final 2002-30 research - minnesota department of...

168
Evaluation of Water Flow Through Pavement Systems 2002-30 Final Report Research

Upload: others

Post on 19-Mar-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Evaluation of Water Flow Through Pavement

Systems

2002-30 Final Report

Res

earc

h

Page 2: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Technical Report Documentation Page 1. Report No. 2. 3. Recipients Accession No. MN/RC - 2002-30 4. Title and Subtitle 5. Report Date

June 2002 6.

EVALUATION OF WATER FLOW THROUGH PAVEMENT SYSTEMS 7. Author(s) 8. Performing Organization Report No. Paola Ariza and Bjorn Birgisson 9. Performing Organization Name and Address 10. Project/Task/Work Unit No.

11. Contract (C) or Grant (G) No.

University of Florida Civil and Coastal Engineering Department P.O. Box 116580 Gainesville, FL 32611-6580 (C) 79080

12. Sponsoring Organization Name and Address 13. Type of Report and Period Covered

Final Report 14. Sponsoring Agency Code

Minnesota Department of Transportation 395 John Ireland Boulevard Mail Stop 330 St. Paul, Minnesota 55155 15. Supplementary Notes http:// www.lrrb.org/PDF/200230.pdf 16. Abstract (Limit: 200 words) Most research on the effects of moisture in pavement has been based on conditions of total saturation with loss of pavement strength calculated using saturated flow assumptions. Yet roadbeds reach full saturation only when positive total heads are present (e.g., surface ponding, etc.) and distributed in such a manner that saturation of the pavement system is reached. Most dense graded pavement base layers tend to be unsaturated most of the time. The flow of water through unsaturated soil (unsaturated flow) is primarily a function of matric suction and gravity. Matric suction varies with saturation and changes over the course of a rainfall event. The authors propose a first step toward a comprehensive approach to drainage and pavement design that integrates the true effects of moisture on pavement moduli and mechanistic-empirical pavement design. The authors used SEEP/W and DRIP software to analyze data collected at the Minnesota Road Research project (Mn/ROAD Cell 33, Cell 34, and Cell 35). The SEEP/W software modeled unsaturated flow under transient conditions through layered systems under complex boundary conditions and material characterizations. Results establish that SEEP/W is a valuable tool for modeling unsaturated flow that can predict results comparable to field studies. The time to drain calculated based on unsaturated flow theory will generally be longer than that evaluated under saturated flow assumptions. The study also looked at the edge drains and under drains and found that under drains alone did not significantly improve drainage, but collector pipes or edge drains used in combination with under drains are very effective in reducing the amount of moisture in the soil.

17. Document Analysis/Descriptors 18. Availability Statement pavements drainage finite elements soil water characteristics curve

hydraulic conductivity curveunsaturated flow edge drains under drains

No restrictions. Document available from: National Technical Information Services, Springfield, Virginia 22161

19. Security Class (this report) 20. Security Class (this page) 21. No. of Pages 22. Price Unclassified Unclassified 166

Page 3: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

EVALUATION OF WATER FLOW THROUGH PAVEMENT SYSTEMS

Final Report

Prepared by

Paola Ariza, Bjorn Birgisson, Ph.D., P.E.

UNIVERSITY OF FLORIDA Civil and Coastal Engineering Department

P.O. Box 116580 Gainesville, FL 32611-6580

June 2002

Published by:

Minnesota Department of Transportation Office of Materials & Road Research

1400 Gervais Avenue MS 645 Maplewood, MN 55109

This report represents the results of research conducted by the authors and does not necessarily represent the views or policy of the Minnesota Department of Transportation. This report does not contain a standard or specified technique. The authors and the Minnesota Department of Transportation and/or Center for Transportation Studies do not endorse products or manufacturers. Trade or manufacturers’ names appear herein solely because they are consid-ered essential to this report.

Page 4: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

TABLE OF CONTENTS

Page CHAPTER 1 INTRODUCTION.......................................................................................... 1 1.1 Objectives ............................................................................................... 2 1.2 Scope of Report....................................................................................... 3 CHAPTER 2 LITERATURE REVIEW............................................................................... 4 2.1 Current State-of-the-Practice .................................................................. 4 2.1.1 Effects of drainage on pavement design life............................... 5 2.1.2 Summary of current practice....................................................... 7 2.2 Positive Drainage Systems...................................................................... 8 2.2.1 Open graded base course ............................................................ 8 2.2.2 Edge drains ................................................................................. 8 2.2.3 Use of geotextiles for drainage improvement............................. 9 2.3 Unsaturated Flow Through Pavements................................................... 9 2.3.1 Soil water characteristic curves ................................................ 11 2.3.2 Hydraulic conductivity models ................................................. 20 2.3.3 Effects of layering on pavement drainage ................................ 22 CHAPTER 3 RESEARCH METHODOLOGY ................................................................. 25 3.1 Selection of Analysis Tools .................................................................. 25 3.1.1 SEEP/W .................................................................................... 26 3.1.2 DRIP ......................................................................................... 26 3.2 Materials and Pavement Sections ......................................................... 28 3.2.1 Pavement sections..................................................................... 28 3.2.2 Materials ................................................................................... 29 CHAPTER 4 EVALUATION OF SEEP/W FOR MODELING UNSATURATED FLOW. . . . 35 4.1 Example 1 ............................................................................................. 33 4.2 Example 2 ............................................................................................. 39 4.3 Example 3 ............................................................................................. 44 4.4 Summary ............................................................................................... 52 CHAPTER 5 VERIFICATION OF RESULTS TO FIELD RESULTS FOR CELLS 33, 34, 35......................................................................................... 54 5.1 Pavement System Description .............................................................. 54 5.2 Finite Element Model ........................................................................... 54 5.2.1 Hot mix asphalt layer ................................................................ 54 5.2.2 Base course ............................................................................... 56 5.2.3 Subgrade soil............................................................................. 56 5.2.4 Initial and boundary conditions ................................................ 57 5.2.5 Finite element model analysis................................................... 58 5.2.6 Measured volumetric water contents ........................................ 59

Page 5: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

5.3 Materials Characterization Adjustment ................................................ 61 5.3.1 Initial calibration results ........................................................... 64 5.3.2 Second calibration results ......................................................... 64 5.3.3 Final matching of predicted and measured volumetric water contents ........................................................................... 76 5.4 Cell 34 ................................................................................................... 79 5.4.1 Calibration of results at locations 101 and 102......................... 82 5.4.2 Calibration to location 103 ....................................................... 82 5.5 Cell 35 ................................................................................................... 85 5.6 Summary ............................................................................................... 90 CHAPTER 6 PARAMETRIC STUDY.............................................................................. 91 6.1 Initial Slope of the Base Material Soil Water Characteristic Curve..... 91 6.2 Air Entry Value of Base Material ......................................................... 94 6.3 Effects of Ksat of Mn/DOT Class 6 Special Crushed Granite Base Material ........................................................................................ 96 6.4 Air Entry Value of Subgrade Material.................................................. 98 6.5 Ksat at Subgrade Material ..................................................................... 98 6.6 Effects of the Type of Base Material .................................................. 100 6.7 Infiltration Effects .............................................................................. 105 6.8 Water Table Influence......................................................................... 106 6.9 Summary ............................................................................................. 108 CHAPTER 7 EFFECT OF EDGE and UNDER DRAINS ON WATER FLOW THROUGH FLEXIBLE PAVEMENTS.................................................... 110 7.1 Description of Case 1: Under Drain ................................................... 111 7.2 Description of Case 2: Under Drains With Collector Pipes ............... 114 7.3 Description of Case 3: Edge Drains.................................................... 116 7.4 Description of Case 4: Combination of Edge and Under Drains........ 117 7.5 Drainage Systems Comparison ........................................................... 117 CHAPTER 8 SUMMARY, CONCLUSIONS, and RECOMMENDATIONS ................ 119 LIST OF REFERENCES…………………………………………………………………….122 APPENDIX A EVALUATION OF TIME TO DRAIN CALCULATIONS FOR PAVEMENTS…………………………………………………………….A-1 APPENDIX B DETERMINATION OF AIR ENTRY VALUE FOR CLASS 6 SPECIAL………………………………………………………………….B-1

Page 6: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

List of Tables Page

Table 2.1 Classification of 50% of Drainage ................................................................. 6 Table 2.2 Recommended m Values for Modifying Structural Layer Coefficients of Untreated Base and Subbase Materials in Flexible Pavements ................. 7 Table 2.3 Soil Water Characteristic Models – Main Features, Advantages, and Disadvantages............................................................................................... 18 Table 2.4 Hydraulic Conductivity Models – Main Features, Advantages, and Disadvantages............................................................................................... 23 Table 3.1 Asphalt Layer Description............................................................................ 29 Table 4.1 Soil Hydraulic Parameters............................................................................ 34 Table 4.2 Maximum Difference Between Predicted and Measured Water Table Results in Figures 4.5 – 4.8 .......................................................................... 39 Table 5.1 Cell 33 – Calibration for Base Layer............................................................ 79 Table 5.2 Cell 33 – Calibration for Subgrade Soil ....................................................... 79 Table 5.3 Summary of SEEP/W Runs Performed to Adjust Predicted Volumetric Moisture Contents to TDR Measured Values at Cell 33.............................. 80 Table 5.4 Cell 34 – Calibration for Base Layer............................................................ 86 Table 5.5 Summary of Runs to Calibrate Cell 34 ........................................................ 86 Table 5.6 Cell 35 – Calibration for Base Layer............................................................ 89 Table 5.7 Summary of Runs to Calibrate Cell 35 ........................................................ 89 Table 6.1 Effects of the Initial Slope of the Base Material Soil Water Characteristic Curve on Maximum Predicted Volumetric Moisture Content ..................... 94 Table 6.2 Effects of Air Entry Value of Mn/DOT Class 6 Special Crushed Granite Base Material on Maximum Predicted Volumetric Moisture Content ........ 96 Table 6.3 Effects of Ksat of Mn/DOT Class 6 Special Crushed Granite Base Material on Maximum Predicted Volumetric Moisture Content ............................... 98

Page 7: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Table 6.4 Effects of Air Entry Value of R-12 Silty Clay Subgrade Soil on Maximum Predicted Volumetric Moisture Content ...................................................... 99 Table 6.5 Effects of Ksat of R-12 Silty Clay Subgrade Soil on Maximum Predicted Volumetric Moisture Content..................................................................... 101 Table 6.6 Maximum Difference Between Saturated Volumetric Water Content and Predicted Volumetric Water Content for Class 3 Special, Class 4 Special, Class 5 Special, and Class 6 Special Mn/DOT Granular Base Material.... 105 Table 6.7 Effects of Infiltration on the Maximum Predicted Volumetric Moisture Content at TDR Location 101 .................................................................... 106 Table 6.8 Water Table Positions – Effects Summary................................................. 108 Table 6.9 Summary of Evaluation of Effects of Unsaturated Soil Hydraulic Proper- ties and SEEP/W Input Parameters on the Sensitivity of Predicted Volu- metric Moisture Content at TDR Location 101.......................................... 109

Page 8: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

List of Figures Page

Figure 2.1 Typical Soil Water Characteristic Curve ..................................................... 12 Figure 2.2 Typical Values for the Brooks and Corey Model for Water Retention........ 14 Figure 2.3 Typical Values for the Van Genuchten Model for Water Retention............ 15 Figure 2.4 Hysteresis Effects – Main Drainage and Wetting Curves............................ 17 Figure 2.5 Variation of the Air-Water Interface Due to Raindrop Effect...................... 17 Figure 3.1 Thickness of Base and Hot Mix Asphalt Layers for Mn/ROAD Cells 33-35. . ............................................................................................... 28 Figure 3.2 Pavement Geometry and Dimensions .......................................................... 29 Figure 3.3 Grain Size Distribution for Mn/DOT Class 6 Special.................................. 30 Figure 3.4 Soil Water Characteristic Curve – Base Material ........................................ 30 Figure 3.5 Hydraulic Conductivity Curve – Base Material ........................................... 31 Figure 3.6 Soil Water Characteristic Curve – Subgrade Material ................................. 32 Figure 3.7 Hydraulic Conductivity Curve – Subgrade Material.................................... 32 Figure 4.1 Example 1 – Geometry and Boundary Conditions....................................... 34 Figure 4.2 Soil Water Characteristic Curve for Example 1........................................... 35 Figure 4.3 Hydraulic Conductivity Curve for Example 1 ............................................. 35 Figure 4.4 Example 1 – Finite Element Model.............................................................. 36 Figure 4.5 Comparison Between Measured and Predicted Water Table at 2 Hours ..... 37 Figure 4.6 Comparison Between Measured and Predicted Water Table at 3 Hours ..... 37 Figure 4.7 Comparison Between Measured and Predicted Water Table at 4 Hours ..... 38 Figure 4.8 Comparison Between Measured and Predicted Water Table at 8 Hours ..... 38 Figure 4.9 Example 2 – Geometry and Boundary Conditions....................................... 40

Page 9: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Figure 4.10 Soil Water Characteristic Curve for Example 2........................................... 40 Figure 4.11 Hydraulic Conductivity Curve for Example 2 ............................................. 41 Figure 4.12 Example 2 – Finite Element Model.............................................................. 42 Figure 4.13 Comparison Between Measured and Predicted Volumetric Water Content at 1.5 Hours for Example 2 .......................................................................... 42 Figure 4.14 Comparison Between Measured and Predicted Volumetric Water Content at 3 Hours for Example 2 ............................................................................. 43 Figure 4.15 Comparison Between Measured and Predicted Volumetric Water Content at 4.5 Hours for Example 2 .......................................................................... 43 Figure 4.16 Comparison Between Measured and Predicted Volumetric Water Content at 6 Hours for Example 2 ............................................................................. 44 Figure 4.17 Example 3 – Geometry and Boundary Conditions....................................... 45 Figure 4.18 Soil Water Characteristic Curve for Base & Subbase for Example 3 .......... 46 Figure 4.19 Hydraulic Conductivity Curve for Base Material in Example 3 .................. 46 Figure 4.20 Hydraulic Conductivity Curve for Subbase Material in Example 3 ............ 47 Figure 4.21 Example 3 – Finite Element Model.............................................................. 47 Figure 4.22 Comparison Between Measured and Predicted Volumetric Water Content in Base Material at 1.5 Hours....................................................................... 48 Figure 4.23 Comparison Between Measured and Predicted Volumetric Water Content in Base Material at 3 Hours.......................................................................... 48 Figure 4.24 Comparison Between Measured and Predicted Volumetric Water Content in Base Material at 4.5 Hours....................................................................... 49 Figure 4.25 Comparison Between Measured and Predicted Volumetric Water Content in Base Material at 6 Hours.......................................................................... 49 Figure 4.26 Comparison Between Measured and Predicted Volumetric Water Content in Subbase Material at 1.5 Hours ................................................................. 50

Page 10: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Figure 4.27 Comparison Between Measured and Predicted Volumetric Water Content in Subbase Material at 3 Hours .................................................................... 50 Figure 4.28 Comparison Between Measured and Predicted Volumetric Water Content in Subbase Material at 4.5 Hours ................................................................. 51 Figure 4.29 Comparison Between Measured and Predicted Volumetric Water Content in Subbase Material at 6 Hours .................................................................... 51 Figure 4.30 Comparison of Measured Volumetric Water Content in Base Material at 6 Hours for Examples 2 and 3 .................................................................. 52 Figure 4.31 Comparison of Predicted Volumetric Water Content in Base Material at 6 Hours for Examples 2 and 3 .................................................................. 53 Figure 5.1 Layer Thicknesses for Mn/ROAD Cells 33, 34, and 35 .............................. 55 Figure 5.2 Pavement Geometry and Dimensions for Cells 33, 34, and 35.................... 55 Figure 5.3 Section of the Finite Element Model Used to Represent Cells 33, 34, and 35. . ................................................................................................. 56 Figure 5.4 Finite Element Model Used to Represent Cells 33, 34, and 35 ................... 57 Figure 5.5 Precipitation Events for Cells 33, 34, and 35............................................... 59 Figure 5.6 TDR Installation at Mn/DOT Cells 33, 34, and 35 ...................................... 60 Figure 5.7 TDR Location Within Finite Element Model for Cells 33, 34, and 35........ 60 Figure 5.8 Measured Data for Different TDR Locations at Cell 33 .............................. 61 Figure 5.9 Measured Data for Different TDR Locations at Cell 34 .............................. 62 Figure 5.10 Measured Data for Different TDR Locations at Cell 35 .............................. 62 Figure 5.11 Volumetric Water Content at Cell 33 – Location 101.................................. 64 Figure 5.12 Volumetric Water Content at Cell 33 – Location 102.................................. 65 Figure 5.13 Volumetric Water Content at Cell 33 – Location 103.................................. 65 Figure 5.14 Soil Water Characteristic Curve (Air Entry – 3 kPa) – Base Material ........ 66 Figure 5.15 Estimated Hydraulic Conductivity Curve – Base Material .......................... 67

Page 11: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Figure 5.16 Final Soil Water Characteristic Curve – Base Material ............................... 67 Figure 5.17 Estimated Hydraulic Conductivity Curve – Base Material .......................... 68 Figure 5.18 Volumetric Water Content at Cell 33 – Location 101.................................. 69 Figure 5.19 Volumetric Water Content at Cell 33 – Location 102.................................. 69 Figure 5.20 Volumetric Water Content at Cell 33 – Location 103.................................. 70 Figure 5.21 Modified Hydraulic Conductivity Curve: 10 Times Ksat - Subgrade Material ........................................................................................ 70 Figure 5.22 Volumetric Water Content at Cell 33 – Location 101.................................. 71 Figure 5.23 Volumetric Water Content at Cell 33 – Location 102.................................. 71 Figure 5.24 Volumetric Water Content at Cell 33 – Location 103.................................. 72 Figure 5.25 Final Hydraulic Conductivity Curve: Ksat = 2.7535E-06 m/s – Subgrade Material. . . .................................................................................................. 72 Figure 5.26 Volumetric Water Content at Cell 33 – Location 101.................................. 73 Figure 5.27 Volumetric Water Content at Cell 33 – Location 102.................................. 74 Figure 5.28 Volumetric Water Content at Cell 33 – Location 103.................................. 74 Figure 5.29 Volumetric Water Content at Cell 33 – Location 101.................................. 75 Figure 5.30 Volumetric Water Content at Cell 33 – Location 102.................................. 75 Figure 5.31 Volumetric Water Content at Cell 33 – Location 103.................................. 76 Figure 5.32 Final Soil Water Characteristic Curve for Location 102 – Base Material ... 77 Figure 5.33 Volumetric Water Content at Cell 33 – Location 102.................................. 77 Figure 5.34 Final Soil Water Characteristic Curve for Location 103 – Base Material ... 78 Figure 5.35 Volumetric Water Content at Cell 33 – Location 103.................................. 78 Figure 5.36 Volumetric Water Content at Cell 34 – Location 101.................................. 81

