Download as pdf or txt
Download as pdf or txt
You are on page 1of 175

BACKCALCULATION OF PAVEMENT LAYERS MODULI USING 3D NONLINEAR EXPLICIT FINITE ELEMENT ANALYSIS

By

Gergis W. William

Thesis Submitted to the College of Engineering and Mineral Resources at West Virginia University in Partial Fulfillment of the Requirements for the Degree of

Master of Science in Civil Engineering

Samir N. Shoukry, Ph.D., Chair David R. Martinelli, Ph.D. W. J. Head, Ph. D.

Morgantown, West Virginia 1999

ACKNOWLEDGMENTS

For his excellent support and advice in both my academic and research works, for many hours spent improving, evaluating, correcting my work, and his friendship, I would like to express my gratitude to Dr. Samir N. Shoukry whose assistance and guidance made this work feasible. Appreciation is also extended to Dr. David Martinelli and Dr. W. J. Head for serving on the examining Committee.

Thanks to Mr. George Hanna, West Virginia Division of Highways, who performed the field tests using FWD. I would like also to thank Professor Per Ullidtz, Technical University of Denmark, who provided some of his results for comparison with my results listed in Chapter 8, Table 8.2.

I would like thank to my parents for their love, support, and understanding and for being there when times were rough. I would like also to thank my sister and my grand father for their love and support Thanks are also extended to my friends for their encouragement and help.

The author gratefully acknowledges the financial support for this research from West Virginia Division of Highways, and Mid Atlantic Universities Transportation Center.

ii

ABSTRACT BACKCALCULATION OF PAVEMENT LAYERS MODULI USING 3D NONLINEAR EXPLICIT FINITE ELEMENT ANALYSIS by: GERGIS W. WILLIAM After reviewing the existing literature on FWD testing and backcalculation algorithms, a new backcalculation approach based on Three Dimensional Finite Element Modeling (3D FEM) was developed. This approach accounts for the transient dynamic nature of FWD load, the three dimensional geometry of the pavement structure, and the friction and bonding characteristics of pavement layers interfaces. The 3D FEM backcalculation approach was used to backcalculate the layers moduli of flexible, rigid, and composite pavement sites located in West Virginia. The layers moduli of each site were also evaluated using three widely used backcalculation algorithms: MODULUS, EVERCALC, and MODCOMP. Comparison of their results with those obtained using 3D FEM revealed that the former should be multiplied by correction factors in order to match the latter. Using 3D FEM backcalculation results as reference values, correction factors were developed for each program and pavement type. The mechanistically evaluated correction factors were found to be in close agreement with the experience-based factors recommended for flexible and rigid pavements in the American Association of State Highway and Transportation Officials (AASHTO) Pavement Design Guide. The 3D FEM approach was also used to predict the apparent depth to bedrock. The decay of vertical stress and displacement in the subgrade layer were examined and used to predict the apparent depths to bedrock for four pavement sites located in Texas. The 3D FEM results were found to be in good agreement with the measured values provided by Texas DOT. A parametric study was conducted to evaluate the effect of the subgrade layer thickness (assumed in the finite element pavement structural model) on the stress and deformation obtained on top of the subgrade layer and on the 3D FEM-generated deflection basin. It was found that a subgrade layer thickness of 6 ft. would produce satisfactory results. The effect of concrete slab length on the deflection basin was examined for both doweled and undoweled concrete slabs. For doweled concrete pavements, slab length has no effect on the deflection basin. For broken slabs or undoweled ones, the minimum slab length required to produce an acceptable deflection basin was found to be 10 ft. KEYWORDS: Falling Weight Deflectometer, Pavement Evaluation, Nondestructive testing, Backcalculation, Finite Element Analysis of Pavements. iii

TABLE OF CONTENTS
ACKNOWLEDGMENT ABSTRACT TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES CHAPTER 1 INTRODUCTION 1.1 1.2 1.3 BACKGROUND BACKCALCULATION PROBLEMS OBJECTIVE OF THIS THESIS 1.3.1 Organization of this Thesis 1 1 4 5 6 ii iii iv vi vii

CHAPTER 2 LITERATURE REVIEW 2.1 2.2 2.3 BACKCALCULATION PROGRAMS PROGRAM SELECTION DESCRIPTION OF SELECTED BACKCALCULATION PROGRAMS 2.3.1 2.3.2 2.3.3 2.3.4 Description of MODCOMP3 Program Description of MODULUS5.0 Program Description of EVERCALC4.0 Program Description of ELMOD4 Program 8 8 9 10 11 12 13 15

CHAPTER 3 COMPARISON OF BACKCALCULATION RESULTS FOR SHRP TEST SECTIONS 20

CHAPTER 4 COMPARISON OF BACKCALCULATION RESULTS FOR SEVEN SELECTED SITES IN WEST VIRGINIA 34

CHAPTER 5 FIELD TESTING 5.1 5.2 PAVEMENT SITES MEASUREMENTS 66 66 67

iv

5.3 5.4 5.5

DEFLECTION TESTING EVALUATION OF BACKCALCULATED MODULI CONCLUSIONS

69 70 70

CHAPTER 6 EVALUATION OF BACKCALCULATION ALGORITHMS THROUGH FINITE ELEMENT MODELING OF FWD TEST 6.1 6.2 6.3 6.4 INTRODUCTION REVIEW OF FINITE ELEMENT MODELING OF PAVEMENTS GUIDELINES FOR 3D-FEM OF PAVEMENT STRUCTURES FINITE ELEMENT MODELS OF EXPERIMENTAL TEST SITES 6.4.1 6.4.2 6.4.3 6.5 6.6 6.7 Flexible Pavement Model Rigid Pavement Model Composite Pavement Model 76 76 76 78 81 81 81 82 83 84 85 85 86 87

STRUCTURAL MATERIAL MODELING LAYERS MODULI EVALUATION FROM FE MODEL RESULTS VERIFICATION OF FINITE ELEMENT MODELS 6.7.1 6.7.2 Deflection basins Displacement time-history

6.8

EVALUATION OF BACKCALCULATION PROGRAMS

CHAPTER 7 SOME FACTORS INFLUENCE BACKCALCULATIONS OF RIGID PAVEMENTS 7.1 7.2 INTRODUCTION FINITE ELEMENT STRUCTURAL MODEL 7.2.1 7.2.2 7.3 7.4 7.5 Model Loading and Material Model Model Verification 103 103 103 105 106 107 108 109

PERFORMANCE ASSESSMENT OF BACKCALCULATION PROGRAMS EFFECT OF SLAB LENGTH AND DOWEL BARS CONCLUSIONS

CHAPTER 8 EFFECT OF 3D FEM MODEL DEPTH ON BACKCALCULATIONS OF FLEXIBLE PAVEMENTS 8.1 8.2 INTRODUCTION FINITE ELEMENT STRUCTURAL MODELS 117 117 121

8.3 8.4 8.5

EVALUATION OF LAYERS MODULI EVALUATION OF DEPTH TO BEDROCK EFFECT OF 3D FE MODEL DEPTH ON 3D FEM RESULTS 8.5.1 8.5.2 8.5.3 8.5.4 Effect of Model Depth on Deflection Basin Effect of Stress Wave Reflection from Model Bottom on Deflection Basin Effect of Model Depth on the 3D FEM-Calculated Depth to Bedrock Effect of Model Depth on Stresses Induced in Subgrade

122 123 125 126 126 126 127 128 128

8.6 8.7

EFFECT OF LOAD DURATION ON THE DEFLECTION BASIN CONCLUSION

CHAPTER 9 CONCLUSIONS 144

REFERENCES

148

APPENDIX A

PARAMETRIC INPUT FILE TO GENERATE FLEXIBLE PAVEMENT MODEL

158

VITA

166

vi

LIST OF TABLES
TABLE 2.1 TABLE 2.2 TABLE 3.1 TABLE 3.2 TABLE 4.1 TABLE 4.2a TABLE 4.2b TABLE 4.2c TABLE 4.3 TABLE 4.4 TABLE 5.1 TABLE 5.2 TABLE 6.1 TABLE 7.1 TABLE 8.1 TABLE 8.2 TABLE 8.3 TABLE 8.4 Available Backcalculation Programs. Comparison of Selected Backcalculation Programs. Moduli Range and Poissons Ratio for Backcalculation Inputs. Backcalculated Layer Moduli (Ksi) for SHRP Sections. Assumptions Used in Different Backcalculation Algorithms. Backcalculated Layer Moduli (Ksi) for Flexible Pavement Sections. Backcalculated Layer Moduli (Ksi) for Composite Pavement Sections. Backcalculated Layer Moduli (Ksi) for Rigid Pavement Sections. Comparison of Pavement Moduli Backcalculated by Different Programs. Number of Backcalculated Out of Range Moduli. Deflection Data Collected from Field Tests. Backcalculated Layer Moduli (Ksi) for Tested Pavement Sections. Correction Factors for Backcalculated Subgrade Modulus. Properties of Pavement Materials. Layer Thicknesses and Temperature Measurements. Backcalculated Pavement Layers Moduli, Ksi. Comparison Between Measured and Calculated Depth to Bedrock. Effect of Assumed Model Depth on the Calculated Depth to Bedrock. 18 19 23 24 38 40 41 42 43 45 72 73 90 105 130 131 132 133

vii

LIST OF FIGURES
FIGURE 1.1 FIGURE 3.1 FIGURE 3.2 FIGURE 3.3 FIGURE 3.4 FIGURE 3.5 FIGURE 3.6 FIGURE 3.7 FIGURE 3.8 FIGURE 3.9 FIGURE 4.1 FIGURE 4.2 FIGURE 4.3 FIGURE 4.4 FIGURE 4.5 FIGURE 4.6 FIGURE 4.7 FIGURE 4.8 FIGURE 4.9 FIGURE 4.10 FIGURE 4.11 FIGURE 4.12 FIGURE 4.13 FIGURE 4.14 FIGURE 4.15 FIGURE 4.16 FIGURE 4.17 FIGURE 4.18 FIGURE 4.19 FIGURE 4.20 FIGURE 5.1 FIGURE 5.2 FIGURE 5.3 FWD setup and schematic presentation of stress bulb. SHRP sections data used for backcalculations. Comparison of backcalculated moduli for SHRP section A. Comparison of backcalculated moduli for SHRP section B. Comparison of backcalculated moduli for SHRP section C. Comparison of backcalculated moduli for SHRP section D. Comparison of backcalculated moduli for SHRP section E. Comparison of backcalculated moduli for SHRP section F. Comparison of backcalculated moduli for SHRP section G. Comparison of backcalculated moduli for SHRP section H. Backcalculated layer moduli (Ksi) for flexible pavement section US52_pre station 16.250. Backcalculated layer moduli (Ksi) for composite pavement section US52_pre station 18.360. Backcalculated layer moduli (Ksi) for composite pavement section US52_pre station 19.179. Backcalculated layer moduli (Ksi) for composite pavement section US60_smi station 1.230. Backcalculated layer moduli (Ksi) for composite pavement section US60_smi station 2.267. Backcalculated layer moduli (Ksi) for composite pavement section US60_smi station 3.393. Backcalculated layer moduli (Ksi) for composite pavement section WV2_frie station 4.296. Backcalculated layer moduli (Ksi) for composite pavement section WV2_frie station 6.100. Backcalculated layer moduli (Ksi) for composite pavement section WV2_frie station 7.333. Backcalculated layer moduli (Ksi) for flexible pavement section WV3_whit station 40.806. Backcalculated layer moduli (Ksi) for flexible pavement section WV3_whit station 41.013. Backcalculated layer moduli (Ksi) for flexible pavement section WV3_whit station 41.221. Backcalculated layer moduli (Ksi) for flexible pavement section WV71_blu station 0.009. Backcalculated layer moduli (Ksi) for flexible pavement section WV71_blu station 1.259. Backcalculated layer moduli (Ksi) for flexible pavement section WV71_blu station 2.740. Backcalculated layer moduli (Ksi) for rigid pavement section US2_moun station 18.250. Backcalculated layer moduli (Ksi) for rigid pavement section US2_moun station 19.028. Backcalculated layer moduli (Ksi) for rigid pavement section US19_hic station 20.856. Backcalculated layer moduli (Ksi) for rigid pavement section US19_hic station 21.563. Backcalculated layer moduli (Ksi) for rigid pavement section US19_hic station 21.839. Flexible pavement section. Rigid pavement section. Composite pavement section. 8 25 26 27 28 29 30 31 32 33 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 65 74 74 74

viii

FIGURE 5.4 FIGURE 6.1 FIGURE 6.2 FIGURE 6.3 FIGURE 6.4 FIGURE 6.5 FIGURE 6.6 FIGURE 6.7 FIGURE 6.8 FIGURE 6.9 FIGURE 6.10 FIGURE 6.11 FIGURE 6.12 FIGURE 7.1 FIGURE 7.2 FIGURE 7.3 FIGURE 7.4 FIGURE 7.5 FIGURE 7.6 FIGURE 7.7 FIGURE 7.8 FIGURE 8.1 FIGURE 8.2 FIGURE 8.3 FIGURE 8.4 FIGURE 8.5 FIGURE 8.6 FIGURE 8.7 FIGURE 8.8 FIGURE 8.9 FIGURE 8.10 FIGURE 8.11

Instrumentation layout. Finite element mesh of the flexible pavement model. Finite element mesh of the rigid pavement model. Finite element mesh of the composite pavement model. Impact loading curves used in different finite element models. Load-Deflection relation for different types of pavements. Fringes of vertical Stress at time of maximum FWD load. Vertical stresses in different types of pavements due to FWD load. Comparison between experimental and FE deflection basins for different pavements models.. Fringes of vertical displacement at time of maximum FWD load. Deflection-time histories for different pavement models. Comparison between backcalculated deflection basins and measured basins. Comparison of backcalculated layer moduli for the three types of pavements. Finite element mesh for a rigid pavement. Cross section in a doweled joint. Impact load curve used in finite element model (Ref. (48)). Model Verification. (Slab length=20 ft) Comparison between the results of backcalculation programs. Effect of slab length on deflection basin. Change of maximum deflection with slab length for undoweled pavement. Effect of slab length on the backcalculation results ( Using MODULUS). Finite element model. Measured FWD impact load curves used in finite element models. Measured and FE-calculated deflection basins. Subgrade vertical displacement versus depth. Subgrade vertical displacement on a logarithmic scale versus depth. Decay of vertical stress and displacement in subgrade. Effect of the depth to bedrock on the deflection basin. Effect of reflective subgrade bottom on deflection basin. Subgrade vertical displacement for different model depths. Effect of model depth on vertical stress distribution for site 3. Effect of load duration on the deflection basin.

75 91 92 93 94 95 96 97 98 99 100 101 102 110 110 111 112 113 114 115 116 134 135 136 137 138 139 140 141 142 143 144

ix

CHAPTER 1

INTRODUCTION

1.1

BACKGROUND

Backcalculation is an analytical procedure in which the deflection data collected during a Falling Weight Deflectometer (FWD) test are used to predict the moduli of different pavement layers. The backcalculation procedure involves theoretical calculations of the deflections produced under a known applied load using an assumed set of layers moduli. The theoretical deflections are then compared with those measured during the test. In case of differences between the theoretical and measured deflections, the assumed pavement layers moduli are adjusted and the process is repeated until the differences between the theoretical and measured values fall within acceptable limits. Techniques like iteration, database searching, regression analysis, and artificial intelligence (neural networks) have been used as backcalculation tools (1,2)1.

Most of the existing backcalculation algorithms use iterative analysis. In this case, the solution for the theoretical deflections is initiated at the distant sensor locations assuming that surface deflections at the distant sensor positions are due to strains or deflections in the subgrade layer only and independent of the overlaying layers (3-5). As shown in Figure 1.1, the stress zone intersects the interface between the subbase and subgrade layers at the radial distance r = ae. This means that any surface deflection value obtained from the deflection basin at or beyond the distance r = ae is due only to the deformation within the subgrade layer. Thus, in-situ modulus of the subgrade can be evaluated from the known values of measured deflections at the outer geophone positions. From the above discussion, ae value is obviously very important. The AASHTO Guide for Design of Pavement Structures (1993), Section 5.4.5, Step4: Deflection Testing points out that the deflection used to backcalculate the subgrade modulus must be measured far enough from the center of load application

Numbers in parentheses refer to numbers of references.

so that it provides a good estimate of the subgrade modulus. The guide suggests that minimum distance may be determined from the following relationship: r

 0.7a e

the following formula is provided for the determination of a e (5):

ae

a 2 D(3

Ep MR

where

ae = a =

radius of a stress bulb at the subgrade-pavement interface, inches; FWD load plate radius, inches;

D = total thickness of pavement layers above the subgrade, inches; EP = effective modulus of all pavement layers above the subgrade, psi; MR = subgrade resilient modulus, psi.

The AASHTO guide gives a graph for the determination of the ratio EP/ MR for a six inch radius loading plate, known total thickness of pavement layers above the subgrade, known maximum deflection under the center of the loading plate, and known load magnitude.

The following aspects should be considered for the evaluation of subgrade modulus from FWD data:

1.

Different types of pavement structures produce a different spread of stress zone (stress bulb). Flexible pavements are characterized by localized stress distribution while rigid pavements have a wider spread of stress zone. The stress zone in composite pavements falls between the above two, depending on: the type of overlay, construction practice, and condition of the second layer (old top layer). Calculation of ae is useful in the determination of geophone location for the evaluation of subgrade moduli. 2

2.

Accuracy of measurement progressively diminishes at distant geophone locations due to the fact that the displacement magnitude is close to the resolution accuracy of the geophone. Measurements below 2.5 mils (0.0025 inch) should be processed with care.

3.

High load magnitude could be an additional source of error in the evaluation of linear resilient subgrade modulus since it is likely to produce nonlinear deformations in the soil layer. AASHTO Guide for Design of Pavement Structures (1993), Section 5.4.5, suggests the use of a load magnitude of approximately 9,000 lb to avoid nonlinearity of the subgrade response.

Subgrade layers are the most complex layers in the pavement structure due to their physical nature and construction practices. Undisturbed natural soil deposits frequently reveal an increase in the value of elastic modulus with depth, due to higher consolidation of geologic material. Therefore, if such a soil sample is divided into layers, the top layers will have lower moduli than the lower ones. The backcalculation process is based on the assumption that surface deflection at a certain offset is characteristic of the elastic modulus at a certain depth. However, deflections measured at distant locations are very small and cannot be measured accurately based on geophone resolution. Higher loads can produce deflections of large magnitudes, but unfortunately they will also cause stresses in the soil that fall into the nonlinear zone. This again alters the accuracy of linear elastic modulus evaluation for deeper layers. The only subgrade modulus that can be accurately predicted by FWD test is the resilient modulus of the top of the subgrade layer positioned right under a base-subgrade interface.

During the 1998 annual Transportation Research Board (TRB) Meeting of the Committee on Backcalculation of Pavement Moduli, a specific question regarding accuracy of subgrade moduli calculation was addressed to L. Irwin, the developer of the MODCOMP3 backcalculation program. He suggested improving accuracy of subgrade modulus evaluation by dividing the soil layer into a larger number of thin sublayers. Regarding the depth of top subgrade layer, he suggested calculating the depth by subtracting the height of pavement structure above the subgrade from the depth of frost penetration value, which is 30 to 36 inches for West Virginia. 3

Assumption of a thicker layer in backcalculations consistently produced higher subgrade modulus value. This observation shows that a thicker layer assumption leads to overestimation of subgrade moduli, which is an undesirable outcome.

In the backcalculation procedure, the modulus of the lowest subgrade layer is changed in the iterative procedures until such value is found that will produce surface deflection at the distant geophone that agrees with the measured one, within a certain tolerance. Once the modulus value for the lowest layer is found, it is assumed to be the true value and is used as a constant in the evaluation of moduli for the upper layers. The solution progresses from the distant geophone locations to the center of load application and the layers moduli are evaluated from the bottom to top layers. Knowledge of existing pavement layer thicknesses, Poissons ratios, and load magnitude are necessary conditions for the backcalculation procedure. The pavement layers moduli derived in the above process are further used for the stress and strain analysis of the pavement structure. The moduli are usually averaged from the results of several load drops in order to lower the weight of the random error that is introduced during every FWD drop.

The major assumption used in backcalculation algorithms is that the amount of surface deflection at any point is dependent on the stress-strain state in subsequent layers. This assumption is true for static loading, but may be violated if the load is dynamic, or if there is discontinuity (separation) at one or more interfaces.

1.2

BACKCALCULATION PROBLEMS

Although FWD load is an impact load by nature (6-9), most of available backcalculation programs are based on the static multilayer elastic theoty. The major limitation of the elastostatic analysis of the FWD load is that it does not account for factors such as material inertia and damping . As noted by Hoffman and Thompson (10), inertial effects in the pavement layers subjected to FWD impact may be significant and therefore need to be considered in the theoretical analysis. 4

As pointed out by Chou et al. (11), none of the programs based on the static multilayered elastic theory could guarantee accurate results for every test section. Chou et al. stated that two independent agencies who used the same backcalculation software to determine moduli for the same pavement section produced different results. Therefore, engineering judgement plays an important role in the evaluation of the test results.

Irwin et al. (12) reported that most errors occur during the evaluation of the surface layer modulus and Huang (4) stated that this is especially true in the case of thin asphalt layers. One reason may be the difference between patterns of dynamic and static deformations of the surface layer.

1.3

OBJECTIVES OF THIS THESIS

Most of the existing backcalculation algorithms are based on simplifying approximations which include that: 1. the pavement structure has infinite extension perpendicular and along the traffic direction, 3. all pavement layers are fully bonded, and 4. the short duration loading impulse (25 millisecond) applied by the FWD may be approximated as a static applied load.

The work presented in this thesis is directed towards developing a backcalculation technique in which the actual pavement geometry, the full FWD loading pulse time-history, and the properties of the interfaces between pavement layers are accounted for in the backcalculation of the pavement moduli profile. The structural moduli profile obtained using such a backcalculation technique could be used to evaluate the performance of existing backcalculation programs. It could also be used in situations where existing backcalculation algorithms would produce unrealistic results such as in the case of composite pavements. To achieve this objective, Three Dimensional Finite Element Modeling (3D FEM) was used to simulate the pavement structure. The use of 3D FEM allows loading the model 5

with the exact FWD loading-time history measured from FWD load cell, as well as representation of the properties of interfaces between different layers. Although elastic material models were assumed for all pavement layers (based on the experimental measurement of FWD load-deflection relation), the 3D FEM approach allows the use of any nonlinear material models including thermo-elastic and thermoplastic material models.

