figure 10: Phase angle of each point at the surface for a specific example (t =0 s) The wavelength is defined as the spatial period of the phase angle, or the distance over 32 which the phase angle’s shape repeats. The wavelength in the above example is around 18.5 m, and the wave velocity (VR) is around 92.5 m/s, as the loading frequency is 5 Hz. The mass density (ρ) of the subgrade layer is assumed as 2000 kg/m3, and the shear modulus (G) is 23.78 MPa, calculated from elastic modulus and Poisson’s ratio (Eq. 12). The shear wave velocity is 109 m/s, from Eq. (25). G ρ Vs = (25) The k-value is 0.8486, which is VR divided by Vs (Eq. 26). k= VR Vs (26) The k-value only depends on the Poisson’s ratio in plane wave, independent of loading frequency and elastic modulus of the subgrade layer. Similar conclusion is expected for the axial symmetric FWD test. Cyclic loading is applied in the SASW method; however, only load impulse is applied in the FWD test. The phase angle for each point at the surface of the subgrade layer can be replaced by the occurrence time of the peak deflection. 3.2.4 Sensitivity Analysis for k-value The typical range of the elastic modulus of subgrade is 35 - 280 MPa (5 to 40 ksi) (NCHRP, 2004). The dominant frequency in the FWD test is 5 to 25 Hz. Since the typical Poisson’s ratio ranges from 0.20 to 0.49, a sensitivity analysis for k-value with different combinations of elastic 33 modulus and frequency was conducted for possible values of Poisson’s ratio (Fig. 11 - 17). R S K−value (V /V ) 1 0.8 0.6 0.4 0.2 0 30 8 20 6 4 10 0 Frequency (Hz) 8 2 x 10 0 Elastic Modulus (MPa) Figure 11: Sensitivity analysis for k-value (ν= 0.20) 0.8 R S k−value (V /V ) 1 0.6 0.4 0.2 0 30 8 20 6 4 10 Frequency (Hz) 2 0 0 Elastic Modulus (MPa) Figure 12: Sensitivity analysis for k-value (ν= 0.25) 34 8 x 10 0.8 R S k−value (V /V ) 1 0.6 0.4 0.2 0 30 8 20 6 4 10 0 Frequency (Hz) 8 2 x 10 0 Elastic Modulus (MPa) Figure 13: Sensitivity analysis for k-value (ν= 0.30) 0.8 R S k−value (V /V ) 1 0.6 0.4 0.2 0 30 8 20 6 4 10 Frequency (Hz) 2 0 0 Elastic Modulus (MPa) Figure 14: Sensitivity analysis for k-value (ν= 0.35) 35 8 x 10 k−value (VR/VS) 1 0.8 0.6 0.4 0.2 0 30 8 20 6 4 10 0 Frequency (Hz) 8 2 x 10 0 Elastic Modulus (MPa) Figure 15: Sensitivity analysis for k-value (ν= 0.40) k−value (VR/VS) 1 0.8 0.6 0.4 0.2 0 30 8 20 6 4 10 Frequency (Hz) 2 0 0 Elastic Modulus (MPa) Figure 16: Sensitivity analysis for k-value (ν= 0.45) 36 8 x 10 k−value (VR/VS) 1 0.8 0.6 0.4 0.2 0 30 8 20 6 4 10 0 Frequency (Hz) 8 2 x 10 0 Elastic Modulus (MPa) Figure 17: Sensitivity analysis for k-value (ν= 0.49) Sensitivity analysis indicates the k-value is almost constant if the elastic modulus of the subgrade is less than 500 MPa (70 ksi); it is not sensitive to the frequency of the loading for the dominant range in the FWD test, but it is slightly affected by Poisson’s ratio. Thus, one average k-value could be used for each Poisson’s ratio. Table 3 shows the average k-value of all the entire sensitivity range as well as for the more practical range of moduli (5 to 40 ksi) and frequencies (5 to 25 Hz). Table 3: The average k-value for different cases of Poisson’s ratio Poisson’s ratio Average Entire range k-value Practical range COV Entire range (%) Practical range 0.20 0.7957 0.7955 1.63 0.83 0.25 0.8051 0.8056 1.58 0.76 0.30 0.8132 0.8140 1.65 0.94 37 0.35 0.8176 0.8189 1.89 1.22 0.40 0.8173 0.8197 2.34 1.78 0.45 0.8102 0.8154 3.32 2.77 0.49 0.7970 0.8073 5.48 4.66 Table 3 shows the average k-value does not significantly affected by Poisson’s ratio. It would be much better to use a single average k-value to simplify the problem; k=0.8109 is suggested. However, it should be mentioned that the k-value may be quite different if there is a stiff layer underneath, as the cases are not verified for subgrade with elastic modulus higher than 700 MPa (100 ksi). 3.3 Proposed Procedure to Estimate Subgrade Elastic Modulus Based on the result in Section 3.2, the following steps are applied to estimate the elastic modulus of the subgrade layer using k-value. 1. Estimate the Poisson’s ratio (ν) and mass density (ρ) of the subgrade layer. The Poisson’s ratio depends on the soil type, and it is well documented in the literature. The soil type, compaction level during construction, and water table level will affect the 3 mass density, ranging from 1500 to 2000 kg/m . However, the estimated elastic modulus is not significantly affected by the mass density, and the engineers’ experience will be helpful for the estimation. 2. Find the slope in the plot of time delay vs. location. Six sensors are located at different positions in the FWD test, and the occurrence time of peak deflection for each sensor is different. Time delay of each sensor is defined as the difference between the time of peak loading and the time of peak deflection. Time delay and locations can be plotted in one figure, and the slope (m) for the farther sensors is found by linear regression. One example of the load impulse and time-history for the FWD test are shown in Fig. 18. The test was conducted at LTPP (Long Term Pavement Performance) station 04-1036 in 1998. The plot of time delay vs. location is shown in Fig. 19. The slope in Fig.19 is 38 calculated as: m = 0.0069 s/m. FWD test history Load (kPa)/deflection (um) 800 load Sensor Sensor Sensor Sensor Sensor Sensor Sensor 700 600 500 400 300 1 2 3 4 5 6 7 200 100 0 -100 0 10 20 30 Time (ms) 40 50 60 Figure 18: Example of FWD load and deflection time histories at LTPP station 04-1036 in 1998 39 Time delay vs. Location 0.01 0.009 Time delay (ms) 0.008 0.007 0.006 y = 0.0069x - 0.0013 2 R = 0.9978 0.005 0.004 0.003 0.002 0.001 0 0.0 0.2 0.4 0.6 0.8 1.0 Location (m) 1.2 1.4 1.6 1.8 Figure 19: Time delay vs. location for the example of FWD test at LTPP station 04-1036 in 1998 3. Calculate the Rayleigh wave velocity (VR). The wave propagation velocity is the inverse of the slope (m) in the plot of time delay vs. location. In the current example, VR = 1/0.0069 = 144.9 m/s. 4. Calculate the Shear wave velocity (Vs). The Vs is VR divided by k-value. In this example, Vs = 144.9/0.8109 = 178.1 m/s. 5. Calculate the shear modulus (G) of the subgrade layer. The mass density is estimated in step 1, as ρ=1800 kg/m3. The relationship between G and Vs is shown in Eq. (27), which is similar to Eq. (25) in Section 3.2. 40 G = Vs2 * ρ (27) The shear modulus is calculated as: G= 57.1 MPa in the current example. 6. Calculate the elastic modulus (E) of the subgrade layer by Eq. (28), with assumed Poisson’s ratio (ν) in Step 1. E = 2 (1 + υ ) G (28) The elastic modulus in this example is E = 165.6 MPa, assuming ν = 0.45. For simplification, a flow chart for this procedure is presented in Fig. 20. Assume ρ and ν Find time delay and location for each sensor in FWD Plot time delay vs. location, as Fig. 16 Obtain the slope (m) for farther sensors Calculate Rayleigh wave velocity, VR=1/m Calculate shear wave velocity, VS=VR/k (k=0.8109) Calculate shear modulus, by Eq. (27) Calculate elastic modulus, by Eq. (28) Figure 20: Flow chart to estimate subgrade elastic modulus using k-value It is difficult to identify how many sensors should be used in step 2. Often there are 6 - 7 41 sensors in the FWD test, so the farther sensors would be sensors 4 - 6 or 5 - 7. There are interactions between interfaces of different layers, so the time delay for close sensors is influenced by the combined effect of different layers. Farther sensors are less affected by interactions, and it is preferable to obtain the m-value from multiple sensors. The slope between the farthest sensors is the dominant factor for m-value, and the time delay for other sensors 2 should be ignored if including them reduces the regression R to below 0.98. The proposed procedure uses the time delay of each sensor in the FWD test, while the SASW method uses the phase angle of each sensor. Although load impulse can be converted to summation of different magnitudes of cyclic loading at different frequencies by Fourier Transform, the phase angle for each frequency is very sensitive in converting. Furthermore, FWD time histories are typically truncated before they fully decay, leaving errors when using field measurements. The pavement response in the frequency domain is not as accurate as in the time domain. Therefore, the time delay instead of phase angle is used in the proposed method. 3.4 Verification of the Proposed Method No field data are available with all the information on the pavement profile, so only numerical examples are used to validate the proposed method. The time-history of the field FWD test is simulated by two well-known dynamic solutions: SAPSI and LAMDA. Both are briefly introduced first, and numerical examples of subgrade layer with and without ground water table are presented later, to illustrate the theoretical error of the proposed method. 3.4.1 SAPSI The forward program SAPSI (Chen, 1987) models the pavement structure as a system of 42 finite layers that are infinite in the horizontal direction and underlain by an elastic half-space with viscous boundaries. The finite layer solution is based on Kausel’s formulation (Kausel, 1981; Doyle, 1997) which subdivides the medium into discrete layers that have a linear displacement function in the vertical direction (finite element method with lumped mass formulation) and satisfy the wave equation in the horizontal direction (exact formulation). The solution is based on the premise that if the sublayer thickness is small relative to the wavelength of interest, it is possible to linearize the transcendental functions and reduce them to algebraic expressions. The materials are assumed to be isotropic and linearly elastic with hysteretic damping. Full interface bonding is assumed at the layer interfaces. The mass densities and elastic moduli change with depth, from layer to layer, but are assumed to be constant within each layer. The top layer represents the asphalt concrete surface, which can be modeled as a viscoelastic material by allowing its (complex) modulus to be a function of frequency. 