Page 12: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Figure 5.37 Volumetric Water Content at Cell 34 – Location 102.................................. 81 Figure 5.38 Volumetric Water Content at Cell 34 – Location 103.................................. 82 Figure 5.39 Final Soil Water Characteristic Curve for Locations 101 and 102 at Cell 34 – Base Material ............................................................................ 83 Figure 5.40 Volumetric Water Content at Cell 34 – Location 101.................................. 83 Figure 5.41 Volumetric Water Content at Cell 34 – Location 102.................................. 84 Figure 5.42 Final Soil Water Characteristic Curve for Location 203 Cell 34 – Base Material. . . .................................................................................................. 84 Figure 5.43 Volumetric Water Content at Cell 34 – Location 103.................................. 85 Figure 5.44 Volumetric Water Content at Cell 35 – Location 101.................................. 87 Figure 5.45 Volumetric Water Content at Cell 35 – Location 103.................................. 87 Figure 5.46 Final Soil Water Characteristic Curve for Location 102 Cell 35 – Base Material. . . . ................................................................................................. 88 Figure 5.47 Volumetric Water Content at Cell 35 – Location 102.................................. 88 Figure 6.1 Soil Water Characteristic Curves for Initial Slope Cases ............................ 92 Figure 6.2 Hydraulic Conductivity Curves for Initial Slope Cases ............................... 93 Figure 6.3 Results for Initial Slope Cases ..................................................................... 93 Figure 6.4 Base Soil Water Characteristic Curves for Air Entry Value Cases.............. 95 Figure 6.5 Results for Air Entry Value Cases at Base Layer ........................................ 95 Figure 6.6 Hydraulic Conductivity Curves for Ksat Cases at Base Material ................. 97 Figure 6.7 Results for Ksat Cases at Base Layer............................................................ 97 Figure 6.8 Results for Air Entry Values at Subgrade Layer.......................................... 99 Figure 6.9 Hydraulic Conductivity Curves for Ksat Cases at Subgrade Material........ 100 Figure 6.10 Results for Ksat Cases at Subgrade Layer .................................................. 101

Page 13: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

Figure 6.11 Soil Water Characteristic Curves for Mn/DOT Aggregate Base Material.... 102 Figure 6.12 Soil Water Characteristic Curves (Air Entry = 3 kPa) for Base Material ..... 103 Figure 6.13 Estimated Hydraulic Conductivity Curves for Base Material....................... 103 Figure 6.14 Results for Different Types of Base Materials.............................................. 104 Figure 6.15 Results for Different Rain Events ................................................................. 106 Figure 6.16 Results for Different Water Table Positions ................................................. 107 Figure 7.1 Under Drain Location for Case 1 in the Pavement System........................... 111 Figure 7.2 Soil Water Characteristic Curve for Under Drain Material .......................... 112 Figure 7.3 Estimated Hydraulic Conductivity Curve for Under Drain Material ............ 113 Figure 7.4 Finite Element Model for Case 1 .................................................................. 114 Figure 7.5 Under Drain and Collector Pipes Location for Case 2 in the Pavement System .......................................................................................... 115 Figure 7.6 Finite Element Model for Case 2 .................................................................. 115 Figure 7.7 Edge Drain Location for Case 3 in the Pavement System ............................ 116 Figure 7.8 Finite Element Model for Case 3 .................................................................. 116 Figure 7.9 Finite Element Model for Case 4 .................................................................. 117 Figure 7.10 Results Comparison for Drainage Systems ................................................... 118

Page 14: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

EXECUTIVE SUMMARY Water in the pavement system can lead to moisture damage, modulus reduction, and loss of

strength. In the past, the approach taken by state agencies has focused on preventing water from

entering the pavement, providing adequate drainage to remove it quickly, or building the pave-

ment strong enough to resist the combined effect of vehicle loads and water, thus presumably

reducing the detrimental effects of water on pavements. The drainage design criteria used in the

past have been based on the assumption that both the flow of water through pavements and the

drainage of pavement layers can be represented with saturated flow assumptions. However, full

saturation of pavement systems can occur only under very specific circumstances, when positive

total heads are present (e.g., surface ponding, etc.) and distributed in such a manner that

saturation of the pavement system is reached. In the absence of positive total heads, the

pavement system will remain unsaturated. When rain starts following a dry period, the system is

usually in an unsaturated state, and may remain in this condition even at the end of the rainfall.

The amount of water that can flow through the soil is primarily a function of matric suction and

gravity forces (7). For a given soil with a given gradation, the permeability is controlled by the

matric suction, which in turn is affected by the degree of saturation. A degree of saturation of 80

percent may result in heightened suction, and a resulting permeability of only half its saturated

value. Hence, the models that use only fully saturated approaches are not adequate for

representing the flow of water through pavements, and thus pavement drainage. Finally, recent

studies by Minnesota Road Research Project (Mn/ROAD) pavement researchers (8, 9) have

shown that most dense graded pavement base layers tend to be unsaturated most of the time, with

the exception of brief periods around major rain events.

This report focuses on the modeling of unsaturated flow through flexible pavement systems to

represent the first step in the development of a comprehensive approach to drainage and pave-

ment design that integrates the true effects of moisture on pavement moduli and mechanistic-

empirical pavement design. Three Mn/ROAD test cells are selected for the evaluation of water

flow through pavement systems, namely Cells 33 to 35. Volumetric moisture contents from

automated time domain reflectometry probes in the granular base courses at Cells 33 through 35,

in combination with measured rain events are used to back-calculate the likely infiltration and

drainage response of the pavement systems for Cells 33 through 35. The resulting unsaturated

Page 15: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

flow patterns through these pavement sections are then used to evaluate how water moves

through pavements, how long the water stays in a pavement structure, what material properties

control how long water stays in a given structure, and what boundary conditions and structure

effects (water table, shoulder construction, edge drains, etc.) most affect the moisture conditions

in the pavement.

Page 16: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

1

CHAPTER 1

INTRODUCTION It is a well-known fact that water in pavement systems is one of the principal causes of prema-ture pavement failure (1). Water in the pavement system can lead to moisture damage, modulus reduction, and loss of strength. Saturation can reduce the dry modulus of both the asphalt layer (30% or more) and the base and subbase modulus (50% or more) (2). Similarly, modulus reduction of up to 30 percent can be expected for asphalt-treated bases, and over 50 percent for saturated fine-grained subgrade soils (2). These detrimental effects can be reduced by preventing water from entering the pavement,

providing adequate drainage to remove infiltration, or building the pavement strong enough to

resist the combined effect of load and water. Pavement service life can be increased by 50% if

infiltrated water can be drained without delay (3). Similarly, pavement systems incorporating

good drainage can be expected to have a design life of two to three times that of undrained

pavement sections (4, 5). The AASHTO pavement design guide (2) and the US Army Corps of

Engineers Pavement Design Guide (6) both account for the positive effects of drainage through

the use of drainage factors.

The magnitude of these factors is directly related to the length of the time that excess moisture

remains in the pavement. With new mechanistic-empirical pavement design procedures be-

coming more common, there is a need to develop an improved understanding of the mechanics

of water flow through pavement systems, as well as the direct quantification of the effects of

modulus with varying water content. Current design criteria are based on only saturated flow

theory and on an over simplification of in situ conditions.

Full saturation of pavement systems can occur only when positive total heads are present (e.g.,

surface ponding, etc.) and distributed in such a manner that saturation of the pavement system

is

reached. In the absence of positive total heads, the pavement system will remain unsaturated.

When rain starts following a dry period, the system is usually in an unsaturated state, and may

remain in this condition even at the end of the rainfall. The amount of water that can flow

Page 17: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

2

through the soil is a function of permeability and gravity forces, as well as material matric suc-

tion (7). For a given soil with a given gradation, the permeability is controlled by the matric

suction, which in turn is affected by the degree of saturation. A degree of saturation of 80 per-

cent may result in heightened suction, and a resulting permeability of only half its saturated

value. Hence, the models that use only fully saturated approaches are not adequate for repre-

senting the flow of water through pavements, and thus pavement drainage. Finally, recent

studies by Minnesota Road Research Project (Mn/ROAD) pavement researchers (8, 9) have

shown that most dense graded pavement base layers tend to be unsaturated most of the time,

with the exception of brief periods around major rain events. Hence, it is of extreme

importance to fully understand: 1) how water moves through pavements, 2) how long the water

stays in a pavement structure, 3) what material properties control how long water stays in a

given structure, and 4) what boundary conditions and structure effects (water table, shoulder

construction, edge drains, layering, etc.) most affect the moisture conditions in the pavement.

Knowledge of the relative effects of these factors should allow for the development and

integration of more direct measures of the effects of moisture in pavements into new pavement

mechanistic-empirical design procedures.

This research focuses on the modeling of unsaturated flow through pavement systems, and

therefore represents a first step in the development of a comprehensive approach to drainage

and pavement design that integrates the true effects of moisture on pavement moduli and

mechanistic-empirical pavement design.

1.1 OBJECTIVES

The objectives of this research include: 1) the modeling of the movement of water through flex-

ible pavement systems that are unsaturated, 2) evaluating which material properties most affect

how long water stays in a given pavement, and 3) evaluating the potential effects of water

table, shoulder construction, layering, edge drains, and open-graded bases on the drainage of

flexible pavement systems.

Page 18: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

3

1.2 SCOPE OF REPORT

The study addresses the following: 1) verification of SEEP/W software for modeling unsatu-

rated flow of water through pavements, 2) modeling of the movement of water through flexible

pavement systems that are variably saturated, 3) identification of which material properties

most affect how long water stays in a given pavement, and 4) evaluating the potential effects of

water table, shoulder construction, layering, edge drains, and open-graded bases on the

drainage of flexible pavement systems.

Chapter 2 provides a literature overview of both the current state-of-the-practice in drainage

design, as well as the current-state-of-the-art. Chapter 3 details the research methodology used

in this study. Chapter 4 evaluates the use of SEEP/W for simulating the flow through

simplified pavement systems. Chapter 5 describes the calibration of predicted results to field

data for different pavement structures. Chapter 6 evaluates the effects of various parameters on

flow under unsaturated conditions. Chapter 7 presents the effects of edge and under drains in a

typical flexible pavement system. Chapter 8 presents conclusions and recommendations based

on the analyses presented in this study.

Page 19: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

4

CHAPTER 2

LITERATURE REVIEW

Current assumptions (2) used in pavement drainage construction and design will be reviewed.

An example will be provided of the effects of poorly drained flexible pavements on pavement

thickness. Some of the more common positive pavement drainage systems will be reviewed,

including the use of open graded base courses, edge drains, and the use of geo-textiles for

drainage improvement. The effects of unsaturated pavement material properties and layering

on pavement drainage will also be discussed.

2.1 CURRENT STATE-OF-THE-PRACTICE

There are two different types of fluid flow, saturated and unsaturated. In general, the fluid flow

through porous soil is governed by the matric suction of the soil, which in turn controls the per-

meability of the soil. The matric suction is affected by the factors that affect how closely

packed the soil particles are, including gradation, density, and particle angularity (10). As any

given soil becomes less and less saturated, the matric suction will in turn increase and become

more and more negative. Similarly, in the extreme case when the soil is saturated, and all the

voids are filled with water, the matric suction is zero, and the hydraulic conductivity is

independent of the pore suction; hence Ksat is considered as a constant value. In this case,

positive total heads are required at every point in the pavement system to allow water to flow

into and through the system. Thus, the driving forces that are required for saturated flow are

gravitational and pressure-potential gradients (11). Under saturated conditions, the flow (flux)

is proportional to the length of the drainage path and the total hydraulic head loss between any

two points of interest, as well as the cross sectional area of flow (Darcy’s Law).

Based on saturated flow conditions, the two different cases that are typically considered for the

hydraulic design of pavement systems include steady-state flow conditions, and time-to-drain

conditions. Under steady state flow conditions, the permeable base should carry the design

flow that infiltrates the pavement surface. However, the difficulty in estimating the proper

Page 20: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

5

precipitation frequency and duration makes the application of steady-state analyses tedious

under most circumstances, except where field measurements are available to determine all

input parameters and boundary conditions. Here the second approach, time-to drain, offers a

more practical solution.

According to the FHWA Pavement Subsurface Drainage Design Manual (12), the “time-to

drain” is a parameter that allows the determination of the drainage performance of pavements.

This approach is based on flow entering the pavement until the aggregate base course is satu-

rated. Excess runoff will not enter the pavement section after it is saturated; this flow will

simply run off on the pavement surface. After the precipitation event, the base will drain to a

drainage system.

Casagrande and Shannon (13) showed that the time for 50% drainage of a base layer can be

calculated as follows: (2.1) where t50 is time for 50% drainage, ne is the effective porosity, L is the length of the drainage layer, k is the permeability of the drainage layer, H is the thickness of the drainage layer, S is the slope of the drainage layer.

2.1.1 Effects of Drainage on Pavement Design Life

A good understanding of the flow of water in the subsurface of the pavement may improve the

design of pavement drainage systems. Cedergren (5) performed a study that shows that 15

billion dollars a year can be saved by designing and building pavements with good pavement

drainage and sub-drainage characteristics. However, a thorough understanding of the factors

influencing the flow of water through pavement systems is required to fully realize these

savings.

2e

50n Lt

2k(H SL)=

+

Page 21: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

6

Currently, the FHWA promotes the use of free draining materials in base and subbase construc-

tion. Using the framework of time-to-drain, the AASHTO (2) pavement design equations show

that pavement performance can be greatly improved if free draining materials are used for base

and subbase construction. The effects of excess moisture and the length of time it is retained

within the pavement system are shown in the 1998 AASHTO Guide for pavement design. This

guide also contains specific structural requirements for pavements, which are weakened due to

effects of moisture. The magnitude of these structural factors is directly related to the length of

time that the moisture is retained in the structure. They apply not only to the design of new

pavements but also to the evaluation of existing pavements. For a pavement to have good

drainability characteristics according to AASHTO (2), the structural section of the pavement

should not be filled with excess water and it should not carry heavy wheel loads during periods

when there is excess moisture under the pavement. For this, the water should be able to flow

out of the pavement faster than it enters. The time required to drain at the end of the inflow

period must be short for the excess water not to remain in the structure long enough to freeze

(in cold places).

Table 2.1 presents the different drainage levels for a pavement structure, according to

AASHTO (2), for 50% of drainage.

Table 2.1 Classification of 50% of drainage

Quality of Drainage Water Removed Within

Excellent 2 hours Good 1 day Fair 1 week Poor 1 month Very poor (Water will not drain)

When calculating the adequate thickness for a layer within a pavement system, AASHTO (2)

uses a layer coefficient “m” to modify the structural number (SN). Table 2.2 presents the

recommended m values, as a function of the quality of drainage and the percent of time during

the year the pavement structure would normally be exposed to moisture levels approaching

saturation.

Page 22: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

7

Table 2.2 Recommended m values for modifying structural layer coefficients of untreated base and subbase materials in flexible pavements

Percent of Time Pavement Structure is Exposed to Moisture

Levels Approaching Saturation Quality of Drainage Less Than 1% 1-5% 5-25% Greater Than 25%

Excellent 1.40-1.35 1.35-1.30 1.30-1.20 1.20 Good 1.35-1.25 1.25-1.15 1.15-1.00 1.00 Fair 1.25-1.15 1.15-1.05 1.00-0.80 0.80 Poor 1.15-1.05 1.05-0.80 0.80-0.60 0.60 Very poor 1.05-0.95 0.95-0.75 0.75-0.40 0.40

2.1.2 Summary of Current Practice

Unfortunately, the current drainage criteria used by the FHWA and AASHTO (2), have all

been performed under the assumption of saturated conditions (13, 14, 15, 16, 17, 18). In these

previous works, saturated permeability and gravity were identified as the controlling factors

for pavement drainability. However, most pavements stay unsaturated most of the time and it

is rare to have fully saturated conditions in pavements. When rain follows a dry period, the

base and the subbase are usually unsaturated. The amount of water that flows through the base

and subgrade is not only a function of gravitational forces, it is also the function of matric

suction of the material, which controls the permeability (7). Because of the relationship

between matric suction and permeability, it is not justified to consider fully saturated condition

for study of pavement drainage. Variation of time is another factor that should be considered in

addition to unsaturated zone. Transient flow problems are much more complex than the steady

state for which classical solutions are available.

The unsaturated characteristics of pavement materials, along with layering and geometry

dictate the flow of water through pavement systems. The next part of this chapter deals with an

overview of unsaturated flow studies.

Page 23: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

8

2.2 POSITIVE DRAINAGE SYSTEMS

2.2.1 Open Graded Base Course

Perhaps the simplest drainage system is the inclusion of an open graded base course in the

pavement structure. This type of layer must consist of sound, clean and open-graded materials.

It must also have a high permeability to allow the free passage of water. It should be protected

from clogging by the use of filters. In order to have a high permeability, the fine portion

should be eliminated from the gradation (19). However, this results in the decrement of the

drainage layer stability. To compensate for this, small amounts of asphalt or Portland cement

are introduced into the base for stability. The material will have only a slightly reduced perme-

ability and be stable at the same time. A key issue in the design of open graded base courses, is

to design the underlying filter material to be fine enough to prevent the adjacent subgrade

(finer) material from piping into the filter material, and course enough to carry water without

any significant resistance (20).

2.2.2 Edge Drains

The design and functioning of subsurface drainage systems involves the consideration of many

details, such as material type, separation layer type, edge drain location, outlet design, and con-

struction. The pipes to be used for subdrainage may be made of concrete, clay, bituminized

fiber, metal, or various plastics with smooth or corrugated surfaces (19). Suitable aggregates or

fabrics as filter materials to prevent the openings from clogging must surround them. The mate-

rial in contact with the pipes must be coarse enough that no appreciable amount of this material

can enter into the pipes.