1.3.1 Organization of this Thesis

The work presented in this Thesis starts in Chapter Two which contains a review of the research studies which aimed at evaluating the performance of existing backcalculation algorithms. Two major studies were reviewed in some detail: the Strategic Highway Research Program ( SHRP) study whose results were first published in 1993 and the MnDOT study published in 1996. The outcome of this review was the selection of four major backcalculation programs that were the most widely used by the pavement community. In Chapters Three and Four, the results from the four programs were compared to each other using SHRP-LTPP data and using data for seven different roads measured in West Virginia. Chapter Five outlines the FWD testing procedures used for testing three pavement sites in West Virginia. 3D FEM backcalculation procedures were presented in Chapter Six. The 3D FEM-backcalculated layer moduli were compared with those obtained using the previously selected conventional backcalculation algorithms and correction factors were obtained. Some factors which influence the results of backcalculation of rigid pavements were examined in Chapter Seven. The factors which influence the backcalculation of flexible pavements were examined in Chapter Eight. Chapter Nine presents the conclusions drawn from this work along with some points for future research.

FIGURE 1.1 FWD setup and schematic presentation of stress bulb.

CHAPTER 2

LITERATURE REVIEW

2.1

BACKCALCULATION PROGRAMS

In 1993, a comprehensive review of existing software for backcalculation procedures was published as a result of Project SHRP-90-P-001B, sponsored by the Strategic Highway Research Program (13). After considering a large number of different backcalculation programs, the study was narrowed down to the following three: MODULUS, WESDEF, and MODCOMP3. The programs were tested by a group of independent researchers and practicing engineers for different types of pavement structures. Evaluation criteria that were considered are the repeatability of results obtained by different users, reasonableness of results, deflection matching errors, ability to match assumed moduli from simulated deflection basins, and versatility.

As a result of this evaluation, the MODULUS program was judged to be the best and was therefore chosen for routine use on SHRP and LTPP data. Nonetheless, the MODULUS program had one drawback that it required deflections to be measured at the specific sensor locations. This means that some of the points in the deflection basin will be excluded from the matching procedure. The results from WESDEF program were found to have the highest level of sensitivity to user input. However, in the same report it was pointed out that this could have been due to the default depth to bedrock that was not overridden by the users in cases of semi-infinite subgrade layers. The authors concluded that the MODCOMP3 program tends to over predict subgrade modulus and under predict base and subbase moduli.

Research was conducted recently by the Minnesota DOT to select a backcalculation software for the Minnesota Road Research Project (14). The strategy of that study was to compare the performance of the candidate programs for flexible pavement evaluation in terms of usability and accuracy of backcalculation results. Four different programs were evaluated: EVERCALC3.3, 8

EVERCALC4.0. WESDEF, and MOCOMP3. All four programs were based on the multilayer linear elasto-static forward calculation subroutines. As a result of this project, the EVERCALC4.0 program was recommended for routine research of the Minnesota Road Research test sections.

2.2

PROGRAM SELECTION

Information about sixteen different backcalculation programs available on the market was collected and is presented in summarized form in Table 2.1. The selection of programs to be evaluated during the current project was based upon the following recommendations:

1. 2.

SHRP Report on Layer Moduli Backcalculation Procedure, 1993 (13). Minnesota Road Research Project Report on Selection of Flexible Pavement Backcalculation Software, 1996 (14).

As a result, four backcalculation programs were selected for evaluation:

   

MODCOMP3 MODULUS5.0 EVERCALC4.0 ELMOD4

Developed at Cornell University Developed by Texas Transportation Institute Used by Washington State DOT Developed by Dynatest

Evaluation of the above programs using data from eight SHRP test sections and twenty-one sections for seven different roads in West Virginia is presented in this report.

2.3

DESCRIPTION OF SELECTED BACKCALCULATION PROGRAMS

2.3.1 Description of MODCOMP3 Program

MODCOMP3 was initially developed by Irwin and Speck for the U.S. Army Cold Regions Research and Engineering Laboratory. Later, MODCOMP3 was modified by Irwin and Szebenyi (15) in 1991 at Cornell University.

The MODCOMP3 program uses CHEVRON computer code as a forward calculation subroutine. This code is based on the multilayer linear elasto-static theory that is traditionally used for flexible pavement analysis. Fitting the calculated results with the experimentally obtained ones is implemented through an iterative analysis approach. The modulus of each layer is assumed to affect the deflection measured at a certain distance from the load. The program first evaluates the modulus of the deepest layer corresponding to the deflection reading at the farthest geophone location specified and then it works upwards to calculate the moduli of near surface layers. The drawback of the program is that the user is responsible for the accurate prediction of the exact distance that would reveal the deflection characterizing the modulus of a particular layer.

The program can handle from two to fifteen layers, including the bottom layer that is treated as a semi-infinite half space. Layers moduli can be treated both as known or unknown. No more than five unknown layers are recommended for use in the analysis. Based on the readings at distant geophone locations, the program has the ability to calculate the depth to bedrock. Up to ten geophone locations can be included in backcalculation. The program can accept up to six different loading cases in the analysis. Layers can be treated both as linear elastic or nonlinear elastic.

The backcalculation analysis is initiated by the set of initial seed moduli specified by the user. Since no moduli range can be specified, therefore the algorithm has the potential of producing out of range moduli.

10

The process of backcalculation terminates when one of the following conditions is met: 1. 2. 3. The deflection fit precision tolerance is satisfied. The modulus convergence tolerance is satisfied. The allowed number of iterations is exhausted.

At the end of each iteration, MODCOMP3 checks two tolerances before proceeding to the next iteration. If either of the tolerances is satisfied, the calculations cease.

The deflection fit precision is a check on the match between the measured deflections and the deflections that have been calculated at the same radii using the backcalculated moduli. Only the sensors that are assigned to the layers are included in this check. This means that the whole deflection basin may not be used in the backcalculation process. For example, if the pavement structure consists of three layers and the FWD test setup has seven geophones, the user will need to choose three out of seven geophone readings to backcalculate pavement moduli. In this case, initially, different geophone selections will produce different sets of moduli; then, the deflection basin built using the backcalculated moduli will fit the measured basin only at the selected points. So the program leaves little opportunity to achieve a full match between measured and calculated basins. The accuracy of the solution is highly dependant on the success in the selection of the right geophones for backcalculation of pavement moduli.

The modulus convergence is a check on the rate of change of the backcalculated moduli from one iteration to the next. A tolerance of 1.5 percent is typically assigned; but if a value close to zero is used, this tolerance check is eliminated because it will not be satisfied before either the deflection fit precision tolerance is achieved, or the allowed number of iterations is exhausted.

2.3.2 Description of MODULUS5.0 Program

MODULUS5.0 was developed by Texas Transportation Institute (16,17). This computer 11

code uses WESLEA program as a forward calculation subroutine. WESLEA is based on the multilayer linear elasto-static theory that is traditionally used for the purposes of flexible pavement analysis. The WESLEA subroutine is used to build a data base for the calculated deflection basins of a given pavement system. A pattern search technique is then used to determine the set of layers moduli that produce a deflection basin that fits the measured one.

The deflection data collected during the FWD test can be read directly from the DYNATEST FWD field data diskette or manually typed by the user. Up to eight different load drops at each station can be considered in the backcalculation analysis, and up to seven geophone locations can be included in the backcalculation.

The maximum number of unknown layers is limited to four in order to minimize the errors and produce acceptable results. The depth to the stiff layer can be automatically calculated in the program. In addition, the user can specify different depth values.

MODULUS5.0 uses weighting factors assigned to each geophone deflection reading. Different values of weighting factors can maximize or minimize the significance of a certain deflection value in the backcalculation process. If the weighting factors are not specified by the user, the program implements an automatic weighting factor determination algorithm. In this case, if a nonlinear subgrade stiffening with depth is detected, the backcalculation program automatically drops a maximum of one sensor reading by assigning to it a zero weighting factor.

The backcalculation program has two analysis options. In the first option called Full Analysis, the program asks the user to specify the moduli range to be used in the backcalculation. The user can specify moduli ranges for up to three layers except the subgrade, and provides a seed modulus value for the subgrade layer. This feature prevents the program from producing out-ofrange moduli for all layers except the subgrade. Although the program allows the user to enter up to seven different locations, the manual has a note stating that in MODULUS5.0 the sensors must be placed at 0, 12, 24, and 36 in.. 12

In the second option called Material Types, the user is asked to input only the material types and layers thicknesses and the program automatically selects acceptable ranges of moduli and Poissons Ratios. Both options produce similar outputs for the same types of pavement layer materials. However, it is worth noting that this program was designed in Texas and oriented to the pavement design practices used there. Thus, the material selection table is limited to those materials that are frequently used in Texas and cannot cover all the wide range of pavement materials used in road construction in the USA. As a result, it is more appropriate to use the first option of the program in the current backcalculation evaluation process.

Both computed and measured deflection basins can be plotted on the screen. This feature provides the user with a visual assessment of the accuracy of the program.

2.3.3 Description of EVERCALC4.0 Program

EVERCALC4.0 was developed by J. Mahoney at the University of Washington for Washington DOT (18). The EVERCALC4.0 uses the WESLEA computer code developed by Waterways Experimental Station, U.S. Army Corps of Engineers, as a forward calculation subroutine and a modified augmented Gauss-Newton algorithm for solution optimization. The WESLEA computer code is based on the multilayer linear elasto-static theory that is traditionally used for purposes of flexible pavement analysis. The fit of the calculated deflection basin with the

experimentally measured one is implemented through the iterative analysis approach.

The program can handle up to seven sensor measurements and eight drops per section. Flexible pavement sections containing up to five layers can be evaluated.

To initialize the backcalculation process, initial moduli, as well as moduli ranges, need to be specified for all the layers. This feature prevents the program from producing out-of-range moduli for all the layers. At the end of each iteration, the deflections calculated by WESLEA are compared 13

with the measured ones. The discrepancies between the calculated and measured deflections are characterized by Root Mean Square (RMS) error. The iterations are terminated when one of the following three conditions are satisfied: (1) RMS falls within the allowable tolerance, (2) the changes in modulus between two successive iterations fall within the allowable tolerance, or (3) the number of iterations has reached the maximum limit. When two or more deflection data sets for the same location are analyzed, the final moduli from the first deflection data is used as a seed moduli for the next one.

If the pavement section contains no more than three layers, the program can assign the seed moduli internally. In this case, a set of regression equations is used to determine a set of seed moduli from the relationships between the layers moduli, surface deflections, applied load, and layer thickness.

A stiff layer that has a known modulus can be included in the analysis. In this case, the depth to the stiff layer will be calculated by the program. Inclusion of the stiff layer in the analysis normally results in a decrease of the subgrade modulus and an increase of the modulus of the layer above the subgrade (subbase or base) layer.

The program is also able to normalize the modulus of the Asphalt Concrete (AC) layer to that evaluated at the standard laboratory conditions (for a temperature of 77o F and 100-millisecond loading time).

A disadvantage of the EVERCALC4.0 program is that the output files are stored in binary format, which complicates the communication of EVERCALC4.0 with any other software when needed (such as an external data base of deflection basins, etc.).

The FWD raw deflection data file can be used for the direct input if DYNATEST FWD model 8000 is used. In this case, the EVERCALC4.0 program will internally convert FWD data to the EVERCALC4.0 deflection data file. Otherwise, deflection data should be entered manually and 14

saved in a deflection data format provided by EVERCALC4.0. The EVERCALC4.0 program provides excellent visualization of the results in a variety of graphs and bar charts. The program can handle English or Metric units.

2.3.4 Description of ELMOD4 Program

ELMOD4 was developed by Dynatest Consulting Inc. ELMOD is an acronym for Evaluation of Layer Moduli and Overlay Design. The program accepts DYNATEST-FWD file format. If such a file does not exist, no means for manual input of deflection data is provided by the program. The ELMOD4 program is capable of calculating pavement layer moduli using one of these two options: the Odemark-Boussinesq transformed section approach or the deflection basin fit backcalculation. The first method can be used for one to four layers pavement systems. The second method can handle up to five layers. In both methods, the dependancy of AC material on temperature conditions is accounted for. In order to use this feature of the program, the 24-hour air temperature should be entered in the Structural Data input screen.

Odemarks layer transformed (structural) section approach is used in conjunction with Boussinesqs equations to calculate deflections. An iterative procedure is used to determine layers moduli which results in the same deflections as measured by a FWD. This method provides the apparent moduli for the as-measured deflections at each FWD test point, it also takes the nonlinearity of the subgrade into consideration. This approach is very reliable for flexible three layered pavement systems that include an unbound base layer plus the bound surface layer . Four layer flexible systems may also be evaluated using the Odemark-Boussinesq mode provided that the ratio E2/E3 is known. The program includes a procedure for the automatic calculation of this ratio for granular materials from the thickness of layers number two and three. For the backcalculation input, a layer can consist of several materials. To better simulate the theoretical conditions on which the backcalculation approach is based, the program developers suggest combining an asphalt layer with an adjacent gravel layer or another stabilized layer. 15

A total of five layers, including the subgrade, may be backcalculated if the deflection basin fit backcalculation option is utilized; however, many layers are not recommended unless one or more of them are assigned some fixed modulus. The deflection basin fit backcalculation method is done either by using the normal Odemark transformed section factors (0.8 for multi-layered systems and 0.9 for two-layered systems) and adjusting the moduli accordingly, or by recalculating the factors by a numerical integration forward calculation procedure using a modified version of WES5 and then performing the deflection basin fit backcalculation using calibrated transformed section factors. The developers of the program recommend the use of the first approach for backcalculation.

ELMOD4 can also provide a theoretical estimate of the equivalent depth to a rigid layer from the measured deflections. If this option is not chosen, ELMOD4 considers an infinitely thick nonlinear subgrade. In case of a stiff layer input, linear elastic behavior of the finite subgrade layer is assumed. ELMOD4 calculates the equivalent depth of an apparently stiff layer and compares this calculated depth with the users maximum depth input; if it is less than the input depth it uses the calculated depth to perform the analysis. If the calculated depth is greater than the input depth, the analysis reverts to the consideration of a semi-infinite (nonlinear) subgrade.

ELMOD4 backcalculates the elastic moduli of up to five layers provided that the following conditions are met: The structure should contain only ONE bound upper (stiff) layer, where E1/Esubgrade 5. If the structure contains more than one bound layer, they should be combined into one analysis for the purpose of structural evaluation.

1.

2.

The moduli should decrease with depth by approximately Ei/Ei+1>2. Where Ei is the modulus of the ith layer counted from the surface.

3.

The thickness of the upper (bound) layer (H1) should be larger than half the radius of the loading plate (generally > 75 mm or 3 inches). 16

4.

For three layered structures, the thickness of the upper layer should be less than the diameter of the loading plate and the thickness of layer one should be less than that of layer two (H1< H2).

5.

When testing on/near a joint, a very wide crack, on gravel surfaces, the structure should be treated as a two-layer system.

Another limitation of the ELMOD4 program is that Poissons Ratios are assumed to be 0.35 for all layers. This is suitable for asphaltic and unbound granular materials, but different from the values for the cohesive soils (0.42-0.45) and concrete (0.15-0.18).

The main features of the above four reviewed programs are summarized in Table 2.2.

17

TABLE 2.1

Available Backcalculation Programs.

Name of the main program BISDEF CHEVDEF CLEVERCALC

Name of subroutine for pavement analysis BISAR CHEVRON CHEVRON

Theoretical method for pavement analysis Multilayer elastic Multilayer elastic Multilayer elastic

Backcalculation method Iterative Iterative Iterative

Source

USACE-WES USACE-WES Royal Institute of Technology, Sweden M. Anderson Texas A&M University, USACE-WES PCS/LAW J. Mahoney W. Uddin University of Illinois University of Illinois R. Stubstud L. Irwin, Szebenyi Texas Transportation Institute S.F.Brown et al. USACE-WES Michigan State University

COMDEF ELSDEF

BISAR ELSYM5

Multilayer elastic Multilayer elastic

Data Base Iterative

EMOD EVERCALC FPEDDI ILLIBACK ILLI-CALC

CHEVRON CHEVRON BASINPT ILLIBACK ILLIPAVE

Multilayer elastic Multilayer elastic Multilayer elastic Plate on elastic foundation theory Nonlinear elastostatic finite element modeling Multilayer elastic Multilayer elastic Multilayer elastic

Iterative Iterative Iterative Closed form solution Iterative

ISSEM4 MODCOMP MODULUS

ELSYM5 CHEVRON WESLEA

Iterative Iterative Data Base

PADAL WESDEF MICHBACK

PADAL WESLEA CHEVRON

Multilayer elastic Multilayer elastic Multilayer elastic

Iterative Iterative Iterative

18

TABLE 2.2
Name of the main program

Comparison of Selected Backcalculation Programs.


Theoretical Backmethod for calculation pavement method analysis Multilayer elastic Iterative Analysis Maximum Number of sensors number of used in analysis unknown layers Five Internally or user specified number of layer-sensor assignments,up to ten sensors Program selects optimum number of sensors to use in backcalculation, can be overridden by user through specification of weighting factors for up to seven sensors Maximum number of different load levels Six Use of bedrock in backcalculation Calculates depth to bedrock upon users request. Temperature Developed corrections by Operating Environment

Forward backcalculation solution CHEVRON

MODCOMP3

No

l. Irwin, Szebenyi, Cornell University Texas Trasportation Institute

DOS

MODULUS5.0

WESLEA

Multilayer elastic

Data Base Search

Four

No limit

Automatically calculates depth to bedrock and includes it in the analysis, can be overridden by user.

No

DOS

EVERCALC4.0 WESLEA

Multilayer elastic

Iterative Analysis

Five

Up to seven sensors Eight

Calculates depth Yes to bedrock only if bedrock modulus is known. Calculates depth to bedrock and includes it in the analysis, can be overridden by user. Yes

J. Mahoney, University of Washington Dynatest

Windows

ELMOD4

ELMOD4, WES5

Odemark- Iterative Boussinesq Analysis

Five

As provided in raw FWD data file

as provided in raw FWD data file.

Windows

19

CHAPTER 3

COMPARISON OF BACKCALCULATION RESULTS FOR SHRP TEST SECTIONS

To evaluate the performance of backcalculation programs, the deflections and loadings measured for eight different SHRP sections were used with each of the four selected programs. The data required for the backcalculation were obtained from the SHRP-P-651 report. The types of materials and thicknesses for all eight sections are given in Figure 3.1. Six of the SHRP sections were for flexible pavement structures and the other two were for composite pavements constructed by overlaying original plain Portland Cement Concrete (PCC) pavement structures with asphaltic concrete. The moduli ranges for the different sections were obtained from the literature. The values were those commonly used to characterize a particular type of geologic material. When the information on a particular material type was not available in the literature, the modulus range for this material was interpolated from the published SHRP backcalculation results. The ranges of pavement layers moduli and Poissons Ratios are presented in Table 3.1.

All of the selected programs, MODCOMP3, MODULUS5.0, EVERCALC4.0, and ELMOD4 produced sets of pavement moduli that were within an acceptable match with the published results of backcalculation obtained from the SHRP project report, as listed in Table 3.2.

During different backcalculation runs, it was found that the inclusion or omission of a stiff layer resulted in different values for subgrade modulus and affected, to some extent, the backcalculated values of other layers. Inclusion of the stiff layer results in evaluation of a lower subgrade modulus. This dependancy was observed mainly in MODCOMP3 and MODULUS5.0 programs.

It was noted that the results from MODCOMP3 program were very sensitive to the assignment of the deflection sensor reading to a particular layer. Different layer-deflection 20

assignments may result in different results. Another disadvantage is that when the number of deflection sensors exceed the number of pavement layers, some of the deflection sensor readings are ignored in the backcalculation; and instead of the whole calculated deflection basin fit, only the chosen deflection sensor readings are fitted with the calculated ones. This limited approach makes the task of deflection basin fitting extremely difficult. A possible solution is to divide one or two layers into sublayers so that the number of sensors and layers will be equal. This approach has two drawbacks. First, it is difficult to decide which layers to subdivide so that the surface deflection will accurately predict their deflections. Second, division of a layer into sublayers may result in the calculation of two completely different moduli for the same layer.

The EVERCALC4.0 program terminates as soon as a moduli tolerance of one percent is satisfied. However, frequently this solution does not satisfy the deflection basin tolerances and has a high Root Mean Square value (RMS) error between measured and calculated deflections. All the attempts to override the default, one percent moduli tolerance, have failed. The program does not save this change upon its exit from the data input file. For most of the eight SHRP sections, subgrade moduli backcalculated by EVERCALC4.0 program were around the upper bound of the moduli range obtained from SHRP report. This might mean that EVERCALC4.0 has a tendency to overestimate the subgrade modulus.

The ELMOD4 program produced good results for the flexible pavement sections with unbound granular bases. However, backcalculations for the sections that include a stabilized or concrete layer under the surface AC layer resulted in over prediction of the surface layer modulus and under prediction of the base layer modulus. This could be due to the limitations of the ELMOD4 forward calculation algorithm that were discussed previously in Section 2.4.

The results of backcalculations for the eight SHRP sections are presented in Figures 3.2 through 3.9 for SHRP sections A through H. Each figure contains bar charts for comparison between pavement layers moduli evaluated by the above four backcalculation programs and the maximum and minimum values of moduli produced for the same layers during the SHRP project evaluation. All the 21

tested programs produced results that fall within the range obtained during SHRP project program evaluation.

Out of the four tested programs, MODCOMP3 program proved to be the most user-sensitive and required more engineering judgement. The approach of using seed moduli only, without specifying the acceptable moduli range, can lead to the calculation of unexpectably high or low moduli in different layers. The accuracy of the solution is highly dependent on the layer-deflection assignments. Totally different sets of moduli can be backcalculated for different layer-deflection assignments.

The ELMOD4 program failed to perform well on the sections that contained a stiff layer under the AC surface. However, it produced good results for flexible pavement sections with unbound granular bases.

Both MODULUS5.0 and EVERCALC4.0 are user-friendly and capable of producing pavement moduli within an acceptable range. MODULUS5.0 program allows to internally assign layer moduli based on the material types specified by the user. This feature can be advantageous if the material type is known and supported by the program. The EVERCALC4.0 program is fitted with a temperature correction option. This feature enables the user to adjust the backcalculated modulus of the AC layer to the modulus value measured at the standard temperature testing conditions.

22

TABLE 3.1

Moduli Range and Poissons Ratio for Backcalculation Inputs. Material Type Moduli Range (Ksi) 1000 - 10000 200 - 2,500 10 - 160 10 - 100 10 - 90 5 - 80 5 - 80 500 - 2,500 5 - 200 3-4 25 - 30 100 - 1000 100 - 400 Poissons Ratio 0.15 0.25 - 0.35 0.35 - 0.40 0.40 - 0.42 0.35 0.35 0.35 0.25 - 0.35 0.25 - 0.35 0.42- 0.45 0.42 - 0.45 0.20 0.25

Portland Cement Concrete Asphalt Concrete (cold->hot) Unstabilized Crushed Stone or Gravel Base Course (well drained) Unstabilized Crushed Stone or Gravel Subbase (poorly drained) Asphalt Treated Base Sand Base Sand Subbase Cement-Stabilized Base and Subbase Lime-Stabilized Base and Subbase Subgrade Soil Cohesive Clay Subgrade Soil Fine-Grained Sands Cement Stabilized Soil and Bedrock Lime Stabilized Soil

23

TABLE 3.2

Backcalculated Layer Moduli (ksi) for SHRP Sections.