3.4.2 LAMDA Al-Khoury et al. (2001) developed an efficient axial-symmetric forward solution, called LAMDA, for the dynamic analysis of flexible pavements using the spectral element method for the simulation of wave propagation in layered systems. The spectral element method developed by Doyle (1997) combines elegantly the exact solution of wave motions with the finite element organization of the system matrices. The system is solved by double summation over the involved frequencies and the wave numbers (Rizzi and Doyle, 1992). The double summation approach using Fourier series is computationally advantageous over Kausel’s formulation, which relies on the numerical evaluation of integrals between zero and infinity. This type of integration involves singularities if the system has no damping or very sharp peaks for small damping, and it requires considerable computational time and capacity. The mass distribution is modeled exactly 43 and hence only one element is sufficient to describe a whole layer without the need for subdivisions. This makes the resulting system of dynamic equations very small and hence computationally efficient. 3.4.3 Case of Subgrade without Ground Water Table One specific numerical example of a pavement without ground water table (GWT) subjected to the FWD test is analyzed in SAPSI, and simulated by the Finite Element Software ABAQUS. Table 4 presents a summary of the pavement parameters. Table 4: Basic information of a three-layer pavement structure without GWT Physical Layer AC Base Subgrade Elastic Modulus (MPa (ksi)) Experimental data 150 (21.8) 100 (14.5) Poisson’s ratio 0.35 0.35 0.45 Mass density 3 (kg/m ) 2300 2000 1500 Thickness (m (in)) 0.1 (4) 0.3 (12) Infinity The FWD load is assumed as a haversine function with duration time of 50 ms, and the maximum pressure is 707 kPa with a loading plate radius of 0.15 m. In SAPSI, 3% damping ratio is applied for the subgrade layer, as it is a typical value for clay and can eliminate the free vibration of the system after dynamic loading. The time delay of each sensor is given in Table 5. Table 5: Time delay of each sensor simulated by SAPSI Sensor 1 Physical distance (m) 0.00 Time delay (s) 0.0039 2 0.41 0.0049 3 0.67 0.0067 4 0.93 0.0089 5 1.19 0.011 6 1.44 0.0129 7 1.70 0.0148 Since the sampling interval is 0.1 ms in the numerical simulation, the error from calculation may be relatively high. For the elastic modulus of the subgrade layer, the time delay and the location of sensors 4 - 7 are used given the thickness of the pavement structure is 0.4 m (16 in). 44 The linear regression for the time delay of these sensors is shown in Fig. 21. Time delay VS. Location 0.016 y = 0.007656 x + 0.001832 2 R = 0.999474 Time delay (s) 0.014 0.012 0.01 0.008 0.006 0.004 0.002 0 0 0.2 0.4 0.6 0.8 1 1.2 Sensor Location (m) 1.4 1.6 1.8 Figure 21: Time delay vs. location for a numerical example of FWD test (without GWT) From the statistical analysis of these sensors, the slope is obtained directly by SAPSI, as m = 0.007656 s/m. The Rayleigh wave velocity is the inverse of the slope m, i.e., the Rayleigh wave velocity VR = 1/m = 130.6 m/s. The Poisson’s ratio of the subgrade layer is assumed as υ = 0.45 , which is a typical value for clay. The average k-value equals 0.8109. Thus, Vs= VR/k = 160.5 m/s. The mass density of the subgrade layer is assumed as ρ = 1500 kg / m3 , so the shear modulus of the subgrade layer is calculated in Eq. (18) as G= 38.6 MPa. From elasticity theory, the elastic modulus E is calculated in Eq. (14) as E= 112.1 MPa. Given the input parameters of the 45 pavement structure (Table 3), this value of E is relatively accurate with an error of 12.1%. This level of error is acceptable in pavement engineering. 3.4.4 Case of Subgrade with GWT One specific numerical example of a pavement with GWT during the FWD test is analyzed in SAPSI, with GWT 2 ft below the surface of the subgrade layer. The material properties of the saturated subgrade layer are calculated by the assumption that the shear wave velocity is the same as that in the dry subgrade layer, while the compression wave velocity equals the sound propagation velocity in water (1450 m/s). Table 6 gives a summary of the parameters of the pavement profile with GWT. Table 6: Basic information for a three-layer pavement structure with GWT Elastic Modulus (MPa (ksi)) AC Experimental data Base 108 (15.0) Dry subgrade 34.5 (5.0) Saturated subgrade 42.9 (6.2) Physical Layer Poisson’s ratio 0.35 0.30 0.45 0.4979 Mass density 3 (kg/m ) 2300 2000 1500 1682 Thickness (m (in)) 0.15 (6) 0.3 (12) 0.6 (24) Infinity The FWD load is assumed as a haversine function with duration time of 35 ms, and the maximum pressure is 707 kPa with a loading plate radius of 0.15 m. In LAMDA, 3% damping ratio is applied for the subgrade layer, as it is a typical value for clay and can eliminate the free vibration of the system after dynamic loading. The simulated time-history of the field FWD test is shown in Fig. 22. 46 Deflection history with GWT (2ft) 6.E-04 sensor 1 sensor 3 Deflection (m) sensor 6 sensor 7 4.E-04 sensor 4 sensor 5 5.E-04 sensor 2 sensor 8 3.E-04 2.E-04 1.E-04 0.E+00 -1.E-04 0.01 0.02 0.03 0.04 0.05 0.06 time (sec) 0.07 0.08 0.09 Figure 22: LAMDA simulated time-history due to the FWD test (with GWT) The time delay of each sensor is given in Table 7. Table 7: Time delay of each sensor by LAMDA Sensor 1 Distance (m) 0.00 Delay (s) 0.0051 2 0.20 0.0055 3 0.30 0.0058 4 0.46 0.0065 5 0.61 0.0074 6 0.91 0.01 7 1.22 0.0137 8 1.52 0.0175 Since the sampling interval is 0.1 ms in numerical simulation, the error from calculation may be relatively high. For the elastic modulus of the subgrade layer, the time delay and the location of sensors 6 - 8 are considered, given the thickness of the pavement structure 0.45 m (18 in). The linear regression for the time delay for these sensors is plotted in Fig. 23. 47 Time delay vs. location (GWT=2ft) Time Delay (s) 0.02 y = 1.2303E-02x - 1.2667E-03 0.015 R2 = 9.9994E-01 0.01 0.005 0 0 0.5 1 1.5 2 Location (m) Figure 23: LAMDA simulated time-delay for the FWD test (with GWT) From the statistical analysis of these sensors, the slope is obtained directly by LAMDA, as m = 0.00123 s/m. The Rayleigh wave velocity is the inverse of the slope m, i.e., the Rayleigh wave velocity VR = 1/m = 81.3 m/s. The Poisson’s ratio of the subgrade layer is assumed as υ = 0.45 , which is a typical value for clay. Using the average k-value of 0.8109, we obtain Vs = VR/k = 99.9 m/s. The mass density of the subgrade layer is assumed as ρ = 1500 kg / m3 , so the shear modulus of the subgrade layer is calculated in Eq. (25) as G = 15.0 MPa. From elasticity theory, the elastic modulus E is calculated in Eq. (11) as E = 43.4 MPa. Given the input parameters of the pavement structure (Table 6), this value of E is almost exactly the elastic modulus of the saturated subgrade layer, with an error of 1.0%. Although the estimated modulus is higher than the modulus of the dry 48 subgrade layer, with an error of about 27.6%, this estimated value is commonly accepted as reasonable in pavement engineering. These two numerical examples show that the proposed method to estimate elastic modulus of the subgrade layer is valid in both cases (without and with GWT), and the validation is checked by field FWD time-histories from the Long-Term Pavement Performance (LTPP) database, although the actual elastic modulus of the subgrade layer is unknown. 3.5 Validating the Proposed Method Using Case Studies from LTPP Three cases are chosen from the LTPP database in this section, and their pavement profiles and time-histories are shown in each sub-section. Two different methods are used to estimate the modulus of the subgrade layer. One is calculated by the most common backcalculation software, MODCOMP5, which is based on minimizing the root mean squared error (RMSE) of the deflection basin. The other is the proposed method using Rayleigh wave velocity, which is based on the time delay of each sensor. Since the proposed method works very well in numerical examples, it is expected that the estimation of the elastic modulus from Rayleigh wave is better than that from the backcalculation software MODCOMP5. 3.5.1 Example of FWD Test at LTPP 04-1036 Station in 1998 One example of a FWD test at 04-1036 station (Arizona) in 1998 was selected for analysis, and the FWD test load and deflection histories are shown in Fig. 18 in section 3.3. It illustrates that the deflection histories yields information on viscoelasticity. There are several overlays above the original pavement; however, they are combined into one layer as they are all thin AC mixtures. The maximum pressure during the FWD test is 749 kPa (108.63 psi). 49 The time delay and maximum deflection of each sensor for this example are summarized in Table 8. Note that the measured time delay values are not as accurate because of the sampling time interval being too coarse (0.4 ms). Table 8: Time delay and peak deflection of each sensor due to the FWD test at station 04-1036 in 1998 Sensor Distance (in) Time delay (ms) Peak deflection (mils) 1 0.00 1.6 21.69 2 8.0 1.6 17.05 3 12.0 2.4 13.86 4 18.0 2.8 9.13 5 24.0 2.8 6.50 6 36.0 5.2 3.35 7 60.0 9.2 1.65 Using the wave velocity method proposed in this chapter, the Rayleigh wave velocity is 144.2 m/s, and the elastic modulus is 156.9 MPa (22.7 ksi), assuming the mass density of the 3 subgrade layer is 1800 kg/m . By MODCOMP5, the elastic modulus of each layer can be calculated, given the pavement structural profile. The pavement structure information and the back-calculated results from MODCOMP 5 are shown in Table 9. Table 9: Structural information and MODCOMP5 backcalculated modulus at LTPP station 04-1036 in 1998 Physical Thickness Layer (m (in)) AC 0.122 (4.8) Base (GB) 0.472 (18.6) Subgrade (SS) Infinity Poisson’s ratio 0.35 0.30 0.45 Mass density 3 (kg/m ) 2300 2000 1800 Elastic Modulus (MPa (ksi)) 4771 (692) 86.9 (12.6) 222.7 (32.3) The deflection basins from the FWD test and the backcalculation result by MODCOMP5 are shown in Fig. 24. It is clear that the deflection basin is simulated reasonably well, and the deflections of close sensors match almost exactly. Hence, the backcalculation program converges, although the result may be unrealistic. 