The edge drain must have the necessary hydraulic capacity to handle water being discharged

from the permeable base. The placement of the edge drain depends on the sequence of con-

struction. The drainage trench should not be placed in the trackline of the paver or in the wheel

path. It should be lined with the geotextile to prevent migration of fines from the surrounding

soil into the drainage trench, but the top of the trench adjacent to the permeable base should be

left open to allow a direct path for the water into the drainage pipe.

Page 24: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

9

The drainage pipes are usually placed at a constant depth below the pavement surface. Longi-

tudinal slope of the edge drain has a significant effect on flow capacity of pipe edge drains

(12).

According to saturated flow theory, the one-dimensional flow of water from a high

permeability soil layer into a lower permeability soil layer can be represented by an effective

permeability for the two layers that is lower than the permeability of the high permeability soil

layer (10, 4). Therefore, the current practice has been to specify backfill material for the edge

drain trench with at least the permeability of the surrounding aggregate base material (19, 21).

The proper compaction of the backfill material is also important to avoid settlement over the

edge drain, as discussed by Birgisson and Roberson (9).

2.2.3 Use of Geotextiles for Drainage Improvement

Geotextiles are filter fabrics that not only retain the soil and allow water to flow, but also

protect the drainage layer from clogging (19). They are made from strong, tough, rot-proof

polymer fibers formed into a fabric of the woven or nonwoven type, and they must have

sufficient opening areas to prevent them from clogging.

The most important dimension of geotextiles is the apparent opening size (AOS), defined as the

size of glass beads when 5% pass through the geotextile. The American Society of Testing

Materials (22) specifies the method for determining AOS, as well as the retention, permeability

and clogging criteria.

Geotextiles are used in pavement subsurface drainage as an envelope around trench drains, a

wrapping of pipe drains, or a filter of drainage layers (23).

2.3 UNSATURATED FLOW THROUGH PAVEMENTS

The unsaturated zone is located above the water table. Within this zone, the pore spaces are

usually only partially filled with water, the reminder of the pores are taken up by air. There-

fore, the volumetric water content is lower than the soil porosity. Due to the fact that water in

this zone is held in the soil pores under surface-tension forces, negative pressures or matric

Page 25: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

10

suction pressures are developed. In addition, in this zone both the volumetric water content

and the hydraulic conductivity are functions of this suction pressure (24, 25). Similarly,

changes in matric suction and volumetric water content, affect the hydraulic conductivity. The

hydraulic conductivity increases with increasing volumetric water content and decreasing

suction, as the matric suction becomes less and less negative (26).

Most water in pavements is introduced to the pavement system through the process of infiltra-

tion into unsaturated pavement layers. Infiltration is the process in which water moves across

the atmosphere-soil interface, i.e., water seeps from the pavement ground surface and enters the

base, subbase, and subgrade soils. The time rate at which water infiltrates across the atmo-

sphere-soil interface is known as infiltration rate (26). The total volume of liquid crossing the

interface over a given period of time is known as cumulative infiltration. Infiltration also signi-

fies soil sorptivity, which is the characteristic that determines how much of the incident rainfall

will run off and how much will enter the soil and either percolate downward or be evapo-trans-

ported (11). Under ponded conditions, infiltration into an initially dry soil profile has a high

rate early in time, decreasing rapidly and then more slowly settling down to nearly a constant

rate.

Water may infiltrate into the pavement from a number of sources, some of the most common

being surface water entering the pavement-shoulder joints, longitudinal and transverse

construction joints and pavement cracks, as well as from capillary rise of water above the water

table from below the pavement. Similarly, seasonal increases in ground water table elevation,

along with a rise in the associated capillary fringe may allow near saturation of pavement

components at various times (4).

Soil characteristics play an important role in the infiltration rate. Total infiltration of any layer

depends upon its porosity, thickness and quantity of water or other liquid present. Soil texture,

structure, organic matter, and other physical properties determine the magnitude of the porosity

of a given soil (11).

Page 26: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

11

2.3.1 Soil Water Characteristic Curves

Water flow in the unsaturated zone primarily is due a gradient in total head potential, which is

dominated by differences in matric potential between one point and another (24). A volumetric

moisture content less than the saturated value leads to the development of suction within the

soil mass, which in turn causes a reduction of hydraulic conductivity. Hence, two material

properties are needed to describe the drainage behavior of soils at any given saturation level,

namely the suction present in the soil at a given saturation level, and the corresponding

hydraulic conductivity. To fully describe the unsaturated behavior of soils, it is necessary to

determine the suction present in the soil at all likely saturation levels, along with the hydraulic

conductivity as a function of the resulting soil suction. The first relationship is generally

referred to as the “soil water characteristic curve” and the second one is called the “hydraulic

conductivity curve.”

According to Gupta and Wang (27), the soil water characteristic curve describes the

relationship between the volume of water in the soil and the energy state of the water present in

the soil. The shape of the soil water characteristic curve is a function of the pore size

distribution, which is determined by the gradation of the soil, as well as the soil particle shape,

and the particle packing (27). Figure 2.1 shows a typical soil water characteristic curve,

displaying the relationship between volumetric water content and soil suction (28). The air-

entry value of the soil is the matric suction at which air starts entering the largest pores in the

soil. The residual water content is the water content of the soil at which a large amount of

suction pressure is required to remove the additional water from the initially saturated soil. The

desorption curve differs from the absorption curve due to hysteresis (26).

Since soil water characteristic curves are represented by relationships between volumetric

water content and suction and degree of saturation and suction, it is possible to also develop a

relationship between volumetric water content and the degree of saturation, which can be

expressed with the following relationship (29):

( )( )

( )( )

( )r r re

r s r

S SS

1 S− θ− θ θ− θ

= = =− θ − θ ∆θ

(2.2)

Page 27: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

12

Figure 2.1 Typical soil water characteristic curve

where: S is the degree of saturation, Sr is the residual saturation corresponding to the value of θr , θ is the volumetric water content θr is the residual water content θs is the volumetric water content at saturation. There are primarily two ways of obtaining hydraulic conductivity curves for soil – either by

direct measurement or by estimation. Direct measurement of hydraulic conductivity curves is

tedious and time consuming. Therefore, hydraulic conductivity curves tend to be estimated

from soil-water characteristic curves, analytical models, and grain size curves. The hydraulic

conductivity models developed over the years vary in complexity from purely empirical

methods to more sophisticated closed-form solutions. Little work has been performed in

identifying appropriate hydraulic conductivity models for roadway materials. In the following

sections, some common models for both soil water characteristic curves and hydraulic

conductivity curves will be described.

0

10

20

30

40

50

60

0.1 1 10 100 1000 10000 100000 1000000

Soil Suction (KPa)

Vol

umet

ric

Wat

er C

onte

nt (%

) s Air-Entry Valueθs

Desorption Curve

Residual Water Content, θ'r

Adsorption Curve

θ's

Residual air content

Page 28: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

13

2.3.1.1 Brooks and Corey model (30)

This model proposed a relationship between the degree of saturation and the matric suction

based on experimental data:

1

eSPB

−νΨ =

for Ψ ≥ PB (2.3)

eS 1= for Ψ< PB

where Se is the effective degree of saturation, Ψ is the matric suction, PB is the bubbling pressure of the soil, which is the height of the capillary fringe, ν is the pore size distribution index parameter, a measure of the soil grain uniformity. Based on Bear’s (31) definition of the water capacity:

wC ∂θ= −

∂ψ (2.4)

and Equations (2.2) and (2.3), a new expression for this water capacity can be obtained (2.5):

ew

SC ∆θ×=

νψ (2.5)

Figure 2.2 shows the influence of the parameter ν on the Se - Ψ response. 2.3.1.2 Van Genuchten model (32)

Van Genuchten proposed an empirical equation to relate the matric suction and the volumetric

water content:

(2.6) for Ψ< 0 where: β and γ are dimensionless coefficients, and γ = 1-1/β, α is a coefficient that has the dimension of the inverse of the piezometric head.

( )( )e1S

1γβ

=+ αΨ

eS 1=

Page 29: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

14

Figure 2.2 Typical values for the Brooks and Corey model for water retention (33) For limited cases, Brooks and Corey’s parameters can be related to those used in Van

Genuchten’s model. However, this relation gets distorted when the water content approaches

saturation. As well as equation (2.5), an expression of the water capacity can be obtained based

on the Van Genuchten model:

( ) ( ) 1 1 1w eC 1 Sβ− + γ= α× ∆θ× β − × α× Ψ (2.7)

Figure 2.3 shows the variation of the matric suction for a range of b values.

2.3.1.3 Brutsaert model (34)

Brutsaert (34) proposed another relationship between the degree of saturation and the matric

suction. It is a relationship that combines the pore size distribution and statistical arguments

(34):

for Ψ>0 (2.8)

for Ψ< 0

( )eSβ

α=

α + Ψ

eS 1=

Page 30: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

15

Figure 2.3 Typical values for the Van Genuchten model for water retention (29)

where α depends on the suction units, β is dimensionless.

The water capacity equation for the Brutsaert’s model is given by:

( )

12 1

w e2C Sβ−

β−

β

∆θ×α×β× Ψ ∆θ×β= = × Ψ

αα + Ψ (2.9)

where α and β are empirical coefficients.

2.3.1.4 Vauclin model (35)

Similar to the model proposed by Brutsaert (34), Vauclin (35) introduced another empirical

equation relating the degree of saturation and the matric suction: for Ψ>1 (2.10)

for Ψ≤ 1 cm

( )eSln( )β

α=

α + Ψ

eS 1=

Page 31: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

16

The water capacity equation for this model is given by:

for Ψ≥1 cm (2.11) where α and β are fitting coefficients.

2.3.1.5 Bear and Verruijt model (36)

The relationship between the degree of saturation and the suction is not unique, but is charac-

terized by hysteresis between wetting and drying curves. Therefore, a value of suction during

drainage is greater than during wetting (Figure 2.4). This phenomenon, called hysteresis, has

been explained through several models. Bear and Verruijt (36) model states that the hysteresis

is due to the ink-bottle effect and the raindrop effect (Figure 2.5), which are related to the

variation of the meniscus radius in the capillary tube cross-section, and the variation of the

contact angle, respectively.

2.3.1.6 Gray and Hassanizadeh model (37)

Based on the thermodynamic principles and momentum balance equations, Gray and

Hassanizadeh (37) proposed a general expression for the capillarity. They also defined the

suction as net energy per unit volume of pore space that would be released per unit change in

saturation. Therefore, a unique suction curve is obtained as a function of water and air density,

temperature, degree of saturation, and area of air-water interface per unit of porous medium.

This allows obtaining the capillary pressure without taking into account the hysteresis effect.

However, the required parameters are not easy to obtain.

2.3.1.7 Mualen model (38)

Mualen presented a model based on the independent domain theory (38), which conceptualizes

the porous medium as a system of pore domains with characteristic wetting and drying pore

radii such that the pore volume can be expressed as: (2.12) where: Ψw and Ψd are independent variables representing the suction on the main wetting and

drying curves, Ψw being always greater that or equal to Ψd . f is a bivariate probability distribution function

( )1

2 1w e2

ln( )C S (ln )(ln )

β−β−

β

∆θ×α×β× Ψ ∆θ×β= = × × Ψ

α× ΨΨ α + Ψ

( )w d w df , d dθ = Ψ Ψ Ψ Ψ∫ ∫

Page 32: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

17

Figure 2.4 Hysteresis effects – Main drainage and wetting curves (31)

Figure 2.5 Variation of the air-water interface due to raindrop effect (31)

Page 33: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

18

2.3.1.8 Summary of soil water characteristic curve models

Table 2.3 summarizes the features of the soil water characteristic curve models discussed

previously.

Table 2.3 Soil water characteristic models – Main Features, Advantages, and

Disadvantages

Model Main Features Advantages Disadvantages Brooks and Corey (30)

Relates the degree of saturation and the matric suction

Based on substantial experimental data

Empirical method

Brutsaert (34) Relates the degree of saturation and the matric suction

Sensitive to suction values close to saturation

It has empirical coefficients

Mualen (38) Determines the volumetric water content as a function of the suc-tion on the main wetting and drying curves

Based on rigorous statistical considerations and hysteresis effect included

Requires wetting and drying test data

Vauclin (35) Relates the degree of saturation and the matric suction

Describes laboratory measurements at small suction well

It has empirical coefficients

Bear and Verruijt (36)

It takes into account the hysteresis phenomenon

Hysteresis effect included Requires wetting and drying data

Gray and Hassanizadeh (37)

Allows obtaining the capillary pressure without taking into account the hysteresis effect

Based on thermodynamic principles that allow for a unique relationship between a suction curve as a function of water and air density, temperature, degree of saturation, and area of air-water interface per unit of porous medium

The required parameters are tedious to obtain

2.3.1.9 Relative permeability models

Under unsaturated conditions, the total permeability is reduced due to the presence of matric

suction in the porous media. Therefore, the concept of relative permeability is used, and it is

the ratio between the unsaturated permeability and the saturated permeability:

er e

K (S )k (S )

K= (2.13)

Page 34: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

19

where kr is the relative permeability ranging from 0 to 1, K is the saturated permeability.

Different models have been proposed to represent the relative permeability as a function of the

effective degree of saturation. Most of these models assume the porous medium is equivalent

to a set of tubes connected randomly in parallel-series, each tube having a different

permeability. As a result, the random variations of the pores sizes normal to and along the

direction of flow can be taken into account (39).

Burdine (40) used the series-parallel model and related the permeability to the soil water char-

acteristic curves. By using the definition of the relative permeability (Equation 2.13), he

obtained:

(2.14)

where Se(r) = dSe/dr is the effective pore size density function.

Mualen (41) derived an analytical expression in which the hydraulic conductivity is not

controlled by the radius of the narrower element:

(2.15) For an n equal to 0.5, Equation 2.15 predicts reasonably well most hydraulic conductivity

curves obtained in the laboratory (41).

In order to determine the saturated permeability of the soil, the falling and constant test perme-

ameter tests are commonly used for sandy soils. For lower hydraulic conductivities, larger

pressure differences need to be applied using pumps. Either flexible or rigid wall permeameters

eS2 ee 20

er e 1 e

20e

dSS(S )k (S ) dS

(S )

Ψ=

Ψ

e2

S e0

n er e e 1 e

0e

dS(S )k (S ) S dS(S )

Ψ = Ψ

Page 35: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

20

can be used. The oedometer test can also be used to determine the hydraulic conductivity of

soils of low permeability (42).

For the determination of the unsaturated permeability of the soil, there are different techniques

such as transient flow in soil columns, pressure plate outflow technique, osmotic permeameter,

and permeameter apparatus (27, 31).

2.3.2 Hydraulic Conductivity Models

Gardner (43) developed one of the first interpolation functions for the hydraulic conductivity

curve, namely:

( ) ( )s

k

kk1 A β

ψ =+ ⋅ψ

(2.16)

where kS is the saturated hydraulic conductivity, ψ is the pore suction, Ak, β are empirical curve fitting coefficients. In order to use this method, the empirical curve-fitting coefficients must be acquired from

experimental hydraulic conductivity data.

Brooks and Corey (30) suggested a hydraulic conductivity relationship that relates hydraulic

conductivity values to the effective saturation of a soil, as follows:

( )n

rs

s r

k k θ − θ

θ = θ − θ (2.17)

where kS is the saturated hydraulic conductivity, θ is the volumetric water content, θs is the saturated volumetric water content, θr is the residual volumetric water content, n = 3 + 2/λ, and is based on the pore size index (λ).

This model allows the inclusion of the soil-water characteristic curve, in a direct way, due to

the defined equation was written in terms of the volumetric water content.

Page 36: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

21

Brooks and Corey’s relationship is based on the analysis performed by Burdine (40), who

applied theories of fluid flow through porous material to develop a relationship that produces

the relative hydraulic conductivity (kr) from the soil-water characteristic curve. The

formulation proposed by Brooks and Corey (30) is simply a reduced form of Burdine’s

equation (40), including an assumed interpolation equation for the soil-water characteristic

curve. Although the Brooks and Corey (30) hydraulic conductivity model should apply nicely

to the coarse base materials, it loses some validity because it is not very efficient at low suction

values (3).

Finally, as can be seen from Equation 2.17, the Brooks and Corey function requires a pore size

distribution value (λ) to characterize the material. The λ value can be calculated using a

regression function from Rawls (44). The regression relationship is based upon the percent

sand, the percent clay, the soil porosity, and the cross product of these values.

Green and Corey (45), based on Childs and Collis-George (46), proposed a more refined

equation to model the hydraulic conductivity, resulting in the following interpolation function:

( ) ( )2 p m

2sj2i

j 1sc

k 30k 2j 1 2ik gn n

=

γ ε θ = ⋅ ⋅ + − ψ ρ ∑ (2.18)

for i = 1, 2, …, m where k(θ)i is the calculated conductivity for a specified water content, θ is the volumetric water content, i denotes the last water content class on the wet end, e.g. i = 1 identifies the pore class

corresponding to the lowest water content for which the conductivity is calculated, ks/ ksc is a matching factor, γ is the surface tension, ρ is the density of water, g is the gravitational constant (cm/sec2), η is the viscosity of water (g/cm/sec-1), ε is the soil porosity (cm3/cm3), p is a parameter that accounts for interaction of pore classes, n is the total number of pore classes between θ = 0 and θs, ψj is the pore pressure for a given class of water filled pores.

Page 37: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

22

As can be seen from Equation (2.18), the range of validity for this relation is based on the use

of many soil and fluid properties. Also, it should be mentioned that with some manipulation,

this method provides the same results as the method outlined by Fredlund et al. (47).

Van Genuchten (32) used a method similar to Brooks and Corey’s (30) model to describe the

hydraulic conductivity at low suction values, resulting in the following relationship:

( )

2m1 12 m

r rs

s r s r

k k 1 1

θ − θ θ − θ θ = − − θ − θ θ − θ

(2.19)

where: ks is the saturated hydraulic conductivity, θ is the volumetric water content, θs is the saturated volumetric water content, θr is the residual volumetric water content, m = λ / (1 +λ), and is based on the pore size index (λ).

Unfortunately, just like the Brooks and Corey model (30), Van Genuchten’s approach also

relies on the pore size index discussed previously. However, the difference is that instead of

basing his work on Burdine’s equation (40), Van Genuchten used Mualen’s model (41).

Table 2.4 summarizes the key features of the hydraulic conductivity models presented.