Surface Layer Base Layer 65.8 70 80.4 92 10 57.4 87 175 155 85 500 447 537 1020 400 678.7 967 870 2510 430 753.3 689 696 890 90 79.4 77 50.3 178 37 8311.8 7470 7130 11400 3900 4272.3 3789 6140 6300 0 19.5 133 30 133 19 51.8 11 157 182 34 Subbase Layer 16.8 14 16.7 92 10 Subgrade 36.4 39 29.6 40 13 27.6 23 24.2 27 22 36.5 36 36.4 37 26 20.1 20 19.1 21 12 33.6 79 63.3 52 32 32.5 44 41.3 45 28 21.6 31 30.6 31 18 26.1 25 15.1 36 19

SHRP Section A MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN SHRP Section B MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN SHRP Section C MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN SHRP Section D MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN SHRP Section E MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN SHRP Section F MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN SHRP Section G MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN SHRP Section H MODULUS EVERCALC MODCOMP3 SHRP MAX SHRP MIN

993 918 1000 1320 720 1009 813 547 1420 500 566 572 613 630 450 2526 1933 2230 3050 1400 991 1034 1090 2310 850 1069 1015 1180 1300 720 1426 1248 1050 2440 820 282 308 239 7761 200

24

FIGURE 3.1 SHRP sections data used for backcalculations.

25

SHRP SECTION A
1400

SHRP SECTION A
100 90

SHRP SECTION A
100 90 80

SHRP SECTION A
45 40 35 30 25 20 15 10 5 0

1200 80

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

1000

70 60 50 40 30 20

70 60 50 40 30 20 10 0

800

600

400

200 10 0 0

Surface layer

Base layer

Subbase layer

Backcalculated moduli, Ksi

Subgrade

FIGURE 3.2

Comparison of backcalculated moduli for SHRP section A.


26

SHRP SECTION B
1600 1400

SHRP SECTION B
200 180 30

SHRP SECTION B

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

1200 1000 800 600 400 200 0

140 120 100 80 60 40 20 0

Backcalculated moduli, Ksi Base layer

160

25

20

15

10

Surface layer

Subgrade

FIGURE 3.3

Comparison of backcalculated moduli for SHRP section B.


27

SHRP SECTION C
900 800

SHRP SECTION C
1200 45 40 1000

SHRP SECTION C

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

600 500 400 300 200 100 0

800

Backcalculated moduli, Ksi Base layer

700

35 30 25 20 15 10 5

600

400

200

surface layer

Subgrade

FIGURE 3.4

Comparison of backcalculated moduli for SHRP section C.


28

SHRP SECTION D
5000 4500

SHRP SECTION D
3000 25

SHRP SECTION D

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

3500 3000 2500 2000 1500 1000 500 0

2000

Backcalculated moduli, Ksi Base layer

4000

2500

20

15

1500

10

1000

500

Surface layer

Subgrade

FIGURE 3.5

Comparison of backcalculated moduli for SHRP section D.


29

SHRP SECTION E
2500

SHRP SECTION E
1000 900

SHRP SECTION E
200 180 160

SHRP SECTION E
90 80 70 60 50 40 30 20 10 0

2000

800

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

700 600 500 400 300 200 100

140 120 100 80 60 40 20 0

1500

1000

500

Surface layer

Base layer

Subbase layer

Backcalculated moduli, Ksi

Subgrade

FIGURE 3.6

Comparison of backcalculated moduli for SHRP section E.


30

SHRP SECTION F
1400 1200 200 180

SHRP SECTION F
50 45 40

SHRP SECTION F

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

160 140 120 100 80 60 40 20 0

1000 800 600 400 200 0

Backcalculated moduli, Ksi Base layer

35 30 25 20 15 10 5 0

Surface layer

Subgrade

FIGURE 3.7

Comparison of backcalculated moduli for SHRP section F.


31

SHRP SECTION G
14000

SHRP SECTION G
12000

SHRP SECTION G
140

SHRP SECTION G
50 45

12000

10000

120 40

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

10000

100

Backcalculated moduli, Ksi Subbase layer

8000

35 30 25 20 15 10

8000

80

6000

6000

60

4000

4000

40

2000

2000

20 5

Surface layer

Base layer

Subgrade

FIGURE 3.8

Comparison of backcalculated moduli for SHRP section G.


32

SHRP SECTION H
14000 12000 7000 6000

SHRP SECTION H
40 35

SHRP SECTION H

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi

Backcalculated moduli, Ksi Base layer

10000 8000 6000 4000 2000 0

5000 4000 3000 2000 1000 0

30 25 20 15 10 5 0

Surface layer

Subgrade

FIGURE 3.9

Comparison of backcalculated moduli for SHRP section H.


33

CHAPTER 4

COMPARISON OF BACKCALCULATION RESULTS FOR SEVEN SELECTED SITES IN WEST VIRGINIA

FWD deflection data from seven different road sites in West Virginia were provided by WVDOT for the backcalculation of layers moduli using three different backcalculation programs: MODULUS5.0, EVERCALC4.0, and ELMOD4. The fourth program, MODCOMP3, was dropped from the evaluation due to its high sensitivity to layer-deflection assignments, and its inability to converge to a solution if a modulus of any layer, during the solution process, was found to be not sensitive to a certain geophone deflection reading. Out of seven road sites, three were composite pavements, two flexible, and two rigid. Pavement layer material types and layer thicknesses were provided by WVDOT. Typical moduli ranges, as provided in Table 3.1, were used as seed moduli values in the backcalculation programs.

In some cases, the backcalculation programs failed to produce a deflection basin that was within the acceptable tolerance with the measured deflection basin. When this situation occurred, the range of acceptable moduli was expanded in order to reach a convergence of the solution. The assumptions used in backcalculations are provided in Table 4.1. When an infinite subgrade option was used in the backcalculation analysis, most backcalculation programs over predicted the subgrade moduli compared with the ones obtained through laboratory testing.

MODULUS5.0 and ELMOD4 programs calculate the depth to bedrock internally and use this value in the backcalculation. Comparison of the backcalculated results obtained for different pavement sections showed that when a section was assumed as infinitely thick, high values of the subgrade moduli were frequently predicted. Inclusion of a finite depth to bedrock to the analysis led to more realistic lower values for the subgrade moduli. EVERCALC4.0 calculates the depth to rigid layer only if the subgrade modulus is provided. However, since the objective of backcalculation is to obtain a subgrade modulus, the subgrade modulus is treated as an unknown. Therefore, it was not 34

possible to include the depth to bedrock in the analysis using EVERCALC4.0.

In practice, subgrades generally display an increase in the stiffness with depth due to overburden pressure. However, laboratory tests are made for specimens of soil obtained immediately after the base course. This resulted in a lower laboratory measured subgrade modulus. To account for this, subgrade layers were divided into two layers and two different subgrade moduli were evaluated for each section. In this case, the moduli of top subgrade layers were in closer agreement with the laboratory measured moduli. This method did not work well with the ELMOD4 program due to the limitation of the Transformed Section Method which is used as a forward calculation algorithm. However, for the flexible pavement sections with granular bases, the ELMOD4 program seems to produce reasonable subgrade moduli even with one layer subgrade system.

The following four step approach for evaluation of subgrade layer modulus was used in backcalculation:

1.

Run the program assuming that the subgrade layer is uniform and infinite and check the subgrade modulus value. If the value is out of the expected moduli range for a given type of subgrade material ( as given in Table 3.1) or the Root Mean Square (RMS) error for the deflection basin fit is too high, proceed with the following steps.

2.

Rerun the program using default depth to the subgrade layer calculated by the program and check the moduli and RMS again.

3.

If the resulting moduli in Step 2 are not acceptable, see if there will be any improvements by increasing or deceasing the default depth value.

4.

If subgrade modulus or RMS is not very sensitive to the changes in Step 3, subdivide the subgrade layer into two sublayers with top subgrade sublayer height H not exceeding

35

H (36" - Height of Pavement Structure above Subgrade) Repeat the backcalculation process starting from Step 1.

The backcalculated moduli obtained for every pavement section using three different backcalculation programs are given in Table 4.2. Table 4.3 shows which of the moduli were lower, higher, or within the range of moduli acceptable for the section materials types. Table 4.4 summarizes the accuracy of the backcalculation results for different programs and pavement types.

From the section by section comparison of the performances of the three backcalculation programs the following conclusions were drawn:

1.

MODULUS5.0 performed well with all three types of pavement structures. This program consistently produced subgrade moduli that were reasonably close to the laboratory evaluated values and within the material range (below 25 psi), as shown in Figures 4.1 through 4.20. During the backcalculation process, the first program run for each section was carried out assuming an infinite subgrade layer. If the percent of error was greater than five, then the finite depth to bedrock calculated during the first run was obtained and another run of the program was carried out assuming bedrock at a finite depth. This procedure always resulted in lowering the percentage of error.

2.

ELMOD4 predictions were very poor for the composite pavement sections. Out of the three different pavement types shown in Figures 4.1 through 4.20, the program worked best with flexible pavements. For flexible pavements, all backcalculated moduli for the surface and base layers were within the acceptable range of pavement moduli. In the case of composite pavements, the program produced very high moduli for the surface and subgrade layers and very low moduli for the base layer, as shown in Figures 4.2 through 4.9. It was concluded that the program works only in cases where the layer moduli gradually decrease with depth. The program performed a little better with the rigid sections; however, the ELMOD4 36

program over predicted subgrade moduli in three out of five rigid pavement cases.

3.

EVERCALC4.0 worked fairly well with flexible and rigid pavements. However, in some flexible and rigid cases, backcalculated moduli for the base layer were on the low margin of acceptance. In three out of seven cases for the flexible pavement sections and in three out of five cases for rigid pavements, the program backcalculated very low modulus values for the base layers. For all composite pavement sections, high moduli values were predicted for the subgrade. The resulting moduli for surface and base layers in the composite pavement sections were either under predicted or over predicted compared to the acceptable range of moduli, and no common trend was possible to track down.

37

TABLE 4.1 Assumptions Used in Different Backcalculation Algorithms.


Moduli Range , Ksi Backcalculation Program Road/Station Number Pavement Type Surface Layer Base Subgrade Second Subgrade Modulus if it is Subdivided into Two Layers
calculated internally 0 2-60 calculated internally 0 2-45 calculated internally 0 2-40 calculated internally 0 4-35 calculated internally 0 4-35 calculated internally 0 4-35 calculated internally 0 4.35 calculated internally 0 4-35 calculated internally 0 4-35 calculated internally 0 2-60

Subgrade Depth to Bedrock (in.).

MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0

US52/16.250 US52/16.250 US52/16.250 US52/18.360 US52/18.360 US52/18.360 US52/19.179 US52/19.179 US52/19.179 US60/1.230 US60/1.230 US60/1.230 US60/2.267 US60/2.267 US60/2.267 US60/3.393 US60/3.393 US60/3.393 WV2/4.926 WV2/4.926 WV2/4.926 WV2/6.100 WV2/6.100 WV2/6.100 WV2/7.333 WV2/7.333 WV2/7.333 WV3/40.806 WV3/40.806 WV3/40.806

flexible flexible flexible composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite flexible flexible flexible

300-3000 n/a 200-3000 200-3000 n/a 200-2500 200-3000 n/a 200-2500 200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000

5-100 n/a 5-100 1000-9000 n/a 1000-10000 1000-9000 n/a 1000-10000 1000-7000 n/a 300-8000 1000-9000 n/a 300-8000 1000-9000 n/a 300-8000 1000-9000 n/a 100-7000 1000-9000 n/a 100-7000 1000-9000 n/a 100-7000 5-160 n/a -5-160

4-7 n/a 2-6 4-10 n/a 3-20 4-10 n/a 3-30 4-6 n/a 2-20 4-10 n/a 2-20 4-10 n/a 2-60 4-10 n/a 2-60 4-10 n/a 2-60 4-10 n/a 2-60 4-18 n/a 2-60

183.7 max 300 infinity 227.2 max 300 infinity 212.4 max 300 infinity 300 max 300 infinity 134.5 max 300 infinity 96.0 max 300 infinity 202.3 max 300 infinity 300 max 300 infinity 121.3 max 300 infinity 183.7 max 300 infinity

38

Moduli Range , Ksi Backcalculation Program Road/Station Number Pavement Type

Surface Layer

Base

Subgrade

Second Subgrade Modulus if it is subdivided into Two Layers


calculated internally 0 2-60 calculated internally 0 2-60 calculated internally 0 2-60 calculated internally 0 2-60 calculated internally 0 2-60 calculated internally 0 3-38 calculated internally 0 3-38 calculated internally 0 3-38 calculated internally 0 3-38 calculated internally 0 3-38

Subgrade Depth to Bedrock (in.)

MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0

WV3/41.013 WV3/41.013 WV3/41.013 WV3/41.221 WV3/41.221 WV3/41.221 WV71/0.009 WV71/0.009 WV71/0.009 WV71/1.259 WV71/1.259 WV71/1.259 WV71/2.740 WV71/2.740 WV71/2.740 WV2/18.250 WV2/18.250 WV2/18.250 WV2/19.028 WV2/19.028 WV2/19.028 US19/20.856 US19/20.856 US19/20.856 US19/21.563 US19/21.563 US19/21.563 US19/21.839 US19/21.839 US19/21.839

flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible rigid rigid rigid rigid rigid rigid rigid rigid rigid rigid rigid rigid rigid rigid rigid

200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000 200-3000 n/a 120-3000 1000-10000 n/a 1000-1000 1000-10000 n/a 1000-1000 1000-10000 n/a 1000-1000 1000-10000 n/a 1000-1000 1000-10000 n/a 1000-1000

5-160 n/a 5-160 5-160 n/a 5-160 5-160 n/a 5-160 5-160 n/a 5-160 5-160 n/a 5-160 10-160 n/a 10-160 10-160 n/a 10-160 10-160 n/a 5-160 10-160 n/a 5-160 10-160 n/a 5-160

4-18 n/a 2-60 4-18 n/a 2-60 4-18 n/a 2-60 4-18 n/a 2-60 4-18 n/a 2-60 4-18 n/a 2-60 5-7 n/a 4-10 5-7 n/a 4-10 5-7 n/a 3-10 5-8 n/a 3-10

108.9 max 300 infinity 254.4 max 300 infinity 61.65 max 300 infinity 66.71 max 300 infinity 49.10 max 300 infinity 182.7 max 300 infinity 279.2 max 300 infinity 300 max 300 infinity 300 max 300 infinity 202.6 max 300 infinity

39

TABLE 4.2a Backcalculated Layer Moduli (Ksi) for Flexible Pavement Sections.
Site/Station US52_pre Station 16.250 WV3_whit Station 40.806 WV3_whit Station 41.013 WV3_whit Station 41.221 WV71_blu Station 0.009 WV71_blu Station 1.259 WV_whit Station 2.740 Program MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 Surface Layer 594.0 537.0 510.0 321.0 315.0 291.0 366.0 360.0 418.0 454.0 540.0 520.0 643.0 487.0 751.0 508.0 498.0 425.0 274.0 247.0 264 Base Layer 42.6 27.0 61.0 14.8 12.7 7.0 43.6 32.3 5.0 66.7 61.7 18.0 17.1 59.6 5.0 6.3 8.1 5.0 9.7 9.9 5.0 Top Subgrade Subgrade Layer 7.0 14.8 7.0 7.3 19.5 12.0 11.0 10.1 3.0 15.8 9.8 16.0 4.9 11.2 4.0 12.5 8.0 3.0 4.9 8.0 4.0 17.1 14.8 20.0 18.9 19.5 26.0 11.8 10.1 24.0 9.3 9.8 12.0 30.4 11.2 35.0 6.7 8.0 27.0 6.2 8.0 50

40

TABLE 4.2b Backcalculated Layer Moduli (Ksi) for Composite Pavement Sections.
Site/Statiom US52_pre Station 18.360 US52_pre Station 19.179 US60_smi Stastion 1.230 US60_smi Program MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 Surface Layer 561.0 621.0 516.0 422.0 654.0 445.0 274.0 973.0 273.0 747.0 4255.0 3000.0 1620.0 8884.0 3000.0 203.0 264.0 192.0 200.0 240.0 137.0 222.0 305.0 301.0 Concrete Slab 6122.9 1214.0 7026.0 1638.7 606.0 1023.0 6196.0 611.0 3920.0 2267.2 84.1 384.0 1823.7 30.2 550.0 1326.7 434.0 1323.0 1910.1 722.0 2473.0 600.0 136.0 121.0 Base Layer 8.0 49.8 25.0 8.8 35.3 17.0 5.2 23.5 7.0 4.9 19.3 14.0 5.3 20.0 5.0 7.7 26.7 4.0 11.0 22.8 22.0 4.0 31.9 11.0 Subgrade 36.1 49.8 45.0 34.9 35.3 36.0 16.6 23.5 24.0 17.7 19.3 30.0 16.2 20.0 35.0 19.8 26.7 29.0 18.6 22.8 22.0 20.9 31.9 35.0

MODULUS5.0 Sration 2.267 ELMOD4 EVERCALC4.0 WV2_frie MODULUS5.0 Station 4.926 ELMOD4 EVERCALC4.0 WV2_frie MODULUS5.0 Station 4.926 ELMOD4 EVERCALC4.0 WV2_frie MODULUS5.0 Station 6.100 ELMOD4 EVERCALC4.0 WV2_frie MODULUS5.0 Station 7.333 ELMOD4 EVERCALC4.0

41

TABLE 4.2c Backcalculated Layer Moduli (Ksi) for Rigid Pavement Sections. Site/Station US2_moun Station 18.250 US2_moun Station 19.028 US19_hic Station 20.856 US_hic Station 21.653 US_hic Station 21.839 Program MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 Surface Layer 5280.0 4840.4 4308.0 6567.0 6157.0 5679.0 8152.0 6706.0 5948.0 5094.0 4117.0 3869.0 7160.0 4565.0 4716.0 Base Layer 21.9 5.2 17.0 34.6 1.5 10.0 25.1 24.5 5.0 20.1 16.0 5.0 39.5 39.3 5.0 Top Subgrade Subgrade Layer 6.2 37.4 4.0 5.5 32.6 4.0 5.6 15.1 3.0 6.4 10.6 3.0 5.3 26.0 3.0 22.4 37.43 38.0 14.8 32.6 23.0 14.2 15.1 18.0 7.0 10.6 12.0 14.7 26.0 34.0

42

TABLE 4.3 Comparison of Pavement Moduli Backcalculated by Different Programs. Backcalculation program
MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0

Road/Station Pavement Type Surface Number Layer


US52/16.250 US52/16.250 US52/16.250 US52/18.360 US52/18.360 US52/18.360 US52/19.179 US52/19.179 US52/19.179 US60/1.230 US60/1.230 US60/1.230 US60/2.267 US60/2.267 US60/2.267 US60/3.393 US60/3.393 US60/3.393 WV2/4.926 WV2/4.926 WV2/4.926 WV2/6.100 WV2/6.100 WV2/6.100 WV2/7.333 WV2/7.333 WV2/7.333 WV3/40.806 WV3/40.806 WV3/40.806 flexible flexible flexible composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite composite flexible flexible flexible acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable very high high acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable low acceptable acceptable acceptable acceptable acceptable acceptable

Base
acceptable acceptable acceptable acceptable acceptable acceptable acceptable low acceptable acceptable low high high very low very low acceptable very low low acceptable low acceptable acceptable low acceptable acceptable low low acceptable acceptable acceptable

Top Subgrade
acceptable acceptable acceptable high high high high high high acceptable high high acceptable high high acceptable high high acceptable high high acceptable high high high high high high high high

43

Backcalculation program MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0 MODULUS5.0 ELMOD4 EVERCALC4.0

Road/Station Number WV3/41.013 WV3/41.013 WV3/41.013 WV3/41.221 WV3/41.221 WV3/41.221 WV71/0.009 WV71/0.009 WV71/0.009 WV71/1.259 WV71/1.259 WV71/1.259 WV71/2.740 WV71/2.740 WV71/2.740 WV2/18.250 WV2/18.250 WV2/18.250 WV2/19.028 WV2/19.028 WV2/19.028

Pavement Type flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible flexible rigid rigid rigid rigid rigid rigid

Surface Layer acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable

Base acceptable acceptable acceptable acceptable acceptable acceptable acceptable acceptable low acceptable acceptable low acceptable acceptable low acceptable acceptable acceptable acceptable acceptable low acceptable acceptable low acceptable acceptable low acceptable acceptable acceptable

Top Subgrade acceptable acceptable high high acceptable acceptable high acceptable high acceptable acceptable high acceptable acceptable high high high high acceptable high high acceptable acceptable acceptable acceptable acceptable acceptable high high high

WV19/20.856 rigid WV19/20.856 rigid WV19/20.856 rigid WV19/21.563 rigid US19/21.563 US19/21.563 US19/21.839 US19/21.839 US19/21.839 rigid rigid rigid rigid rigid

44

TABLE 4.4 Number of Backcalculated Out of Range Moduli.


Backcalculation Program Surface layer MODULUS5.0 ELMOD4 EVERCALC4.0 0 0 0 0 0 3/7 3/7 1/7 5/7 Flexible Pavement Sections Base Subgrade Surface layer 0 2/9 3/9 1/8 7/8 4/8 3/8 8/8 8/8 Composite Pavement Sections Base Subgrade Surface layer 0 0 0 0 0 3/5 2/5 3/5 3/5 Base Subgrade Rigid Pavement Sections

45

620

70

16

25

600

60

14 20

580 50 560
Modulus (Ksi) Modulus (Ksi) Modulus (Ksi)

12

10 40
Modulus (Ksi)

15

540

30

520 20 500

10

4 5 10

480

460

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.1

Backcalculated layer moduli (Ksi) for flexible pavement section US52_pre station 16.250.

46

700

8000

60

60

600

7000 50 6000 50

500
40 5000 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi)
Modulus (Ksi)

40

400

4000

30

30

300

3000 20 20

200
2000 10 1000 10

100

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

PCC Slab
ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.2

Backcalculated layer moduli (Ksi) for composite pavement section US52_pre station 18.360.

47

700

1800

40

36.2 36 35.8

1600

600
1400

35

30

500
1200 25 1000 Modulus (Ksi) Modulus (Ksi)
Modulus (Ksi)

35.6 35.4 35.2 35 34.8 400 10 34.6

400

Modulus (Ksi)

20

300

800

15 600

200

100

200

34.4 34.2

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

PCC Slab
ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

Subgrade
ELMOD4

FIGURE 4.3

Backcalculated layer moduli (Ksi) for composite pavement section US52_pre station 19.179.