50 Deflection Basin Maximum Deflection (mils) 0 5 ModComp FWD test 10 15 20 25 0 10 20 30 40 Sensor Location (in) Radius distance(in) 50 60 70 Figure 24: Deflection basin from FWD test and MODCOMP5 Simulation at LTPP station 04-1036 in 1998 From the LTPP database, the resilient modulus of the subgrade layer is tested at different confining pressures and different axial loads, and the summary of the test results is given in Table TST_UG07_SS07_WKSHT_SUM in the Test Module of the LTPP database. Typically, the confinement of the subgrade layer is 41.3 kPa (6 psi) in design (Federal Highway Administration, 1996), and the corresponding resilient modulus can be calculated as 102 MPa (14.8 ksi) by interpolation. The elastic modulus from the backcalculation by MODCOMP5 is 222.7 MPa (32.3 ksi), and the modulus from the wave velocity method proposed here is 156.9 MPa (22.7 ksi). Thus, the difference between the experimental data and the backcalculation result is roughly 118.3%, while the difference between the experimental data and the result from the proposed wave velocity 51 method is 53.4%. These values are summarized in Table 10. Table 10: Elastic modulus of the subgrade layer at LTPP station 04-1036 in 1998 Lab_ measured MR @ 6psi confinement 102 MPa (14.8 ksi) Predicted by k-value (Ek-value) 156.9 MPa (22.7 ksi) Difference MODCOMP5 Difference backcalculation between MR and between MR Ek-value (EMODCOMP) and EMODCOMP 222.7 MPa 53.4 % 118.3 % (32.3 ksi) 3.5.2 Example of FWD Test at LTPP 32-0101 Station in 1996 Another example is chosen from LTPP station 32-0101 in 1996. The FWD test load and deflection histories are shown in Fig. 25. There are 5 relatively thick physical layers. For the backcalculation analysis, they are separated into each physical layer first, and then they are combined together if the results from backcalculation are unrealistic. The test was done in Nevada in early winter. The maximum test pressure was 454 kPa (65.85 psi). 52 FWD test history Load (kPa)/deflection (um) 500 load Sensor Sensor Sensor Sensor Sensor Sensor Sensor 400 300 200 1 2 3 4 5 6 7 100 0 -100 0 10 20 30 Time (ms) 40 50 60 Figure 25: Example of FWD test results at LTPP station 32-0101 in 1996 The time delay and maximum deflection of each sensor at station 32-0101 in 1996 are summarized in Table 10. Again, it is worth noting that the measured time delays are not as accurate, because of the large sampling time intervals (0.4 ms). Table 10: Time delay and maximum deflection of sensors during the FWD test at LTPP station 32-0101 in 1996 Sensor Distance (in) Time delay (ms) Peak Deflection (mils) 1 0.00 1.2 4.84 2 8.0 1.6 4.06 3 12.0 1.6 3.39 4 18.0 2.0 2.56 5 24.0 2.0 1.89 6 36.0 5.2 1.18 7 60.0 8.8 0.79 Using the proposed method, the Rayleigh wave velocity is 138.5 m/s, and the elastic modulus is 120.6 MPa (17.5 ksi), assuming the mass density of the subgrade layer is 1500 3 kg/m . 53 The pavement structural information and the back-calculated results from MODCOMP 5 are shown in Table 11. Table 11: Basic information and MODCOMP5 backcalculated modulus at LTPP station 32-0101 in 1996 (Trial: I) Physical Layer AC Base (GB) Subbase (GS) Subbase (TS) Subgrade (SS) Thickness (m (in)) 0.183 (7.2) 0.216 (8.5) 0.579 (22.8) 0.305 (12) Infinity Poisson’s ratio 0.35 0.30 0.45 0.45 0.45 Mass density 3 (kg/m ) 2300 2000 1800 1800 1500 Elastic Modulus (MPa (ksi)) 7997 (1160) 60.7 (8.8) 5405 (784) 250969 (36400) 165 (24.0) The result of backcalculation for the subbase (TS) seems to be unreasonable. Another trial was conducted, in which all the base and subbase layers were combined into one single base layer with a total thickness of 1.10 m (44.3 in). The results are shown in Table 12. Table 12: Basic information and MODCOMP5 backcalculated modulus at LTPP station 32-0101 in 1996 (Trial: II) Physical Layer AC Base (combined) Subgrade (SS) Thickness (m (in)) 0.183 (7.2) 1.10 (43.3) Infinity Poisson’s ratio 0.35 0.40 0.45 Mass density 3 (kg/m ) 2300 2000 1500 Elastic Modulus (MPa (ksi)) 6640 (963) 298 (43.2) 281 (40.8) The modulus is more reasonable in this trial, although the deflection basin may generate a higher RMSE. The deflection basins from the FWD test and the backcalculation results obtained by MODCOMP5 are shown in Fig. 26. It illustrates that the deflection basin is simulated reasonably well. 54 Deflection Basin Maximum Deflection (mils) 0 1 ModComp-1 FWD test ModComp-2 2 3 4 5 6 0 10 20 30 40 Sensor distance (in) Radius Location (in) 50 60 70 Figure 26: Deflection basin due to the FWD test and two cases of MODCOMP5 simulation at LTPP station 32-0101 in 1996 Similar to the previous example, the subgrade resilient modulus is estimated from the triaxial laboratory test results to be 64.0 MPa (9.3 ksi) at 6 psi confinement. The elastic modulus obtained from backcalculation by MODCOMP5 is 281 MPa (40.8 ksi), and the elastic modulus by the wave velocity method is 120.6 MPa (17.5 ksi). Thus, the difference between the laboratory experimental data and the backcalculation is 339.1%, as compared to 88.2% when using the wave velocity approach. These values are summarized in Table 13. Table 13: Elastic modulus of the subgrade layer at LTPP station 32-0101 in 1996 Lab_ measured MR @ 6psi confinement 64.0 MPa (9.3 ksi) Predicted by k-value (Ek-value) 120.6 MPa (17.5 ksi) Difference MODCOMP5 Difference backcalculation between MR and between MR and (EMODCOMP) Ek-value EMODCOMP 281 MPa 88.2% 339.1% (40.8 ksi) 55 The difference between the experimental data and the result from proposed wave velocity method is still high, because the total thickness of the pavement structure is 50.5 inches, and the distance for the farthest sensor in the FWD test is only 60 inches, so the time delay for sensor 7 does not directly indicate the elastic modulus of the subgrade layer, rather it is the combination of the subbase and the subgrade layers. The freezing Index for this station is 246.63 degree-day, and the frost penetration depth is about 25 inches by interpolation from the chart (Yoder and Witczak, 1975). The thickness of the total pavement structure is more than 50 inches, so the frost problem should not affect the subbase (TS) and subgrade layer. Since only one FWD test is input in MODCOMP5, the nonlinear property of the subgrade layer cannot be obtained, and the subgrade layer is assumed as linear elastic material, or the modulus is independent of the stress condition. This simplification would significantly increase the error from MODCOMP5. 3.5.3 Example of the FWD Test at LTPP Station 06-0565 in 1999 Another example is chosen from LTPP station 06-0565 in California in 1999. The FWD test load and deflection histories are shown in Fig. 27. There are 5 relatively thin AC layers, so they are combined into one AC layer for analysis, although they were constructed at different times. The maximum applied pressure is 532 kPa (77.16 psi). 56 FWD test history Load (kPa)/deflection (um) 1000 load Sensor Sensor Sensor Sensor Sensor Sensor Sensor 800 600 400 200 1 2 3 4 5 6 7 0 -200 0 10 20 30 Time (ms) 40 50 60 Figure 27: Example of the FWD test history at LTPP station 06-0565 in 1999 The time delay and maximum deflection of each sensor are summarized in Table 14. Table 14: Time delay and maximum deflection of the sensors due to the FWD test at LTPP station 06-0565 in 1999 Sensor Distance (in) Time delay (ms) Peak deflection (mils) 1 0.00 3.6 32.56 2 8.0 4.4 26.18 3 12.0 5.2 20.75 4 18.0 6.4 15.04 5 24.0 8.0 11.02 6 36.0 11.2 7.24 7 60.0 16.0 4.41 Following the proposed method, the Rayleigh wave velocity is 116.0 m/s, and the elastic modulus is 101.5 MPa (14.7 ksi), assuming the mass density of the subgrade layer is 1800 kg/m3. The pavement structure information and the back-calculated results from MODCOMP 5 (trial I) are shown in Table 15. 57 Table 15: Structural information and MODCOMP5 backcalculated modulus at LTPP station 06-0565 in 1999 (Trial: I) Physical Thickness Layer (m (in)) AC 0.208 (8.2) Base (TB) 0.119 (4.7) Subbase (GS) 0.544 (21.4) Subgrade (SS) Infinity Poisson’s ratio 0.35 0.30 0.45 0.45 Mass density 3 (kg/m ) 2300 2000 1800 1800 Elastic Modulus (MPa (ksi)) 834.3 (121.0) 6.9 (1.0) 344738 (50000) 9.7 (1.4) The result of backcalculation for the subbase (GS) is unreasonable, partially due to the soft subgrade, and partially due to the soft base (TB) layer. Another trial is conduced by combining all the base and subbase layers into one single base layer. The results for trial II are shown in Table 16. Table 16: Structural information and MODCOMP5 backcalculated modulus at LTPP station 06-0565 in 1999 (Trial: II) Physical Layer AC Base (combined) Subgrade (SS) Thickness (m (in)) 0.208 (8.2) 0.663 (26.1) Infinity Poisson’s Mass density 3 ratio (kg/m ) 0.35 2300 0.40 2000 0.45 1800 Elastic Modulus (MPa (ksi)) 703.3 (102) 55.8 (8.1) 57.2 (8.3) The modulus is more reasonable in trial II, although the deflection basin may generate a higher RMSE. The deflection basins from the FWD test and the backcalculation results by MODCOMP5 are shown in Fig. 28. It shows that the deflection basin is simulated reasonably well. 58 Deflection Basin Maximum Deflection (mils) 0 5 ModComp-1 FWD test ModComp-2 10 15 20 25 30 35 0 10 20 30 40 Radius distance (in) Sensor Location (in) 50 60 70 Figure 28: Deflection basin during the FWD test and two cases of MODCOMP5 simulation