2.3.3 Effects of Layering on Pavement Drainage

In coarse-grained materials the soil water characteristic curve shows that almost all the volu-

metric water content held in the pore space is removed at low suction values. These results are

reasonable because of the low capillary effects associated with large pores. Conversely, in

fine-grained materials the soil water characteristic curve shows that much of the water is held

in the soil at higher suction. These features determine where the majority of the volumetric

water content is stored in the infiltration system. The other important factor in unsaturated

flow behavior through layered systems is the hydraulic conductivity function for each material.

In general, due as matric suctions increase and become more negative, the hydraulic

conductivity decreases with

Page 38: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

23

Table 2.4 Hydraulic conductivity models – Main Features, Advantages, and Disadvantages

Model Main Features Advantages Disadvantages

Gardner (43)

A two-parameter empirical model

Permeability is represented as a smooth function of suction instead of degree of saturation

Empirical curve-fitting coefficients must be acquired from experimental hydraulic conductivity data

Brooks and Corey (30)

Allows the inclusion of the soil-water characteristic curve in a direct way

Based on substantial experimental data

Empirical equation, not very efficient at low suction values, relies on a difficult to obtain λ value

Green and Corey (45)

Refined equation to model the hydraulic conductivity

Based on substantial experimental data

Use of many soil and fluid properties

Van Genuchten (32)

Equation based on the volumetric water content

Based on substantial experimental hydraulic conductivity data at low suction values

Relies on a difficult to obtain λ value

suction. Due to these effects, when a fine material is overlying a coarser material under

unsaturated conditions, the effect is a reduction in flow in the coarse-grained material.

Although the coarse-grained material starts with a higher hydraulic conductivity, the larger

pore size and reduced surface tension effects cause the hydraulic conductivity to drop rapidly at

lower pore suction levels. The hydraulic conductivity of the fine-grained material however

tends to not decrease nearly as rapidly and hence the hydraulic conductivity can actually

become lower in the coarse-grained material, resulting in a capillary barrier between the two

different materials. The intersection of these two curves represents the matric suction that

causes equivalent hydraulic conductivities.

A combination of these two effects causes the majority of the water to remain in the fine-

grained material where the hydraulic conductivity is much higher at the crossover suction.

Therefore, the water tends to flow primarily through the fine-grained material. However, this

only happens if the pore suction values are higher than the suction at which there are equivalent

hydraulic conductivities.

Page 39: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

24

A proper understanding and use of the capillary barrier effect in multiple layer systems can be

effective in reducing the amount of flow through the system. Similarly, the capillary barrier

effect may result in adverse drainage conditions if not accounted for properly in the design of

pavements.

Page 40: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

25

CHAPTER 3

RESEARCH METHODOLOGY In this Chapter, the research methodology used in this study will be discussed. First, the

analysis tools used will be described, followed by a description of pavement sections and

materials, and scope of study.

3.1 SELECTION OF ANALYSIS TOOLS

Based on the established fact that drainage through pavements is generally governed by un-

saturated flow conditions, it was necessary to select a software analysis tool that could perform

unsaturated flow analysis. In addition, other desired properties included seamless transition

between unsaturated and saturated flow analysis, full transient analysis capabilities, user

friendly pre- and post-processor interfaces, established track record, user-friendliness,

flexibility in terms of changing unsaturated material properties, and solid user support. Almost

immediately, the software package that was identified as possible candidate was SEEP/W from

Geoslope International (48). The SEEP/W package is formulated to analyze both saturated and

unsaturated flow, and includes user-friendly pre- and post-processors. SEEP/W also contains

options for choosing and inputting soil water characteristic curves and hydraulic conductivity

curves. A database of typical unsaturated hydraulic properties is included, and soil water

characteristic curves can be predicted from grain size distribution curves. Similarly, Van

Genuchten (32) and Fredlund and Xing (28) type hydraulic conductivity curves can be

predicted from soil water characteristic curves. Therefore, it was decided to select SEEP/W as

the software analysis tool to be used for the rest of this study.

Finally, to provide a baseline comparison between time-to-drain using unsaturated flow theory

and SEEP/W and traditional saturated time-to-drain calculations, the FHWA computer program

entitled “Drainage Requirements in Pavements,” or DRIP, was selected (49)

Page 41: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

26

3.1.1 SEEP/W

The finite element-based SEEP/W is formulated to analyze both saturated and unsaturated

flow. Flow in unsaturated soil follows Darcy's Law in a similar manner to flow in saturated

soil. The flow is proportional to the hydraulic gradient and the hydraulic conductivity

(coefficient of permeability). The major difference between saturated and unsaturated flow in

SEEP/W is that in a saturated soil, the hydraulic conductivity is insensitive to the pore-water

pressure, while in an unsaturated soil, the hydraulic conductivity varies greatly with changes in

pore-water pressure.

Although SEEP/W assumes that the flow of water follows Darcy’s law for both unsaturated and

saturated flow, the governing equation for all the calculations is Richards’ equation, shown

below, which represents in a more suitable form the unsaturated flow:

x yH Hk k Q

x x y y t ∂ ∂ ∂ ∂ ∂Θ + + = ∂ ∂ ∂ ∂ ∂

(3.1)

where H is total head, kx is hydraulic conductivity in the x-direction, ky is hydraulic conductivity in the y-direction, Q is the applied boundary flux, Θ is volumetric water content, t is time.

Richards’ equation (Eqn. 3.1) is based on the balance of flow into and out of an element

volume of soil. This fundamental partial differential equation simply states that the difference

between the flow entering and leaving an elemental volume at a point in time is equal to the

change in volumetric water content. Richards’ equation can be used for both saturated and

unsaturated conditions, since the right part of the equation becomes zero when the first case is

applied or in steady-state conditions.

3.1.2 DRIP

The microcomputer program Drainage Requirements In Pavements (50) can be used to design

subsurface drainage for highway pavements. Among the drainage design elements, DRIP

allows for the calculation of the time to drain in the drainage layer of a pavement system. This

Page 42: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

27

program is based on simple analytical prediction methods, which assume that pavement

systems are only exposed to saturated conditions.

The calculations are based on two methods: Barber and Sawyer method (51) and Casagrande

and Shannon method (13).

The Barber and Sawyer (51) equations are:

kh

LTnt

2Re= (3.2)

where t is time required to drain U% of water from the drainage layer, U is the percent of water drained, and it is a function of the porosity, saturation and

effective porosity of the drainage layer, ne is the effective porosity of the drainage layer, LR is the resultant length of the drainage path. It is a function of the width longitudinal

slope and resultant slope (SR) of the drainage path, k is the permeability of the drainage layer, h is the thickness of the drainage layer, T is a factor determined by, (3.2a)

for U > 0.5 (3.2b) for U ≤ 0.5 Casagrande and Shannon’s method (13) is based on the same equation (3.2), but T factor is

determined by:

(3.3a)

for U>0.5

( )( )

++−

⋅+

+⋅−=

4.2SU-12.1USSlogS15.1

S4.21logS48.0S5.0T

R

RRR

R

2RR

+⋅−=

R

2RR S

U8.41logS48.0UST

( )( )2R R R

R R RR R

2S 2US 1 S 1cT S S ln S ln2 2 2U S 1 S

− + + = + ⋅ − ⋅ − +

Page 43: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

28

(3.3b)

for U ≤ 0.5,

(3.3c)

The amount of water present in the drainage layer material as a percentage of the available

volume is defined by the percent of saturation. In the “Time to drain” method, DRIP assumes

that saturation is 100% because it is considered that the drainage layer is saturated at the time

to drain and that there is no additional inflow to this layer once the rainfall has ceased. Thus,

the hydraulic conductivity is considered as a constant value.

3.2 MATERIALS AND PAVEMENT SECTIONS

3.2.1 Pavement Sections

The pavement systems that were used for this study correspond to testing sections constructed

by Minnesota Department of Transportation (Mn/DOT) as a part of the Minnesota Road

Research Project (Mn/ROAD). These sections are denominated Mn/ROAD Cell 33, Cell 34,

and Cell 35. They consist of a layer of hot mix asphalt, a Mn/DOT Class 6 Special base

course, consisting of 100 percent crushed granite, an R-12 silty clay subgrade. Figure 3.1

shows the thickness of each of base and hot mix asphalt layers for Cells 33-35.

Figure 3.1 Thickness of base and hot mix asphalt layers for Mn/ROAD Cells 33-35 Figure 3.2 shows pavement geometry, materials and dimensions of the complete system used to

simulate Cells 33, 34, and 35.

Asphalt Layer

Cell 33

4.04 in

Cell 34

3.92 in

Cell 35

3.96 in

Base Layer

12 in 12 in 12 in

2 RR R

R

S 2UcT 2US S ln2 S

+ = − ⋅

1 3R

0.8c 2.4S

= −

Page 44: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

29

Figure 3.2 Pavement geometry and dimensions

3.2.2 Materials

3.2.2.1 Hot mix asphalt layers

For this study the hot mix asphalt layer was considered as an impervious material, therefore its

properties were not taken into account for the material characterization for the different finite

element models. However, Table 3.1 presents the asphalt binder description and thickness for

each cell.

Table 3.1 Asphalt layer description

Cell Asphalt Binder Thickness (in)33 PG 58-28 4.04 34 PG 58-34 3.92 35 PG 58-40 3.96

3.2.2.2 Base materials

The granular base course material for Cells 33-35 consisted of Mn/DOT Class 6 Special

crushed granite aggregate base material. The grain-size distribution curve is shown in Figure

3.3.

4.53 m

3.43 m 4.27 m 4.27 m4.53 m

CLBase Hot Mix Asphalt

Subgrade

30405060708090

100

Page 45: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

30

Figure 3.3 Grain size distribution for Mn/DOT Class 6 special

The soil water characteristic curve and the estimated hydraulic conductivity curve, shown in

Figures 3.4 and 3.5, respectively were obtained from suction plate measurements performed by

Mn/DOT. For modeling purposes, the air entry value for the soil water characteristic curve

was shifted to 3 kPa (52). The verification of the resulting soil water characteristic curve is

shown in Appendix B, along with a verification of the final air entry value to be used.

Figure 3.4 Soil water characteristic curve – Base material (Class 6 special)

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 0.1 1 10 100

Suction (kPa)

Page 46: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

31

Figure 3.5 Hydraulic conductivity curve – Base material (Class 6 special)

A saturated hydraulic conductivity (Ksat) of 1.54E-6 m/s was obtained from the estimated

hydraulic conductivity curve (Figure 3.5) provided by Mn/DOT. In this case, the saturated

hydraulic conductivity was simply obtained as the hydraulic conductivity value at zero suction.

This saturated hydraulic conductivity corresponds to fine sand.

3.2.2.3 Subgrade soil

The subgrade soil can be characterized as an R-12 sandy clayey silty soil. Figures 3.6 and 3.7

show the soil water characteristic curve and the hydraulic conductivity curve for the subgrade,

obtained from Mn/DOT. Figure 3.6 shows that this sandy clayey silt has little variation of

volumetric water content. For example, there is only 9 percent drop for the first 10 kPa. The

total change is 15%. This means that the subgrade has the capacity to hold water for a long

time, implying that its drainage will be slow.

As well as the soil water characteristic curve, the hydraulic conductivity curve also implies that

the subgrade soil has a high resistance to drainage. Within the first 5 kPa, this material goes

from 2.75E-08 m/s to a value close to 2E-13 m/s.

1.0E-15

1.0E-13

1.0E-11

1.0E-09

1.0E-07

1.0E-05

0.01 0.1 1 10 100

Suction (kPa)

k (m

/s)

Page 47: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

32

20.0

25.0

30.0

35.0

40.0

45.0

50.0

55.0

0.001 0.01 0.1 1 10 100 1000 10000

Suction (kPa)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

Figure 3.6 Soil water characteristic curve – Subgrade material (R-12 Silty clay)

1.0E-15

1.0E-13

1.0E-11

1.0E-09

1.0E-07

0.01 0.10 1.00 10.00 100.00

Suction (kPa)

k (m

/s)

Figure 3.7 Hydraulic conductivity curve – Subgrade material (R-12 Silty clay)

Page 48: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

33

CHAPTER 4

EVALUATION OF SEEP/W FOR MODELING UNSATURATED FLOW

In this Chapter, SEEP/W predictions are compared to a series of experimental and analytical

results reported in the literature to evaluate its use for modeling the flow of water through soils

and simplified pavement systems. Results from three laboratory and analytical experiments

reported in the literature are used to provide a comparison to SEEP/W predictions. The first

example is based on a laboratory experiment performed by Vauclin (35), in which unsaturated

soil is infiltrated with a constant flux, and the resulting changes in water table within the soil

are monitored. The second example is based on a modification of Vauclin’s (35) experiment,

where now the soil in Vauclin’s (35) experiment is replaced by a typical dense graded base

material and the flux is set to be representative of the infiltration process that pavement struc-

tures may go through during rainfall events (29). Similarly, the third example is based on a

modification of Example 2, in which materials corresponding to typical granular base and

subbase layers are introduced to simulate the case of a subdrainage layer with high

permeability (29).

4.1 EXAMPLE 1

The first example is based on an experimental setup performed by Vauclin (35) in the labora-

tory with a single layer of soil 3 m long, 2 m high and 0.05 m thick, in order to study the

changes of water content and water pressure occurring in the flow. The soil was packed as

homogeneously as possible between two walls supported by a frame resting on an impervious

horizontal boundary. One of the vertical ends of the slab was connected to a constant head

reservoir, and a water table was imposed at the depth of Ho = 1.35 m. There was no flow

through the vertical left hand side of the slab. A constant flux corresponding to qo = 4.1111 E-

5 m/s was applied on the soil surface over a width Lo = 0.50 m. Figure 4.1 describes the

geometry and conditions for this example.

Page 49: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

34

Figure 4.1 Example 1 – Geometry and boundary conditions

As part of the experiment, suction and hydraulic conductivity curves for the soil material were

determined in the laboratory and fitted by regression analysis to the Brutsaert (34) and Gardner

(43) models, which are also presented in Equations 2.18 and 2.16, respectively. Table 4.1

shows the parameters obtained with the regression analysis.

Based on the hydraulic parameters in Table 4.1 for the Brutsaert (34) and Gardner (43) models,

the soil water characteristic and the hydraulic conductivity curves were obtained (Figures 4.2

and 4.3).

Table 4.1 Soil hydraulic parameters (35)

Soil Water

Characteristic

Curve Parameters

(Brutsaert,1966)

Hydraulic Conductivity

Curve Parameters

(Gardner, 1956)

α β θs A β Ks (m/s)

40,000 2.9 0.30 2.99E+06 5.0 9.72E-05

The finite element mesh used in SEEP/W to simulate this experiment is presented in Figure 4.4.

The mesh is composed of 601 uniformly sized 0.1 m by 0.1 m quadrilateral elements. As

shown previously in Figure 4.1, the boundary conditions are modeled as impervious at the left,

right,

qo = 4.1111 E-05 m/s

2 m

3 m

0.65 m

0.50 m

Impervious boundary

H = 0.65 m

Total Head (H)

Page 50: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

35

Figure 4.2 Soil water characteristic curve for Example 1 (34)

Figure 4.3 Hydraulic conductivity curve for Example 1 (43)

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 0.10 1.00 10.00 100.00

Suction (kPa)

1.0E-12

1.0E-10

1.0E-08

1.0E-06

1.0E-04

0.01 0.10 1.00 10.00 100.00

Suction (kPa)

Hyd

raul

ic C

ondu

ctiv

ity (m

/s) s

Page 51: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

36

Figure 4.4 Example 1 – Finite element model and bottom edges of the mesh, as well as the top 2.5 m portion of the boundary that is not sub-

jected to infiltration.

A comparison between the measured water table data (35) and the predicted water table results

(SEEP/W) is presented in Figures 4.5 to 4.8. A comparison of Figures 4.5 to 4.8 shows that the

variation in the water table elevation with time and elevation obtained from the SEEP/W model

follows the same trend as the measured data.

Table 4.2 summarizes the maximum and minimum differences between the measured and

predicted results in Figures 4.5 through 4.8.

The SEEP/W simulations resulted in a slightly higher water table position under the section

where the constant flux was applied. However, over time, the differences between the

measured and predicted water table are reduced, until the water table reached its steady state

position, after the 8th hour.

Page 52: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

37

Figure 4.5 Comparison between measured and predicted water table at 2 hours

Figure 4.6 Comparison between measured and predicted water table at 3 hours

0.00

0.10

0.20

0.30

0.40

0.50

0.60

0.70

0.80

0.90

0.00 0.50 1.00 1.50 2.00 2.50 3.00

Distance (m)

Vauclin (1979)SEEP/W

0.00

0.20

0.40

0.60

0.80

1.00

1.20

0.00 0.50 1.00 1.50 2.00 2.50 3.00

Distance (m)

Ele

vatio

n (m

)

Vauclin (1979)SEEP/W

Page 53: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

38

Figure 4.7 Comparison between measured and predicted water table at 4 hours

Figure 4.8 Comparison between measured and predicted water table at 8 hours

0.00

0.20

0.40

0.60

0.80

1.00

1.20

0.00 0.50 1.00 1.50 2.00 2.50 3.00

Distance (m)

Ele

vatio

n (m

)

Vauclin (1979)SEEP/W

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

0.00 0.50 1.00 1.50 2.00 2.50 3.00

Distance (m)

Vauclin (1979)SEEP/W

Page 54: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

39

Table 4.2 Maximum difference between predicted and measured water table results in Figures 4.5 – 4.8

Time (h)

Minimum difference (percent)

Maximum difference (percent)

2 0 4.0 3 0 4.0 4 0 4.0 8 0 3.0

4.2 EXAMPLE 2

The second example is based on a modification of Vauclin’s (35) experiment (Example 1),

where now the soil in Vauclin’s (35) experiment is replaced by a typical dense graded granular

base material and the flux is set to be representative of the infiltration process that pavement

structures may go through during rainfall events (29). This case is represented by two-

dimensional infiltration into a simplified pavement structure.

The simplified pavement system consists of an impermeable wearing course overlaying homo-

geneous base material (Figure 4.9). The bottom and vertical boundaries were impermeable, as

well as the wearing course. Therefore, the only source of water was the “shoulder,” which was

modeled with as a 0.5 m wide strip, with a uniform infiltration rate to simulate rainfall. As an

initial condition, the effective degree of saturation was specified as 50%, and a “rainfall” of

constant intensity was applied (qo = 2.7778 E-06 m/s) on the “shoulder” area.

The granular base material was characterized as an anisotropic material. The horizontal

conductivity was 10 times greater than the vertical. The vertical value was taking into account

in SEEP/W by introducing a ratio of 10 between the horizontal and vertical hydraulic conduc-

tivities.