48

1200

7000

25

30

1000

6000 20 5000

25

800 15 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) 4000 600 Modulus (Ksi)

20

15

3000

10

400 2000 5 200 1000

10

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

PCC Slab
ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.4

Backcalculated layer moduli (Ksi) for composite pavement section US60_smi station 1.230.

49

4500

2500

25

35

4000
30

3500

2000

20 25

3000 1500
Modulus (Ksi) Modulus (Ksi) Modulus (Ksi)

15
Modulus (Ksi)

2500

20

2000

1000

10

15

1500
10

1000

500

5 5

500

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

PCC Slab
ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.5

Backcalculated layer moduli (Ksi) for composite pavement section US60_smi station 2.267.

50

10000 9000 8000 7000 6000 Modulus (Ksi) Modulus (Ksi) 5000 4000 3000 2000 1000 0

2000 1800 1600 1400 1200 Modulus (Ksi) 1000 800 600

25

40

35 20 30

15 Modulus (Ksi)

25

20

10

15

10 400 200 0 0 5 5

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

PCC Slab
ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.6

Backcalculated layer moduli (Ksi) for composite pavement section US60_smi station 3.393.

51

300

1400

30

35

250

1200

25

30

1000 200 Modulus (Ksi) 800 Modulus (Ksi) 20 Modulus (Ksi)

25

Modulus (Ksi)

20

150

15

600

15

100 400

10 10

50

200

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

PCC Slab
ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.7

Backcalculated layer moduli (Ksi) for composite pavement section WV2_frie station 4.926.

52

300

3000

25

25

250

2500 20 20

200 Modulus (Ksi) Modulus (Ksi)

2000 15 Modulus (Ksi) Modulus (Ksi) 15

150

1500

10

10

100

1000

5 50 500

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

PCC Slab
ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.8

Backcalculated layer moduli (Ksi) for composite pavement section WV2_frie station 6.100.
53

350

700

40

40

300

600

35

35

30 250 500 25 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) 200 400 Modulus (Ksi)

30

25

20

20

150

300

15 100 200 10

15

10

50

100

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

PCC Slab
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.9

Backcalculated layer moduli (Ksi) for composite pavement section WV2_frie station 7.333.

54

325 320 315

16

25

30

14 25 20 12

310 20 305 Modulus (Ksi) Modulus (Ksi) 300 295 290 4 285 280 275 2 5 5 10 Modulus (Ksi) 15 Modulus (Ksi) 15

10

10

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.10

Backcalculated layer moduli (Ksi) for flexible pavement section WV3_whit station 40.806.

55

430 420 410 400 390 Modulus (Ksi) Modulus (Ksi) 380 370 360 350 340 330

50 45

12

30

10 40 35 8 30 Modulus (Ksi) 25 20 15 10 2 5 0 0 6 Modulus (Ksi)

25

20

15

10

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.11

Backcalculated layer moduli (Ksi) for flexible pavement section WV3_whit station 41.013.

56

560

80

18

14

540

70

16 12 14

520

60 10 12

500 Modulus (Ksi) Modulus (Ksi)

50 Modulus (Ksi) Modulus (Ksi) 10 8

480

40

460

30 6 4

440

20 4

420

10

2 2

400

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.12

Backcalculated layer moduli (Ksi) for flexible pavement section WV3_whit station 41.221.

57

800

70

12

40

700

60

35 10 30

600 50 8 500 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) 40 Modulus (Ksi)

25

400

20

30

300 4 20 200 2

15

10

100

10

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.13

Backcalculated layer moduli (Ksi) for flexible pavement section WV71_blu station 0.009.

58

520

14

30

8 500 7 480 6 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) 460 8 Modulus (Ksi) 10 20 12 25

15

440

3 420 2 400 1 2 4

10

380

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4

FIGURE 4.14

Backcalculated layer moduli (Ksi) for flexible pavement section WV71_blu station 1.259.

59

6000

25

40

40

35 5000

35

20
30 4000 Modulus (Ksi) Modulus (Ksi)
Modulus (Ksi)

30

Modulus (Ksi)

15
3000

25

25

20

20

10

15

15

2000 10 10

5
1000 5 5

PCC Slab
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0 EVERCALC4.0

Subgrade
ELMOD4

FIGURE 4.15

Backcalculated layer moduli (Ksi) for flexible pavement section WV71_blu station 2.740.

60

6000

25

40

40

35 5000 20 30 4000 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) 15 25

35

30

25

3000

20

20

10

15

15

2000 10 5 1000 5 5 10

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0

Subgrade
ELMOD4 EVERCALC4.0

FIGURE 4.16

Backcalculated layer moduli (Ksi) for rigid pavement section US2_moun station 18.250.

61

6800

40

35

35

6600

35

30

30

6400

30 25 25

6200 Modulus (Ksi) Modulus (Ksi)

25 Modulus (Ksi) 20 Modulus (Ksi) 20

6000

20

15

15

5800

15 10 10

5600

10

5400

5200

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0

Subgrade
ELMOD4 EVERCALC4.0

FIGURE 4.17

Backcalculated layer moduli (Ksi) for rigid pavement section US2_moun station 19.028.

62

9000

30

16

20 18 16

8000 25 7000

14

12 14 6000 Modulus (Ksi) Modulus (Ksi) 20 10 Modulus (Ksi) 5000 Modulus (Ksi) 12 10 8 6 2000 5 1000 2 2 0 4 4

15

4000

6 3000 10

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0

Subgrade
ELMOD4 EVERCALC4.0

FIGURE 4.18

Backcalcualted layer moduli (Ksi) for rigid pavement section US19_hic station 20.856.

63

6000

25

16

14

14 5000 20 12

12

10 4000 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) 15 10 8

3000

10

6 4 4

2000

5 1000 2 2

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0

Subgrade
ELMOD4 EVERCALC4.0

FIGURE 4.19 Backcalculated layer moduli (Ksi) for rigid pavement section US19_hic station 21.563.

64

8000

45

30

40

7000

40 25 35

35

6000 30 5000 Modulus (Ksi) Modulus (Ksi) Modulus (Ksi) 25 Modulus (Ksi) 20

30

25

4000

15

20

20

3000 15 2000 10

15

10 5

10

1000

Surface Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Base Layer
MODULUS5.0 EVERCALC4.0 ELMOD4

Top Subgrade
MODULUS5.0 EVERCALC4.0 ELMOD4 MODULUS5.0

Subgrade
ELMOD4 EVERCALC4.0

FIGURE 4.20

Backcalculated layer moduli (Ksi) for rigid pavement section US19_hic station 21.839.

65

CHAPTER 5 FIELD TESTING

5.1

PAVEMENT SITES

Field tests were performed at three different sites located in Morgantown, West Virginia. The selected sites were chosen to be representative of the three types of pavement structures: rigid, flexible, and composite (rigid pavement overlaid by asphalt layer). In each tested site, the pavement surface was checked so that it was free from visible signs of distress such as transverse or longitudinal cracks.

1. Flexible Pavement Site This site was the traffic lane of a divided four-lane road on Route 857, opposite Mountaineer Mall. The lane width is 12 ft. and the asphalt concrete layer is 9 in. thick constructed over a 7 in. base of crushed stone. The subgrade soil is a clay material. Figure 5.1 illustrates the different layers in this section.

2. Rigid Pavement Site This site was located on Route 857. The traffic lane of a divided four-lane jointed Portland cement concrete pavement with untied shoulder. The spacing between the transverse joints in this site varied between 20 and 45 ft. The lane is 12 ft. wide and the slab is 9 in. thick. The slabs are supported by an 8 in. thick granular base of crushed stone. The subgrade soil in this site was silty clay. Figure 5.2 provides a section of this site showing the different layers.

3. Composite Pavement Site This site was located on Interstate I-68, station 2.644. The test was performed on the traffic lane of the eastbound side. The pavement section consists of a 9.5 in. concrete slab overlaid by a 5 in. asphalt concrete layer. The concrete slab is supported by a 3 in. base of crushed stone. The 66

subgrade soil is clay material. Figure 5.3 illustrates the different layers of this pavement section.

Each site was fitted with a set of sensors to measure the temperature gradient and displacements of the layers due to the application of FWD load. Two boreholes were drilled in each site for the placement of these gauges. The two boreholes were placed at a distance of 6 ft from each other as shown in Figure 5.4. Each borehole was drilled to a depth of 32 in. and core samples were taken for laboratory testing by WVDOH. All boreholes were placed at a distance of 1 ft from the shoulder edge of the traffic lane. During borehole sampling the correct thickness of each pavement structure was identified and recorded. After placing the sensors in each borehole, it was filled with fine sand and covered on top with a concrete mix. All lead cables of geophones and thermocouples from each borehole were routed into a 0.75 in. deep groove sawn in the surface layer leading to the end of the shoulder and then covered with concrete grouting. Each trench terminates with a small hole to keep the ends of the cables accessible for measurements.

5.2

MEASUREMENTS

Temperature gradient through the depth of pavement structure was measured using thermocouples. The thermocouples were placed such that one was at the interface between the base and subgrade, and another at the interface between the surface layer and the base. For the composite pavement site, an additional thermocouple was placed at the interface between the asphaltic layer and the concrete. The connecting cables of each thermocouple were covered with Teflon. Additionally, the cables were environmentally shielded using a 0.25 in. diameter polyethylene tube that extended the full length of the cable. The connections of the tube with the thermocouple-sensing element were sealed using a heat shrink water-proof tube. Each thermocouple was placed in position inside a 3 in. diameter borehole that was drilled adjacent to the shoulder edge of the pavement site. Each thermocouple had a sensitivity of 0.1F. The temperature readings were acquired using a hand held digital readout unit, type HH-21 produced by the OMEGA company. At the time of the Falling Weight Deflectometer test, acquired temperature measurements were as follows: 67

Rigid Pavement Site On top of the concrete slab At the concrete-base interface At the base-subgrade interface To = 47.0oF T1 = 49.5oF T2 = 51.5oF

Flexible Pavement Site On top of the asphaltic layer At the asphalt-base interface At the base-subgrade interface To = 38.5oF T1 = 41.0oF T2 = 42.3oF

Composite Pavement Site On the top the of asphaltic layer At the asphalt-concrete interface At the concrete-base interface At the base-subgrade interface To = 38.3oF T1 = 40.3oF T2 = 41.4oF T3 = 41.6oF

None of the temperature measurements showed a significant temperature gradient, due to the fact that the measurements were carried out during the months of March and April from 10:00 am to 2:00 pm. The absence of a significant thermal gradient was considered advantageous from the point of view of finite element modeling.

It was the intention in this study to measure the displacement of every pavement layer in order to observe the possibility of separation between layers and to provide additional confirmation of the accuracy of the finite element models. For this reason, a set of twelve geophones were acquired with the following specifications:

Geophone type: PE-8-SM 6UB Manufacturer: SENSOR, Netherlands. Measuring direction: Vertical 68

Resistance: 375 Ohm Resonant frequency: 4.5 Hz

Each geophone was placed in an environmentally isolated casing to protect the sensing element from being damaged by underground water or humidity. During the installation of geophones in each borehole, care was taken to ensure that they were vertically aligned inside each hole. The displacement data measured from the geophones contained significantly high levels of noise, which hindered any reliable readings. Considering the effort of arranging for traffic control and because of the time span of the project, no attempts were made to replace the geophones with more sophisticated sensors. Thus, additional geophone measurements were omitted on the basis that verification of the FE models could be achieved through an additional comparison of the deflection-time histories with those measured using FWD geophones.

5.3

DEFLECTION TESTING

The Falling Weight Deflectometer (FWD) test was performed using WVDOH equipment. The equipment is manufactured by DYNATEST and operated by WVDOH Materials Division personnel. The FWD was calibrated in accordance with SHRP procedures in May 1997. The calibration results demonstrated that both the FWD load cell and sensors were in excellent condition, and that their readings fell within the tolerance range set by the manufacturer. During each FWD test, surface deflections were measured using nine geophone sensors located at distances of 0, 8, 12, 18, 24, 36, 48, 60, and 72 in. away from the center of the 11.8 in. diameter steel loading plate. The sensors positions remained unchanged for all three pavement sites.

For every pavement site, the FWD deflection data were collected at three drop heights. At each height two drops were made in accordance with ASTM-Standard D 4694 - 87, article 9.4. Since the temperature measurements did not show appreciable thermal gradient, no temperature corrections were considered necessary during the backcalculation of layers moduli. 69 Table 5.1

contains a list of the load and deflection data measured using FWD for the three tested sites. Additionally, the data were acquired in graphical format and the loading curves were digitized for use with the finite element models, as will be discussed in the next chapter.

5.4

EVALUATION OF BACKCALCULATED MODULI

FWD deflection data obtained from testing the three sites were analyzed using three different backcalculation programs: MODULUS5.0, EVERCALC4.0, and MODCOMP3. The information required for the evaluation of layers moduli for each site includes the deflection data and the pavement profile. The pavement profile consists of layer thicknesses and material types. The material types were determined from visual examination of the layers material extracted during borehole drilling. All backcalculation programs require the user to supply seed values for the moduli of each layer. Those values were selected by referring to the ranges listed in Table 3.1. In some circumstances, the backcalculation program failed to produce a deflection basin that fits the measured one within the specified tolerance. In this case, the moduli ranges were expanded to reach a convergence of the solution. As discussed in Chapter 4, the subgrade layers were divided into two layers to take into account the change of the value of subgrade modulus with depth. Consequently, two different subgrade moduli were evaluated for each section.

The results of backcalculated moduli obtained for each pavement section are given in Table 5.2. For every pavement site, an average modulus was estimated for each layer. This average was obtained by taking the mean of the three moduli produced for the same layer by the three backcalculation programs. Next, the average was used to estimate the Percent Deviation From Average (PDFA) in modulus produced by each program (shown in Table 5.2 as percent error).

5.5

CONCLUSIONS

Based on the results listed in Table 5.2, the following conclusions were reached: 70

1.

MODULUS5.0 program performed well with all types of pavement structures and resulted in moduli values that seemed reasonable and close to the average.

2.

EVERCALC4.0 program overestimated the values of subgrade moduli for the three types of pavements compared to the other two programs. For the composite pavement structure, the program overestimated the modulus of the asphalt layer.

3.

MODCOMP3 program resulted in moduli values which seemed to be reasonable and within the material range. However, the resulting modulus value of the concrete layer in the rigid pavement section seemed to be high compared with those resulting from the other two programs. This program requires more engineering judgement and experience from the user in introducing the values of seed moduli for different layers to obtain reasonable results. The high sensitivity of the program to the deflection reading assignments may cause the termination of the program if the values of seed moduli are not compatible with the deflection readings.

4.

Comparison of the backcalculated moduli for the pavement layers reveals that for each pavement type the results obtained using MODULUS5.0 were always close to the average obtained from all three programs. The values of the layers moduli obtained for all pavement structures using MODULUS5.0 seemed reasonable and the mean percent difference did not exceed eleven percent.

5.

None of the three programs tested in this study produced extremely unreasonable values. This is reflected in the fact that the error relative to the average calculated modulus never exceeded 100 percent for any layer.

6.

It should be remembered that all three programs evaluated in this study are designed primarily to handle flexible pavements. The capability of predicting layers moduli for rigid and composite pavements is an added advantage. 71

TABLE 5.1

Deflection Data Collected from Field Tests. Radial offset from center of loading plate (in.) 0.00 8.00 12.00 18.00 24.00 36.00 48.00 60.00 72.00

Applied load lbf psi

RIGID PAVEMENT SITE: 10124 10135 13110 13186 17877 17899 92.6 92.7 119.9 120.6 163.5 163.7 3.46 3.46 4.63 4.66 6.37 3.36 3.23 3.19 4.30 4.26 5.89 5.85 3.04 3.02 4.05 4.03 5.56 5.54 2.74 2.74 3.66 3.66 5.04 5.03 2.43 2.43 3.25 3.26 4.47 4.46 1.90 1.89 2.52 2.52 3.47 3.47 1.40 1.40 1.87 1.87 2.59 2.59 0.98 0.96 1.35 1.35 1.89 1.89 0.69 0.69 0.98 0.98 1.38 1.38

FLEXIBLE PAVEMENT SITE: 10037 10048 12793 12815 17853 17998 91.8 91.9 117.0 117.2 164.2 164.6 10.11 10.13 13.33 13.40 18.23 18.08 8.82 8.84 11.73 11.79 16.07 15.96 7.97 7.99 10.63 10.69 14.60 14.50 6.78 6.79 9.10 9.17 12.58 12.47 5.70 5.72 7.67 7.72 10.62 10.53 3.88 3.91 5.21 5.25 7.41 7.35 2.45 2.46 3.40 3.42 4.81 4.77 1.49 1.46 2.15 2.15 2.97 2.95 1.02 1.02 1.29 1.30 1.96 1.95

COMPOSITE PAVEMENT SITE: 9873 9873 12596 12617 17330 17384 90.3 90.3 115.2 115.4 158.0 159.0 2.96 2.97 3.99 4.00 5.51 5.50 2.45 2.46 3.32 3.35 4.63 4.62 2.31 2.31 3.14 3.15 4.37 4.36 2.17 2.19 2.96 2.97 4.14 4.13 2.03 2.03 2.75 2.76 3.83 3.82 1.66 1.68 2.28 2.29 3.19 3.19 1.32 1.32 1.84 1.83 2.54 2.55 1.03 1.03 1.41 1.41 1.99 1.98 0.77 0.79 1.10 1.09 1.54 1.54

72

Table 5.2 BACKCALCULATED LAYERS MODULI (Ksi).


PROGRAM MODULUS Surface PDFA* Base 837 338 909 694.6 3278 2997 4540 3755 805 2000 832 1212.3 -0.7% 20.5% 34.9 PDFA -32.1% 94.4% -62.3% Subgrade PDFA Top** 10.7 4.7 11.6 9.00 11.2% -64.3% 53.1% 13.4 20 2.8 12.1 9.7% -6.7% -3.0% 156 100 80 112 39.3% -10.7% -28.6% 10.9% 65.6% -76.5% 18.9% -48.1% 28.8% Subgrade PDFA Bottom 9.3 34.5 12 18.6 14.5 50 13.5 26.0 13.7 30 17.6 20.4 -33.0% 46.8% -13.9% -44.2% 92.3% -48.1% -50.0% 85.5% -35.5%

FLEXIBLE

EVERCALC MODCOMP LAYER AVERAGE MODULUS

-51.3% 100 30.9% 19.4 51.4 162

RIGID

EVERCALC MODCOMP LAYER AVERAGE MODULUS

-20.2% 52 20.9% 223 145.7 -33.6% 5880 65.0% 5000

COMPOSITE EVERCALC
MODCOMP LAYER AVERAGE *

-31.4% 5200 5360

PDFA= Percent Deviation from Average.

** The subgrade was divided into two layers for the purpose of backcalculation.

73

74

FIGURE 5.4 Instrumentation layout.

75

CHAPTER 6 EVALUATION OF BACKCALCULATION ALGORITHMS THROUGH FINITE ELEMENT MODELING OF FWD TEST

6.1

INTRODUCTION

In this chapter, a new approach for the evaluation of the performance of backcalculation algorithms is developed using Finite Element (FE) analysis. Currently, the only means of assessing the performance of a backcalculation algorithm is to compare the layers moduli results with laboratory measurements of core samples. The wide discrepancies between the backcalculated and measured moduli are often attributed to the difficulty in obtaining undisturbed soil samples. The end result is that backcalculated moduli for the same pavement structure may differ widely, depending on the backcalculation algorithm used and the assumptions made by the program operator. In this Chapter, Three Dimensional Finite Element (3D-FE) approach is used to simulate the response of the three sections, tested in Chapter 5, to FWD load. For each pavement structure, the deflection basin obtained from the finite element model is compared with the one experimentally measured. The elastic moduli used in the model are changed until a satisfactory agreement between the Finite Element-Calculated and the experimental basins is reached. Next, the layers moduli are evaluated from the FE-generated deflection basin using the three backcalculation programs MODULUS5.0, EVERCALC4.0 and MODCOMP3, and the results are compared with the layers moduli used in the finite element model. Correction factors for backcalculated moduli were developed and found to be in close agreement with the values recommended by SHRP Pavement Design Guide.

6.2

REVIEW OF FINITE ELEMENT MODELING OF PAVEMENTS

A number of 2D finite element programs such as ILLI-PAVE and MICH-PAVE were developed to analyze flexible pavements. In these programs, the 3D pavement structure is idealized 76

using 2D axially symmetric elements. To account for material nonlinearity, the unbounded nature of granular soils, and locked-in lateral stresses produced by compaction, stress-dependant resilient moduli were incorporated for granular and cohesive soils (4,19). In MICH-PAVE, a flexible boundary was utilized at the bottom of the FE model to reduce computer memory and processing time (19). Limitations of the axi-symmetric assumptions used in these programs include: inability to simulate asymmetrical loading condition, assumption of full contact with the base layer, inability to simulate cracks and rutting conditions, and the use of static loading. Other 2D finite element programs for rigid pavements include ILLI-SLAB, JSLAB, KENSLABS, WESLIQUID, FEACONS III, KENSLABS, and WESLAYER (20-25). In these programs the concrete slab was treated as a 2D medium-thick plate. To accommodate the presence of the base layer, the slab and base are transformed into one equivalent layer. Although the programs are specifically designed for rigid pavement analysis, the actual behavior of concrete pavements is more complex than the 2D model idealization . The capability of handling a moving load was introduced to ILLI-SLAB by Chatti et al. (26, 27); recently, Roesler et al. (28) modified the program to allow for partial-depth crack analysis (ILSL97). The major limitation of the 2D FE modeling approach is its inability to handle the geometrical features near dowel bars without a significant degree of approximation. The need for developing a deep understanding of pavement behavior motivated researchers to use the 3D FE approach.

None of the available 3D finite element codes is specifically designed for pavement analysis. Ioannides et al. (29) studied stress-dependant foundation using the GEOSYS program which is primarily designed for geotechnical problems. Starting the late 1980's, general purpose finite element codes such as ABAQUS, DYNA3D, and NIKE3D were introduced in pavement engineering research (30-38). Zagloul and White studied the dynamic response of flexible pavements to FWD loading and moving loads using ABAQUS. The modeling results were found to be in close agreement with the field measurements (39,41). Seaman et al. (34) examined the response of airport runways using NIKE3D. This work was later extended by Kennedy et al. (35-38) who developed a user interface to DYNA3D and NIKE3D specially designed for pavement structural analysis.