at LTPP station 06-0565 in 1999 Using laboratory measured values of 6 psi confinement, the subgrade resilient modulus is 83.2 MPa (12.1 ksi). The elastic modulus obtained by the backcalculation by MODCOMP5 is 57.2 MPa (8.3 ksi), and the modulus by the wave velocity method is 101.5 MPa (14.7 ksi). Thus, the difference between the experimental data and the backcalculation is 31.3%, compared to 21.5% when using the proposed wave velocity method. These values are summarized in Table 17. Table 17: Elastic modulus of the subgrade layer at station 06-0565 in 1999 Lab_ measured Predicted by k-value MR @ 6psi (Ek-value) confinement 83.2 MPa 101.5 MPa (12.1 ksi) (14.7 ksi) MODCOMP5 backcalculation (EMODCOMP) 57.2 MPa (8.3 ksi) 59 Difference between MR Difference between MR and and Ek-value EMODCOMP 21.5 % 31.3 % 3.5.4 Discussion of Case Studies Obtained from LTPP The three examples illustrate that the proposed method generates reasonable results. The results obtained by wave propagation method are closer (20% to 80%) to the experimental data than the backcalculation results obtained by MODCOMP5 (30% to 300%). However, this observation is based on the lab-measured resilient modulus at the confinement of 6 psi, which may not be right for the case studies above. Another experimental model is applied to estimate the field elastic modulus of each subgrade layer, based on the possible compaction level during construction and the load applied by FWD test. 3.5.4.1 Effect of Confinement From experimental data, the resilient modulus (kPa), or elastic modulus, of the subgrade layer is usually expressed as the Uzan model (Chen et al., 2007), in Eq. (29): M R = k1θ k2τ k3 , where: θ is the bulk stress (kPa); τ= 1 3 (29) (σ 1 − σ 2 )2 + (σ 1 − σ 3 )2 + (σ 3 − σ 2 )2 is the octahedral shear stress (kPa), and σ1, σ2 , and σ3 are the principal normal stresses; and k1, k2, k3 are the regression coefficients. If the axial stress is almost identical to the confinement pressure, the octahedral shear stress is around zero, and the regression is not easy, so the modified model, Cornell model (Irwin, 1994), is also implemented to better capture the above relationship, as in Eq. (30). 60 M R = k1θ k2 (τ + 1) k3 (30) The experimental results of the above three examples are summarized for different cases of axial load and confinement based on the LTPP database, and the RMSE technique is applied to find the regression coefficients. There are no laboratory data for station 06-0565; however, the experimental data for 06-0564 are available with two locations (BA9, BA10). One can assume the subgrade should be the same since the two stations are separated for only 100 ft. The results of data regression are shown in Table 18. Table 18: Results of data regression of resilient modulus on stress parameters using Uzan and Cornell Models Station ID 04-1036 32-0101 06-0564-BA9 06-0564-BA10 Uzan model k1 29295 4991 45994 24216 k2 0.1963 0.5341 0.1444 0.2177 k3 -0.0278 -0.0708 -0.0322 -0.0493 2 R 0.631 0.597 0.364 0.277 Cornell model k1 29696 5224 44761 26140 k2 0.2068 0.5427 0.1690 0.2240 k3 -0.0521 -0.1027 -0.0665 -0.0910 R2 0.712 0.614 0.424 0.331 With the exception of data from station 06-0564 at location BA10, all other data are 2 reasonable with regression R >0.40 (Baladi, 2010). For simplification and better estimation, the experimental data from 06-0564-BA9 were used for station 06-0565 for further analysis. The actual confinement of the subgrade layer is unknown, although it was assumed to be 41.3 kPa (6 psi) in the design procedure. The coefficient of lateral earth pressure at rest (K0) for subgrade may vary from 0.4 to 2.0, depending on the compaction level during construction. The overburden stress in the Z-direction can be approximately calculated based on the mass density of each layer, as the thickness of each layer is known. The stress status inside the subgrade layer 61 can be calculated based on the coefficient of lateral earth pressure at rest (K0). For the FWD test, the stress inside the pavement can be directly obtained by the layered elastic program JULEA, and the total stress is the summation of the stress at rest added to the stress from the FWD load. Since the subgrade layer is semi-infinite, only the stress at 1 ft below the subgrade surface (the compaction depth in construction) is considered, and the horizontal distances are chosen for the location of sensors 5, 6 and 7. The calculation variables and outcomes at station 04-1036 are shown in Table 19, given the maximum FWD test pressure is 749 kPa (108.63 psi). Table 19: The stress status of the subgrade layer in the FWD test for different K0 values Sensor Horiz. No. dist. (in) 5 6 7 24.0 36.0 60.0 Depth At rest FWD FWD FWD (in) σ (psi)σ (psi) σ (psi) σ (psi) z x y z 35.4 2.52 -0.14 0.22 1.54 35.4 2.52 -0.10 0.35 0.82 35.4 2.52 -0.04 0.27 0.20 Octa. k0=0.4 k0=1.0 k0=2.0 τ (psi) 6.2 9.2 14.2 0.98 5.6 8.6 13.7 0.67 5.0 8.0 13.0 0.28 Total Bulk Stress (psi) Based on the bulk stress and octahedral shear stress ranges, the range of the resilient modulus (MR) can be determined, given the regression coefficients in Table 18. A similar procedure was done for stations 32-0101 and 06-0565. The range of MR calculated using the non-linear model, the elastic modulus from forward calculation by Rayleigh wave method, and backcalculation by MODCOMP5 for the three cases in the LTPP database are shown in Table 20. 62 Table 20: The range of elastic moduli predicted by nonlinear models, measured in the lab, estimated by the proposed method, and backcalculated by MODCOMP5 for the case studies (MPa) Lab-Measured Estimated by the MODCOMP5 MR @ 6 psi proposed method Backcalculation Minimum Average Maximum confinement 04-1036 58.4 64.6 72.7 102.0 156.9 222.7 32-0101 43.9 59.8 79.0 64.0 120.6 281.0 06-0565 79.0 86.2 93.8 83.2 101.5 57.2 The difference in subgrade moduli predicted by the three methods was smallest at station Station ID Nonlinear model prediction 06-0565, and largest at station 04-1036. As the bulk stress affects the MR more than the octahedral shear stress, the bulk stress, or the confinement is more important in determining the MR of the subgrade layer. The maximum bulk stress at station 04-1036 is 13 psi or less, as shown in Table 16, meaning the confinement is much less than 6 psi. At station 06-0565, the bulk stress is much higher than that at 04-1036 (values are not shown here), as the pavement structure is thicker in the base layer, and the confinement pressure may be around 6 psi. The confinement of the subgrade layer is unknown, and the error may be less if the actual confinement is known. The Poisson’s ratio and mass density are assumed in the proposed method, and that have a linear effect on the final result. In the above examples, the mass density is estimated from Stubstad (2002). If the mass density were more accurate, the error would be smaller. 3.5.4.2 Effect of Sampling Intervals Theoretically, it is possible to accurately predict the modulus of subgrade layers, if the sampling interval is small enough and more sensors are available. Practically, the sampling time is around 0.4 ms, and there are typically 7 sensors within 72 inches from the FWD load, so the error may be significant. Assuming the S-wave velocity is 150 m/s, and R-wave velocity is 142.4 63 m/s, the time required for each wave to arrive at each sensor in the FWD test is listed in Table 21. Table 21: Arrival time for S-wave and R-wave for each sensor due to the FWD test Sensor Location (m) S-wave (ms) R-wave (ms) 1 0 0 0 2 0.3 2.0 2.1 3 0.6 4.0 4.2 4 0.9 6.0 6.3 5 1.2 8.0 8.4 6 1.5 10.0 10.5 7 1.8 12.0 12.6 If the sampling interval is 0.4 ms in the FWD test, the time difference between S-wave and R-wave propagation cannot be clearly identified for close sensors, which causes a significant error for subsequent sensors. The vibrations of R-wave and S-wave are in opposite directions (Fig. 9 in section 3.2.1). Because the two waves interact, the occurrence time of the peak deflection is not accurate, and thus the error can be as high as 30%. The error difference between numerical examples and case studies from the LTPP can be explained by differences in sample time intervals. In order to reduce the error in estimation of subgrade elastic modulus from wave propagation, the frequency of sampling should be increased. 3.6 Numerical Example of Bedrock under Subgrade layer As shown in section 3.2.4, k-values vary if the elastic modulus of the subgrade layer is higher than 500 MPa (70 ksi). However, in some cases, a shallow bedrock may underlie the subgrade. In this section, the effect of existing shallow bedrock is investigated by numerical simulation. Bedrock is assumed to be located 5 ft below the surface of the subgrade layer, and the time-history is simulated by SAPSI. Table 22 is a summary of the material properties of the 64 pavement profile. Table 22: Material properties of a three-layer pavement structure with bedrock Physical Layer AC Base Subgrade Bedrock Elastic Modulus (MPa (ksi)) Experimental data 108 (21.8) 34.5 (5.0) 7000 (1000) Poisson’s ratio 0.35 0.30 0.45 0.25 Mass density 3 (kg/m ) 2300 2000 1500 2800 Thickness (m (in)) 0.15 (6) 0.3 (12) 1.5 (60) Infinity The FWD load is assumed as a haversine function with a duration time of 35 ms, and the maximum pressure is 707 kPa with a loading plate radius of 0.15 m. The simulated FWD time-history is plotted in Fig. 29, and the time delay of each sensor is given in Table 23. Deflection (m) Deflection history with Bedrock (5 ft) 7.E-04 6.E-04 5.E-04 4.E-04 3.E-04 2.E-04 1.E-04 0.E+00 -1.E-04 -2.E-04 sensor 1 sensor 3 0.04 sensor 6 sensor 7 0.02 sensor 4 sensor 5 0 sensor 2 sensor 8 0.06 0.08 0.1 time (sec) Figure 29: SAPSI simulated time-history of FWD test with bedrock underneath 65 Table 23: Time delay of each sensor by SPASI simulation with bedrock Sensor 1 Distance (m) 0.00 Time delay (s) 0.0053 2 0.20 0.0056 3 0.30 0.0059 4 0.46 0.0064 5 0.61 0.0069 6 0.91 0.0083 7 1.22 0.01 8 1.52 0.0122 Fig.28 shows there is free vibration after the FWD test, due to the stiff layer underneath. The time delay vs. location is plotted in Fig. 30. Time delay vs. location (Bedrock 5 ft) Time Delay (s) 0.014 0.012 0.01 0.008 0.006 y = 5.7743E-03x + 3.1900E-03 0.004 R2 = 9.8965E-01 0.002 0 0.00 0.40 0.80 1.20 1.60 Location (m) Figure 30: Time delay vs. location for a numerical example of FWD test with bedrock The slope can be obtained by linear regression, as m = 0.005774 s/m. Following the proposed procedure in section 3.3, the elastic modulus of the subgrade layer can be calculated as: E = 368.0 MPa, which is quite different from the modulus of the subgrade layer or the bedrock layer. It is the combination of the two layers, as the more distant sensors are significantly 66 affected by the interaction at the interface of the subgrade layer and bedrock. Therefore, the proposed method is not appropriate in the case of shallow bedrock. In wave propagation theory, the path of wave propagation changes if there is a stiff layer underneath and the wave is called Love wave instead of Rayleigh wave. A detailed explanation can be found in reference (Richart et al., 1970). However, the existence of shallow bedrock can be easily identified by the free vibration response of the pavement under FWD testing. 