The Brooks and Corey models (30) were used to represent both the soil water characteristic

curve and hydraulic conductivity curve (Figures 4.10 and 4.11).

Page 55: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

40

Figure 4.9 Example 2 – Geometry and boundary conditions

Figure 4.10 Soil water characteristic curve for Example 2 (30)

Kh=3.50 E-05 m/s Kv=3.50 E-06 m/s θo=0.005, ∆θ=0.382 Si=0.5 (initial saturation)

0.025 m

0.20 m

2 m

0.50 m

qo = 2.7778 E-06 m/s Impervious boundary

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

45.0

0.01 0.10 1.00 10.00

Suction (kPa)

Page 56: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

41

Figure 4.11 Hydraulic conductivity curve for Example 2 (30)

This example was modeled in SEEP/W, using the finite element mesh shown in Figure 4.12,

with 580 0.025 m by 0.025 m uniformly sized quadrilateral elements. Due to the fact that the

initial degree of saturation was 50%, an equivalent water table was set, resulting in a maximum

achievable suction pressure of 2.12 kPa above the water table, which was the suction required

to have an effective saturation (Se) of 50%. The suction value of 2.12 kPa is developed for the

entire base layer to reproduce the desired effective saturation.

Figures 4.13 through 4.16 show the variation of the volumetric water content with the distance

at times 1.5 hour, 3 hours, 4.5 hours, and 6 hours, respectively, for different periods of time.

The areas of maximum volumetric water content are located under the pavement shoulder,

within the first 0.5 m away from the only source of water for the system. As expected, the

volumetric water content (θ) changes most rapidly in the transition zone between the shoulder

and the edge of the wearing course. Also, as expected, the variation of θ is higher in the hori-

zontal direction than in the vertical one, due to the anisotropic hydraulic conductivity

conditions. Finally, the SEEP/W predicted results at the first two points closest to the

infiltration zone are

1.0E-10

1.0E-09

1.0E-08

1.0E-07

1.0E-06

1.0E-05

1.0E-04

0.01 0.10 1.00 10.00

Suction (kPa)

Page 57: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

42

Figure 4.12 Example 2 – Finite Element Model

Figure 4.13 Comparison between measured and predicted volumetric water content at 1.5 hours for Example 2

0.0

5.0

10.0

15.0

20.0

25.0

30.0

0.00 0.40 0.80 1.20 1.60 2.00

Distance (m)

Espinoza et al. (1993)SEEP/W

Page 58: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

43

Figure 4.14 Comparison between measured and predicted volumetric water content at 3 hours for Example 2

Figure 4.15 Comparison between measured and predicted volumetric water content at 4.5 hours for Example 2

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

0.00 0.40 0.80 1.20 1.60 2.00

Distance (m)

Espinoza et al. (1993)SEEP/W

0.05.0

10.015.020.025.030.035.040.045.0

0.00 0.40 0.80 1.20 1.60 2.00

Distance (m)

Espinoza et al. (1993)SEEP/W

Page 59: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

44

Figure 4.16 Comparison between measured and predicted volumetric water content

at 6 hours for Example 2 slightly lower than the solution obtained by Espinoza, et al. (29), which used a simple first

order finite difference scheme to simulate the infiltration response. Due to the limitations of

the finite difference scheme used by Espinoza et al. (29), it may not have adequately predicted

the response in the zone where the volumetric water content is changing most rapidly at early

times (Figure 4.13), hence resulting in slight differences from the SEEP/W predictions.

In summary, the results show that SEEP/W provides reasonable predictions of unsaturated flow

behavior under controlled conditions using materials that are typical for dense graded granular

base courses used in pavements.

4.3 EXAMPLE 3

The third Example is based on a modification of Example 2, in which materials corresponding

to typical granular base and subbase layers, shown in Figure 4.17, are introduced to simulate

the case of a subdrainage layer with high permeability (29) As in Example 2, the pavement

slab, both lateral boundaries, and the bottom boundary were assumed to be impervious. The

subbase was

0.05.0

10.015.020.025.030.035.040.045.0

0.00 0.40 0.80 1.20 1.60 2.00

Distance (m)

Espinoza et al. (1993)SEEP/W

Page 60: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

45

Figure 4.17 Example 3 – Geometry and boundary conditions considered to be 100 times more permeable than the base course. Similar to Example 2, the ini-

tial moisture conditions were defined by assuming 50% saturation. The only source of infiltra-

tion was the shoulder, with a rate qo equal to 2.7778 E-06 m/s.

Espinoza, et al. (29) obtained the unsaturated soil properties for the base and subbase materials

from the Brooks and Corey model (30). Figures 4.18 to 4.20 show the resulting soil water

characteristic curve and the hydraulic conductivity curve.

Figure 4.21 presents the SEEP/W Finite Element Model for this example, which has 580

0.025 m by 0.025 m uniformly sized quadrilateral elements to represent both layers. As well as

in Example 2, the water table elevation was fixed at 0.7 m below the surface, to obtain suctions

that produced a state of initial saturation of 50%.

The resulting variation in volumetric water content with distance is presented in Figures 4.22

through 4.29, at times 1.5, 3, 4.5 and 6 hours for the granular base and subbase materials.

0.20m

0.025 m Kh=Kv=3.50 E-06 m/s

Kh=Kv=3.50 E-04 m/s 0.10 m

θo=0.005, ∆θ=0.382 Si=0.5 (initial saturation)

0.50 m qo = 2.7778 E-06 m/s

2 m Impervious Boundary

Page 61: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

46

Figure 4.18 Soil water characteristic curve for base and subbase for Example 3 (30)

Figure 4.19 Hydraulic conductivity curve for base material in Example 3 (30)

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

45.0

0.01 0.10 1.00 10.00

Suction (kPa)

1.0E-12

1.0E-11

1.0E-10

1.0E-09

1.0E-08

1.0E-07

1.0E-06

1.0E-05

0.01 0.10 1.00 10.00

Suction (kPa)

Page 62: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

47

Figure 4.20 Hydraulic conductivity curve for subbase material in Example 3 (30)

Figure 4.21 Example 3 – Finite Element Model

1.0E-10

1.0E-09

1.0E-08

1.0E-07

1.0E-06

1.0E-05

1.0E-04

1.0E-03

0.01 0.10 1.00 10.00

Suction (kPa)

Page 63: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

48

Figure 4.22 Comparison between measured and predicted volumetric water content in base material at 1.5 hours

Figure 4.23 Comparison between measured and predicted volumetric water content in base material at 3 hours

0.05.0

10.0

15.0

20.0

25.0

30.0

35.040.0

0.00 0.40 0.80 1.20 1.60 2.00

Distance (m)

Espinoza et al. (1993)SEEP/W

0.05.0

10.015.020.025.030.035.040.0

0.00 0.40 0.80 1.20 1.60 2.00

Distance (m)

Espinoza et al. (1993)SEEP/W

30.0

35.0

40.0

onte

nt (%

) s

Page 64: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

49

Figure 4.24 Comparison between measured and predicted volumetric water content in base material at 4.5 hours

Figure 4.25 Comparison between measured and predicted volumetric water content

in base material at 6 hours

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

0.0 0.4 0.8 1.2 1.6 2.0

Distance (m)

Espinoza et al. (1993)SEEP/W

Page 65: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

50

Figure 4.26 Comparison between measured and predicted volumetric water content in subbase material at 1.5 hours

Figure 4.27 Comparison between measured and predicted volumetric water content in subbase material at 3 hours

5.0

10.0

15.0

20.0

25.0

0.0 0.4 0.8 1.2 1.6 2.0

Distance (m)

Espinoza et al. (1993)SEEP/W

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.0 0.4 0.8 1.2 1.6 2.0

Distance (m)

Espinoza et al. (1993)

SEEP/W

Page 66: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

51

Figure 4.28 Comparison between measured and predicted volumetric water content in subbase material at 4.5 hours

Figure 4.29 Comparison between measured and predicted volumetric water content in subbase material at 6 hours

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.0 0.4 0.8 1.2 1.6 2.0

Distance (m)

Espinoza et al. (1993)

SEEP/W

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.0 0.4 0.8 1.2 1.6 2.0

Distance (m)

Espinoza et al. (1993)

SEEP/W

Page 67: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

52

The effect of the subbase layer can be appreciated by comparing the changes in volumetric

water content in the base between Examples 2 and 3 (Figure 4.30 and 4.31). Due to the higher

hydraulic conductivity of the subbase, the base drains faster for the system described in

Example 3.

4.4 SUMMARY

In Example 1, SEEP/W predictions were compared to laboratory testing results to establish that SEEP/W could be used to model unsaturated flow under transient conditions. In Example 2, it was shown that SEEP/W could also be used to accurately model the infiltration process that pavement structures may go through during rainfall events. Finally, Example 3 showed that SEEP/W could also be used to model more complicated structures consisting of layers of soil. The SEEP/W results in Example 3 showed a good agreement with previous predictions by Espinoza et al. (29) using first order finite differences. The results from all three examples show that SEEP/W is a useful tool for the modeling of unsaturated flow through layered systems under complex boundary conditions and material characterization. However, it is important to correctly specify the initial and boundary condi-tions in order to accurately predict the unsaturated flow of water through materials. Therefore, material characterization (soil water characteristic curve and hydraulic conductivity curve), initial water table position, and geometry, are some important key conditions.

Figure 4.30 Comparison of measured volumetric water content in base material

at 6 hours for Examples 2 and 3

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

45.0

0.00 0.40 0.80 1.20 1.60 2.00Distance (m)

Vol

umet

ric

Wat

er C

onte

nt (%

)s

Example 2 -Espinoza et al. (1993)Example 3 - Espinoza et al. (1993)

Page 68: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

53

Figure 4.31 Comparison of predicted volumetric water content in base material at 6 hours for Examples 2 and 3

0.05.0

10.015.020.025.030.035.040.045.0

0.00 0.40 0.80 1.20 1.60 2.00

Distance (m)

Example 2 - SEEP/WExample 3 -SEEP/W

Page 69: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

54

CHAPTER 5

VERIFICATION OF RESULTS TO FIELD RESULTS FOR CELLS 33, 34, 35

In order to develop an understanding of the behavior of flow of water through flexible pave-

ments under unsaturated conditions, a finite element based numerical model of flow through

flexible pavements was constructed based on actual Mn/ROAD pavement geometries and

material characteristics, along with results from automated time domain reflectometry (TDR)

probes placed in the base layers of the sections studied. The measured TDR volumetric water

contents were used as benchmarks for adjusting laboratory-based unsaturated soil hydraulic

properties. The results show that measured field volumetric moisture contents can be matched

well with SEEP/W, through only minor adjustments in the unsaturated soil hydraulic

properties. In the following, the finite element model will be described, followed by a

discussion of the backcalculation of unsaturated soil hydraulic properties that finally resulted in

the matching of predicted and measured volumetric moisture contents.

5.1 PAVEMENT SYSTEM DESCRIPTION

The pavement systems that were used for this study are Mn/ROAD Cells 33, 34, 35. They all

consist of a layer of hot mix asphalt, a Mn/DOT Class 6 Special crushed granite base, and an

R-12 silty clay subgrade. Figure 5.1 shows the depth of each layer for these sections.

5.2 FINITE ELEMENT MODEL

Based on Figure 5.2, idealized finite element-based cross sections of Mn/ROAD cells 33, 34,

and 35 were constructed using SEEP/W. An overview will be provided of the construction of

the Finite Element Model. Figure 5.3 presents a section of the mesh utilized to model Cell 33,

34 and 35, that included 8046 quadrilateral and triangular elements.

5.2.1 Hot Mix Asphalt Layer

The hot mix asphalt layer was considered as an impervious material, therefore its

characteristics were not required as input for the Finite Element Model. This layer was not

Page 70: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

55

represented with

Page 71: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

56

Figure 5.1 Layer thicknesses for Mn/ROAD Cells 33, 34, and 35

Figure 5.2 Pavement geometry and dimensions for Cells 33, 34, and 35

12 in

Cell 33

4.04 in

12 in

Cell 34

3.92 in

12 in

Cell 35

3.96 in

HMA

Class 6 Special

1.83 m 4.27 m 4.27 m3.05 m

CL

0.3 m Class 6 Special 0.1 m Hot Mix Asphalt

R-12 silty clay

4:1 4:1

4.0 m

16.5 m

Page 72: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

57

Figure 5.3 Section of the Finite Element Model used to represent Cell 33, 34 and 35 finite elements. Only its geometry was taken into account. This assumption was made in order

to simplify the modeling process, since field measurements of infiltration through the hot mix

asphalt layer were not provided.

5.2.2 Base Course

The base layer was represented with 1926 quadrilateral and triangular finite elements. The

material properties, including the soil-water characteristic curve and hydraulic conductivity

curve corresponded to Class 6 Special material, discussed in Chapter 3.

5.2.3 Subgrade Soil

The subgrade was represented with a coarser mesh of finite elements. The finite element model

is composed of 6120 quadrilateral and triangular elements. The subgrade was extended laterally

10 m beyond the area covered by the asphalt and base layer (Figure 5.4), on each side, in order

Hot Mix Asphalt Base Finite

Subgrade Finite Elements

Page 73: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

58

Figure 5.4 Finite Element Model used to represent Cells 33, 34, and 35 to represent real conditions more accurately, and provide continuity to the extension of the

subgrade, for consistency with field conditions. The subgrade corresponds to silty clay,

classified as R-12, with the soil water characteristic curve and hydraulic conductivity curve

presented previously in Chapter 3.

5.2.4 Initial and Boundary Conditions

Applying a time-dependent flux boundary condition on the shoulder simulated infiltration due

to precipitation. The rain event selected for this simulation is presented later in the final

element analysis.

In Figure 5.3, the triangular nodes represent boundary conditions. The ones on top of the base

simulate rain events (time dependent flux conditions), and the ones underneath the hot mix

asphalt layer represent no infiltration through this material (q = 0).

An initial water table position was set as well, as part of the boundary conditions. The initial

water table was chosen so as to match the suction value in the pavement to obtain the initial

volumetric water contents of the measured data. In order to induce lateral and vertical drainage

in the system, points of total head equal to 0 m were applied at the bottom corners (subgrade

layer) of the model. Figure 5.4 shows the resulting Finite Element Model.

Extended Subgrade

H = 0 m

Extended Subgrade

H = 0 m

4.0 m

36.5 m

Page 74: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

59

5.2.5 Finite Element Model Analysis

To adequately model the non-steady unsaturated flow of water through pavements, a transient

analysis was performed, in which the pavement system at initial equilibrium was subjected to a

transient “rain event,” resulting in time-dependent changes in the volumetric moisture content

throughout the pavement system. Ideally, a detailed set of water table measurements at discrete

well points throughout the pavement cross-section would provide the initial head conditions for

the system. However, the only water table measurements available at the time this study was

performed are in the unpaved shoulder area, which is at an offset from the paved central part of

the pavement. These water table measurements should be expected to be somewhat different

than those obtained under the paved part of the pavement, due to the impervious nature of the

hot mix asphalt layer. Therefore, before starting the transient analysis, a steady-state analysis

was performed to obtain the initial head conditions for the system.

For the steady-state analysis, an initial total head value was set on the vertical sides of the sub-

grade to obtain a water table elevation that provided the required suction values to match the

initial moisture contents measured in the field. Therefore, a total head of 3.82 m was used for

that purpose. Material characterization, boundary conditions (rain event), geometry and mesh

configuration were not different from those used later on the transient analysis. After running

the steady-state analysis, a water table at 0.76 m depth was obtained, as measured from the top

of the pavement at the centerline cross section. The suction corresponding to this depth of the

water table resulted in a match with the initial volumetric water contents of the measured data.

The water table obtained from the steady-state analysis under the impervious paved part of the

pavement is higher than the measured water table in the unpaved shoulder. The water table in

the shoulder was measured at 2.98 m and 3.00 m depths on 2 August 2001 and 7 August 2001,

respectively. Water table measurements using field pressure transducers at Cell 34 under the

centerline of the pavement resulted in water table measurements at a depth of 1.57 m in August

2000. However, a detailed cross section of water table measurements for Cell 33 was not

available at the time the analysis for this report was performed.

Page 75: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

60

Finally, transient analysis was performed, in which actual rain events were infiltrated into the

pavement in a time-dependent fashion and the resulting time-dependent changes in heads and

volumetric water content were monitored throughout the pavement model.

The precipitation events input into the numerical model correspond to real measurements

gathered from July 9th through September 30th, 2000. The original data had units of inches per

day, but due to requirements of SEEP/W, it was converted to units of flux (m/s per m2), as

shown in Figure 5.5.

5.2.6 Measured Volumetric Water Contents

Automated TDR volumetric moisture content readings with time in the base material were

obtained at three different locations at each of the pavement sections studied. TDR probes are

located within the base layer at 0.13 m, 0.25 m, and 0.38 m depth (Figure 5.6), at a horizontal

offset of 1.83 m from the centerline. They are designated as locations 101, 102, and 103,

respectively for all the Cells. The corresponding locations of the TDRs in the Finite Element

Model are shown in Figure 5.7.

Figure 5.5 Precipitation events for Cells 33, 34, and 35

-5.0E-08

0.0E+00

5.0E-08

1.0E-07

1.5E-07

2.0E-07

2.5E-07

3.0E-07

3.5E-07

190 200 210 220 230 240 250 260 270

Time (Julian day)

Prec

ipita

tion

(m/s

)

Page 76: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

61

Figure 5.6 TDR installation at Mn/DOT Cells 33, 34 and 35

Figure 5.7 TDR location within Finite Element Model for Cells 33, 34, and 35

Hot Mix AsphaltBase

Subgrade Finite Elements

CL

Offset (-1.83 m)

Centerline

0.13 m0.25m 0.38m

0.10 m

0.30 m

3.6 m

Automated *TDR

HMA

Class 6 Special R-12 silty clay

Location 101Location 102Location 103

Page 77: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

62

To allow a more detailed and definitive comparison between measured and predicted results,

the span of time from July 31st to September 30th, 2000 was chosen for this study. This range

is 22 days shorter than the one used for the precipitation data. However, the first 22 days, from

July 9th to July 30th, were used to establish an initial volumetric water content baseline in the

model, similar to that observed in the field by July 31st, 2000.

Finally, Figures 5.8 to 5.10 show the measured TDR water contents for each cell.