The earlier finite element codes developed for the purpose of modeling FWD test were based 77

on static interpretation of FWD load (41-43). Many studies were conducted to compare deflection basins resulting from dynamic analysis and those from static analysis. Mamlouk and Davies developed a multi-degree of freedom model based on the principle of elasto-dynamics; it accounted for the three-dimensional response properties and the inertia effects. Transient loading was represented by a series of steady-state harmonic loadings with different frequencies and magnitudes (44, 45). Sebaaly modified the program to include the calculation of stresses and strains in pavements caused by harmonic and impulsive loading (46). The results showed that the surface deflections obtained using elasto-dynamic analysis of FWD tests were within 3 to 15 percent from field measurements. Static representation of FWD load was found to produce surface deflections that are up to 40 percent larger than field measurements (46, 47). Mallela et al. (48) developed a threedimensional response model for a rigid pavement structure subjected to FWD load. Linear elastic materials were used to characterize all pavement layers. This assumption was based on the argument that the stresses induced in each layer under standard 18 Kip equivalent single axle load are not likely to produce stresses which exceed the elastic limits of each layer. The deflection basin resulting from dynamic FEM showed a good agreement with the measured one while the deflection basin resulting from the static FEM was 80 percent larger than that measured experimentally. Nazarian et al. (49) conducted an investigation to assess the significance of layer stiffnesses, thicknesses, and the depth to bedrock on the measured and backcalculated deflections. They found that the dynamic nature of the FWD load significantly affects the deflections measured away from the load. The depth to bedrock and load duration interacted to produce significantly different static and dynamic deflections. Uddin et al. (50, 51) reported an increase of 22 percent in maximum deflections due to the presence of multiple cracks. The role of pavement layers interfaces in transmitting FWD load was examined by Shoukry et al. (52, 53) who reported stiffer FWD response due to the lack of adequate interface representation.

6.3

GUIDELINES FOR 3D FEM OF PAVEMENT STRUCTURES

From the above review, it can be seen that 3D FE modeling of pavements has reached a degree of maturity which permits its use in pavement structural evaluation problems. In building a 78

finite element structural model, there are guidelines which help produce theoretical results close to the experimentally measured ones. These guidelines can be summarized as follows:

1.

Finite element model loading should have the same dynamic characteristics and time duration as that applied in practice. This means that traffic and/or impact loads should be dynamically applied in the model with the same duration or speed encountered in practice.

2.

Structural interfaces between different parts should be represented in the model. Allowing separation of the interfaces under inertia and/or thermal effects is a primary factor which influences the deflections and stresses obtained from the model. Approximate values of coefficients of friction at the interface help produce more realistic results.

3.

For short time duration studies and transient analysis, it seems practical to assume that all pavement layers behave elastically. The need for using nonlinear material models should be assessed carefully by the engineer. In most cases, nonlinear material models require the use of constants that may not be available and must be assumed. The benefits gained from using nonlinear material models may be lost due to incorrect assumptions of material constants. In most cases which do not involve repeated application of loads in the model, the use of nonlinear material models for base and/or subgrade may not be necessary. On the other hand, studies of rutting development in flexible pavements would certainly require the use of inelastic material models.

4.

The continuity of the stresses and deformation in the model should be checked using fringe plots1. The changes in fringe intensity from one element to another should be smooth and continuous. If this is not the case, the finite element mesh should be refined until smooth patterns of fringes are obtained. This provides a confirmation of the adequacy of the FE mesh

Fringe plots are colored visualization of the deformations and stresses in the finite element model. Each color represents a certain region of stress or the deformation. For example, see Figures 6.6 and 6.9.

79

but does not guarantee that the values are correct. The values of the stresses and deflection throughout the model are primarily dependent on the material constants used and accuracy in representing different geometrical features.

5.

In Explicit Finite Element analysis, the model deformation should be displayed using a large scale factor to ensure that zero energy modes did not develop anywhere in the model and that none of the interfaces penetrate each other.

6.

Finite element model results should be verified experimentally. One simple experimental proof of the correct operation of pavement structural models can be obtained from comparing deflection basins produced by FWD loading with those obtained from the model.

7.

Finite element solutions are numerical solutions of sets of partial differential equations which combine materials constitutive laws with the geometry of the structure. The equations are based on the same fundamental stress-strain relations used in developing any closed form solution. However, while closed form solutions are limited to geometrically simple structures, finite element analysis can handle any geometry. Thus FE solutions are dependent to a great extent on the accuracy of modeling the structural geometry. Thus, if the true structure contains a crack, finite element modeling cannot produce deflection results that mimic the experimentally measured ones unless the crack is accounted for in the model. Engineering judgement should be exercised in deciding the level of structural details that should be included in the model in order to produce reliable results.

8.

Use should be made of computer visualization and animation capabilities to display model results. The selection of a powerful post processor is a key factor in understanding the model results and drawing conclusions.

80

6.4

FINITE ELEMENT MODELS OF EXPERIMENTAL TEST SITES

6.4.1 Flexible Pavement Model

Flexible pavements are continuous in the direction of traffic and jointless in the transverse direction. Due to the softness of the asphaltic material, it is expected that the response of pavement to the FWD load will be localized. For this reason, the pavement structure was modeled as a multilayered system consisting of asphaltic concrete layer, base, and subgrade as shown in Figure 6.1. The model width was chosen to be the full lane width of 12.0 ft. The flexible pavement section has a shoulder made from the same material as the surface layer. This allows for the assumption of a continuity of the propagation of in-plane stress waves generated by impact load. To simulate this continuity, non-reflective boundaries were modeled along the pavement sides. The model length in the traffic direction is 20.0 ft. The center of the loading plate was located at a distance 5.0 ft from the lane edge adjacent to the shoulder. Due to the geometrical symmetry around a vertical plane passing through the center of the loading plate perpendicular to traffic direction, only one half of the pavement was meshed as shown in Figure 6.1. A tied interface between base and subgrade was assumed while a sliding interface with a coefficient of friction of 0.9 was assumed between the asphaltic layer and the base.

6.4.2 Rigid Pavement Model

The pavement structure was modeled as a multilayered system consisting of a Portland Cement concrete slab, base, and subgrade as shown in Figure 6.2. The model width was chosen to be the full lane width of 12.0 ft. The center of the loading plate was located at the center of the concrete slab. Due to the geometrical symmetry around the transverse plane passing through the center of the loading plate, only one half of the pavement was meshed as shown in Figure 6.2. The influence of dowel bars on the deflection basin was assumed to be negligible due to two reasons: (1) large distance of the transverse joint from the center of FWD load application, and (2) the relatively 81

small magnitude of FWD load. Therefore, dowel bars at the transverse joint were not included in the model. Figure 6.2 also illustrates the boundary conditions used in the model. A sliding interface with a coefficient of friction of 0.9 was assumed between the concrete layer and subgrade while a fully tied interface was assumed between the subgrade and base.

6.4.3 Composite Pavement Model

The pavement structure was modeled as a multilayered linear elastic system consisting of a Portland Cement concrete slab overlaid by an asphaltic concrete layer, a base, and a subgrade as shown in Figure 6.3 . The model width is 12.0 ft and its length in the traffic direction was chosen to be a full slab length of 27.0 ft. The center of the loading plate was located at the center of the model. As a result of loading and geometrical symmetry around the transverse plane passing through the center of the loading plate, only one half of the pavement was meshed as shown in Figure 6.3. A tied interface was assumed between the asphalt and concrete layers. A sliding interface with a coefficient of friction of 0.9 was assumed between the concrete and subgrade while the base/subgrade interface was assumed to be fully tied.

In all models, bedrock was assumed to lie at a depth which will not introduce reflections within the loading time duration under investigation (30 ms). This was achieved by applying nonreflective boundaries which simulate the semi-infinite extent of the subgrade at the bottom of each model. All models were meshed using 8-node brick element with 24 degrees of freedom per element. The mesh sizes varied through each model to assure the accuracy of the results within the regions of interest. Thus, a refined mesh was necessary in regions of high stress intensity such as concrete slabs. A coarser mesh was used for the base and subgrade. A one in. thick and 12 in. diameter steel loading plate (modulus of elasticity = 30,000 Ksi, unit weight = 0.2831 lb/in.3) was added to each pavement model for more accurate simulation of the FWD setup.

The impact loads used in this investigation were obtained from the measured FWD loads 82

applied during tests. The impact loading-time relations recorded by the FWD load cell were digitized over one millisecond time intervals and the pressure-time history was calculated by dividing the load by the area of the loading plate. Figure 6.4 illustrates the digitized pressure-time curves used for modeling the three types of pavements.

6.5

STRUCTURAL MATERIAL MODELING

The surface deflections measured at various FWD sensor locations can be useful in determining if the pavement structure displays a significant degree of nonlinearity due to the application of the FWD load. FWD tests were carried out at three different loading levels, as listed in Table 5.1. The deflections versus FWD load recorded at each sensor location are shown in Figure 6.5. The plots reveal that the deflections measured at all sensor locations are directly proportional to the applied load. In the case of the flexible pavement, a negligible nonlinearity can be observed for sensors located at 0 in. and 8 in. from the center of load application. For all other sensor locations, the experimentally measured deflections fell close to a straight line obtained through least squares fitting. This linearity indicates that the structural materials of all three pavements behaved elastically under the loading levels applied in each case. Since in the present analysis FWD load will be limited to 10,000 lb, it is safe to assume that all materials used in the three models can be represented using linear elastic models. That is, each layer material is characterized by its modulus of elasticity, density, and Poissons Ratio. This choice of material model has been further confirmed from studying the amount of stresses induced in different layers due to the application of the FWD load. Figure 6.6 illustrates by fringes the distribution of vertical stresses for different pavements. Furthermore, the vertical stress along the vertical line passing through the center of the loading plate was plotted versus depth in each pavement structure as shown in Figure 6.7. It can be noticed that the stresses induced in base and subgrade layers are very small which validate the assumption of linear elastic material models.

83

6.6

LAYERS MODULI EVALUATION FROM FE MODEL RESULTS

Since the purpose of this study is to evaluate the structural capacity of the three pavement structures, the elastic moduli of the pavement materials are all unknown. An iterative procedure similar to that commonly used in backcalculation algorithms was employed to determine those moduli. Iterative procedure steps are as follows:

1.

FWD test is conducted and the experimental deflection basin is obtained.

2.

The experimental basin is used with any backcalculation program to evaluate a set of layers moduli.

3.

The layers moduli obtained from backcalculation are used in the finite element model and a theoretical deflection basin is obtained.

4.

The FE generated deflection basin is compared with the experimental basin. If the two basins are in good match, the moduli used in the finite element program are taken to be the correct moduli.

5.

If the basins are different from each other, the layers moduli are adjusted and inserted in the finite element program to produce a new deflection basin.

The above procedures are repeated until the condition in step number 4 is satisfied. This method of determining layers moduli was considered to be superior to backcalculation for the following reasons:

1.

FE does not contain assumptions about pavement geometry as traditional backcalculation algorithms do. The model accounts for the layers interfaces and the inertial properties of the 84

materials.

2.

The load used in the finite element programs is the experimentally measured load-time history applied by FWD. This permits comparing the experimental and theoretical deflection time histories at all sensor locations. All existing backcalculation programs reach convergence based only on the maximum deflection measured at each sensor location.

6.7

VERIFICATION OF FINITE ELEMENT MODELS

6.7.1 Deflection Basins

The initial seed moduli used in the FE models produced highly stiff responses in all cases. Therefore the moduli of different layers were reduced. After several cycles of adjusting the values of layer moduli used in each model, a satisfactory match between the FE-generated and the FWDmeasured basins was reached for every pavement structure. The final basins are plotted together with those experimentally measured in Figure 6.8. In all cases, the FE deflection basins fell close to those measured experimentally. Figure 6.8 (c) for the composite pavement shows deviation between the experimental and theoretical basins near the center of FWD load application. This deviation may be attributed to one or both of the following reasons: 1. Existence of unrepaired cracks in the concrete slab. 2. The center of FWD loading plate fell near a transverse joint in the overlaid slab.

Additionally, the asphalt material under the loading plate, being confined between the loading plate and the underlying concrete, has suffered local excessive deformation that was recorded by the FWD sensor located at zero offset. At the locations of other sensors, the surface of the asphalt layer is free to follow the deformation of underlying layers. As indicated in Figure 6.8 (c) the FE model successfully simulated this behavior, however lack of representation of a cracked concrete layer resulted in the small deviation between the FE-generated and the FWD-measured deflections up to 85

24 in. offset from the center of load application.

Figure 6.9 illustrates the fringes of vertical deformation at the time of maximum FWD load through two sections; one is perpendicular to the traffic direction and the other passes through the FWD sensors line. The patterns of vertical displacement fringes illustrate the uniformity of displacements through different layers.

6.7.2 Displacement-Time History

The FE-generated and experimentally measured deflection-time histories are compared in Figure 6.10 for three different sensor positions. The plots reveal that as the FWD load is applied, both the theoretical and experimental deflections increase at almost the same rate. After peak deflections are reached, the FWD measured deflection curves rebound much faster than the FE-generated deflections. This behavior could be attributed to one of two reasons:

1.

The top layer separates from the base. If this is the case, then the FWD sensors should have recorded different values of positive deflections at different sensor positions. As seen from Figure 6.10 (a), this cannot be true since the sensor located at 48 in. recorded almost the same positive deflections as the sensor located at the center of load application. Furthermore, Figure 6.10 (b) shows that the magnitude of rebound at the center of FWD load application exceeds that recorded by the sensor located at 12 or 48 in. offsets, which excludes layer separation as a reason.

2.

Either all the sensors, or the FWD structure (which supports all sensors) rebounded following the application of FWD load. It is unlikely that all sensors rebound as the FWD manufacturer makes sure that the magnitude of spring loading is large enough to keep the sensors in intimate contact with the pavement surface. The rebound of the whole FWD structure is more plausible and is observed during the FWD test. When the sensor supporting-structure 86

rebounds, the spring loaded sensors are lifted slightly up and the deflection recovery portion of each deflection-time history takes place at a faster rate than reality. This explains the positive deflections shown in Figure 6.8.

According to the above discussion only the Downward Portion (DP) of the experimental deflection-time history is suitable for comparison with 3D-FEM results. Of all sensors, the DP of the deflection-time history recorded at the center of FWD load application is the least affected by possible rigid body rebounding of the FWD structure. Figures 6.10 (a), (b), and (c) reveal that the FWDrecorded deflection-time history at zero sensor position coincides with the FE-generated deflectiontime history. Additionally, the major features of the experimentally measured response are also seen in the FEM response. These features are a time delay of displacements measured at different offsets, and a time delay between maximum loading and maximum displacement response (recorded at the center of FWD load application) of around three to five milliseconds in both the theoretical and experimental data.

6.8

EVALUATION OF BACKCALCULATION PROGRAMS

Backcalculation programs terminate as soon as either the deflection basin fit precision tolerance or the moduli convergence is satisfied. Although all the theoretical deflection basins from different backcalculation programs match the FWD measured ones, this doesnt mean that the layers moduli profiles reached by these programs are correct. Backcalculation programs interpret the measured deflection basin as being produced by an applied static load. The short load duration of FWD load does not allow the structural materials to deflect fully in response to the maximum load magnitude. This means that if the same FWD load is statically applied for a longer time duration, the deflection basin measured would be larger. Previous studies(45-48) reported that experimentally measured deflection basins produced by the application of static load (whose magnitude is equal to FWD load) are up to 80 percent larger than those measured during FWD tests. Thus backcalculation programs that are based on static interpretation of FWD load may overestimate the layers moduli. 87

The 3D-FEM approach accounts for both the inertia properties of the structural materials and the dynamic nature of FWD load. When the backcalculated moduli were used as seed moduli in the FE models they resulted in highly stiff basins. As the values of layers moduli were reduced, the FE deflection basins converged to those experimentally measured. The measured, backcalculated, and FE-generated deflection basins are compared in Figure 6.11. All deflection basins are very close to each other. However, deviations were found between the backcalculated moduli using different programs and those obtained using 3D-FEM as shown in Figure 6.12.

Using the FE-backcalculated subgrade modulus, a set of correction factors were computed for the subgrade modulus value obtained from each program when used for different types of pavements and the results are given in Table 6.1. The correction factor is computed by dividing the FE-backcalculated subgrade modulus by the value obtained from the specific backcalculation program. From the values listed in Table 6.1, the following remarks could be drawn:

1.

MODULUS has the best consistency among the three programs. The correction factors obtained for this program fell very close to the values recommended by the American Association of State and Transportation Officials (AASHTO) Pavement Design Guide (5).

2.

MODCOMP behaved similar to MODULUS with flexible pavements, however it underestimated the value of the subgrade modulus of the rigid pavement.

3.

EVERCALC produced a subgrade modulus for the flexible pavement that is very close to the FE-backcalculated value. For rigid and composite pavements it under estimated the subgrade modulus. The correction factors found for rigid pavements for this program fell within the 0.25 recommended for rigid pavements by AASHTO Pavement Design Guide (5).

4.

The FE-backcalculated modulus for the three pavements tested in this study did not significantly change. This may be due to the fact that the three pavement sections were located in Morgantown, West Virginia within a circle of radius less than seven miles. 88

The AASHTO Pavement Design Guide recommends (based on experimental observations) that a correction factor less than 0.33 is used with backcalculated subgrade modulus of flexible pavements. For concrete pavements the recommended factor is less than 0.25. It can be seen from table 6.1 that the mechanistic 3D-FEM approach in backcalculating layers moduli produced almost the same result. That is, the FE-evaluated moduli dont require correction and therefore can be used as a reference for assessing the performance of backcalculation algorithms.

89

TABLE 6.1 Correction Factors for Backcalculated Subgrade Modulus.


PAVEMENT TYPE Flexible Rigid AVERAGE 0.35 0.24 0.22

FE-Backcalculated Modulus (kPa) 27.56 26.18

MODCOMP 0.34 1.34* 0.23

MODULUS 0.37 0.28 0.29

EVERCALC 0.85* 0.19 0.13

27.56 Composite * Value excluded for the calculation of average

90

FWD loading plate Non-reflective ends of all layers 144 in. 120 in.

Non-reflective edge of all layers

Concrete 9 in. Base 8 in.

60 in. Subgrade 72 in.

Symmetry plane

Non-reflective boundaries for all layers

Non-reflective bottom

FIGURE 6.1 Finite element mesh of the flexible pavement model.

91

FWD loading plate Non-reflective edge ends of layers 162 in.

Non-reflective ends of layers

144 in.

PCC slab 9 in. Base 8 in. Subgrade 72 in. 72 in.

Symmetry plane

Reflective side of concrete slab

Non-reflective bottom

Non-reflective base & subgrade

FIGURE 6.2 Finite element mesh of the rigid pavement model.

92

FWD loading plate 144 in. 162 in.

Non-reflective ends of all layers

Asphalt 5 in. PCC slab 9.5 in. Base 3 in.

Subgrade 72 in. 72 in.

Reflective side of concrete slab

Non-reflective base & subgrade Non-reflective bottom

FIGURE 6.3 Finite element mesh of the composite pavement model.

93

a. Flexible Pavement

b. Rigid Pavement

c. Composite Pavement

FIGURE 6.4

Impact loading curves used in different finite element models.

94

a. Flexible Pavement

b. Rigid Pavement

c. Composite Pavement

FIGURE 6.5

Load-Deflection relation for different types of pavements.

95

-6.000E+01> -5.500E+01> -5.000E+01> -4.500E+01> -4.000E+01> -3.500E+01> -3.000E+01> -2.500E+01> -2.000E+01> -1.500E+01> -1.000E+01> -5.000E+00> 0.000E+00>

-7.000E+01> -6.417E+01> -5.833E+01> -5.250E+01> -4.667E+01> -4.083E+01> -3.500E+01> -2.917E+01> -2.333E+01> -1.750E+01> -1.167E+01> -5.833E+00> 0.00E+00>

a. Flexible Pavement

b. Rigid Pavement
-1.500E+02> -1.375E+02> -1.250E+02> -1.125E+02> -1.000E+02> -8.750E+01> -7.500E+01> -6.250E+01> -5.000E+01> -3.750E+01> -2.500E+01> -1.250E+01> 0.000E+00>

c. Composite Pavement Figure 6.6 Fringes of vertical stresses at time of maximum FWD load.

96

0 20 40 60 80 100 0 15

Asphalt Layer Base Layer

0 20

Concrete Base Layer Subgrade

Subgrade

40 60 80 90 100 0 15 60 45 30 Vertical Stress, psi

60 45 30 Vertical Stress, psi

75

75

90

a. Flexible Pavement.
0 20 40 60 80 100 0 15 60 45 30 Vertical Stress, psi 75 Asphalt Concrete Base Layer Subgrade

b. Rigid Pavement

90

c. Composite Pavement.
FIGURE 6.7 Vertical stresses in different types of pavement due to FWD load.

97

a. Flexible Pavement

b. Rigid Pavement

c. Composite Pavement

FIGURE 6.8 Comparison between experimental and FE deflection basins for different pavements models.

98

FIGURE 6.9 Fringes of vertical displacement at time of maximum FWD load. (Display scale factor 2500) 99

FIGURE 6.10 Deflection-time histories for different pavement models.

100

a. Flexible Pavement

b. Rigid Pavement

c. Composite Pavement

FIGURE 6.11 Comparison between backcalculated deflection basins and measured basins.

101

Surface layer (Ksi)


1000 Modulus, Ksi Modulus, Ksi 800 600 400 200 0 120 100

Base layer (Ksi)


14 12 Modulus, Ksi 10 8 6 4 2 0

Subgrade (Ksi)

80 60 40 20 0

a. Backcalculated layer moduli for flexible pavement


Surface Layer (Ksi)
5000 Modulus, Ksi 4000 Modulus, Ksi 3000 2000 1000 0 120 100 80 60 40 20 0 Modulus, Ksi

Base Layer (Ksi)


25 20 15 10 5 0

Subgrade (Ksi)

b. Backcalculated layer moduli for rigid pavement


Surface Layer (Ksi) 2500 2000

PCC Layer (Ksi)


7000 6000 Modulus, Ksi 5000 4000 3000 2000 1000 0
Modulus, Ksi

Base Layer (Ksi) 200 150 Modulus, Ksi 100 50 0 35 30 25 20 15 10 5 0

Subgrade (Ksi)

Modulus, Ksi

1500 1000 500 0

c. Backcalculated layer moduli for composite pavement

FIGURE 6.12 Comparison of backcalculated layer moduli for the three types of pavements.

102

CHAPTER 7

SOME FACTORS INFLUENCE BACKCALCULATIONS OF RIGID PAVEMENTS

7.1

INTRODUCTION This chapter focuses on examining the behavior of rigid pavement layers during the Falling

Weight Deflectometer (FWD) test; factors affecting the design of a concrete slab, such as whether the joints are doweled or undoweled and the spacing between transverse joints were considered. Explicit finite element analysis was employed to investigate the response of pavement layers to the action of FWD impulse load. The accuracy of the finite element models developed in this investigation was verified by comparing the finite element-generated deflection basin with that experimentally measured during an actual test. The results showed that the measured deflection basin can be reproduced through finite element modeling of the pavement structure. The resulting deflection basins from different models which simulate different pavement design features were processed using several backcalculation programs. The results reveal the effect of different pavement design features on the backcalculated moduli profile. It was found that ignoring the dynamic nature of the FWD load may lead to crude results, especially during backcalculation procedures.