3.7 Summary Both numerical examples and case studies obtained from the LTPP database showed that the subgrade elastic modulus calculated by the proposed method based on wave propagation theory was acceptable for pavement engineering application, compared to the static backcalculation, no matter whether there is GWT underneath or not. The estimation of the subgrade elastic modulus can be improved in field FWD tests by increasing the sampling frequency of each sensor. When shallow bedrock presents, the subgrade elastic modulus cannot be estimated accurately, as the Rayleigh wave is theoretically based on one layer of semi-infinite medium, and it is only valid for layered elastic systems with higher modulus of the surface layers. The existence of shallow bedrock underneath the subgrade layer is indicated by free vibration in the pavement response due to the FWD test; in this case, the elastic modulus of subgrade layer cannot be obtained by the proposed method. 67 Chapter 4. Layered Viscoelastic Forward Solution 4.1 Introduction In each loop of any backcalculation algorithm, there is a forward solution that predicts the theoretical response, which is compared to the field measured response. Since a backcalculation program often converges after many loops of iteration, time-efficiency and accuracy are critical factors. The forward viscoelastic solution is presented in this chapter. It utilizes the concept of ‘quasi-elastic’ approximation suggested by Schapery (1974). Both time-efficiency and accuracy are compared, and the viscoelastic solution is chosen for further analysis, mainly based on the consideration of time-efficiency. The error and limitation of the proposed solution is investigated later in this chapter. 4.2 Layered Viscoelastic Forward Solution Algorithm 4.2.1 Theoretical Background The time-dependent response of linear viscoelastic material subjected to a random loading history can be computed using the following Boltzmann’s superposition integral (Schapery, 1974): 68 t ve RH (t − τ ) dI (τ ) ∫ ve R (t ) = τ =0 where R ve (t ) is the linear viscoelastic response, unit step function of input ( I (t ) = (31) , ve RH (t ) is the viscoelastic response to a H (t ) , where H(t) is the Heaviside step function) and dI (τ ) is the change in input at time τ . In the case of a layered system with a circular load as shown in Fig. 31, the time dependent input is the contact stress (i.e., I(t) = σ(t)) and the response of interest may be the deflections (i.e., R ve (t ) = u ve (t ) ) at certain locations. Then we can rewrite Eq. (31) as follows: t u ve (t ) = ve u H (t − τ ) dσ (τ ) ∫ τ =0 where ve uH (32a) , is the deflection due to a unit contact stress (i.e., σ(t)=1). It is noted that ve uH can be either the vertical or radial deflection, although in this paper, only the vertical surface deflection is considered. The generalized formula for Eq. (32a) can be written, in cylindrical coordinates, as follows: t ve uvertical (t,z,r) = ∫u τ =0 ve H − vertical (t − τ,z,r) dσ (τ ) (32b) t ve uradial (t,z,r) = ∫u τ =0 ve H − radial 69 (t − τ,z,r) dσ (τ ) (32c) ve uvertical (t,z,r) and ve uradial (t,z,r) , respectively, are the displacements in vertical and radial directions, observed at time t and at location (r,z); u ve−vertical (t − τ,z,r) H and uve−radial (t − τ,z,r) , respectively, are the viscoelastic deflections due to a unit contact H stress (i.e., σ(t)=1). r σ H 1 H2 H=Inf Ε(t), ν1 Ε ,ν AC Base 2 Subgrade 2 Εn, νn Figure 31: Typical geometry of a pavement structure ve The viscoelastic deflection due to a unit contact stress ( u H ) can efficiently and accurately be computed by using Schapery’s ‘quasi-elastic’ approximation (Schapery 1965; 1974). Quasi-elastic theory states that unit viscoelastic response (e.g., u ve (t) = surface deflection) at H any time t can be approximated by the unit elastic response (e.g., u e (E ve (t)) H = elastic surface deflection), which is calculated using the modulus ( E ) equal to viscoelastic modulus E ve (t ) evaluated at time t, i.e.: ve u H (t) ≅ u e (E ve (t)) , H 70 (33) where e uH is the elastic deflection due to a unit step load computed using a layered elastic solution where E = E ve (t) is utilized for the AC layer. The detailed derivation of Eq. (33) can be found in Levenberg (2008) and will not be repeated here for brevity. In this implementation, the unit response ve e u H (t ) ≅ u H (t ) values at the points of interest were computed using the CHEVLAY2 layered elastic analysis program. Then the convolution integral in Eq. (32b) is used to calculate the viscoelastic deflection u ve (t ) . The algorithm is described in the following section. 4.2.2 Algorithm Steps Based on the theoretical formulation, the following steps are applied to implement the Viscoelastic (VE) solution from the linear elastic layered solution of CHEVLAY2. 1. Define the geometry (layer thicknesses, contact pressure…etc) of a layered system similar to the one in Fig. 31. 2. Select a stress versus time history, σ(t), and divide the data into Ns discrete intervals as shown in Fig. 32. 71 120 Figure 2 Stress (psi) 100 80 60 Number of time intervals: Ns =50 40 20 0 0 0.02 0.04 0.06 0.08 0.1 t(s) Figure 32: Discretization of stress history 3. Divide the relaxation modulus E(t) mastercurve into NE number of time steps (Fig. 33). 8 10 Figure 3 6 E(t) psi 10 4 10 Number of time intervals: NE=100 2 10 -6 10 -4 10 -2 10 t(s) 0 10 Figure 33: Discretization of the relaxation modulus mastercurve 72 2 10 4. Calculate the elastic response of the structure to a unit stress (step function of the load input) using the E(ti) evaluated at different times (i.e., t1, t2, t3 ….tNE). In this present implementation, the surface deflections at several radial distances to a circular plate load shown in Fig. 31 were of interest. Therefore, these surface deflections were computed using the CHEVLAY2 program using the modulus value corresponding to different times in Fig. 34; i.e., E(t1), E(t2), E(t3), E(t4)… E(tNE): ve e u H (ti ) ≅ u H calculated using E(ti) where i=1,2,3 …NE (34) -1 10 rc = 0 in uVE(t) = uE (t) (inches) H H rc = 13 in rc = 21 in -2 10 rc = 35 in rc = 49 in rc = 63 in -3 10 -4 10 -6 10 -4 -2 10 10 t(s) 0 10 2 10 Figure 34: Deflections calculated for points at different distances from the centerline of the circular load at the surface 73 Fig. 34 shows the e uH values calculated for points at different distances from the centerline of the circular load at the surface. These curves are herein called “unit response mastercurves”. 5. Calculate the viscoelastic response using the discrete form of Eq. (33) given in Eq. (35) below. Eq. (35) is evaluated at each discrete time Fig. 32. Fig. 35 below illustrates the dσ (τ j ) ti using the stress history shown in in Eq. (35) for each time step τj. i ve u (t i ) = ∑ u H (t i − τ j ) dσ(τ j ), ve (35) j= 0 i = 1, 2,...Ns where 120 Stress (psi) 100 80 60 dσ (τ j ) 40 δσ(τ) 20 0 τj 0 0.02 τ Figure 35: Illustration of the 0.04 0.06 0.08 0.1 t(s) dσ (τ j ) 74 in Eq. (32b) for each time step τj In order to illustrate an example, the viscoelastic surface deflections of the three-layered pavement structure shown in Fig. 36 are computed. Fig. 37 shows the vertical surface deflections at points located at different radial distances from the centerline of the load, and clearly shows the relaxation behavior of deflection at each point. σ=102.6 psi H1=3.94” H2=11.8” 1 r=5.9” 2 3 6 5 E(t )= Fig. 15, ν1=0.35 AC Base H3=Inf Subgrade 4 E2=1567.5 psi ν2=0.35 E3=1451.4 psi, ν3=0.35 Figure 36: Example of a pavement structure used for VE forward calculation 0.15 rp1= 0 in rp2= 13 in rp3= 21 in 0.1 uVE(t) in rp4= 35 in rp5= 49 in rp6= 63 in 0.05 0 0 0.02 0.04 0.06 0.08 0.1 t(s) Figure 37: Examples of computed VE surface deflections at different radial distances from the centerline of the load 75 4.3 Verification of the Layered Viscoelastic Forward Solution First, verification of ‘quasi-elastic’ approximation was performed. Chen (2009) introduced a theoretical solution for response of viscoelastic (VE) pavement layers. The unit responses computed using Chen’s method (Chen and Pan, 2007) was compared against the ‘quasi-elastic’ approximation shown in Eq. (33). Second, the proposed VE solution is compared against the well-known dynamic solutions by SAPSI and LAMDA. The effect of wave propagation was eliminated from these dynamic solutions. Later, the time-efficiencies of both solutions are compared in the next section. 