5.3 MATERIALS CHARACTERIZATION ADJUSTMENT

As discussed, Figures 3.4 and 3.5 present the soil water characteristic curves and the hydraulic

conductivity curves for the base material. The hydraulic conductivity curve shows that Class 6

special is a material with a Ksat value equal to 1.54E-06 m/s. This corresponds to fine sand.

However, within the first 10 kPa the K value has dropped down to about 1E-13 m/s.

Figure 5.8 Measured data for different TDR locations at Cell 33

2.0

4.0

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

TDR Location 101TDR Location 102TDR Location 103

Page 78: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

63

Figure 5.9 Measured data for different TDR locations at Cell 34

Figure 5.10 Measured data for different TDR locations at Cell 35

0.0

5.0

10.0

15.0

20.0

25.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

TDR Location 101TDR Location 102TDR Location 103

0.0

5.0

10.0

15.0

20.0

25.0

210 220 230 240 250 260 270

Time ( Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

TDR Location 101TDR Location 102TDR Location 103

Page 79: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

64

The subgrade characterization is presented in Figures 3.6 and 3.7. The soil water characteristic

curve shows that this silty clay subgrade has a high water holding capacity, with a total change

is 15% in the volumetric moisture content over the range of measured suction. The hydraulic

conductivity curve for the subgrade soil also shows that this material has a high resistance to

drainage. Within the first 5 kPa, this material goes from a permeability of 2.75E-08 m/s to a

value close to 2E-13 m/s.

The laboratory-based soil water characteristic curves resulted in predicted volumetric water

contents that were slightly different from the measured field volumetric moisture contents.

Therefore, it was necessary to adjust or “tune” the measured soil water characteristic curves to

fit the measured TDR data from the field. Another reason for the differences in predicted and

measured volumetric moisture contents may be that the TDR volumetric moisture contents may

have a plus or minus 1 to 2 percent differences in volumetric water contents for the same

material, simply due to the limitations of the TDR technology and the required empirical

calibration of results.

The results from the original TDR data are shown, followed by calibration to field results.

Adjusting the air entry values slightly and changing the slope of the soil water characteristic

curves easily achieved the calibration to field results. Finally, the saturated volumetric water

content was adjusted slightly to reflect the variability in the TDR results, due to their limits of

accuracy.

Once calibrated, soil water characteristic curves were obtained, and SEEP/W was used to gen-

erate the appropriate hydraulic conductivity curves, using the Green and Corey (45) approach.

Since the geometry and material properties were the same for Cells 33, 34, and 35, Cell 33 was

chosen for the “tuning” or backcalculation of the soil water characteristic curve and the

resulting hydraulic conductivity curves.

For simplicity during the initial calibration, the infiltration was assumed to be 100 percent of

the measured precipitation. Once the soil water characteristic curve and hydraulic conductivity

Page 80: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

65

curves were “tuned,” the infiltration rate was decreased to a percentage of the measured pre-

cipitation as a part of the final calibration to field results. Since it is unlikely that all of the

measured precipitation will infiltrate into the pavement system, only a percentage of the total

infiltration was assumed to infiltrate the pavement.

5.3.1 Initial Calibration Results

The initial calibration results were obtained from the original (uncalibrated) soil water charac-

teristic curve and hydraulic conductivity curve results for Cell 33. Figures 5.11 through 5.13

show the predicted versus measured volumetric water content results. No change was observed

in the volumetric water content, not even during the application of the highest precipitation

events. For all three locations (101, 102, 103), the predicted volumetric water content remained

constant at 32 percent, as shown in Figures 5.11 through 5.13.

5.3.2 Second Calibration Results

In this part of the study, the air entry potential of the soil water characteristic curve for Cell 33

was adjusted to better represent likely field density conditions. Since the density in the field is

Figure 5.11 Volumetric water content at Cell 33 – Location 101

0.0

4.0

8.0

12.0

16.0

20.0

24.0

28.0

32.0

36.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

MeasuredPredicted

24 0

28.0

32.0

36.0

onte

nt (%

) s

Page 81: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

66

Figure 5.12 Volumetric water content at Cell 33 – Location 102

Figure 5.13 Volumetric water content at Cell 33 – Location 103

4.0

8.0

12.0

16.0

20.0

24.0

28.0

32.0

36.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

Measured

Predicted

Page 82: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

67

likely to be slightly different from that in the laboratory, due to the differences in compaction processes, a small variation in the air entry value of the soil water characteristic curve in the unsaturated region can be assumed (52). Based on a recommendation from Mn/DOT (52), a new air entry value of 3kPa was used. Subsequently, SEEP/W was used to obtain a corresponding hydraulic conductivity curve. The Ksat value used was the same as the original one, 1.54E-06 m/s. The new characterization curves are presented in Figures 5.14 and 5.15. The modifications didn’t change the initial results significantly. The volumetric water content stayed at 32%. Hence, it was decided that some more modifications were needed in order to see a predicted response with the same trends as the observed field response. For this reason, the slopes of the soil water characteristic curve for the Class 6 Special base material (Figure 5.16) were changed slightly. In order to maintain all other aspects of the soil water characteristic

curve, this change in slope resulted in a change in saturated volumetric moisture content (θsat)

of 2 percent. Subsequently, SEEP/W was used to estimate the corresponding hydraulic conduc-tivity curve, shown in Figure 5.17.

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 1.00 100.00

Suction (kPa)

Figure 5.14 Soil water characteristic curve (air entry = 3 kPa) – Base material (Class 6 special)

Page 83: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

68

1.0E-12

1.0E-10

1.0E-08

1.0E-06

1.0E-04

0.0 1.0 100.0

Suction (kPa)

k (m

/s)

Figure 5.15 Estimated hydraulic conductivity curve – Base material (Class 6 special)

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 1.00 100.00

Suction (kPa)

Figure 5.16 Final soil water characteristic curve – Base material (Class 6 special)

Page 84: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

69

Figure 5.17 Estimated hydraulic conductivity curve – Base material (Class 6 special)

Figures 5.18 to 5.20 show the resulting Finite Element Model predictions using the updated soil

water characteristic curve and hydraulic conductivity curve. The results now show a stronger

response that is more consistent with the observed field results. However, the post-rain event

part of the volumetric water content curves tends to be higher than the observed field response,

implying a resistance to drainage in the numerical model that is not present in the field.

The resistance to vertical flow is heavily determined by the properties of the subgrade soil.

Hence, the Ksat value for the subgrade was increased by one order of magnitude (i.e. 10 times),

and the hydraulic conductivity curve was re-estimated using SEEP/W with the existing soil

water characteristic curve and the new Ksat value (Figure 5.21).

Figures 5.22 through 5.24 show the resulting Finite Element Model predictions for TDR 101,

102, and 103. Subsequently, the process was repeated, until correspondence with field results

was achieved.

The final Ksat value for the subgrade was taken as Ksat = 2.7535E-6 m/s (Figure 5.25).

1.0E-11

1.0E-10

1.0E-09

1.0E-08

1.0E-07

1.0E-06

1.0E-05

0.0 1.0 100.0

Suction (kPa)

k (m

/s)

Page 85: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

70

0.0

4.0

8.0

12.0

16.0

20.0

24.0

28.0

32.0

210 220 230 240 250 260 270

Time (Julian day)

Measured

Predicted

Figure 5.18 Volumetric water content at Cell 33 – Location 101

6.0

10.0

14.0

18.0

22.0

26.0

30.0

34.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.19 Volumetric water content at Cell 33 – Location 102

Page 86: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

71

6.0

10.0

14.0

18.0

22.0

26.0

30.0

34.0

210 220 230 240 250 260 270

Time (days)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

MeasuredPredicted

Figure 5.20 Volumetric water content at Cell 33 – Location 103

1.0E-10

1.0E-09

1.0E-08

1.0E-07

1.0E-06

0.01 1.00 100.00

Suction (kPa)

k (m

/s)

Figure 5.21 Modified hydraulic conductivity curve: 10 times Ksat – Subgrade material (R-12 Silty clay)

Page 87: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

72

0.0

4.0

8.0

12.0

16.0

20.0

24.0

28.0

32.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

MeasuredPredicted

Figure 5.22 Volumetric water content at Cell 33 – Location 101

6.0

10.0

14.0

18.0

22.0

26.0

30.0

34.0

210 220 230 240 250 260 270Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.23 Volumetric water content at Cell 33 – Location 102

Page 88: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

73

6.0

10.0

14.0

18.0

22.0

26.0

30.0

34.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.24 Volumetric water content at Cell 33 – Location 103

1.0E-10

1.0E-08

1.0E-06

1.0E-04

0.01 1.00 100.00

Suction (kPa)

Figure 5.25 Final hydraulic conductivity curve: Ksat = 2.7535E-06 m/s – Subgrade material

(R-12 Silty clay)

Page 89: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

74

Figures 5.26 through 5.28 show the predicted Finite Element Model results. Now, the overall

shape of the trends in the predicted and measured results at TDR at locations 101 and 103

(Figures 5.26 and 5.28) show close correspondence with each other. It is important to keep in

mind that no attempt has still been made to adjust the actual infiltration rate, hence the

predicted results should show a large response due to a rain event. However, the Finite

Element Model predictions for TDR location 102 (Figure 5.27) are still slightly different from

the field results.

Finally, all the water coming from precipitation events is not going to infiltrate the soil. Hence,

an adjustment is needed for the infiltration function that represents these events in the Finite

Element Model. For the first 22 days, 70% of the precipitation was applied, simply to achieve

the initial measured field values within a reasonable amount of time. Subsequently, starting

with day 23, that percentage was reduced to 30 percent, which seemed to result in a good

correspondence with observed field values, as shown in Figures 5.29 through 5.31.

It can be see that for all three locations the trend in the predicted Finite Element Model results

now follows that of the measured data. In particular, TDR location 101 predicted Finite

Element Model volumetric moisture content values correspond closely with the field data.

6.0

7.0

8.0

9.0

10.0

11.0

12.0

13.0

14.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.26 Volumetric water content at Cell 33 – Location 101

Page 90: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

75

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

22.0

24.0

26.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.27 Volumetric water content at Cell 33 – Location 102

6.0

10.0

14.0

18.0

22.0

26.0

30.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.28 Volumetric water content at Cell 33 – Location 103

Page 91: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

76

6.0

7.0

8.0

9.0

10.0

11.0

12.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.29 Volumetric water content at Cell 33 – Location 101

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

210 220 230 240 250 260 270

Time (Julian day)

Measured

Predicted

Figure 5.30 Volumetric water content at Cell 33 – Location 102

Page 92: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

77

8.0

12.0

16.0

20.0

24.0

28.0

210 220 230 240 250 260 270

Time (Julian day)

Measured

Predicted

Figure 5.31 Volumetric water content at Cell 33 – Location 103

5.3.3 Final Matching of Predicted and Measured Volumetric Water Contents

To obtain a final fit to measured TDR volumetric moisture contents for TDR location 102, the

saturated volumetric moisture content was increased by 3 percent (Figure 5.32), and the

hydraulic conductivity curve was re-estimated using the adjusted soil water characteristic

curve. Due to the small percentage change in the volumetric moisture content, the estimated

hydraulic conductivity curve did not change from the previous one.

As expected, the slight increase in volumetric water content improved the correspondence

between the predicted Finite Element Model and the measured volumetric moisture contents

for at TDR location 102 (Figure 5.33). Using the same approach, the saturated volumetric

moisture content for the Class 6 material around TDR location 103 was reduced, resulting in

the soil water characteristic curve shown in Figure 5.34.

The resulting Finite Element Model prediction shown in Figure 5.35 now shows a close match

with the field results.

Page 93: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

78

0.0

10.0

20.0

30.0

40.0

0.01 1.00 100.00

Suction (kPa)

Figure 5.32 Final soil water characteristic curve for Location 102 – Base material

(Class 6 special)

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

22.0

24.0

210 220 230 240 250 260 270

Time (Julian day)

Measured

Predicted

Figure 5.33 Volumetric water content at Cell 33 – Location 102

Page 94: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

79

0.0

5.0

10.0

15.0

20.0

25.0

30.0

0.01 1.00 100.00

Suction (kPa)

Figure 5.34 Final soil water characteristic curve for Location 103 – Base material

(Class 6 special)

8.0

10.0

12.0

14.0

16.0

18.0

20.0

22.0

24.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.35 Volumetric water content at Cell 33 – Location 103

Page 95: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

80

Tables 5.1 and 5.2 summarize the changes made in the saturated volumetric water content and

the saturated hydraulic conductivity to match field conditions, and Table 5.3 summarizes the

changes made between SEEP/W runs for all cases studied. In summary, it has to be concluded

that minor changes in the volumetric moisture contents, saturated hydraulic conductivity of the

subgrade soil, air entry value and slope of the soil water characteristic curve, with subsequent

adjustments in the hydraulic conductivity curves were needed to match the field data.

5.4 CELL 34

The adjustments in the unsaturated hydraulic properties obtained for Cell 33 were applied for

TDR locations 101, 102, and 103 in Cell 34.

As expected, Figures 5.36 through 5.38 show that the predicted finite element model results

match the shape for the measured field volumetric moisture contents well.

Table 5.1 Cell 33 – Calibration for base layer (Class 6 special)

Final Adjusted Values Parameters Initial data

Location 101 Location 102 Location 103

θ sat (%) 32.02 30.19 33.37 26.97

k sat (m/s) 1.5479E-06 1.5479E-06 1.5479E-06 1.5479E-06

Table 5.2 Cell 33 – Calibration for subgrade soil (R-12 Silty clay)

Parameters Initial data Final Calibration

θ sat (%) 49 49

k sat (m/s) 2.7535E-08 2.7535E-06

Page 96: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

81

Table 5.3 Summary of SEEP/W runs performed to adjust predicted volumetric moisture contents to TDR measured values at Cell 33

Parameter Baseline

Run 1 (Adjustment in Air Entry

Value)

Run 2 (Adjustment in Slope

of Soil Water Characteristic

Curve)

Run 3 (Adjustment in Subgrade

Ksat)

Run 4 (Adjustment in

Precipitation and Predictions at TDR Location

101)

Run 5 (Adjustment in Predictions at TDR Location

102)

Run 6 (Adjustment is Predictions

at TDR Location 103)

Base

Air entry

(kPa)

0 3 3 3 3 3 3

Initial Slope 0.24 0.24 0.08 0.08 0.08 0.08 0.08

Subgrade

Ksat (m/s) 2.75E-08 2.75E-08 2.75E-08 2.75E-06 2.75E-06 2.75E-06 2.75E-06

Precipitatio

n (%) 100 100 100 100 70 70 70

θ sat (%) 32.02 32.02 30.19 30.19 30.19 33.37 26.97

Page 97: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

81

6.0

7.0

8.0

9.0

10.0

11.0

12.0

13.0

14.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.36 Volumetric water content at Cell 34 – Location 101

4.0

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.37 Volumetric water content at Cell 34 – Location 102

Page 98: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

83

8.0

12.0

16.0

20.0

24.0

28.0

32.0

36.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.38 Volumetric water content at Cell 34 – Location 103

5.4.1 Calibration of Results at Locations 101 and 102

Similar to Cell 33, a slight shift in the predicted and measured volumetric water content trends

was observed, which was subsequently adjusted by shifting the soil water characteristic curve

by an increment of 3 percent in the same manner as described previously for Cell 33. Figure

5.39 shows the updated soil water characteristic curve. The hydraulic conductivity curve

stayed unchanged due to the small change in the soil water characteristic curve.

The resulting comparison between the predicted finite element model and measured volumetric

moisture contents for TDR locations 101 and 102 is shown in Figures 5.40 and 5.41. In both

cases, a good match between predicted and measured TDR moisture contents was obtained.

5.4.2 Calibration to Location 103

Finally, for TDR location 103, a nominal reduction of 0.5 percent in the soil water

characteristic curve saturated volumetric moisture content was needed to match the field data.

Page 99: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

84

Figure 5.42 shows the new soil water characteristic curve for this location at cell 34, and Figure

5.43 shows

Page 100: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

85

Figure 5.39 Final soil water characteristic curve for Location 101 and 102 at Cell 34 –

Base material (Class 6 special)

11.0

11.5

12.0

12.5

13.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.40 Volumetric water content at Cell 34 – Location 101

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 1.00 100.00

Suction (kPa)

Page 101: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

86

10.0

12.0

14.0

16.0

18.0

210 220 230 240 250 260 270

Time (Julian day)

MeasuredPredicted

Figure 5.41 Volumetric water content at Cell 34 – Location 102

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 1.00 100.00Suction (kPa)

Figure 5.42 Final soil water characteristic curve for Location 103 Cell 34 – Base material (Class 6 special)

Page 102: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

87

8.0

12.0

16.0

20.0

24.0

28.0

32.0

36.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.43 Volumetric water content at Cell 34 – Location 103

the resulting overall close comparison between the predicted and measured moisture contents

for TDR location 103.

However, small abnormalities at days 231, 249, and 266 in the measured TDR data were not

captured in the finite element model. These small variations in the measured TDR data might

be due to a problem with the TDRs or a transient problem with the automated data acquisition

system.

Tables 5.4 summarizes the modifications made to the base material to achieve calibration for

Cell 34, and Table 5.5 summarizes the changes made between SEEP/W runs for all cases

studied.

5.5 CELL 35

The same approach that was used for Cells 33 and 34 to calibrate the numerical model to the

field results was used for Cell 35. The same adjustments to material properties (soil water

char-

Page 103: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

88

Page 104: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

89

Table 5.4 Cell 34 – Calibration for base layer (Class 6 special)

Final calibration Parameters Initial data

Location 101 Location 102 Location 103

θ sat (%) 32.02 33.10 33.10 29.70

k sat (m/s) 1.5479E-06 1.5479E-06 1.5479E-06 1.5479E-06

Table 5.5 Summary of runs to calibrate Cell 34

Parameter Baseline

Run 1 (Unsaturated

Soil Hydraulic Properties

from Cell 33)

Run 2 (Adjustment in Predictions at TDR Location

102)

Run 3 (Adjustment in Predictions at TDR Location

103) Base

Air entry (kPa) 0 3 3 3

Initial Slope 0.24 0.08 0.08 0.08

Subgrade

Ksat (m/s) 2.75E-08 2.75E-06 2.75E-08 2.75E-06

Precipitation (%) 100 70 70 70

θ sat (%) 32.02 30.19 33.10 29.70

acteristic curve and hydraulic conductivity curve) that were used for Cell 34 resulted in rea-

sonable comparisons of predicted and measured volumetric moisture contents for TDR

locations 101 and 103, as shown in Figures 5.44 and 5.45.