7.2

FINITE ELEMENT STRUCTURAL MODEL

A rigid pavement (SHRP section No. 285823) located in Mississippi was selected for this study. The pavement structure was modeled as a multilayered linear elastic system consisting of a Portland Cement Concrete (PCC) slab, base, and subgrade as shown in Figure 1. The model dimensions were chosen to be the full lane width, that is 12 ft and extended in the longitudinal direction from both sides to include two quarter parts of the adjacent slabs. The slab length was taken to be 20 ft . To study the effect of the slab length on the performance of a rigid pavement under the 103

effect of FWD load, four other models were developed using slabs of lengths 16 ft, 15 ft , 12 ft, and 10 ft. The slab width was kept constant, 12 ft for all models. The center of the loading plate was located at the center of the middle slab for all models. Due to the geometrical symmetry around the longitudinal plane passing through the center of the loading plate, only one half of the pavement was meshed. In absence of data about the depth to bedrock, it was assumed to be typical of what can be found in West Virginia, i.e. from 5-15 ft under the bottom of the base layer. In this study, it was taken as 6 ft measured from the bottom of the base layer. Additionally, nonreflective boundaries were applied at the bottom of the subgrade to eliminate reflection of the stress wave from affecting the surface displacements. To account for the effect of model size, nonreflective boundaries which simulate a semi-infinite extension of layers were applied at all sides of base and subgrade as well as the transverse ends of the two half slabs as shown in Figure 7.1. A sliding interface was assumed between the concrete slab and the base layer, and a fully bonded interface was assumed between the base layer and the subgrade. All layers were meshed using 8-node brick elements having 24 degrees of freedom per element. The mesh sizes varied through the model to assure the accuracy of the results within the regions of interest. Thus a refined mesh was necessary in regions of high stress intensity such as concrete slabs especially at transverse joints. A coarser mesh was used for the base and subgrade.

The steel dowel bars (modulus of elasticity=30, 000 psi, unit weight= 488.81 pcf) in the transverse joints were modeled using brick elements. The main function of dowel bars is to transfer the load across the joint while allowing the slabs to move longitudinally relative to each other in order to relieve the tensile or thermal stresses due to slab contraction. Therefore the dowel bar may be bonded to one concrete slab while its other end is free to slide in the adjacent slab. Consequently, one end of the dowel bar was modeled with a tied interface with the concrete slab, while the other end was modeled with a sliding interface with the adjacent concrete slab as illustrated is Figure 7.2.

104

7.2.1 Model Loading and Material Model

The FWD impact load was applied to the model through a 12 in. diameter steel loading plate. The impact load used in this investigation was obtained from the experimental results of SHRP section No. 285803 reported by J. Mallela, et al. (48). The impact pressure on the loading plate surface was assumed to be uniformly distributed due to the semi-rigid behavior of the plate. The pressure-time relation, Figure 7.3, was digitized over time increments of 0.25 milliseconds and the values were used in the finite element program . The different load magnitudes employed in the FWD tests are designed so that all pavement layers behave within their elastic limits. Thus the assumption of linear elastic behavior of all layers is realized for this study. This assumption is further confirmed by research studies (48,49) which indicate that the stresses induced in different pavement layers under a maximum impact pressure of 92.8 psi (used in this study) are likely to be within the elastic range. Therefore linear elastic materials models were used for all layers and the material parameters are given in Table 1.

TABLE 7.1
Material Concrete Slab, thickness= 8 in. Base Layer, thickness= 6 in. Subgrade, thickness =72 in.

Properties of Pavement Materials.


Property Modulus (Ksi) Poissons Ratio Unit Weight (pcf) Modulus (Ksi) Poissons Ratio Unit Weight (pcf) Modulus (Ksi) Poissons Ratio Unit Weight (pcf) Values published in Ref. (48) 4650 0.18 149.8 1700 0.40 133.8 11.87 0.30 129.85

105

7.2.2 Model Verification

Figure 7.4 (a) illustrates a comparison between the experimental deflection basin obtained from SHRP data base for section No. 285803 and the corresponding deflection basin obtained from the 3D-FEM model. Initially, the value of 11.87 ksi for the subgrade modulus, backcalculated by Mallela (48), was used in the model. The FEM model produced a deflection basin which was on average 16% less then the experimentally measured one. However, both the theoretical and experimental basins showed a remarkable agreement in their slopes, this agreement indicated that the FEM model is slightly stiffer than the actual pavement. The higher stiffness of the model may be reduced by decreasing the layers moduli values used in the model. When the modulus of the subgrade was reduced to 8.0 ksi, while keeping the moduli of the surface and base layers as listed in Table 7.1, a remarkable agreement between the theoretical and experimental results was realized as illustrated in Figure 7.4 (a). Since the original modulus value of 11.87 ksi was evaluated using backcalculation (Mallela (48)), a 32% reduction in the modulus value will be within the expected error associated with most backcalculation algorithms. In fact, it is not uncommon for the error in evaluating subgrade modulus to exceed 100% (see Reference13). Even in well controlled laboratory testing of subgrade modulus, errors more than 50% are expected.

The most important thing to consider is the ability of the model to simulate the relative displacement of any point to another one on the structure. The success of this model is demonstrated in the good agreement between the slopes of the deflection basins as we move away from the point of load application. Only at a distance of 48 in. did the slopes of the experimental and theoretical basins start to show some deviation. One reason for this divergence may be the approximate values adopted for the moduli of the base and subgrade layers which were evaluated using backcalculation, the value of subgrade modulus is the only one adjusted in this study. Another reason may be due to unreported dowel bar looseness or loss of aggregate interlock. The third reason may be simply the experimental error encountered in measuring very small displacements away from the center of the loading plate. In order to illustrate the effect of dowel bars on the deflection basin, another FEM model was constructed without dowels at the joints and its deflection basin was compared with the 106

experimentally measured one in Figure 7.4 (b).

7.3

PERFORMANCE ASSESSMENT OF BACKCALCULATION PROGRAMS

In this study, finite element modeling was used

for the evaluation of conventional

backcalculation programs. Since a FWD measured deflection basin could be reproduced through finite element modeling of the pavement structure, the FEM generated deflection basin can be used to backcalculate the layers moduli. To evaluate the performance of different backcalculation algorithms the following procedures were followed:

1.The rigid pavement structural model results were verified by comparing the FEM-generated deflection basin with the FWD measured one as shown in Figure 7.4.

2. The theoretical deflection basin obtained from the finite element model was used together with different backcalculation programs to evaluate the moduli of different layers, using different backcalculation programs.

3. The backcalculated layers moduli were compared with the moduli used in the finite element model as shown in Figure 7.5.

Figure 7.5 (a) illustrates the measured deflection basin together with those resulting from using different backcalculation programs. It can be noticed that a remarkable agreement was achieved between all the backcalculated basins and the measured one. However, the backcalculated layers moduli obtained from each program were found to be different from those used in the finite element model as shown in Figure 7.5 (b). This is primarily due to the dynamic nature of the FWD which was accounted for in FE analysis while the backcalculation programs are based on static analysis. Figure 7.5 (b) also indicates that all programs have produced top layer moduli close to that used in the FE model. The values of the base modulus evaluated using conventional backcalculation programs are 107

significantly less than those used in the FE model. The subgrade layer modulus produced by all three backcalculation programs was significantly larger than that used in the finite element model. However, if we apply a correction factor of 0.28 (as recommended in Chapter 6 for MODULUS program), in average, on the backcalculated subgrade modulus, the results become almost identical to the modulus used in the finite element model. This further confirms the validity of the correction factors reached in Chapter 6.

7.4

EFFECT OF SLAB LENGTH AND DOWEL BARS

The FE-generated deflection basins, for the models provided with dowel bars, are plotted for different slab lengths in Figure 7.6 (a). The deflection basins are congruent indicating that, in the presence of dowel bars, the slab length doesnt affect the surface deflection values resulting from FWD impact. This can be explained by the fact that the dowel bars transfer the load to the adjacent slabs. Comparison of the deflection basins of the doweled models in Figure 7.6 (a) with those of undoweled ones in Figure 7.6 (b) shows that dowel bars affect the value of the maximum deflection measured at the center of FWD loading plate. Away from the center of load application, dowel bars have the effect of introducing continuity of deformation at the transverse joint. Therefore no effect of slab length on the measured deflection basin can be observed in Figure 7.6 (a) for doweled pavement structures. In absence of dowel bars, Figure 7.6 (b), the slab length becomes an important factor which affects the FWD deflection basin.

The plots in Figure 7.6 (b) for undoweled pavements indicate that as the slab length increases, the surface deflection increases then begins to decrease at a slab length 15 ft. This is further illustrated in Figure 7.7 which shows the change of the deflection under the center of FWD loading plate versus slab length. Referring to Figure 7.6 (b), the difference between the deflections recorded at the locations of the first and last sensors increases as the slab length increases. This causes backcalculation programs to produce values of layers moduli which are slab length dependent for undoweled pavement sections. 108

The deflection basins plotted in Figure 7.6-b were processed using MODULUS5.0 program to study the effect of slab length on the backcalculated layers moduli. The resulting layers moduli for each slab length together with the estimated depth to bedrock are plotted together with the value used in FEM as shown in Figure 7.8. The plots show a significant change in the backcalculated moduli values obtained for both the surface and the base layers from those used in FEM. However, the values for subgrade modulus are found to be close to each other but still significantly larger than that used in FEM. Therefore it seems that any change in the surface deflection affects the moduli of the top layers more than they do the subgrade. This makes the use of a correction factor of 0.28 for the subgrade modulus satisfactory for all slab lengths.

7.5

CONCLUSIONS

The 3D FEM approach used in this study provides a powerful tool for evaluating the performance of existing backcalculation programs with rigid pavements under different situations encountered in the field. Based on the presented results, the following conclusions can be drawn:

1.

When testing an aged rigid pavement section in which cracks have developed, the FWD testing engineer should check that the length of the tested slab part is not less than 10 ft to assure that this part can produce a reliable deflection basin.

2.

The correction factor of 0.28 reached in Chapter 6 is suitable for adjusting the backcalculated subgrade modulus in all cases of doweled and undoweled joints. This value was found to be in close agreement with the AASHTO experience-based correction factors.

109

Loading Plate, 12 in. Diameter 240 in. 60 in. Concrete Slab, 8 in. Base, 6 in. Subgrade 72 in.

Symmetry plane Non-reflective sides 60 in. of concrete, base and subgrade layers Quarter slab

Reflective slab sides

Nonreflective sides 144 in. Non-reflective side of base & subgrade Non-reflective bottom

Figure 7.1 Finite element mesh for a rigid pavement.

Figure 7.2 Cross section in a doweled joint.

110

90

60

30

0.01 Time, sec

0.02

0.03

Figure 7.3 Impact load curve used in finite element model (Ref. (48)).

111

Figure 7.4 Model verification. (Slab length=20 ft)

112

1.6

2.4

3.2

Measured FEM, Esub=8.00 ksi MODULUS5. EVERCALC4. 0 MODCOMP3 0 16 32 48 Distance from loading plate center, in. 64

4.0

a. Deflection Basins.
50000 45000 40000 35000

14000 Concrete Slab Modulus, MPa 12000 10000 8000 6000 4000 2000 0
Modulus, MPa

300

Base Layer

250 200 150 100 50

Subgrade

Modulus, MPa

30000 25000 20000 15000 10000 5000 0

Published, Ref. (48)

FEM

MODULUS 5

EVERCALC

MODCOMP 3

b. Layers Moduli. Figure 7.5 Comparison between the results of backcalculation programs.

113

Figure 7.6 Effect of slab length on deflection basin.

114

0 10 12 14 16 18 20 22

FIGURE 7.7 Change of maximum deflection with slab length for undoweled pavement.

115

60000 50000 Modulus, MPa 40000 30000 20000 10000 0 FE

Surface Layer
Moulus, MPa Depth to bedrock, m.

16000 Base Layer 14000 12000 FE 10000 8000 6000 4000 2000 0

160 140 120 100 80 60 40 20 0

10

Subgrade
8 6 4 2 0

Depth to Bedrock

Modulus, MPa

FE

1.0 MPa=0.14504 ksi)

Figure 7.8 Effect of slab length on the backcalculation results ( Using MODULUS).

116

CHAPTER 8

EFFECT OF 3D FEM MODEL DEPTH ON BACKCALCULATIONS OF FLEXIBLE PAVEMENTS

8.1

INTRODUCTION Many researchers (Uddin et al.(54) , Yang et al. (55), Briggs et al. (56), Rhode et al. (57), and

Uzan (58)) reported that mechanistic analysis of pavement response to FWD impact load shows a dependancy on the thickness of subgrade layer. Such a conclusion was reached under the assumption that pavement response to a static load is equivalent to its response to a FWD impact load of the same amplitude. Thus the elastic layer theory can be used to backcalculate the layers moduli. However, Mamlouk (45), Seebaly (46), and Mallela (48) reported that the deflections produced under impact loads may be from 40 to 80 percent less than those observed if the load was static. Under the assumption of a statically applied load, each pavement layer, including the full subgrade depth, has enough time to fully deflect, which is not the case if the applied load is a short duration impact. On the other hand, the use of elastic layer theory and a static loading assumption required a knowledge or estimate of the thickness of subgrade layer that can be used in backcalculation programs in order to produce the same deflection measured from FWD test. The theoretical dependancy of FWD surface deflections (predicted from the elastic layer theory) on the assumed thickness of the subgrade layer is not supported in the literature by any known experimental measurements where sites that have the same layers moduli profile but different depths to bedrock are tested using FWD. Such measurements would be extremely difficult because of the variation of subgrade conditions from one location to another. On the other hand, if FWD testing is to achieve its objective, a means of nondestructive testing has to be developed to predict the appropriate depth of subgrade layer that should be used in backcalculation algorithms.

Chang et al. (59) attempted to establish an analytical correlation between the results obtained using Dynaflect (harmonic load) and those using FWD (impact load) to predict a depth to bedrock.

117

His results indicate that only the free vibration part of the FWD displacement-time history may be correlated to the depth to bedrock; none of his results show that the FWD deflection basin is influenced by the depth to the stiff layer. Bush (61) suggested using an arbitrary subgrade thickness such as 20 ft in elastic layer-based backcalculations. Uddin (54) recommended that the actual subgrade thickness should be used. Seng et al. (61) found that the depth to bedrock could be related to the frequency of the free vibration portion of FWD sensors and the shear wave velocity in the subgrade layer. For flexible pavements: DB = Td 1.08 Vs 1.13 / (6.33 -5.04 ) and for rigid pavements: DB = Td 1.11 Vs 1.14 / (6.21 - 3.88 ) Where: Td = the period of the free vibration portion of the FWD sensors, seconds. Vs = the shear wave velocity in the subgrade layer, ft/s. (2)

(1)

 = Poissons ratio of the subgrade material.


The major shortcoming in Sengs approach is that the free vibration amplitudes from FWD sensors are extremely small, which makes accurate measurement of the free vibration period difficult if not impossible for many sites. If the layers interfaces are not fully bonded, the free vibration part of surface displacements may not be related to the depth to bedrock.

Ullidtz (62) used Boussinesqs equation to establish the concept of surface modulus and compute the depth to bedrock. The surface modulus is defined (63) as the stiffness modulus of an equivalent half space pertaining to a specific spacing from the load center that produces a deflection identical to the deflection actually measured on the layered structure at the same spacing. The surface modulus at a specific radius r from the center of FWD load application is given by: Esm(r) = 2 (1-2) P/(% r dr ) 118

(3)

where: Esm(r) = Surface modulus at distance r from load center.

 = Poissons ratio.
P = Applied load. dr = Deflection at distance r.

Ullidtz pointed out that if a stiff layer is found at a certain depth, it will have a consequence that no surface deflection will occur beyond the offset at which the stress zone intercepts the stiff layer. Thus the depth to bedrock can be obtained from the radial distance from the center of FWD loading plate at which the surface deflection reduces to zero.

Rohde et al. (57) used Boussinesques analysis and the subsequent development by Ullidtz to develop a set of regression expressions which accounted to the overall shape of the deflection basin in the estimation of the apparent depth to bedrock. For flexible pavement sites having an AC layer thickness greater than six inches and tested using 9000 lb FWD load, the apparent depth to bedrock B is obtained from: 1/B= 0.0409 + 0.5669 ro + 3.0137 ro2 +0.0033 BDI - 0.0665 log (BCI)

(4)

where: ro= 1/r intercept by extrapolating the steepest part of the inverse radial distance (1/r) versus deflection curve, 1/ft. BDI= Base Damage Index defined as the difference in surface deflection measured at radial distances of 12 in. and 24 in., mils. BCI= Base Curvature Index defined as the difference in surface deflection measured at radial distances of 24 in. and 36 in., mils.

Other relations similar to Equation (4) were developed for different thicknesses of asphalt layers and were subsequently implemented in MODULUS (16) backcalculation program. 119

Zaghloul et al. (41) examined the dependency of FWD deflection basin on the depth to bedrock using 3D Finite Element Modeling (3D FEM). The FWD load was applied to the model as an impact load of the same duration as measured from FWD load cell. The surface deflections were computed and compared with those experimentally measured. He went on to change the depth of the subgrade layer in the model from 95 in. to 140 in. while keeping the layers moduli and the applied load the same. He reported that the variation of the depth to bedrock in the 3D FEM model had no effect on the deflection basin, a conclusion that contradicts the findings in references 54-58.

Analysis of pavement structural response using 3D FEM analysis offers many advantages not normally available using any other analytical approach. Perhaps one of the most important advantages offered by 3D FEM is the ability to accurately simulate the nature and distribution of the applied load. Other major advantages are: 1. accurate modeling of the pavement geometry including discontinuity at joints and modeling of the surface layer with finite width bounded by shoulders, 2. the friction and separation at pavement layers interfaces are accounted for in the analysis, and 3. any form of material behavior could be easily included in the model.

Realizing these advantages, Shoukry et al. (64) developed a new backcalculation algorithm using 3D FEM. The method was shown to produce excellent results when tested on Flexible, Rigid, and Composite pavement sites. The backcalculated moduli using 3D FEM were compared, for several sites, with those obtained using three backcalculation programs: MODCOMP, MODULUS, and EVERCALC. The 3D FEM results were considered to be more accurate than traditional backcalculation algorithms. Correction factors were developed to adjust the values of layers moduli evaluated using the MODULUS backcalculation program to the corresponding moduli values obtained using 3D FEM. The mechanistically evaluated correction factors were found to be in close agreement with the experience-based values recommended in the American Association of State Highway Officials and Transportation (AASHTO) Pavement Design Guide.

120

The work presented in this chapter was initiated as a result of a round-robin test administered by the Transportation Research Board (TRB) A2B05 subcommittee on Backcalculations of layers moduli. In this test, ten different research groups and individuals1 were requested to predict the depth to bedrock from the FWD measurements for two of four flexible pavement sites located in Texas, USA. Texas DOT who provided the FWD data provided also the thicknesses of the surface and subgrade layers of every site and the temperature of top, middle, and bottom of the asphalt layer for every site. They also provided the measured depth to bedrock for sites No. 2 and 4. The author participated in the test as it provided a good opportunity to evaluate the reliability of the 3D FEM based backcalculation algorithm that was developed in Chapter 6 and to achieve the following objectives:

1. Develop procedure for the evaluation of the apparent depth to bedrock from 3D FEM pavement structural models which simulate FWD testing.

2. Investigate the effect of finite element model depth and the reflection from the model bottom on the calculated deflection basin.

8.2

FINITE ELEMENT STRUCTURAL MODELS

The 3D FE models used for the backcalculation of the four Texas sites were identical to the flexible pavement model that was used to calculate the flexible pavement site of West Virginia, described in Chapter 6 and shown in Figure 8.1. The same input file that was used to generate the flexible pavement model of chapter 6 was modified to be a parametric one. The advantage of using a parametric input file is that it allows the user to change only layer thicknesses, applied load, and

1 The participants at this round-robin test were: 1. F. Emanuel, University of Texas A& M. 2. D. Alexander, Waterways Experimental Station. 3. G. M. Rowe and M. J Sharrock, England. 4. L. Irwin, Cornell University. 5. P. Ullidtz, Technical University of Denmark.

6. J. Mahoney, University of Washington. 7. Y. R. Kim, North Carolina State University. 8. W. Uddin, University of Mississippi. 9. S. Shoukry& G. William, West Virginia University. 10. J. Uzzan, Technion, Israel.

121

material properties of different layers while the number of nodes, number of elements, model boundaries, layers interface properties, conditions at model boundaries, and the subgrade layer depth remained unchanged from those reported in Chapter 6. The parametric input file is listed in Appendix A. In each model, the thicknesses of pavement layers were set to those of Texas sites listed in Table 8.1. The model was loaded using the measured FWD load provided by Texas DOT and shown in Figure 8.2. In absence of any prior knowledge of the expected depth to bedrock , the thickness of the subgrade layer in the model was left at 72 in.

8.3

EVALUATION OF LAYER MODULI

The iterative procedures described in Chapter 6 were used to backcalculate the moduli profile of every site. The final FEM-computed deflection basins were plotted together with those experimentally measured in Figure 8.3. For site 1, Figure 8.3 (a) shows a deviation between the experimental and theoretical deflection basins at the position of the second FWD sensor. The reason for this deviation is the presence of a large thermal gradient (16.6 o C) in the asphalt layer, and the simplifying assumption (used in 3D FEM calculation) of an average elastic modulus for the asphalt layer. Although the use of viscoelasto-plastic material model for the asphalt layer may improve the convergence between the theoretical and experimental deflection basins shown in Figure 8.3 (a), this would result in longer execution time. Thus for practical purposes, it was decided to use an average elastic modulus for the asphalt layer and perform a temperature correction on the backcalculated modulus similar to the practice used in elastic backcalculation programs.

Referring to the temperature measurements listed in Table 8.1, the backcalculated moduli values of the asphalt layer should be adjusted to the standard temperature of 20 o C (68 o F). The following relation (65 ) used by the Japanese Highway agencies was used:

Estandard= Eas * 10

-0.0184 (20-T )

(5)

122

where: Estandard = Modulus at temperature 20 o C (68F). Eas T = 3D FE-based backcalculated modulus at field temperature. = The mean temperature of asphalt (o C).