4.3.1 Verification of Quasi-Elastic Approximation A semi-analytical solution for a multilayered viscoelastic pavement under surface loading was derived by Chen (2009; Chen et al, 2009). This solution was used to verify the accuracy of the quasi-elastic approximation. The structural properties of the pavement used in this verification are listed in Table 24. Table 24: Material properties of the pavement structure used in the verification example Physical Layer AC Base Subgrade Elastic Modulus (MPa (ksi)) Experimental data 108 (15.6) 34.5 (5.0) Poisson’s Ratio 0.35 0.30 0.45 Thickness (m (in)) 0.15 (6.0) 0.30 (12.0) Infinity Only 11 terms of the Prony series of E(t) can be implemented by the solution developed by Chen (2009). The input E(t) curve was therefore approximated with a 11-term Prony series, as shown in Fig. 38. It is noted that 11-term Prony series did not accurately approximate the actual E(t) curve. 76 Implemented VS. actual E(t) of one AC layer E(t) (MPa) 1.E+05 Actual 1.E+04 Implemented 1.E+03 1.E+02 1.E+01 1.E+00 1.E-15 1.E-10 1.E-05 1.E+00 1.E+05 1.E+10 1.E+15 Reference time (s) Figure 38: Implemented and actual E(t) for one AC mixture Comparison between analytical and quasi-elastic solution for pavement response under step load 4.E-02 deflection (m) 3.E-02 3.E-02 2.E-02 2.E-02 1.E-02 5.E-03 0.E+00 0.E+00 sensor 1 sensor 5 1.E-02 sensor 2 sensor 6 2.E-02 3.E-02 sensor 3 Sensor 7 4.E-02 5.E-02 sensor 4 6.E-02 7.E-02 time (ms) Figure 39: Implemented and actual E(t) for one AC mixture (Solid lines represent the analytical solution, symbols represent the quasi-elastic approximation) 77 Fig. 39 illustrates that the quasi-elastic approximation did a very good job in predicting unit response. The maximum error was around 3% for sensor 1, and error decreased as sensor number increased. This difference might be due to the 11-term Prony series approximation of E(t) used in Chen’s method. 4.3.2 Comparison between the Layered VE Solution and Dynamic Solutions The layered VE forward solution is also verified against the well-known dynamic solutions, SAPSI and LAMDA. As discussed previously, wave propagation is ignored in the layered VE solution. Therefore, the deflection time histories given by SAPSI and LAMDA were shifted, so that the peaks of all sensors coincide. In order to illustrate the absence of wave propagation, a FWD load pulse of duration 35 ms was hypothetically apply to an infinitely deep AC layer, and the response was simulated using the layered VE solution. The time-histories of the pavement response due to the FWD test simulation are shown in Fig. 40. 78 FWD deflection history Deflection (in) 1.2E-03 0 in 12 in 24 in 48 in 72 in 1.0E-03 8.0E-04 6.0E-04 8 in 18 in 36 in 60 in 4.0E-04 2.0E-04 0.0E+00 0.00 0.01 0.02 0.03 0.04 0.05 0.06 0.07 Time (s) Figure 40: FWD test simulation by layered VE solution for a half-space AC layer Delayed recovery of the displacements/deflection, which is a clear indication of viscoelastic response, is visible in Fig. 40. However, there is no time delay of the peaks between the sensors, i.e., the wave propagation cannot be simulated by layered VE solution. Since the focus of this study is on the viscoelasticity property, not the dynamic response of the system, the wave propagation response is eliminated by shifting the time histories of all sensors such that their peaks occur at the same time. . The layered VE solution is superimposed on the time-shifted dynamic solutions, in order to better compare the shape of the deflection time histories, and to show the accuracy of the layered VE solution. Table 25 shows that the properties of a pavement structure used in comparing dynamic and 79 layered VE solutions. Fig. 41 and Fig. 42 show the response of the pavement under a FWD test calculated by SAPSI and LAMDA, respectively. Table 25: Properties of a pavement used in comparison of dynamic and VE solutions Physical Layer AC Base Subgrade Elastic Modulus (MPa (ksi)) Experimental data 108 (15.6) 34.5 (5.0) Poisson’s Ratio 0.35 0.40 0.45 Density 3 (kg/m ) 2300 2000 1500 Thickness (m (in)) 0.10 (4.0) 0.3 (12.0) Infinity FWD deflection history 0.700 r=0 m r=0.33 r=0.52 r=0.88 r=1.25 r=1.61 Deflection (mm) 0.600 0.500 0.400 0.300 0.200 m m m m m 0.100 0.000 -0.100 0.000 0.020 0.040 0.060 0.080 0.100 Time (second) Figure 41: Sensor deflection time histories predicted by SAPSI 80 0.120 FWD deflection history r=0 m r=0.33 r=0.52 r=0.88 r=1.25 r=1.61 0.700 Deflection (mm) 0.600 0.500 0.400 0.300 0.200 m m m m m 0.100 0.000 -0.100 0.000 0.020 0.040 0.060 0.080 0.100 0.120 Time (second) Figure 42: Sensor deflection time histories predicted by LAMDA The layered VE solution and the time-shifted dynamic solutions for the same example are shown in Fig. 43 and Fig. 44. As shown in these figures, layered VE solution and dynamic solutions by SAPSI and LAMDA match very well. It is noted that the computational efficiency of VE solution is much better than SAPSI and LAMDA, where it takes around 20 minutes to obtain a solution from SAPSI and LAMDA, while it only takes 1 minute, or even a few seconds, to reach a solution using layered VE solution. 81 Deflection (mm) VE solution VS. SAPSI VE Sensor 0 VE Sensor 2 VE Sensor 4 SAPSI Sensor 0 SAPSI Sensor 2 SAPSI Sensor 4 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 0 0.01 0.02 0.03 0.04 0.05 0.06 Time (second) 0.07 VE Sensor 1 VE Sensor 3 VE Sensor 5 SAPSI Sensor 1 SAPSI Sensor 3 SAPSI Sensor 5 0.08 0.09 0.1 Figure 43: Comparison of deflection time histories from VE and SAPSI solutions (Solid lines represent the VE solution, symbols represent the SAPSI result) VE solution VS. Lamda Deflection (mm) VE sensor 0 VE sensor 2 VE sensor 4 0.02 0.04 Lamda Sensor 1 Lamda Sensor 3 Lamda Sensor 4 0 VE sensor 3 VE sensor 5 Lamda sensor 0 Lamda Sensor 2 0.7 0.6 0.5 0.4 0.3 0.2 0.1 0 -0.1 VE sensor 1 Lamda Sensor 5 0.06 0.08 Time (second) Figure 44: Comparison of deflection time histories from VE and LAMDA solutions (Solid lines represent the VE solution, symbols represent the LAMDA result) 82 0.1 4.3.3 Verification Examples using Pavement Structures from the LTPP Database Table 26: Several examples of pavement structure in the SPS-1 project Case No. Layer No. 4 3 116 2 1 Layer Description Original Surface Layer (Layer Type:AC)1.8 Inch AC Layer Below Surface (Binder Course) (Layer Type:AC)2.1 Inch Base Layer (Layer Type:TB)12 Inch Subgrade (Layer Type:SS) Layer No. 6 5 4 3 2 1 Layer Description Original Surface Layer (Layer Type:AC)1.7 Inch AC Layer Below Surface (Binder Course) (Layer Type:AC)1.5 Inch AC Layer Below Surface (Binder Course) (Layer Type:AC)3.2 Inch Base Layer (Layer Type:TB)5.2 Inch Base Layer (Layer Type:GB)4 Inch Subgrade (Layer Type:SS) Layer No. 5 4 3 2 1 Layer Description 117 120 Original Surface Layer (Layer Type:AC)1.8 Inch AC Layer Below Surface (Binder Course) (Layer Type:AC)1.8 Inch Base Layer (Layer Type:PATB)4 Inch Base Layer (Layer Type:GB)8 Inch Subgrade (Layer Type:SS) Layer No. Layer Description 123 6 5 4 3 2 1 Original Surface Layer (Layer Type:AC)1.8 Inch AC Layer Below Surface (Binder Course) (Layer Type:AC)2 Inch AC Layer Below Surface (Binder Course) (Layer Type:AC)2.4 Inch Base Layer (Layer Type:TB)8 Inch Base Layer (Layer Type:PATB)4 Inch Subgrade (Layer Type:SS) 83 In order to further verify the layered VE solution, four pavement structures are selected from the LTPP database. Table 26 shows the selected pavement structures from the SPS-1 experiment of the LTPP database. As shown in Table 26, there are several sublayers of AC. For computational efficiency, similar layers were combined into one layer, as shown in Table 27. Table 27: Layer information for each pavement structure Case No. 116 Physical Layer AC Base Subgrade Elastic Modulus Thickness (MPa (ksi)) (m (in)) Experimental Data 0.099 (3.9) 200 (29.0) 0.305 (12.0) 100 (14.5) Infinity 117 AC Experimental Data Base 200 (29.0) Granular Base 150 (21.8) Subgrade 100 (14.5) 0.163 (6.4) 0.132 (5.2) 0.102 (4.0) Infinity 120 AC Experimental Data 0.091m (3.6) PATB* 180 (26.1) 0.102 (4.0) Granular Base 150 (21.8) 0.204 (8.0) Subgrade 100 (14.5) Infinity AC Experimental Data 0.157 (6.2) Base 200 (29.0) 0.204 (8.0) 123 PATB* 180 (26.1) 0.102 (4.0) Subgrade 100 (14.5) Infinity Note: PATB = Permeable Asphalt Treated Base Poisson’s Ratio 0.35 0.40 0.45 Mass density (kg/m3) 2300 2000 1800 0.35 0.40 0.40 0.45 2300 2000 2000 1800 0.35 0.40 0.40 0.45 2300 2000 2000 1800 0.35 0.40 0.40 0.45 2300 2000 2000 1800 The deflection time histories predicted by SAPSI, LAMDA (after eliminating the time delay, or time-shifted dynamic solution) and by the VE solution are shown in Figures 45 to 48. 