For location 102 a small change (2 percent) in the saturated volumetric moisture content was

needed to adjust the results to density variations around the TDRs, with the resulting soil water

characteristic curve shown in Figure 5.46.

The predictions obtained using calibrated data for TDR location 102 are shown in Figure 5.47.

Page 105: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

90

8.0

9.0

10.0

11.0

12.0

13.0

14.0

15.0

16.0

17.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

Measured

Predicted

Figure 5.44 Volumetric water content at Cell 35 – Location 101

6.0

8.0

10.012.0

14.0

16.0

18.0

20.022.0

24.0

26.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Measured

Predicted

Figure 5.45 Volumetric water content at Cell 35 – Location 103

Page 106: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

91

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 1.00 100.00Suction (kPa)

Figure 5.46 Final soil water characteristic curve for Location 102 Cell 35 – Base material (Class 6 special)

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) F

Measured

Predicted

Figure 5.47 Volumetric water content at Cell 35 – Location 102

Page 107: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

92

Table 5.6 summarizes the modifications made to the base material for Cell 35 to calibrate the

predicted Finite Element Model results to the observed TDR volumetric moisture contents, and

Table 5.7 summarizes the changes made between SEEP/W runs for all cases studied.

Table 5.6 Cell 35 – Calibration for base layer (Class 6 special)

Final calibration Parameters Initial data

Location 101 Location 102 Location 103

q sat (%) 32.02 33.10 32.32 29.70

k sat (m/s) 1.5479E-06 1.5479E-06 1.5479E-06 1.5479E-06

Table 5.7 Summary of runs to calibrate Cell 35

Parameter Baseline

Run 1 (Unsaturated Soil

Hydraulic Properties from Cell 33)

Run 2 (Adjustment in Predictions at

TDR Location 102) Base Air

Entry (kPa) 0 3 3

Initial Slope 0.24 0.08 0.08

Subgrade

Ksat (m/s) 2.75E-08 2.75E-06 2.75E-06

Precipitation

(%) 100 70 100

θ sat (%) 32.02 32.32 29.70

Page 108: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

93

5.6 SUMMARY

In this Chapter, a finite element based numerical model of flow through flexible pavements was

constructed based on actual Mn/ROAD pavement geometries and material characteristics.

Measured TDR volumetric water contents in the Mn/DOT Class 6 Special crushed granite layer

were used as benchmarks for adjusting laboratory-based unsaturated soil hydraulic properties.

The results show that measured field volumetric moisture contents can be matched well with

SEEP/W, through only minor adjustments in the unsaturated soil hydraulic properties. In

particular, the results obtained show that SEEP/W is a valuable tool for modeling the

unsaturated flow of water through pavement systems, and can be used to predict results that

compare favorably to measured field results. It is not surprising that some adjustment in

unsaturated hydraulic soil properties was required to obtain a match between predicted and

measured TDR volumetric moisture contents. For example, the variability in TDR readings

alone may result in differences in volumetric moisture contents of up to plus or minus 1 to 2

percent. Other factors may also potentially influence the results. These factors include the

possibility of slight differences in the soil packing arrangement between the field and the

laboratory due to differences in field and laboratory compaction processes, and the potential

presence of residual stresses and anisotropic stress conditions in the field. However, due to

lack of data at this point in time, it is premature to speculate about the presence and effects of

these various on measured unsaturated hydraulic properties of granular base material.

The numerical results presented in this Chapter show the importance of considering that only a

part of the water coming from precipitation events is going to infiltrate the pavement. Hence,

an adjustment is needed for the infiltration function that represents these events in the finite

element model. Based on the adjusted numerical predictions and comparison to measured field

volumetric water contents, it appears that about 30 percent of the precipitation made its way

into the pavement system, meaning that the presence of a flexible impervious pavement surface

does not preclude water from entering into the base and possibly affecting the modulus and

strength of the base material.

Page 109: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

94

CHAPTER 6

PARAMETRIC STUDY Chapters 4 and 5 showed that SEEP/W can be used to model unsaturated flow through layered

pavement systems under complex boundary conditions and material characterization. In this

Chapter, the sensitivity of predicted volumetric moisture contents due to variations in SEEP/W

input parameters and unsaturated soil hydraulic properties are evaluated. The key parameters

evaluated are: 1) the initial slope of the soil water characteristic curve, 2) the air entry value of

the Mn/DOT Class 6 Special crushed granite, 3) the saturated hydraulic conductivity of the

Mn/DOT Class 6 Special crushed granite, 4) the air entry value of the R-12 silty clay subgrade

soil, 5) the saturated hydraulic conductivity for the subgrade soil, 6) the type of granular base

material, 7) variations in infiltration characteristics, and 8) the water table location. Since Cells

33-35 are similar, except for the composition of the hot mix asphalt layer, Cell 33 was selected

as a representative pavement configuration, with TDR location 101 being typical for

unsaturated flow of water, due to its high elevation within the base material. The original

conditions for this system were presented in Chapters 3 and 5, Figures 5.16 and 5.17 for the

base layer, and Figures 3.6 and 5.25 for the subgrade.

6.1 INITIAL SLOPE OF THE BASE MATERIAL SOIL WATER CHARACTERISTIC CURVE

Variation of the characteristics of the soil water characteristic curve can lead to significant

variation in the predicted volumetric water content. The first characteristic to be studied was

the initial slope of the soil water characteristic curve. The slope dictates the rate at which volu-

metric water content can change during effective saturation conditions (i.e. before passing the

air entry potential). Three cases were considered: 1) no change, meaning that the original slope

is described by θ sat = 30.19% and a drop of 4.6% up to the air entry value (3 kPa), 2) con-

sidering the initial slope equal to zero (0), taking the air entry volumetric moisture content

(25.59 percent) as representative of all effectively saturated conditions (i.e. with suction values

lower than the 3 kPa air entry value), and 3) considering the slope as being twice as steep as

that for Case 1, resulting in θ sat = 34.79%, and a drop of 9% between 0 kPa and 3 kPa suction

values. Figure 6.1 shows the resulting soil water characteristic curves for the three cases.

Page 110: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

95

Figure 6.1 Soil water characteristic curves for initial slope cases

Due to the changes applied to the soil water characteristic curve, the hydraulic conductivity

curve is also affected. Figure 6.2 presents the new curves for these cases, as estimated from

SEEP/W. Although faster drainage could be expected with steeper slopes, the hydraulic con-

ductivity also affects the water flow. The new estimated hydraulic conductivity curves show

that case 1 would likely drain quicker than case 3, which has a steeper slope at the soil water

characteristic curve.

The resulting Finite Element Model predictions are shown in Figure 6.3. It is apparent from

the results that the initial slope of the soil water characteristic curve effects only slightly how

fast the material can drain. Case 1 shows that the base would have a rate of drainage of 0.13 %

within the first 17 days, Case 2 a rate of 0.14%, and Case 3 a rate of 0.09%, showing that the

results are not significantly affected by the change in initial slope of the soil water

characteristic curve.

0.0

10.0

20.0

30.0

40.0

0.01 1.00 100.00

Suction (kPa)

Case 1 - Slope 4.6%Case 2 - Slope 0 %Case 3 - Slope 9 %

Page 111: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

96

Figure 6.2 Hydraulic conductivity curves for initial slope cases

8.9

9.0

9.1

9.2

9.3

9.4

9.5

9.6

9.7

9.8

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Predicted - Case 1 (Slope = 4.6%)

Predicted - Case 2 (Slope = 0%)

Predicted - Case 3 (Slope = 9%)

Figure 6.3 Results for initial slope cases

1.0E-11

1.0E-09

1.0E-07

1.0E-05

0.0 1.0 100.0

Suction (kPa)

k (m

/s)

Case 1 - Slope 4.6%Case 2 - Slope 0 %Case 3 - Slope 9%

Page 112: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

97

Table 6.1 summarizes the effects, in terms of the maximum volumetric water content, of the

change in the initial slope of the soil water characteristic curve for the base material. Overall it

can be concluded that the predicted volumetric moisture content is relatively insensitive to

changes in the in the initial slope of the soil water characteristic curve.

Table 6.1 Effects of the initial slope of the base material soil water characteristic curve on maximum predicted volumetric moisture content

Parameter Case 1 Case 2 Case 3

Initial slope – Base material (%)

4.6 0 9

Maximum volumetric water content (%)

9.67 9.69 9.65

6.2 AIR ENTRY VALUE OF BASE MATERIAL

Another important parameter to evaluate is the air entry value. It represents the transition suc-

tion value between unsaturated and saturated conditions. Figure 6.4 presents three cases consid-

ered for the soil water characteristic curve with different air entry values. In Case 1, the air

entry value is 3 kPa, which is the original case, in Case 2, the air entry value is 4 kPa, and

finally in Case 3, the air entry value is set to 5 kPa.

As expected, the higher the air entry value the longer the material will retain water. Therefore,

higher volumetric water contents are observed for Case 3 than for Cases 1 and 2, as shown in

Figure 6.5.

Table 6.2 summarizes the effects, in terms of the maximum volumetric water content, of the

change in the air entry value of the soil water characteristic curve for the base material. The

results show that the predicted volumetric moisture content is sensitive to changes in the air

entry value, resulting in almost a doubling of the predicted maximum volumetric water content

from 12.20 percent to 23.41 percent, as the air entry value is increased from 4 kPa to 5 kPa.

The results show the importance of an accurate determination of air entry values for modeling

purposes.

Page 113: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

98

Figure 6.4 Base soil water characteristic curves for air entry value cases

0.0

4.0

8.0

12.0

16.0

20.0

24.0

210 220 230 240 250 260 270

Time (Julian day)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Predicted - Case 1 (Air Entry = 3 kPa)Predicted - Case 2 (Air Entry = 4 kPa)Predicted - Case 3 (Air Entry = 5 kPa)

Figure 6.5 Results for air entry value cases at base layer

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 1.00 100.00Suction (kPa)

Case 1 - Air Entry = 3 kPaCase 2 - Air Entry = 4 kPaCase 3 - Air Entry = 5 kPa

Page 114: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

99

Table 6.2 Effects of air entry value of Mn/DOT Class 6 special crushed granite base material on maximum predicted volumetric moisture content

Parameter Case 1 Case 2 Case 3

Air entry value – Base

material (kPa) 3 4 5

Maximum volumetric

water content (%) 9.67 12.20 23.41

6.3 EFFECTS OF KSAT OF Mn/DOT CLASS 6 SPECIAL CRUSHED GRANITE BASE MATERIAL Another important parameter is the saturated hydraulic conductivity of the base material (Ksat).

Three different cases were considered: 1) Ksat = 1.55E-6 m/s (original case), 2) a Ksat which

was 10 times larger than the original value (1.55E-5 m/s), and 3) a Ksat which was 100 times

larger than the original value (1.55E-4 m/s). Again, the resulting hydraulic conductivity curves

shown in Figure 6.6 were estimated from the existing soil water characteristic curve and the

Ksat values studies using SEEP/W.

Figure 6.7 shows the resulting Finite Element Model predictions. The higher the hydraulic

conductivity the faster the material will drain. Also, Case 3 with the highest Ksat value, results

in much more abrupt changes in the predicted volumetric moisture content than for the other

two cases with lower Ksat values. Table 6.3 summarizes the effects, in terms of the maximum

volumetric water content, of the change in the Ksat for the base material. The overall change in

the predicted volumetric moisture content from 9.67 percent to 9.87 percent as the hydraulic

conductivity is increased by two orders of magnitude is only about 2.1 percent, meaning that

overall the predicted volumetric moisture content is fairly insensitive to large variations in

saturated hydraulic conductivity.

Page 115: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

100

Figure 6.6 Hydraulic conductivity curves for Ksat cases at base material

9.209.259.309.359.409.459.509.559.609.659.709.759.809.859.90

210 220 230 240 250 260 270

Time (Julian day)

Predicted - Case 1 ( Ksat = 1.55 E-06 m/s)

Predicted - Case 2 ( Ksat = 1.55 E-05 m/s)

Predicted - Case 3 ( Ksat = 1.55 E-04 m/s)

Figure 6.7 Results for Ksat cases at base layer

1.0E-11

1.0E-09

1.0E-07

1.0E-05

1.0E-03

0.0 1.0 100.0Suction (kPa)

Case 1 - Ksat = 1.55 E-06 m/sCase 2 - Ksat = 1.55 E-05 m/sCase 3 - Ksat = 1.55 E-04 m/s

Page 116: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

101

Table 6.3 Effects of Ksat of Mn/DOT Class 6 special crushed granite base material on maximum predicted volumetric moisture content

Parameter Case 1 Case 2 Case 3

Ksat – Base material

(m/s) 1.55 E-06 1.55 E-05 1.55 E-04

Maximum volumetric

water content (%) 9.67 9.79 9.87

6.4 AIR ENTRY VALUE OF SUBGRADE MATERIAL

Three distinct air entry values for the R-12 silty clay subgrade soil were considered, namely

Case 1: 0 kPa, which is the original case, Case 2: 5 kPa, and Case 3: 10 kPa. Figure 6.8 shows

the resulting Finite Element Model predictions. Interestingly, the base material drainage

wouldn’t be affected significantly by shifting the air entry values of the soil water characteristic

curve. Case 2 and 3 resulted in only small changes in the volumetric water content.

Table 6.4 summarizes the effects, in terms of the maximum volumetric water content, of the

change in the air entry value of the soil water characteristic curve for the subgrade material. In

summary, the results show that the maximum predicted volumetric moisture content as TDR

location 101 is insensitive to variations in the air entry value of the R-12 silty clay subgrade

soil.

6.5 Ksat AT SUBGRADE MATERIAL

In order to evaluate the effects of resistance to drainage through the base/subgrade interface

and into the subgrade soil, the effects of the subgrade Ksat were evaluated. Figure 6.9 presents

the hydraulic conductivity curves considered for the three cases studied. In Case 1: Ksat =

2.7535E-8 m/s, in Case 2: Ksat = 2.7535E-7 m/s, and finally in Case 3: Ksat = 2.7535E-6 m/s

(original case).

Page 117: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

102

9.0

9.1

9.2

9.3

9.4

9.5

9.6

9.7

9.8

210 220 230 240 250 260 270

Time (days)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Predicted - Case 1 (Air Entry = 0 kPa)Predicted - Case 2 (Air Entry = 5 kPa)Predicted - Case 3 (Air Entry = 10 kPa)

Figure 6.8 Results for air entry values at subgrade layer

Table 6.4 Effects of air entry value of R-12 silty clay subgrade soil on maximum predicted volumetric moisture content

Parameter Case 1 Case 2 Case 3

Air entry value –

Subgrade (kPa) 0 5 10

Maximum volumetric

water content (%) 9.67 9.71 9.71

Page 118: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

103

Figure 6.9 Hydraulic conductivity curves for Ksat cases at subgrade material

Figure 6.10 shows the resulting finite element predictions of volumetric water content at TDR

location 101. As expected, the results show that the higher the subgrade Ksat value, the faster

the drainage. Table 6.5 summarizes the effects, in terms of the maximum volumetric water

content, of the change in the Ksat for the subgrade. As the saturated hydraulic conductivity of

the subgrade soil is increased from 2.7535E-8 m/s in Case 1 to 2.7535E-6 m/s in Case 3, the

maximum predicted volumetric moisture contents at TDR location 101 changes from 10.25

percent to 9.67 percent, which corresponds to a change of about 5.6 percent. Hence, as the

Ksat for the subgrade soil is increased by two orders of magnitude, the resulting predicted

volumetric moisture content increases only by about 5.6 percent, implying a moderate

sensitivity to the Ksat of the subgrade soil.

6.6 EFFECTS OF THE TYPE OF BASE MATERIAL

To evaluate the effects of the type and gradation of granular base course material, Finite

Element Model predictions using Mn/DOT Class 3 Special, Class 4 Special, and Class 5

Special materials were compared to those for Class 6 Special. Figure 6.11 presents the soil

water characteristic curve for these new base materials.

1.0E-13

1.0E-11

1.0E-09

1.0E-07

1.0E-05

0.01 0.10 1.00 10.00 100.00Suction (kPa)

Case 1 - K sat = 2.75 E-08 m/sCase 2 - K sat = 2.75 E-07 m/sCase 3 - K sat = 2.75 E-06 m/s

Page 119: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

104

6.06.57.07.58.08.59.09.5

10.010.511.011.512.0

210 220 230 240 250 260 270

Time ( Julian day)

Predicted - Case 1 (K sat = 2.75 E-08 m/s)Predicted - Case 2 (K sat = 2.75 E-07 m/s)Predicted - Case 3 (K sat = 2.75 E-06 m/s)

Figure 6.10 Results for Ksat cases at subgrade layer

Table 6.5 Effects of Ksat of R-12 silty clay subgrade soil on maximum predicted

volumetric moisture content

Parameter Case 1 Case 2 Case 3

Ksat – Subgrade

material (m/s) 2.75 E-08 2.75 E-07 2.75 E-06

Maximum volumetric

water content (%) 10.25 10.20 9.67

Page 120: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

105

0.0

5.0

10.0

15.0

20.0

25.0

0.01 1 100 10000

Suction (kPa)

Class 3 specialClass 4 specialClass 5 special

Figure 6.11 Soil water characteristic curves for Mn/DOT aggregate base materials

Since the air entry value for these materials was unknown, they were all assigned the same air

entry value as the Mn/DOT Class 6 Special, namely 3.0 kPa. SEEP/W was used to estimate the

hydraulic conductivity curves for all the granular base course materials. Figures 6.12 and 6.13

show the resulting soil water characteristic curves and hydraulic conductivity curves for the Cl.

3 Sp., 4 Sp., and 5 Sp. base materials.

Figure 6.14 shows the resulting Finite Element Model predictions of the volumetric moisture

content at TDR location 101. The Cl 3 Sp, Cl 4 Sp, and Cl 5 Sp. base materials show higher

equilibrium water content than the Class 6 Special, and therefore less drainage than the Class 6

Special. This may be due to the fact that their soil water characteristic curves do not have

slopes that are steep enough to induce quick change in the volumetric water content for a small

range of suction values.

Interestingly, although Class 4 has the highest ksat value, it results in the highest volumetric

water contents, since its saturated volumetric water content θsat (24.45 %) is the highest one.