The 3D FEM-backcalculated layers moduli for each site are listed in Table 8.2 together with three other sets of results that were independently computed by Ullidtz (66-68) who participated in the TRB-A2B05 round robin test. Ullidtz used three different backcalculation algorithms:

1. A two dimensional (axially symmetric) finite element (2D FE) backcalculation program with non-linear material models. 2. Waterways Experiment Station (WES) backcalculation program that is based on the elastic layer theory. 3. A backcalculation program based on the Method of Equivalent Thickness (MET) with a non-linear subgrade material model.

Examination of values listed in Table 8.2 reveal that the 3D FEM approach developed in this Thesis independently produced layer moduli values that were close to the values provided by Ullidtz for all sites. This indicates that the use of elastic material models in 3D FEM approach has no effect on the backcalculated moduli specially for the subgrade layer. Examination of the subgrade layer moduli backcalculated using different approaches reveals that the use of a nonlinear material model for subgrade in the MET method resulted in subgrade moduli values which are close to those obtained using the 3D FEM method and elastic material model. It should be remembered that the correction factors computed in Chapter 6 are specific to MODULUS backcalculation program and dont apply to any of the three backcalculation algorithms used by Ullidtz.

8.4

EVALUATION OF DEPTH TO BEDROCK

Examination of the profile of displacement decay with depth at the location of the central 123

FWD sensor position shown in Figure 8.4 reveals that such a decay relation may be used to define an apparent depth to bedrock. Thus, the same 3D finite element model used to backcalculate the layers moduli can also be used to evaluate the depth to bedrock. The maximum vertical displacement that propagates along the vertical line passing through the center of the FWD loading plate can be obtained from the finite element model. From Figure 8.4, it seems plausible to assume that the maximum vertical displacement measured at the center of the FWD loading plate decays exponentially with depth. This assumption can be verified by plotting the natural logarithm of the deflection versus depth as shown in Figure 8.5. The relation is a straight line over a significant portion of the curve. Deviation from straight line relation is due to the effect of the fixed boundary at the model bottom. The exponential decay of the maximum displacement in the subgrade layer with increasing subgrade depth is assumed to be:

z =  ez
where: z = Deflection at Depth z

(6)

 = FE calculated deflection at the top of subgrade.


Z = the depth of subgrade measured from top of subgrade to the point under consideration. The displacement decay constant  is determined in this analysis from the plots of Figure 8.5. From Equation (6), the decay constant  can be calculated as

=1/Z . ln( z/ o)

(7)

In order to minimize the error in calculation of the constant , it was calculated for all the points which fall on the straight line portion of ln ( ) versus depth graph and the average value was used to calculate the depth to bedrock. The values of the decay constant  obtained from Equation (7) for the four sites are listed in Table 8.3. The apparent depth to bedrock is defined as the subgrade depth at which the vertical 124

displacement becomes very small. The definition of a small displacement can be related to the decay of stress in subgrade layer. As it is common in the calculation of instantaneous settlement of a foundation resting on a deep layer, the soil depth beneath the foundation should not be less than the depth at which the maximum vertical stress decays to 10 percent of its value at the foundation level (66). This depth can be identified by plotting the ratio between the vertical stress induced at any depth in the subgrade to its value at the subgrade top as shown in Figure 8.6. The apparent depth to bedrock can be predicted using the stress decay relation. However, the use of displacement decay seems to be more convenient since the accuracy of calculating the displacement in FEM is higher than that of stress calculation since the latter is dependent on the differentiation of nodal displacement with respect to nodal position, so it is mesh dependant.

The plots in Figure 8.5 reveal that as the stress decayes to 10 percent of its value at the top of subgrade, there is an 80 to 85 percent decay in subgrade displacement. Thus the deflection ratio

/ o found at 10 percent decay of vertical stress is substituted in Equation (7) to calculate the
apparent depth to bedrock, that is: Apparent depth to bedrock =[ ln( / o) at 10% stress decay]/

(8)

The final values obtained after adjusting the computed depths to bedrock, so that their values are measured from the top of the surface layer of each site, are listed in Table 8.3.

8.5

EFFECT OF 3D FE MODEL DEPTH ON 3D FEM RESULTS

Modeling of pavement structure in 3D-FEM requires adopting a pre-assumed depth of the subgrade. The effect of this assumed depth on the model results should be examined. After evaluation of pavement layers moduli and the depth to bedrock for each site using a model depth of 95 in. (subgrade thickness= 72 in.), each 3D FE model was modified by changing the subgrade thickness first to 40 in. and then to 144 inch. After each modification, the model for each site was loaded with

125

the corresponding FWD load and the deflection basin was obtained.

8.5.1 Effect of Model Depth on Deflection Basin

The 3D FEM-calculated deflection basins for different model depths are shown in Figure 8.7. In each case, the subgrade depth was changed while keeping the layers moduli and the applied load the same. As the depth to bedrock increases, the deflection values also slightly increase. However, the amount of increase in deflection diminishes with the increase of the model depth. This means that after a certain model depth, the deflection basin will remain the same and the subgrade below that depth will not contribute to the surface deflection. This depth is the apparent depth to bedrock. The differences observed between the deflection basins resulting from different model depths are very small. Such differences are not expected to affect the accuracy of the backcalculated moduli. This agrees with Zaghloul et al. (41) who found that the variation of depth to bedrock doesnt affect the peak surface deflections predicted by 3D-FEM.

8.5.2 Effect of Stress Wave Reflection from Model Bottom on Deflection Basin

To examine the effect of stress wave reflection from the model bottom on the 3D FEMgenerated deflection basin, reflective boundaries were applied to the bottom of the shallowest model of each site (i.e. model with 40 inch subgrade layer thickness). Figure 8-8 illustrates that the wave reflection causes a slight increase in the surface deflection measured at the center of the FWD loading plate and decreases the deflections at sensor positions away from the center. The small deviations observed in Figure 8-8 are not expected to have any significant effect on the backcalculated moduli.

8.5.3 Effect of Model Depth on the 3D FEM-Calculated Depth to Bedrock

The maximum vertical displacement in the subgrade at the position of the FWD loading plate

126

for each pavement site is plotted on a horizontal logarithmic scale versus depth in Figure 8.9. The plots reveal that the change of the model depth slightly affects the maximum vertical displacement observed on the top of subgrade ( o). The upper portions of the plots for model depths 95 in. and 167 in. are almost congruent to each other. On the other hand, the vertical displacements of the shallow models differ significantly from those obtained for the other two deeper models which indicate that their results are influenced by the boundary conditions set on their subgrade bottoms.

The depth to bedrock was calculated for each model depth for every sites and the results are listed in Table 8.5. The results indicate that the predicted values for the depth to bedrock resulted from the models of depths 95 in. and 167 in. are very close to each other. The values obtained from the shallow model are up to 25% less than those of the deeper models because of the effect of the fixed boundary at the model bottom.

8.5.4 Effect of Model Depth on Stresses Induced in Subgrade

The above investigation of the effect of model depth is expanded to include the effect on the stress level induced in the subgrade layer. For this reason, the vertical stress distribution is plotted for the three different model depths as shown in Figure 8.10. The Figure shows the plots only for site 3 as an example and the other sites are similar. The plots show that at the upper one foot of the subgrade, the vertical stresses obtained from different models are congruent to each other. For the models of depths 95 in. and 167 in., the stress plots remain congruent up to depth of 2 ft from the subgrade top, after which the boundary condition at the model bottom influence the stress profile in the remaining subgrade depth.

From the above discussion, the use of subgrade layer thickness of 72 in. seems to be suitable for use in finite element modeling of flexible pavements. A good degree of accuracy can be realized since an increase in the model depth will not result in a significant change in the model results. Limiting the subgrade thickness to 72 in. from the surface can significantly reduce the model size

127

which decreases the required computer memory and the computational time.

8.6

EFFECT OF LOAD DURATION ON THE DEFLECTION BASIN

To illustrate the influence of FWD loading time duration on the deflection basin, the flexible pavement model used in site 3 was processed after modifying the load-time history so that the peak of the load acts for a duration of 40 millisecond as shown in Figure 8.11 (a). The FEM-generated deflection basin was plotted together with the one obtained using the shorter duration FWD load (as measured) as shown in Figure 8.11 (b). The results in Figure 8.11 (b) illustrate that the surface deflection under the center of the FWD loading plate increased by 150 percent. The increase in the surface deflection decreases with the increase of radial distance from the center of load application. The increase in the maximum deflection due to the increase of maximum load-time duration confirms the findings of references (45-48) which reported that the deflection basins produced by an impact load are 40 to 80 percent less than those produced by a static load of the same magnitude.

The most interesting point in Figure 8.11 (b) is that the two deflection basins intersect at an offset of 48 in. from the center of the FWD load. This agrees with the experience-based choice of this distance by some researchers (18,54) to calculate the subgrade modulus. The results indicate that the deflection value at this sensor location is not affected by the assumption that FWD is acting as a static load.

8.7

CONCLUSION

Based on the work presented in this chapter, a new mechanistic approach for evaluating the apparent depth to bedrock has been suggested. This method can be applied for all types of pavements. The method was verified by comparison with field measurements, as well as with the results of other existing methods.

128

The effect of the pre-assumed FE model depth on the finite element results has been studied. From this study, the following conclusions can be made:

1. The FE-generated deflection basin is insignificantly affected by the assumed model depth. 2. The FE-generated deflection basin is not sensitive to stress wave reflections at the bedrock. 3. The apparent depth to bedrock is the depth under which the subgrade will not deflect due to the application of the FWD load. This indicates that the zone influenced by the FWD load is in the load vicinity, which makes any prior knowledge about the depth to bedrock unnecessary in backcalculation using 3D FEM. 4. For research purposes, modeling a flexible pavement structure with a subgrade thickness of 6 ft seems to be suitable for producing satisfactory results for both displacements and stresses at the top of the subgrade.

129

TABLE 8.1

Layer Thicknesses and Temperature Measurements. Site 1 Site 2 8.00 15 30.6 28.7 27.5 Site 3 7.25 15 24.2 24.3 24.4 Site 4 7.25 15 22.8 21.1 19.4

Asphalt layer thickness, in. Base layer thickness, in. Temperature Measurements: in. from asphalt top (o C) Mid depth (o C) in. from asphalt bottom (o C)

8.00 15 57.2 50.5 40.6

130

TABLE 8.2

Backcalculated Pavement Layers Moduli, Ksi.

A. Surface Layer Method 3D FEM 2D FE WES MET Computed by This study Ullidtz Ullidtz Ullidtz Site 1 135 160 165 154 Site 2 56 140 144 123 Site 3 160 790 852 669 Site 4 170 669 703 560

B. Base Layer Method 3D FEM 2D FE WES MET Computed by This study Ullidtz Ullidtz Ullidtz Site 1 450 33.5 31.5 32.5 Site 2 90 28.6 27.8 38.1 Site 3 120 32.8 23.8 35 Site 4 47 24.5 19.5 30.6

C. Subgrade Method 3D FEM 2D FE WES MET Computed by This study Ullidtz Ullidtz Ullidtz Site 1 18 36.8 32.1 20.0 Site 2 10 10.5 11.2 8.5 Site 3 12 18.2 18.8 11.9 Site 4 8.5 11 12.3 8.5

131

TABLE 8.3

Comparison Between Measured and Calculated Depth to Bedrock.


Site 1 Site 2 -0.03478123 0.1443 78.65 Site 3 -0.03488303 0.1446 77.69 Site 4 -0.03333101 0.1617 76.91

Decay Constant (), Model Depth= 95 in. ( / o) at 10% stress decay Calculated depth to bedrock, in.

-0.03059682 0.1903 76

Measured depth to bedrock, in. (By Texas DOT)

30.5

76

63

103

132

TABLE 8.4

Effect of Assumed Model Depth on the Calculated Depth to Bedrock. Site 1 Site 2 Site 3 Site 4

Decay Constant () Model depth= 63 in. Model depth= 95 in. Model depth= 167 in. -0.04471782 -0.03059682 -0.03007312 -0.046714987 -0.034781234 -0.034196126 -0.045303345 -0.034883027 -0.033572694 -0.034748846 -0.033331011 -0.032929402

Calculated Depth to Bedrock, in. Model depth= 63 in. Model depth= 95 in. Model depth= 167 in. 60.10 77.22 78.17 64.44 78.65 79.61 64.94 77.69 79.85 74.68 76.91 77.58

133

- Sliding interface between the asphalt layer and the base (Coefficient of friction=0.90). - Fully tied interface between the base and the subgrade. FIGURE 8.1 Finite element model.

134

140 120 100 80 60

120 100 80 60

40 20
0 -20 0 0.01 0.02 0.03 0.04 Time, sec. 0.05 0.06

40 20
0 -20 0 0.01 0.02 0.03 0.04 Time, sec. 0.05 0.06

a. Site 1
120 100 80 60 120 100 80 60

b. Site 2

40 20
0 -20 0 0.01 0.02 0.03 0.04 Time, sec. 0.05 0.06

40 20
0 -20 0 0.01 0.02 0.03 Time, sec. 0.04 0.05 0.06

c. Site 3

d. Site 4

FIGURE 8.2 Measured FWD impact load curves used in finite element models. 135

0 0

20

40

60

80 0 -5 -10

20

40

60

80

-5

-10

Exp
-15

-15 -20 -25

Exp FE

FE

-20

0 0 -2 -4 -6 -8 -10 -12

20

40

60

80 0 -2 -4 -6 -8

20

40

60

80

Exp FE

-10 -12 -14 -16

Exp FE

FIGURE 8.3 Measured and FE-calculated deflection basins.

136

FIGURE 8.4 Subgrade vertical displacement versus depth.

137

FIGURE 8.5 Subgrade vertical displacement on a logarithmic scale versus depth.

138

40 80

40 80

120 Stress decay 160 0 Displacement decay 20 40 60 80 Percentage of reduction 100

120 Stress decay 160 0 20 Displacement decay 40 60 80 Percentage of reduction 100

a. Site 1

b. Site 2

40 80

40 80

120 Stress decay 160 0 Displacement decay 20 40 60 80 Percentage of reduction 100

120 Stress decay 160 0 Displacement decay 20 40 60 80 Percentage of reduction 100

c. Site 3

d. Site 4

FIGURE 8.6 Decay of vertical stress and displacement in subgrade.

139

0 0

20

40

60

80 0 -5

20

40

60

80

-5 -10 -10 Measured Depth=63 in. -15 Depth= 95 in. Depth=167 in. -20 -25 -20 -15 Measured Depth=63 in. Depth= 95 in. Depth=167 in.

0 0 -2 -4 -6 -8 -10 -12 -14

20

40

60

80 0 -2 -4 -6 -8

20

40

60

80

Measured Depth=63 in. Depth= 95 in. Depth=167 in.

-10 -12 -14 -16

Measured Depth=63 in. Depth= 95 in. Depth=167 in.

FIGURE 8.7 Effect of the depth to bedrock on the deflection basin.

140

0 0

20

40

60

80 0 -5 -10

20

40

60

80

-5

-10 -15 -15

Non-Reflective Reflective
-20 -25

Non-Reflective Reflective

-20

0 0 -2 -4 -6 -8 -10 -12 -14

20

40

60

80

0 0 -2 -4 -6 -8 -10

20

40

60

80

Non -Reflective Reflective

-12 -14 -16

Non-Reflective Reflective

FIGURE 8.8 Effect of reflective subgrade bottom on deflection basin. 141

0 20 40 60 80 100 120 140 0.0001 Depth 63 in. Depth 95 in. Depth 167 in. 0.001 Vertical displacement, in. 0.01

0 20 40 60 80 100 120 140 0.0001 Depth 63 in. Depth 95 in. Depth 167 in. 0.001 Vertical displacement, in. 0.01

a. Site 1
0 20 40 60 80 100 120 140 0.0001 Depth 63 in. Depth 95 in. Depth 167 in. 0.001 Vertical displacement, in. 0.01 0 20 40 60 80 100 120 140 0.0001

b. Site 2

Depth 63 in. Depth 95 in. Depth 167 in. 0.001 Vertical displacement, in. 0.01

c. Site 3

d. Site 4

FIGURE 8.9 Subgrade vertical displacement for different model depths.

142

FIGURE 8.10 Effect of model depth on vertical stress distribution for site 3.

143

120 100 80 60 40 20 0 -20

0.01

0.02

0.03

0.04 Time, sec.

0.05

0.06

0.07

0.08

a. Impact load curves.


Distance, inch. 40 50

0 0 -5 -10 -15 -20 -25 -30

10

20

30

60

70

80

80 msec 40 msec

b. Deflection basins.

FIGURE 8.11

Effect of load duration on the deflection basin.

144

CHAPTER 9

CONCLUSIONS AND SUGGESTED RESEARCH

The importance of considering the dynamic nature of the FWD load in the deflection analysis has been emphasized in this work. In this study, different pavement sections were evaluated with different backcalculation programs that are currently in use. 3D Explicit Finite Element Analysis was used to evaluate the moduli profile of different pavement structures and the resulting moduli were compared with those obtained using three existing backcalculation programs: MODULUS5.0, MODCOMP3, and EVERCALC4.0. The following conclusions can be made:

1.

Comparison of the results obtained from the three backcalculation programs: MODULUS5.0, EVERCALC4.0, and MODCOMP3 reveals that MODULUS5.0 has a consistent

performance for all types of pavement structures. EVERCALC4.0 and MODCOMP3 programs produce acceptable results; however, the moduli values may be overestimated especially for the subgrade layer. The performance of the three programs decreases when evaluating composite pavements.

2.

The change of the seed moduli or the moduli range used as input for the backcalculation programs may significantly alter the resulting moduli profile. In this case, the deflection fit precision criterion used in backcalculation is not sufficient to judge the solution accuracy. Thus the experience in analysis, with materials and with deflections, is essential to check that the backcalculation process yields acceptable results.

3.

A new mechanistic method for backcalculation of pavement layer moduli and estimating the apparent depth to bedrock using 3D Explicit Finite Element approach was developed. The method was used successfully to backcalculate pavement layer moduli for all types of pavements: flexible, rigid, and composite. Backcalculation using 3D FEM offers the following advantages: 145

a. The dynamic nature of the FWD load and the effects of inertia and material damping were accounted for. b. Layer interfaces and the 3D geometry of the pavement structure were taken into account.

4.

The finite element approach enabled backcalculating the modulus of base layer in a composite pavement structure which is extremely difficult, if not impossible, using traditional algorithms.

5.

The method of evaluating the depth to bedrock is not specific for a particular type of pavement. Therefore it may be used for rigid and composite pavements for which no method is currently available.

6.

The backcalculated modulus of subgrade using the threeconventional backcalculation programs: MODULUS5.0, EVERCALC4.0, and MODCOMP3 requires multiplication by a correction factor. Correction factors of 0.35, 0.24, and 0.22 were calculated using 3D FEM for flexible, rigid, and composite pavement structures respectively.

7.

The values of the correction factors obtained using 3D FEM are in good agreement with those found by experience and recommended in the AASHTO Pavement Design Guide. This suggests that 3D FEM backcalculation may be used as a reference for assessing the accuracy of conventional backcalculation algorithms.

8.

For doweled rigid pavements, the spacing between the transverse joints has no effect on the deflection basin.

9.

For undoweled rigid pavements, the slab length is an important factor which influences the results of backcalculation programs.

10.

For aged pavement sections in which transverse cracks developed, the FWD test should not 146

be carried out on a part of the slab that is less than 10 ft to assure that this part can produce a reliable deflection basin provided that the slab width is 12 ft.

11.

Provided that a subgrade layer thickness greater than 6 ft is used in modeling flexible pavement structure using 3D FEM, any further increase in the subgrade layer thickness has insignificant effect on the 3D FE-generated deflection basin.

12.

Similarly, reflection of stress waves from the FE model bottom has little influence on the FE-generated deflection basin provided that the subgrade layer thickness is greater than 4 ft.

13.

When studying a flexible pavement structure using 3D FEM, limiting the subgrade thickness to 6 ft was found to produce satisfactory results for both the stresses and displacements on the top of subgrade layer.

14.

Due to the dynamic nature of the FWD load, the pavement surface response depends on the magnitude the FWD load pulse, shape, and duration. The assumption of static loading in backcalculation programs may produce unrealistic results.

147

SUGGESTED FUTURE RESEARCH

The work presented in this study is one step towards a better understanding of the dynamic response of pavement to FWD impact load. Future research studies should aim at:

1.

Automation of the 3D FEM backcalculation procedures.

2.

Utilization of the 3D FEM models developed for backcalculation in overlay design or in pavement design.

3.

Testing the reliability of the models developed in this thesis on a large number of pavement sites under different environmental conditions.

148

REFERENCES

1.

Meier, R.W., and G.J. Rix. Backcalculation of Flexible Pavement Moduli Using Artificial Neural Networks. Transportation Research Record 1448, TRB, National Research Council, Washington, D.C., 1994, pp. 75-82.

2.

Meier, R.W., and G.J. Rix. Backcalculation of Flexible Pavement Moduli from Dynamic Deflection Basins Using Artificial Neural Networks. Transportation Research Record 1482, TRB, National Research Council, Washington, D.C., 1995, pp. 72-81.

3.

R.W. May, and H.L. Von Quintus.

The Quest For a Standard Guide to NDT

Backcalculation. Nondestructive Testing of Pavements and Backcalculation of Moduli (Second Volume), ASTM STP 1198, Philadelphia, 1994, pp. 505-520.

4.

Huang, Y.H. Pavement Analysis and Design. Prentice Hall, Englewood Cliffs, New Jersey, 1993.

5.

AASHTO Guide for Design of Pavement Structures, Chapter 3, Guides for Field Data Collection. American Association of State Highway and Transportation Officials, 1993, pp. 32-37, 96-97.

6.

SHRP. SHRPs Layer Moduli Backcalculation Procedure: Software Selection . Contract No. SHRP-90-P-001B, Prepared by PCS/Law Engineering for SHRP, 1991.

7.

Stubstad, R.N., and B. Connor. Use of the FWD to Predict Damage Potential to Alascan Highways during Spring Thaw. Transportation Research Record 930, TRB, National Research Council, Washington, D.C., 1983, pp. 46-51.

8.

Irwin, L.N. Determination of Pavement Layer Moduli from Surface Deflection Data for 149

Pavement Performance Evaluation. Proceedings, Fourth International Conference on Structural Design of Asphalt Pavements, No. 1, Univ. Of Michigan, Aug. 1977.

9.

McCullough, B.F., and A. Taute. Use of Deflection Measurements for Determining Pavement Material Properties. Transportation Research Record 852, TRB, National Research Council, Washington, D.C., 1982, pp. 8-14.

10.

Seaman, L., J.W. Simons, D.A. Shockey, R.F. Carmichael, and B.F. McCullough. Unified Airport Pavement Design Procedure. Unified Airport Pavement Design and Analysis Concepts Workshops, Federal Aviation Administration, Washington, D.C., July 16-17, 1991, pp. 447-537.

11.

Chou, Y.J., J. Uzzan, and R.L. Lytton.