84 VE solution VS. SAPSI for Case 116 VE Sensor 0 VE Sensor 1 0.5 Deflection (mm) 0.6 VE Sensor 2 VE Sensor 3 VE Sensor 4 VE Sensor 5 SAPSI Sensor 0 SAPSI Sensor 1 0.3 SAPSI Sensor 2 SAPSI Sensor 3 0.2 SAPSI Sensor 4 SAPSI Sensor 5 0.4 0.1 0 -0.1 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (a) VE solution and SAPSI VE solution VS. Lamda for Case 116 VE Sensor 0 VE Sensor 1 VE Sensor 2 VE Sensor 3 VE Sensor 4 VE Sensor 5 0.4 Lamda Sensor 0 Lamda Sensor 1 0.3 Lamda Sensor 2 Lamda Sensor 3 0.2 Lamda Sensor 4 Lamda Sensor 5 0.6 Deflection (mm) 0.5 0.1 0 -0.1 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (b) VE solution and LAMDA Figure 45: Time-shifted dynamic solution and VE solution for case 116 (Solid lines represent the VE solution, symbols represent the SAPSI / LAMDA result) 85 VE solution VS. SAPSI for Case 117 VE Sensor 0 VE Sensor 1 0.4 VE Sensor 2 VE Sensor 3 0.35 VE Sensor 4 VE Sensor 5 0.3 SAPSI Sensor 0 SAPSI Sensor 1 0.25 0.2 SAPSI Sensor 2 SAPSI Sensor 3 SAPSI Sensor 4 SAPSI Sensor 5 Deflection (mm) 0.45 0.15 0.1 0.05 0 -0.05 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (a) VE solution and SAPSI VE solution VS. Lamda for Case 117 VE Sensor 0 Deflection (mm) VE Sensor 3 VE Sensor 4 VE Sensor 5 Lamda Sensor 0 Lamda Sensor 1 Lamda Sensor 2 Lamda Sensor 3 Lamda Sensor 4 0.35 0.3 0.25 0.2 0.15 VE Sensor 1 VE Sensor 2 0.45 0.4 Lamda Sensor 5 0.1 0.05 0 -0.05 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (b) VE solution and LAMDA Figure 46: Time-shifted dynamic solution and VE solution for case 117 (Solid lines represent the VE solution, symbols represent the SAPSI / LAMDA result) 86 VE solution VS. SAPSI for Case 120 VE Sensor 0 Deflection (mm) 0.2 SAPSI Sensor 3 SAPSI Sensor 4 0.3 SAPSI Sensor 1 SAPSI Sensor 2 0.4 VE Sensor 5 SAPSI Sensor 0 0.5 VE Sensor 3 VE Sensor 4 0.6 VE Sensor 1 VE Sensor 2 0.7 SAPSI Sensor 5 0.1 0 -0.1 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (a) VE solution and SAPSI VE solution VS. Lamda for Case 120 VE Sensor 0 VE Sensor 1 0.5 Deflection (mm) 0.6 VE Sensor 2 VE Sensor 3 0.4 VE Sensor 4 VE Sensor 5 Lamda Sensor 0 Lamda Sensor 1 Lamda Sensor 2 Lamda Sensor 3 Lamda Sensor 4 Lamda Sensor 5 0.3 0.2 0.1 0 -0.1 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (b) VE solution and LAMDA Figure 47: Time-shifted dynamic solution and VE solution for case 120 (Solid lines represent the VE solution, symbols represent the SAPSI / LAMDA result) 87 VE solution VS. SAPSI for Case 123 VE Sensor 0 VE Sensor 1 VE Sensor 2 VE Sensor 3 VE Sensor 4 VE Sensor 5 0.3 SAPSI Sensor 0 SAPSI Sensor 1 0.25 0.2 SAPSI Sensor 2 SAPSI Sensor 3 SAPSI Sensor 4 SAPSI Sensor 5 0.45 0.4 Deflection (mm) 0.35 0.15 0.1 0.05 0 -0.05 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (a) VE solution and SAPSI VE solution VS. Lamda for Case 123 VE Sensor 0 VE Sensor 1 0.4 0.35 Deflection (mm) 0.45 VE Sensor 2 VE Sensor 3 0.3 0.25 0.2 VE Sensor 4 VE Sensor 5 Lamda Sensor 0 Lamda Sensor 1 Lamda Sensor 2 Lamda Sensor 3 Lamda Sensor 4 Lamda Sensor 5 0.15 0.1 0.05 0 -0.05 0 0.02 0.04 0.06 Time (second) 0.08 0.1 (b) VE solution and LAMDA Figure 48: Time-shifted dynamic solution and VE solution for case 123 (Solid lines represent the VE solution, symbols represent the SAPSI / LAMDA result) 88 Figures 45 - 48 show that the typical difference between the results by the VE solution and those by time-shifted dynamic analysis is about 6%, with the maximum difference as high as 10%. The difference appears to be slightly higher when the AC layer is thicker, or when there are more physical layers in the pavement structure. Most importantly, the shapes of the response pulses from dynamic and VE solutions are consistent with each other. Therefore, the four numerical examples show that the difference between the responses from the dynamic solutions (SAPSI or LAMDA) and from the VE solution is minimal. 4.3.4 A Numerical Example with the Bedrock The existence of bedrock may cause problems in numerical simulation. The previous chapter shows that the proposed method to estimate the elastic modulus of the subgrade layer by Rayleigh wave is invalid if there is shallow bedrock underneath. In addition, the reflection of the waves and resulting response can overshadow the effects of viscoelasticity. Fig. 49 shows the responses calculated by SAPSI and layered VE solution. As expected, dynamic response by SAPSI is significant, while layered VE solution does not show the effect of wave propagation. 89 Comparison of deflection history with Bedrock (5 ft) Deflection (m) 8.E-04 Sensor Sensor Sensor Sensor 6.E-04 4.E-04 1 3 5 7 Sensor Sensor Sensor Sensor 2 4 6 8 2.E-04 0.E+00 -2.E-04 0 0.02 0.04 0.06 0.08 0.1 time (s) Figure 49: Time-shifted SAPSI and VE solution for shallow bedrock (Solid lines represent the SAPSI result, symbols represent the VE solution) Fig. 49 illustrates that the layered VE solution cannot accurately simulate the pavement response if there is shallow bedrock underneath. Therefore, the layered VE solution should not be used for simulating the pavement response if there is a shallow, stiff layer underneath. The magnitude of error of the simulation depends on the relative level of stiffness between physical layers, and the relative depth of the bedrock A sensitivity analysis for the depth of the bedrock is done for the numerical example, and the only input variable is the thickness of the subgrade layer, or the depth of the bedrock, and the output variable is the relative error of the maximum deflection for each sensor between dynamic solution SAPSI and VE forward solution. The result is shown in Fig. 50. Depending on the 90 accuracy of the simulation requirement, the effect of bedrock is insignificant if the depth is 15 ft or more. Relative Error of Maximum Deflection Relative Error vs. Bedrock Depth 70% 60% 50% 40% 30% 20% 10% 0% -10% Sensor Sensor Sensor Sensor 0 5 10 1 3 5 7 15 Sensor Sensor Sensor Sensor 2 4 6 8 20 Bedrock Depth (ft) Figure 50: Sensitivity analysis for relative error of maximum deflection between VE solution and SAPSI with bedrock depth 4.4 Time Efficiency for Layered VE Forward Solution One of the primary reasons for implementing Schapery’s “quasi-elastic” approximation is its extreme computational efficiency. It takes more than 25 minutes to calculate the pavement response under step load by semi-analytical solution, and it takes around 20 minutes to simulate the pavement response under FWD test by dynamic solution SAPSI or LAMDA, while it takes less than one minute to calculate the response by VE solution. Many repetitions are required for the forward simulation, so time efficiency is a very important factor. Using a Pentium (4) 3.20 GHz computer with 1.99 GB ram, the computation time of the results shown in Fig. 37 is 44.72 seconds, where the solution was for the 3-layered system shown 91 in Fig. 31 and NS = 50, NE=50. Table 28 shows the computation times for different numbers of discrete time steps for the 3-layered system shown in Fig. 35. The difference in Table 28 is defined as square root of the relative difference of the peak deflection compared to the most difficult (reference) case with NS = 200 and NE = 200. Mathematically the difference is expressed as: 2 1 N ⎛ ωi ( peak in case ) − ωi ( peak reference ) ⎞ ⎟ , ∑⎜ ⎜ ⎟ ωi ( peak reference ) N i =1 ⎝ ⎠ Difference = (36) where: N = No. of Sensors. Table 28: The computation times for different numbers of discrete time steps Ns NE Elapsed time for computation (seconds) Difference (%) 50 24 50 100 100 200 50 100 100 100 200 200 22.60 45.41 44.72 45.35 89.38 89.73 0.069 0.360 0.077 0.011 0.018 0.000 Table 28 shows that NS = 100 and NE = 100 is enough to generate a reasonable result, i.e., the sampling time for the FWD test is around 0.5 ms, and the time cost of each forward simulation is around 45 seconds. Therefore, the VE forward solution is accurate and time efficient, and is chosen as the numerical simulation of the pavement under FWD test in future analysis. 92 4.5 Summary A computation procedure for layered VE solution has been provided in this chapter. It was simple in understanding, efficiency in calculation, and easy for implementation in MATLAB. The accuracy of the proposed forward solution was checked against a semi-analytical solution for a step load, and was checked against dynamic solutions SAPSI and LAMDA for FWD test simulation. The time delay of each sensor cannot be simulated, and the deflection time-histories can be calculated within error around 5% for maximum deflection. The proposed layered VE forward solution was chosen for future backcalculation because of its accuracy and time-efficiency. If there is shallow bedrock underneath the pavement, the effect of mass inertia cannot be ignored, and the proposed VE solution is not appropriate. 93 Chapter 5. Backcalculation Algorithm 5.1 Introduction In this chapter, the outputs of the backcalculation program are identified, and the boundaries of each output variable are measured from the lab as guidance for the convergence of the program. Then the basic theory of inverse problem is briefly introduced. A detailed flowchart of the backcalculation program is presented, and one numerical example of the backcalculation result is shown in the end. 5.2 Variable Identification The input variables are thickness of each layer, Poisson’s ratio of each layer, applied load history and deflection time history of each sensor during a FWD test. The output variables are the elastic modulus of base and subgrade layers, and a function of relaxation modulus E(t) or dynamic modulus mastercurve |E*| of the AC layer. The elastic modulus of base and subgrade layers is noted separately as Ebase and Esubgrade. Mathematically E(t) is commonly expressed as Prony series with more than 30 variables, however, it is almost impossible for any backcalculation to converge with so many variables. A pre-described function must be required to reduce the unknown parameters. The sigmoid function is preferred for |E*| in frequency domain and E(t) in time domain, as Eq. (1) for |E*| and Eq. (2) for E(t) respectively. The backcalculation is used to calculate the parameters of the 94 sigmoidal function (a, b, c, d) for |E*| in frequency domain or E(t) in time domain respectively, although the forward calculation is done in time domain. 5.3 Variable Boundaries A reasonable range for the value of each output variable helps to converge the backcalculation. If the intermediate output value crosses the boundary, it is forced to be within the range by the algorithm. The elastic modulus of base and subgrade layer is suggested to be less than 700 MPa (100 ksi). If not, it is not appropriate for the proposed wave propagation method to estimate the elastic modulus of the subgrade layer (as discussed in Chapter 3) and for the proposed layered VE solution to simulate the pavement response. The boundaries of the parameters for the AC layer in the sigmoid function are found in the experiment data of many mixtures. The seed value of the four parameters can be set as the average value, and the boundary limit is identified via the minimum or maximum value of each variable. The corresponding |E*| or E(t) components are discussed in the following separate sections. 5.3.1 Boundary Limits for |E*| About 30 sets of |E*| were obtained from FHWA (Kutay, 2008) and other sources (Mogawer et al, 2010). Their coefficients for the dynamic modulus mastercurve |E*| are shown in Table 29. 95 Table 29: The |E*| parameters for different AC mixtures Mixture Name PG64-28 Control no-PPA (AI) ALF Control 70-22 SBS 64-40 Terpolymer CR-AZ SBS LG CR-TB Air Blown Fiber Control 70-22 KIM Advera Control Sasobit PPA PPA + Elvaloy SBS SBS + PPA SBS PG64-34 PG64-28 Control no-PPA (AI) PG64-28 with PPA (Hudson) PG64-34 (SEM) PG76-22 (Citgo) PG64-22 with 12% GTR (Gorman) PG64-28 no-PPA(AI)with2.0%Latex WAM Control WAM Foam Asphamin Control Sasobit-1 Sasobit-2 Evatherm-1 Evatherm-2 a 1.699 0.797 1.338 1.468 -0.202 1.341 0.807 0.553 0.504 1.585 2.156 0.458 -0.26 1.552 1.579 1.72 1.568 0.851 1.569 1.49 1.689 1.646 1.245 1.507 0.92 1.13 1.261 1.195 1.423 1.266 1.112 1.191 b 2.765 3.593 2.843 2.799 4.695 2.857 3.519 3.868 3.838 2.955 1.801 3.772 4.907 2.408 2.317 2.21 2.505 2.391 2.65 2.677 2.394 2.476 2.851 2.678 3.371 3.168 2.985 3.127 2.791 3.012 3.162 3.144 c 0.823 1.372 0.164 0.491 1.122 0.876 1.066 1.285 1.535 1.077 0.546 0.948 0.787 0.602 0.582 0.519 0.464 0.157 0.855 0.777 0.156 0.954 1.151 0.905 1.387 1.244 1.284 1.333 1.423 1.375 1.245 1.258 d 0.483 0.503 0.561 0.583 0.346 0.556 0.407 0.405 0.504 0.558 0.854 0.389 0.268 0.702 0.543 0.73 0.643 0.682 0.668 0.639 0.681 0.705 0.483 0.572 0.595 0.641 0.598 0.552 0.578 0.543 0.546 0.497 The average, minimum and maximum values of each variable are listed in Table 30. 96 Table 30: The average and boundaries for the four parameters of |E*| Average value Minimum value Maximum value a 1.192 -0.26 2.156 b 3.016 1.801 4.907 c 0.93 0.156 1.535 d 0.563 0.268 0.854 5.3.2 Boundary Limits for E(t) The relaxation modulus E(t) of the same 30 examples of AC mixture is expressed via the sigmoid function, and values of the four parameters of the function are listed in Table 31. 97 Table 31: The E(t) parameters for different AC mixtures Mixture Name PG64-28 Control no-PPA (AI) ALF Control 70-22 SBS 64-40 Terpolymer CR-AZ SBS LG CR-TB Air Blown Fiber Control 70-22 KIM Advera Control Sasobit PPA PPA + Elvaloy SBS SBS + PPA SBS PG64-34 PG64-28 Control no-PPA (AI) PG64-28 with PPA (Hudson) PG64-34 (SEM) PG76-22 (Citgo) PG64-22 with 12% GTR (Gorman) PG64-28 no-PPA(AI)with2.0%Latex WAM Control WAM Foam Asphamin Control Sasobit-1 Sasobit-2 Evatherm-1 a 1.721 0.841 1.342 1.472 0.045 1.351 0.895 0.682 0.561 1.598 2.155 0.558 0.150 1.553 1.586 1.720 1.570 0.851 1.571 1.493 1.689 1.647 1.278 1.515 0.934 1.136 1.272 1.216 1.438 1.289 1.132 b 2.735 3.540 2.832 2.790 4.400 2.842 3.411 3.717 3.772 2.937 1.800 3.645 4.378 2.405 2.306 2.208 2.501 2.388 2.646 2.671 2.391 2.473 2.810 2.666 3.352 3.159 2.971 3.101 2.772 2.984 3.137 c 0.326 0.860 -0.399 -0.095 0.731 0.313 0.634 0.855 1.025 0.512 -0.307 0.532 0.436 -0.100 0.030 -0.211 -0.180 -0.523 0.185 0.135 -0.523 0.247 0.653 0.326 0.789 0.601 0.681 0.774 0.837 0.823 0.692 d -0.491 -0.515 -0.563 -0.584 -0.380 -0.559 -0.428 -0.431 -0.519 -0.562 -0.845 -0.411 -0.314 -0.699 -0.546 -0.725 -0.641 -0.679 -0.666 -0.638 -0.678 -0.702 -0.494 -0.574 -0.599 -0.642 -0.601 -0.559 -0.583 -0.551 -0.552 Evatherm-2 1.227 3.100 0.749 -0.508 The average, minimum and maximum values of each variable are listed in Table 32. 98 Table 32: The mean and boundaries for the four parameters of E(t) Average value Minimum value Maximum value a 1.234 0.045 2.155 b 2.964 1.800 4.400 c 0.356 -0.523 1.025 d -0.570 -0.845 -0.314 5.4 Inverse Solution Theory The objective of any backcalculation solution is to find a set of parameters. In this case, the parameters are the set (a, b, c, d) of the AC mixture, the elastic modulus of base (Ebase), and the elastic modulus of the subgrade layer (Esubgrade). In the end, the calculated deflection history will match the measured values within a specified tolerance. To accomplish this goal, the algorithm repeatedly adjusts the parameter values until a suitable match is obtained. 5.4.1 Basic Theory Background The VE backcalculation algorithm in this research is an extension of the solution used in the MICHBACK program (Harichandran et al., 1994). It uses the modified Newton method to obtain a least squares solution of an over determined set of equations. In the MICHBACK solution, these sets are real-valued and correspond to the peak deflection values. In this research, the author will use deflection time histories, or many deflection basins (corresponding to different times after eliminating the time delay of all sensors), since the proposed backcalculation scheme uses a layered VE solution to predict the time-dependent deflection basins. Ebase and Esubgrade are not known, and the AC layer can be expressed as a function of four parameters (a, b, c, d). The unknown vectors become 99 {X } = {a b c d Ebase Esubgrade } T , assuming all other information for base, sub-base and roadbed layers are known. The vector of measured responses is therefore expressed as: ⎧[w1 (t s ) ... wm (t s )],..., ⎫ {U (t )} = ⎪[w1 (t k ) ... wm (tk )],...,⎪ , ⎬ ⎨ ⎪ w (t ) ... w (t ) ⎪ m f ⎭ ⎩ 1 f T [ (37) ] where m is the number of sensors in the FWD test, wj(ts) is the deflection of sensor j at the starting time and wj(tf) is the deflection of sensor j at the final time of the specified range. Following the derivation by Harichandran et al. (1994), the increment to the unknown parameters in iteration i, {Δx}i , is obtained by solving the linear set of equations: { } ˆ (t ) i + [G ]i {ΔX } = { (t )}, U U i {Uˆ (t )} (38) i where: is the vector of deflections at individual time steps within the specified time range, computed using the estimates of the parameters ˆ {x}i at iteration i, and gradient matrix at iteration i given by: ⎡ ∂ {U ( t )} ⎤ ⎥ [G ] = ⎢ ⎢ ∂ {X} ⎥ {X}={x}i ⎣ ⎦ ˆ i 100 (39) [G ]i is the The partial derivatives in the gradient matrix must be evaluated numerically using: ( ∂w j (t ) ∂xk ˆ {X }={x}i ) ( ) ˆ}i − w j {x}i ˆ w j [R ]{x = , i ˆ rx k j = 1,2,..., m, k = 1,2,3,4,5,6 (40) , th where: [R] is a diagonal matrix with the k diagonal element being (1+r) and all other diagonal elements being one. A separate call to the forward calculation program is required to compute the partial derivatives in each column of the gradient matrix. If there are n individual time steps, Eq. (38) represents a set of m by n equations for 6 unknowns. Since there are more equations than unknown, a more robust method for solving the problem is to use the singular value decomposition (SVD). This method will be briefly introduced later. After the increments for {X } {Δx}i are obtained by solving Eq. (38), the revised parameters are obtained from Eq. (40), as below: ˆ}i +1 = {x}i + {Δx}i ˆ {x (41) The iteration is terminated when the changes in the six parameters are smaller than a set of specified tolerances: 101 ⎛ xk+1 − xk ˆi ˆi abs⎜ ⎜ x i +1 ⎝ ˆk ⎞ ⎟≤ε ⎟ ⎠ k = 1,2,3,4,5,6 (42) 5.4.2 SVD Method SVD is a very powerful set of techniques for dealing with sets of equations or a matrix that are either singular or numerically very close to singular. SVD methods are based on the following theorem of linear algebra (Press et al, 1989). Any M×N matrix [A] whose number of rows M is greater than or equal to its number of columns N, can be written as the product of an M×N column-orthogonal matrix [U], an N×N diagonal matrix [W] with positive or zero elements (the singular values), and the transpose of an N×N orthogonal matrix [V]. The matrix [A] can be decomposed as three matrices as following: [A] = [U ] ∗ [W ] ∗ [V ]T ⎡ω1 ⎢ ω2 ⎢ = [U ]* ⎢ ⎢ ⎣ ⎤ ⎥ ⎥ ∗ [V ]T ⎥ , ⎥ ωN ⎦ (43) [U ]* [U ]T = I , [V ]* [V ]T = I , and [W] is diagonal matrix. where: For the following ill conditioned system of equations: [A]* {X } = {b} Inversing [A] by SVD method, {X } can then be expressed as: 102 (44) ⎡1 ⎢ω ⎢ 1 ⎢ ⎡1⎤ {X } = [V ]* ⎢ ⎥ * [U ]* {b} = [V ]* ⎢ ⎣W ⎦ ⎢ ⎢ ⎢ ⎢ ⎣ T T where: [U ] * [U ] = I , [V ] * [V ] = I , and 1 ω2 ⎤ ⎥ ⎥ ⎥ ⎥ * [U ]* {b} ⎥ , ⎥ 1 ⎥ ωN ⎥ ⎦ (45) ⎡1⎤ ⎢W ⎥ is a diagonal matrix. ⎣ ⎦ 5.4.3 Truncating Singular Values Reference (Press et al., 1989) defines the condition number of a matrix as a ratio of the largest ωi of to the smallest ωj. ωmax Condition number = ωmin (46) A matrix [A] is singular or ill conditioned if its condition number is too large. Reference (Press et al., 1989) suggested this number should be adjusted according to the experiment with the specific problem. Define the threshold of truncation as: ω j < ωmax *10−Threshold , j = 1,2,..., N 103 (47) After selecting the threshold of a condition number, the SVD algorithm will simply 1 replace ωj with zeros. To reduce the possibility of ill conditioning in the inverse problem, the technique of scaling can be applied. In this program, the magnitude of Ebase and Esubgrade is within the magnitude of 100 MPa (14.5 ksi), but the magnitude of the four parameters in the sigmoid function for AC layer is around 0 – 5, as shown in Table 30 & 32. The condition number would be very high, thus, scaling technique is required. Ebase and Esubgrade are scaled by 100 MPa in this backcalculation program. More information can be found in the references (Ji, 2005). 5.5 VE backcalculation of E(t) Besides the above proposed Newton’s method, MATLAB internal function “fminsearch” works very well for backcalculation problems, if the seed value is close to the actual solution. Thus, the proposed Newton’s method is to scan all possible values and eliminate the local minimum values for the error in Eq. (42). Avoiding the local minimum problem in backcalculation is essential, so a random function inside the backcalculation algorithm is used to prevent the solution converging to local minimum th values. A simplified flow chart is shown in Fig. 51, for the i loop of the backcalculation. 104 (i) Input: X Loop: k=1 k=k+1 ( ) ⎡ abs X (i ) ( j ) − X (k ) ( j ) ⎤ (k ) Error = ∑ ⎢ ⎥ X (i ) ( j ) ⎢ ⎥ ⎣ ⎦ 2 Y Error N (k) <ε k