Backcalculation of Layer Moduli from

Nondestructive Pavement Deflection Data Using Expert System. Nondestructive Testing of Pavements and Backcalculation of Moduli, ASTM STP 1026, Philadelphia, 1994, pp. 341354.

12.

Irwin, L.H., W.S. Yang, and R.N. Stubstad.

Deflection Reading Accuracy and Layer

Thickness Accuracy in Backcalculation of Pavement Layer Moduli. Nondestructive Testing of Pavements and Backcalculation of Moduli, ASTM STP 1026, Philadelphia, 1994, pp.229244.

13.

SHRP-P-651. Layer Moduli Backcalculation Procedure: Software Selection. Contract No. SHRP-90-P-001B, Prepared by PCS/Law Engineering for SHRP, Washington, D.C., 1993.

14.

Deusen, D.V.

Selection of Flexible Backcalculation Software for the Minnesota Road

Research Project. Report No. MN/PR-96/29, MnDOT, St. Paul, Minnesota, 1996.

15.

Irwin, L. and T. Szebenyi. Users Guide to MODCOMP3, Version 3.2". CLRP Report 150

Number 91-4, Cornell University Local Roads Program, Ithaca, New York, March 1991.

16.

Michalak, C.H. and T. Scullion.

MODULUS5.0: Users Manual.

Report No. TX-

96/1987-1, Texas DOT, Austin, Texas, 1995.

17.

Scullion, T., J. Uzan, and M. Paredes. MODULUS: Microcomputer-Based Backcalculation System. Transportation Research Board, Record 1260, TRB, National Research Council, Washington, D.C., 1990, pp. 180-191.

18.

Lee, S.W., J.P. Mahoney, and N.C. Jackson.

Verification of a Backcalculation of

Pavement Moduli. Transportation Research Board, Record 1196, TRB, National Research Council, Washington, D.C., 1988, pp. 85-95.

19.

Harichandran, R.S., M.S. Yen, and G.Y. Baladi. MICH-PAVE: A Nonlinear Finite Element Program for Analysis of Flexible Pavements. Transportation Research Record 1286, TRB, National Research Council, Washington, D.C., 1990, pp. 123-131.

20.

Tabatabaie, A.M., and E.J. Barenberg.

Finite-Element Analysis of Jointed or Cracked

Concrete pavements. Transportation Research Record 671, TRB, National Research Council, Washington, D.C., 1978, pp. 11-19.

21.

Hammons, M.I., and A.M. Ioannides.

Developments in Rigid Pavement Response

Modeling. US Army Corps of Engineers. Technical Report Gl-96-15, August 1996.

22.

Hammons, M.I. Development of an Analysis System for Discontinuities in Rigid Airfield Pavements. US Army Corps of Engineers. Technical Report Gl-97-3, April 1997.

23.

Ioannides, A.M. E.J. Barenberg, and M.R. Thompson. Finite-Element Model With StressDependant Support. Transportation Research Record 954, TRB, National Research 151

Council, Washington, D.C., 1984, pp. 10-16.

24.

Tayabji, S.D., and B.E. Colley. Analysis of Jointed Concrete Pavements. Fedral Highway Adminsteration, McClean, VA , Report No. FHWA/RD-86/041, February 1986.

25.

Tayabji, S.D., and B.E. Colley.

Improved Rigid Pavement Joints.

Transportation

Research Record 630, TRB, National Research Council, Washington, D.C., 1983, pp. 69-78.

26.

Chatti, K., J. Lysmer, and C.L. Monismith. Dynamic Finite-Element Analysis of Jointed Concrete pavements. Transportation Research Record 1449, TRB, National Research Council, Washington, D.C., 1994, pp. 79-90.

27.

Chatti, K.

Dynamic Analysis of Jointed Concrete Pavements Subjected to Moving

Transient Loads. Ph. D. Dissertation, University of California at Berkeley, 1992.

28.

Roesler, J.R., and L. Khazanovich.

Finite Element Analysis of PCC Pavements with

Cracks. Transportation Research Board, TRB, National Research Council, Washington, D.C., 1997, Paper No. 970751.

29.

Ioannides, A.M., and J.P. Donnelly. Three-Dimensional Analysis of Slab on StressDependant Foundation. Transportation Research Record 1196, TRB, National Research Council, Washington, D.C., 1988, pp. 72-84.

30.

Kuo C.M., K.T. Hall, and M.I. Darter. Three-Dimensional Finite Element Model for Analysis of Concrete Pavement Support. Transportation Research Record 1505, TRB, National Research Council, Washington, D.C., 1995, pp. 119-127.

31.

Kuo, C.M.

Three-Dimensional Finite Element Analysis of Concrete Pavement.

Department of Civil Engineering, University of Illinois, Ph. D. Thesis, 1994. 152

32.

Zaghloul, S.M., and T.D. White. Use of a Three-Dimensional, Dynamic Finite Element Program for Analysis of Flexible Pavement. Transportation Research Record 1388, TRB, National Research Council, Washington, D.C., 1993, pp. 60-69.

33.

Zaghloul, S.M., T.D. White, and T. Kuczek.

Use of Three-Dimensional, Dynamic

Nonlinear Analysis to Develop Load Equivalency Factors for Composite Pavements. Transportation Research Record 1449, TRB, National Research Council, Washington, D.C., 1994, pp. 199-208.

34.

Seaman, L., J.W. Simons, D.A. Shockey. Unified Airport Pavement Design Procedure. Proceedings: Unified Airport Pavement Design and Analysis Concepts Workshops. DOT/FAA/Rd-92/17, Volpe National Transportation Systems Center, July 1992, pp. 446537.

35.

Kennedy, T.C., R.D. Everhart, T.P. Forte, and J.A. Hadden. Development of a Governing Primary Response Model (GPRM) For Airport Pavement Design and Analysis. Final Report to Volpe National Transportation Systems Center, Contract No. VA2043 DTRS-57-89-D00006, January, 1994.

36.

Kennedy, T.C., and R.D. Everhart. Dynamic Complaint Boundary. Final Report to Turner Fairbanks Highway Research Center, Contract No. DIFH 61-93-C-00055, May, 1995.

37.

Kennedy, T.C., and R.D. Everhart. Comparison of Predicted Pavement Structural Response With Field Measurments Data. Final Report to Turner Fairbanks Highway Research Center, Mclean, VA 22101, Contract No. DTFH 61-93-C-00055, February 1996.

38.

Kennedy, T.C., and R.D. Everhart. Thermal Effects on Pavement Response. Final Report to Turner Fairbanks Highway Research Center, Mclean, VA 22101, Contract No. DTFH 61153

93-C-00055, March 1997.

39.

Zaghloul, S.M., T.D. White, and T. Kuczek. Evaluation of Heavy Load Damage Effect on Concrete Pavements Using Three-Dimensional, Dynamic Nonlinear Analysis.

Transportation Research Record 1449, TRB, National Research Council, Washington, D.C., 1994, pp. 123-133.

40.

Zaghloul, S.M., T.D. White, J.A. Ramirez, and NBR Prasad. Computerized Overload Permitting Procedure for Indiana. Transportation Research Record 1448, TRB, National Research Council, Washington, D.C., 1994, pp. 40-52.

41.

Zaghloul, S.M, T.D. White, V.P. Drenvich, and B. Coree. Dynamic Analysis of FWD Loading and Pavement Response Using a Three-Dimensional Finite Element Program. Nondestructive Tesing of Pavements and Backcalculation of Moduli (Second Volume), ASTM STP 1198, American Society of Testing and Materials, Philladelphia, 1994, pp. 105138.

42.

Huang, Y. H., and S. T. Wang. Finite-Element Analysis of Rigid Pavements with Partial Subgrade Contact. Highway Research Record No.485 , National Research Council,

Washington, D.C., 1974, pp. 39-54.

43.

Huang, Y. H., and S. T. Wang. Finite-Element Analysis of Concrete Slabs and Its Implications for Rigid Pavement Design. Highway Research Record No.466 , National Research Council, Washington, D.C., 1974, pp. 39-54.

44.

Davis, T.G., and M.S. Mamlouk. Theoretical Response of Multilayer Pavement Systems to Dynamic Nondestructive Testing. Transportation Research Record No. 1022, TRB,

National Research Council, Washington, D.C., 1985, pp. 1-7.

154

45.

Mamlouk, M.S., and T.G. Davis. Elasto-Dynamic Analysis of Pavement Deflections. American Society of Civil Engineers, Journal of Transportation Engineerring,Vol. 110, No. 6, 1994, pp. 536-550.

46.

Sebaaly, B.E. , M. S. Mamlouk, and T. G. Davis. Dynamic Analysis of Falling Weight Deflectometer Data. In Transportation Research Record 1070, TRB, National Research Council, Washington, D.C., 1986, pp. 63-68.

47.

Sebaaly, B.E. Dynamic Models for Pavement Analysis. Department of Civil Engineering, Arizona State University, Ph. D. Dissertation, 1987.

48.

Mallela, J., and K. P. George. Three-Dimensional Dynamic Response Model for Rigid Pavements. In Transportation Research Record 1448, TRB, National Research Council, Washington, D.C., 1994, pp. 92-99.

49.

Nazarian, S., and K. M. Boddapti. Pavement-Falling Weight Deflectometer Interaction Using Dynamic Finite-Element Analysis. In Transportation Research Record 1482, TRB, National Research Council, Washington, D.C., 1995, pp. 33-43.

50.

Uddin, W., P. Noppakunwijai, and T. Chung. Performance Evaluation of Jointed Concrete Pavement Using Three-Dimensional Finite-Element Dynamic Analysis. Transportation Research Board, 76th Annual Meeting, Washington, D.C., 1997 , Paper No. 1414.

51.

Uddin, W., D. Zhang, and F. Fernandez.

Finite Element Simulation of Pavement

Discontinuities and Dynamic Load Response. Transportation Research Record 1448, TRB, National Research Council, Washington, D.C., 1994, pp. 69-74.

52.

Shoukry, S.N., D.R. Martinelli, and O. Selezneva. Dynamic Response of Composite Pavements Under Impact. Transportation Research Record 1570, TRB, National Research 155

council, Washington, D.C., 1997, pp. 163-171.

53.

Shoukry, S.N., D.R. Martinelli, and O.I. Selezneva.

Dynamic Considerations in Pavement Proceedings of the

Layers Moduli Evaluation Using Falling Weight Deflectometer.

International Society for Optical Engineering, Nondestructive Evaluation of Bridges and Highways, Vol. 2946, 4-5 December 1996, pp. 109-120.

54.

Uddin, W., A. H. Meyer, and W. R. Hudson. Rigid Bottom Considerations for Nondestructive Evaluation of Pavements. Transportation Research Record No. 1070, TRB, National Research Council, Washington, D.C., pp. 21-29, 1986.

55.

Yang, N. C. Mechanistic Analysis of Nondestructive Pavement Deflection Data. Cornel University, Ithaca, New York, Ph. D. Dissertation, 1988.

56.

Briggs, R. C., and S. Nazzarian. Effects of Unknown Rigid Subgrade Layers on Backcalculation of Pavement Moduli and Projections of Pavement Performance. Transportation Research Record No. 1227, TRB, National Research Council, Washington, D.C., pp. 183-193, 1989.

57.

Rohde, G.T., and R. E. Smith. Determining Depth to Apparent Stiff Layer From FWD Data. Research Report 1159-1, Texas Transportation Institute, The Texas A&M University System, College Station, Texas, 1991.

58.

Uzan, J. Advanced Backcalculation Techniques. Nondestructive Testing of Pavements and Backcalculation of Moduli (Second Volume), ASTM STP 1198, American Society for Testing and Materials, Philadelphia, pp. 3-37, 1994.

59.

Chang, D. W., Y. V. Kang, J. M. Rosset, and K. H. Stokoe II. Effect of Depth to Bedrock on Deflection Basins Obtained with Dynaflect and FWD Tests. Transportation Research 156

Record No. 1355, TRB, National Research Council, Washington, D.C., pp. 8-17, 1991.

60.

Bush, A. J. Nondestructive Testing for Light Aircraft Pavements; Phase II, Development of the Nondestructive Evaluation Methodology. Report FFA-RD-80-9-II, US Department of Transportation, Washington, D.C., 1980.

61.

Seng, C., K. H. Stokoe II, and J. M. Roesset. Effect of Depth to Bedrock on the Accuracy of Backcalculated Moduli obtained with Dynaflect and FWD Tests. Research Report 11755, Center for Transportation Research, The University of Texas at Austin, Texas, 1993.

62.

Ullidtz, P., and R. N. Stubstad. Analytical-Empirical Pavement Evaluation Using the Falling Weight Deflectometer. Transportation Research Record No. 1022, TRB, National Research Council, Washington, D.C., pp. 36-43, 1985.

63.

CROW. Deflection Profile- Not a Pitfall Anymore Record 17, CROW Information and Technology Center for Transport and Infrastructures, The Netherlands, 1998.

64.

Shoukry, S.N., and G. W. William. Performance Evaluation of Backcalculation Algorithms Through 3D Finite Element Modeling of Pavement Structures. Paper No. 991277,

presented at the 78th Annual Transportation Research Board Meeting, Washington, D.C., January, 1999.

65.

Private Communications. Email message from E. Beauving -CROW-NL to Dr. Smair Shoukry, sent on March 31, 1999.

66.

Naval Facilities Engineering Command. Soil Mechanics: Design Manual 7.01, September 1986.

66.

Ullidtz, P. Pavement Analysis, Development in Civil Engineering 19". Elsevier, 1987. 157

67.

Ullidtz, P. Modeling Flexible Pavement Response and Performance Polytecknisk Forlag, Denmark, 1998.

68.

Private Communications. Email message from Professor Per Ullidtz to the author, sent on March 24, 1999.

158

APPENDIX A PARAMETRIC INPUT FILE TO GENERATE FLEXIBLE PAVEMENT MODEL

title FLEXIBLE PAVEMENT MODEL SUBJECTED TO FWD LOAD . c Global Control Commands. dyna3d lsdyopts hgen 2; endtim 0.040; glstat 0.001 matsum 0.001; d3plot dtcycl 0.001; qh 0.11; parameters c Enter the surface Layer thickness (t1) with a negative sign, in. t1 -7.25 c Enter the thickness of the Base Layer (t2) with a negative sign, in. t2 -15; c Sliding Interface with friction Between the Surface Layer & Base. sid 1 fric 0.9 lsdsi 3; C Sliding Interface with Friction Between the Loading Plate & the Surface Layer.. sid 2 fric 0.8 lsdsi 3; plane 1 0.0 0.0 0.0 -1 0 0 0.001 symm; c Enter the Material Properties for Pavement Layers. dynamats 1 1 rho 2.240e-4 e 12e+4 pr 0.35; dynamats 2 1 rho 2.200e-4 e 120e+3 pr 0.30; dynamats 3 1 rho 1.950e-4 e 12.0e+3 pr 0.40; dynamats 4 1 rho 7.324e-4 e 29.0e+6 pr 0.30; c Enter the measured Load Curve. lcd 1 0 0.0 2e-4 0.0 0.4e-3 0 0.6e-3 0 159

0.8e-3 0 1.0e-3 0 1.2e-3 0 1.4e-3 0 1.6e-3 0 1.8e-3 0 2.0e-3 1 2.2e-3 2 2.4e-3 4 2.6e-3 6 2.8e-3 10 3e-3 14 3.2e-3 19 3.4e-3 23 3.6e-3 30 3.8e-3 37 4e-3 47 4.2e-3 58 4.4e-3 72 4.6e-3 88 4.8e-3 107 5e-3 129 5.2e-3 152 5.4e-3 177 5.6e-3 203 5.8e-3 230 6e-3 257 6.2e-3 282 6.4e-3 306 6.6e-3 331 6.8e-3 354 7e-3 374 7.2e-3 392 7.4e-3 408 7.6e-3 422 7.8e-3 436 8e-3 449 8.2e-3 461 8.4e-3 473 8.6e-3 483 8.8e-3 493 9e-3 503 9.2e-3 514 160

9.4e-3 525 9.6e-3 536 9.8e-3 546 10e-3 558 10.2e-3 568 10.4e-3 579 10.6e-3 590 10.8e-3 600 11e-3 610 11.2e-3 620 11.4e-3 629 11.6e-3 638 11.8e-3 647 12e-3 656 12.2e-3 664 12.4e-3 673 12.6e-3 682 12.8e-3 691 13e-3 699 13.2e-3 708 13.4e-3 718 13.6e-3 726 13.8e-3 734 14e-3 742 14.2e-3 748 14.4e-3 753 14.6e-3 757 14.8e-3 759 15e-3 761 15.2e-3 760 15.4e-3 759 15.6e-3 755 15.8e-3 753 16e-3 746 16.2e-3 737 16.4e-3 729 16.6e-3 719 16.8e-3 708 17e-3 695 17.2e-3 682 17.4e-3 669 17.6e-3 657 17.8e-3 643 161

18e-3 630 18.2e-3 615 18.4e-3 603 18.6e-3 589 18.8e-3 575 19e-3 561 19.2e-3 548 19.4e-3 534 19.6e-3 520 19.8e-3 505 20e-3 491 20.2e-3 478 20.4e-3 465 20.6e-3 451 20.8e-3 439 21e-3 426 21.2e-3 413 21.4e-3 402 21.6e-3 390 21.8e-3 378 22e-3 368 22.2e-3 356 22.4e-3 344 22.6e-3 333 22.8e-3 322 23e-3 310 23.2e-3 298 23.4e-3 285 23.6e-3 273 23.8e-3 261 24e-3 249 24.2e-3 237 24.4e-3 226 24.6e-3 213 24.8e-3 201 25e-3 189 25.2e-3 179 25.4e-3 167 25.6e-3 157 25.8e-3 146 26e-3 137 26.2e-3 128 26.4e-3 119 162

26.6e-3 110 26.8e-3 102 27e-3 95 27.2e-3 88 27.4e-3 82 27.6e-3 75 27.8e-3 69 28e-3 63 28.2e-3 58 28.4e-3 52 28.6e-3 47 28.8e-3 43 29e-3 38 29.2e-3 32 29.4e-3 28 29.6e-3 25 29.8e-3 21 30e-3 17 30.2e-3 14 30.4e-3 10 30.6e-3 7 30.8e-3 5 31e-3 2 31.2e-3 0 31.4e-3 0 31.6e-3 0 31.8e-3 0 32e-3 0;

c Define Asphalt-Concrete Layer. block 1 3 31; 1 16 19 22 37; 1 4; 0.0 8.0 120.0; -72.0 -6 0 6 72.0; 0.0 [%t1]; sii 1 2 ;2 4;1 1;2 s; sii 1 3;1 5;2 2;1 s; 163

nr 1 1 1 3 1 2 nr 1 1 1 1 5 2 nr 3 1 1 3 5 2 b 1 1 1 3 1 2 dy 1 rx 1; b 1 1 1 1 5 2 dx 1 ry 1; b 1 5 1 3 5 2 dy 1; b 3 1 1 3 5 2 dx 1 dy 1; mate 1 endpart c Define Base Layer and subgrade . block 1 41; 1 25; 1 5 15; 0.0 120; -72 72; [%t1] [%t1+%t2] [%t1+%t2-40]; mt 1 1 1 2 2 2 2 mt 1 1 2 2 2 3 3 sii 1 2;1 2;1 1;1 m; nr 1 1 1 1 2 3 nr 2 1 1 2 2 3 nr 1 2 1 2 2 3 nr 1 1 1 2 1 3 c nr 1 1 3 2 2 3 c Boundary Conditions b 1 1 1 1 2 3 dx 1 dy 1 ry 1; b 2 1 1 2 2 3 dx 1 dy 1; b 1 2 1 2 2 3 dx 1 dy 1; b 1 1 1 2 1 3 dx 1 dy 1; b 1 1 3 2 2 3 dx 1 dy 1 dz 1 rx 1 ry 1 rz 1; endpart c Define Loading Plate (R=6 in). block 164

1 6 11 16 21;1 6 11 16 21;1 3; -3. -3. 0. 3. 3.; -3. -3. 0. 3. 3.; 0. 1.; de 1 1 1 2 2 2 de 4 1 1 5 2 2 de 1 4 1 2 5 2 de 4 4 1 5 5 2 de 1 0 0 3 0 0 sf 2 1 1 4 1 2 cy 0. 0. 0. 0. 0. 1. 6 sf 1 2 1 1 4 2 cy 0. 0. 0. 0. 0. 1. 6 sf 2 5 1 4 5 2 cy 0. 0. 0. 0. 0. 1. 6 sf 5 2 1 5 4 2 cy 0. 0. 0. 0. 0. 1. 6 sii ; ; 1 1; 2 m; c Ddfine applied pressure region. orpt - 0.0 0.0 -1000000 pr 1 1 2 5 5 2 1 0.145624; mate 4 endpart merge stp 0.01

165

VITA Gergis W. William was born in Alexandria, Egypt, on August 1, 1973. He received a Bachelor of Science in Civil Engineering with honors from Alexandria University, Egypt. After his graduation in June 1995, he joined the Consultative Bureau for Civil Constructions, Alexandria, as a structural design engineer. During his employment, he took charge of several design projects which included design of different types of hydraulic structures (Barrages, Syphons, Aqueducts, and pipelines), bridges (over 20 bridges), roads (over 20 rural roads), three high-rise buildings, and Three steel structures. In January 1998, he joined West Virginia University as a graduate research assistant in the Department of Civil and Environmental Engineering, where he pursued a Master Degree in Civil Engineering.

Publications: 1. Shoukry, S. N., G. W. William. Performance Evaluation of Backcalculation Algorithms Through 3D Finite Element Modeling of Pavement Structures. Paper No. 991277, Accepted for presentation and publication at the 78th Annual Transportation Research Board Meeting, Washington, D.C., January, 1999. 2. Shoukry, S.N., and G. W. William. Dynamic Backcalculation of Pavement Layer Moduli. Proceedings: ASNT Spring Conference and 8th Annual Research Symposium, American Society of Nondestructive Testings, Florida, March 1999. 3. Shoukry, S.N., Gergis W. William, and D. R. Martinelli. Assessment of the Performance of Rigid Pavement Backcalculation Through Finite Element Modeling. Proceedings of the SPIE conference on Nondestructive Evaluation of Bridges and Highways III, California, March, 1999, pp. 146-156. 4. Shoukry, S. N., and G. William. 3D FEM Analysis of Load Transfer Efficiency.

Proceedings of the First National Symposium on 3D Finite Element Modeling for Pavement Analysis & Design. Charleston, West Virginia, November, 1998, pp. 40-50. 5. Shoukry, S.N., D.R. Martinelli, G.W. William, and M. R. Fahmy. Applications of LS-DYNA in Pavement Analysis and Design. Proceedings: Fifth International LS-DYNA Users Conference, Detroit, Michigan, 1998. 166

You might also like