Page 121: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

106

The Mn/DOT Cl. 3 Sp., Cl. 4 Sp., and Cl. 5 Sp. granular base materials all remain close to their

θsat

Figure 6.12 Soil water characteristic curves (air entry = 3 kPa) for base materials

Figure 6.13 Estimated hydraulic conductivity curves for base materials

0.0

5.0

10.0

15.0

20.0

25.0

30.0

0.01 1 100

Suction (kPa)

Class 3 specialClass 4 specialClass 5 special

1.0E-12

1.0E-10

1.0E-08

1.0E-06

0.01 1.00 100.00Suction (kPa)

Class 3 specialClass 4 specialClass 5 special

Page 122: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

107

2.04.06.08.0

10.012.014.016.018.020.022.024.026.0

210 220 230 240 250 260 270

Time (Julian days)

Vol

umet

ric

Wat

er C

onte

nt (%

) f

Predicted - Class 6Predicted - Class 3Predicted - Class 4Predicted - Class 5

Figure 6.14 Results for different types of base materials

value. On the other hand, the Mn/DOT Class 6 Special drains fast enough to ensure that the

base course never reaches the θsat of 30%. Rather, the predicted saturated volumetric water

contents at location 101 stay around 9% equilibrium water content throughout the rain event, as

well as at later times after the rain event. This means that focusing solely on Ksat may be com-

pletely misleading in evaluating the drainage response of base materials. Table 6.6 shows the

maximum difference between saturated volumetric water content obtained from the soil water

characteristic curves in Figure 6.12 and the minimum predicted water content shown in Figure

6.14, reached after the rain event. Again, the results in Table 6.6 show that Cl. 3 Sp., Cl. 4 Sp.,

and the Cl. 5 Sp. do not result in significant drainage and tend to stay at elevated volumetric

water contents that are close to their saturated values, as compared to the Cl. 6 Sp., which

shows a maximum difference of 20.95 percent from its saturated volumetric water content.

This means that the Cl. 6 Sp. crushed granite is significantly more efficient in draining the

infiltrated water out of the base course than the Cl. 3 Sp., Cl. 4 Sp., or the Cl. 5 Sp. granular

materials.

Page 123: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

108

Table 6.6 Maximum difference between saturated volumetric water content and predicted volumetric water content for Class 3 special, Class 4 special, Class 5 special, and Class 6 special Mn/DOT granular base materials

Material Class 3 Special

Class 4 Special

Class 5 Special

Class 6 Special

θsat (%) 14.07 24.45 18.97 30.19 Max.

Difference (%) 1.17 2.14 1.38 20.95

6.7 INFILTRATION EFFECTS

In order to evaluate the effects of infiltration on the drainage of base materials, three different

cases are considered, namely Case 1: infiltration is equal to the total precipitation

measurements (100%); Case 2: infiltration is 70 percent of the total precipitation, and Case 3:

initial infiltration is 70% of the total precipitation, followed by a reduction down to 30 percent

after the 22nd day. This last case was the final rain event function used in the numerical

simulations for Cells 33-35, discussed previously.

Figure 6.15 shows the resulting Finite Element Model predictions. As expected the volumetric

moisture content for Case 1 remained the highest, since the greater the volume of infiltrated

water, the more time is needed to drain the water.

Table 6.7 summarizes the effects of infiltration on predicted volumetric moisture contents in

terms of the maximum predicted volumetric water content at TRD location 101 As expected,

Case 1 resulted in the greatest predicted volumetric moisture content of 11.89 percent, as com-

pared to Case 3 (the baseline case) of 9.67 percent, which corresponds to a difference of about

23 percent, meaning that the predicted volumetric moisture content is moderately sensitive to

changes in infiltration.

Page 124: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

109

8.0

8.5

9.0

9.5

10.0

10.5

11.0

11.5

12.0

12.5

210 220 230 240 250 260 270

Time (Julian day)

Predicted - Case 1 (100%)Predicted - Case 2 (70%)Predicted - Case 3 (70 and 30%)

Figure 6.15 Results for different rain events

Table 6.7 Effects of infiltration on the maximum predicted volumetric moisture content at TDR location 101

Parameter Case 1 Case 2 Case 3

Rain event (%) 100 70 70 and 30

Maximum volumetric

water content %) 11.89 10.27 9.67

6.8 WATER TABLE INFLUENCE

The initial position of the water table in the Finite Element Model is used to establish the

pressure head conditions for the system at the beginning of each simulation. Therefore, if the

water table is set at different elevations, the system will be under different initial suction and

volumetric moisture conditions.

Page 125: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

110

Again, three cases were evaluated. In Case 1, the water table is set at a depth of 0.76 m below

the surface, which corresponds to the original case. In Case 2, the water table is dropped by

0.20 m down to 0.96 m, and in Case 3 the water table was set at a depth of 1.10 m, from the top

of the hot mix asphalt surface. In all cases, the water table depths were located at a horizontal

offset of 1.83 m from the centerline of the pavement, which is the same offset as for TDR loca-

tions 101, 102, and 103. Again, the predicted volumetric moisture content at TDR location 101

was used to compare results from the three different cases.

When the water table is lowered, the suction increased for the points above it. Therefore, the

lower the water table, the lower the volumetric water content. This effect is due to the fact that

the soil water characteristic curve controls the unsaturated moisture conditions, so that when

the suction increases the volumetric water content decreases. Figure 6.16 presents the results

of the Finite Element Model simulations for the different water table positions, and Table 6.8

summarizes the effects, in terms of the maximum volumetric water content, of the change in

water table position.

4.04.55.05.56.06.57.07.58.08.59.09.5

10.0

210 220 230 240 250 260 270

Time (Julian day)

Predicted - Case 1 (water table at 0.76 m)Predicted - Case 2 (water table at 0.96 m)Predicted - Case 3 (water table at 1.10 m)

Figure 6.16 Results for different water table positions

Page 126: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

111

Table 6.8 Water table positions – Effects summary

Parameter Case 1 Case 2 Case 3

Water table position (m) 0.76 0.96 1.10

Maximum volumetric water

content (%) 9.67 7.68 6.08

In summary, as the water table is changed by 6.2 percent from 0.76 meters depth to 0.96 meters

depth, the resulting predicted maximum volumetric moisture content changes from 9.67 percent

to 7.68 percent, or by 20.6 percent. This means that the predicted volumetric moisture content

at TDR location 101 is sensitive to changes in water table, as expected.

6.9 SUMMARY

In this Chapter, the sensitivity of predicted volumetric moisture contents due to variations in

key unsaturated soil hydraulic properties and SEEP/W input parameters was evaluated. The

key parameters evaluated are: 1) the initial slope of the soil water characteristic curve, 2) the air

entry value of the Mn/DOT Class 6 Special crushed granite, 3) the saturated hydraulic conduc-

tivity of the Mn/DOT Class 6 Special crushed granite, 4) the air entry value of the R-12 silty

clay subgrade soil, 5) the saturated hydraulic conductivity for the subgrade soil, 6) the type of

granular base material, 7) variations in infiltration characteristics, and 8) the water table

location.

Overall, the most sensitive parameters were: 1) the air entry value of the granular base course

material, which if changed from 4 kPa to 5 kPa, changed the predicted volumetric moisture

content at TDR location 101 from 12.20 percent to 23.41 percent, and 2) water table location,

and 3) gradation and type of Mn/DOT Class granular base material, infiltration effects,

similarly, moderate sensitivity of the predicted volumetric moisture content at TDR location

101 was observed due to changes in: 1) the saturated hydraulic conductivity of the subgrade,

and 2) time history of infiltration. Finally, the predicted volumetric moisture content at TDR

location 101 was found to be relatively insensitive to changes in: 1) the initial slope of the soil

water characteristic curve, 2) the saturated hydraulic conductivity of the granular base material,

and 3) the air entry value for the subgrade soil. Table 6.9 provides a summary of the sensitivity

analysis.

Page 127: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

112

Table 6.9 Summary of evaluation of effects of unsaturated soil hydraulic properties and SEEP/W input parameters on the sensitivity of predicted

volumetric moisture content at TDR location 101

Effect Degree of Sensitivity (Low, Moderate, High)

Initial slope of the soil water characteristic curve Low

Air entry value of the Mn/DOT Class 6 Special crushed granite High

Saturated hydraulic conductivity of the Mn/DOT Class 6 Special crushed granite

Low

Air entry value of the R-12 silty clay subgrade soil Low

Saturated hydraulic conductivity of the subgrade soil Moderate

Type of granular base material High

Variations in infiltration characteristics Moderate

Water table location High

The results of the sensitivity study show that it is important to establish good estimates of the

air entry value for granular base course materials, as the water table location throughout the

cross section to be analyzed. The sensitivity to material type (gradation) was expected.

However, the results imply the importance of obtaining actual measured soil water

characteristic curves.

Similarly, the other effects that should be considered important in unsaturated finite element

modeling of unsaturated flow through pavements include establishing the saturated hydraulic

conductivity of the subgrade, as well as a detailed time history of infiltration that not only

includes, but starts well before and extends beyond the rain event of interest.

Finally, the results show that it may be justified in the absence of detailed measurements to

evaluate the following effects: 1) initial slope of the soil water characteristic curve, 2) the

saturated hydraulic conductivity of the granular base course material, and 3) the air entry value

for the subgrade soil, if the soil is a silty clay soil, as evaluated in this study. Other subgrade

soil types would have to be evaluated in the same manner as presented in this Chapter.

Page 128: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

110

CHAPTER 7

EFFECT OF EDGE AND UNDER DRAINS ON WATER FLOW THROUGH FLEXIBLE PAVEMENTS

Edge drains typically consist of a backfilled trench with a collector pipe that is placed longi-

tudinally next to the outer traffic lane, under the shoulder. The collector pipe is hooked up to

transverse drain pipes periodically along the highway. The backfill material used is typically

coarse graded gravel (9). Under drains simply consist of a layer of a woven or a non-woven

geotextile that extends all the way under the traffic lanes. The geotextile used is designed so as

to replace an equivalent sand drainage layer, and therefore the design of the geotextile system

is based on common drainage criteria. The FHWA Geotextile Engineering Manual (23)

provides a good overview of the design of geotextiles for drainage purposes. The main

advantage with an under drain is that the drainage path is significantly shortened. Another

feature of under drains is that they may provide a break in capillary suction. Finally, most

under drain systems today are designed to connect to longitudinal collector pipes, which may

also connect to transverse drainage pipes periodically along the highway. However, these

collector pipes are often embedded into the granular base material directly, rather than placed

in a trench, as in the case of edge drains.

In this Chapter, the relative benefits of edge drains and under drains using geotextiles are

evaluated. The pavement system described previously in Chapter 5 for Cell 33 was used as a

baseline for comparisons. Four different drainage designs were Case 1 consisted of the original

pavement section for Cell 33, but now with a 0.02 m thick geotextile underdrain located

between the Mn/DOT Class 6 Special crushed granite base course and the subgrade soil. Case

2 was a modification of Case 1, in which collector pipes were placed directly in the Mn/DOT

Class 6 Special crushed granite coarse base material under the shoulder. In Case 3, a typical

edge drain configuration from Cell 10 is introduced into the Cell 33 pavement system. Case 4

simply consists of a combination of an edge drain and a geotextile under drain system, in which

the under drain now connects to a backfilled trench containing a collector pipe. Thus, the main

Page 129: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

111

difference between Case 2 and Case 4 is that the material around the collector pipe (Case 2) is

Class 6 Special crushed granite, whereas the material around the edge drain consists of well

draining gravel (pea gravel).

In all cases, the base and subgrade characterization remained unchanged from that presented in

Figures 5.16 and 5.17 for the base layer, and Figures 3.6 and 5.25 for the subgrade material.

Finally, for simplicity of presentation, the comparison between different drainage systems was

limited to an evaluation of the volumetric water content at TDR location 101.

7.1 DESCRIPTION OF CASE 1: UNDER DRAIN

Figure 7.1 shows the pavement system used in the finite element model, along with the location

of the under drain that was included in this part of the study. The thickness of the geotextile

under drain is 0.02 m.

Figure 7.1 Under drain location for Case 1 in the pavement system

1.83 m 4.27 m 4.27 m3.05 m

CL

0.3 m Class 6 Special 0.1 m Hot Mix Asphalt

R-12 silty clay

4:1 4:1

4.0 m

16.5 m

0.02 m Geotextile

Page 130: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

112

The saturated hydraulic conductivity Ksat of the under drain was assumed to be 1.0 E-3 m/s,

which is an order of magnitude higher than the Ksat value for the base material. A hydraulic

conductivity of 1.0 E-3 m/s corresponds to that for uniform sand (26). The corresponding soil

water characteristic curve shown in Figure 7.2 for uniform sand was selected from the SEEP/W

database of soil water characteristic curves, Figure 7.3 shows the corresponding hydraulic

conductivity curve, which was again estimated with SEEP/W based on Green and Corey’s

approach (45).

0.00

5.00

10.00

15.00

20.00

25.00

30.00

0.01 1.00 100.00

Suction (kPa)

Figure 7.2 Soil water characteristic curve for under drain material

Page 131: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

113

1.00E-08

1.00E-06

1.00E-04

1.00E-02

0.01 1.00 100.00

Suction (kPa)

Figure 7.3 Estimated hydraulic conductivity curve for under drain material

Page 132: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

114

As shown in Figure 7.2, the volumetric water content for the under drain drops down to about

21 percent around 10 kPa suction.

The under drain geotextile layer was represented with a row of 304 quadrilateral elements.

Figure 7.4 shows a section of the left side of the system with this new layer.

Figure 7.4 Finite Element Model for Case 1 (Under drain system)

7.2 DESCRIPTION OF CASE 2: UNDER DRAINS WITH COLLECTOR PIPES

Figure 7.5 describes the pavement system used in the finite element model, along with the loca-

tion of the under drain and collector pipes that were used for this case. The thickness of the

geotextile under drain is 0.02 m and the collector pipes have a diameter of 0.1m.

The material characterization for the geotextile is the same as presented in Figures 7.2 and 7.3.

Figure 7.6 presents a section of the left side of the system with the underdrain and collector

pipes. The geotextile layer was represented with a row of 186 quadrilateral elements that is

extended along the Hot Mix Asphalt layer length up to 0.1 m diameter collector pipes. In order

to simulate the collector pipes, a total pressure head equal to 0 m is set around the pipe circum-

ference as a boundary condition.

Page 133: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

115

Figure 7.5 Under drain and collector pipes location for Case 2 in the pavement system

Figure 7.6 Finite Element Model for Case 2 (Under drain with collector pipes)

1.83 m 4.27 m 4.27 m3.05 m

CL

0.3 m Class 6 Special 0.1 m Hot Mix Asphalt

R-12 silty clay

4:1 4:1

4.0 m

16.5 m

0.02 m Geotextile

Collector Pipe

Page 134: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

116

7.3 DESCRIPTION OF CASE 3: EDGE DRAINS

The third case has edge drains at the ends of the asphalt layer. The drains were represented with

0.1 m diameter pipes, around which there was free draining material. The free draining

material was assumed to have the properties shown previously in Figures 7.2 and 7.3, which

correspond to uniform sand. The geometry is shown in Figure 7.7.

Figure 7.7 Edge drain location for Case 3 in the pavement system The material around the edge drain was represented with triangular elements in the finite ele-

ment model. Figure 7.8 shows the left side of the system. In order to simulate the drainage

pipes, a pressure head equal to 0 m is set around the pipe circumference as a boundary condi-

tion.

Figure 7.8 Finite Element Model for Case 3 (Edge drain)

1.83 m 4.27 m 4.27 m3.05 m

CL

0.3 m Class 6 Special 0.1 m Hot Mix Asphalt

R-12 silty clay

4:1 4:1

4.0 m

16.5 m

Drainage Material

Edge Drain

Page 135: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

117

7.4 DESCRIPTION OF CASE 4: COMBINATION OF EDGE AND UNDER DRAINS

The last case consists of the combination of Cases 1 and 3, in which an edge drain system is

combined with an under drain system. The pavement system and materials are the same as in

Cases 1 and 3, with under drains connecting to an edge drain system surrounded by well

draining material, whose hydraulic properties can be represented with Figures 7.2 and 7.3.

Figure 7.9 presents a snapshot of the finite element mesh for this case.

Figure 7.9 Finite Element Model for Case 4 (Combination of drain systems)

7.5 DRAINAGE SYSTEMS COMPARISON

Figure 7.10 shows a comparison of volumetric water content versus time for the four drainage

schemes presented. Interestingly, although Case 1 had an under drain layer, the resulting

volumetric water content is about the same as the reference case without any positive drainage

systems, implying that the sole presence of under drains in combination with Mn/DOT Class 6

Special crushed granite aggregate base course material may not be very effective in reducing

the time to drain or the equilibrium moisture content in the base.

In Cases 2, 3 and 4, the equilibrium moisture contents as compared to Case 1 (no drainage sys-

tem) were lowered from about 9.0 to 9.5 percent down to about 5.0 percent at TDR location

101. However, Figure 7.10 also shows that Cases 2, 3, and 4 resulted in about the same rate of

drainage. This means that whether the under drain is drained into a collector pipe that is placed

Page 136: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

118

0.01.0

2.03.04.05.0

6.07.08.0

9.010.0

210 220 230 240 250 260 270

Time (Julian Day)

No drainage systemCase 1:Under DrainCase 2: Under Drain with Collector PipesCase 3: Edge DrainsCase 4: Combination

Figure 7.10 Results comparison for drainage systems

in the granular base course, or to a backfilled trench with a collector pipe (edge drain), the

resulting drainage is about the same. Also, the benefits of under drains with a collector system

are about the same as those of a traditional edge drain system.

In summary, the results show that under drains by themselves do not significantly improve the

drainability of dense graded bases, whereas the introduction of either collector pipes or edge

drains in combination with under drains is very effective in reducing the amount of moisture in

a Mn/DOT Class 6 Special crushed granite base course material. These benefits are likely due

to the effects of zero head boundary conditions around the collector pipes and edge drains,

which in combination with the shortened drainage path due to the under drains affect the

distribution of suctions significantly throughout the base material, thus promoting better

drainage.

Page 137: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

119

CHAPTER 8

SUMMARY, CONCLUSIONS AND RECOMMENDATIONS

With the evaluation of SEEP/W for modeling unsaturated flow (Chapter 4), it was possible to

conclude that this program can be used to model unsaturated flow under transient conditions

through layered systems under complex boundary conditions and material characterization.

However, it is important to correctly specify the initial and boundary conditions in order to

accurately predict the unsaturated flow of water through materials. Therefore, material charac-

terization (soil water characteristic curve and hydraulic conductivity curve), initial water table

position, and geometry, are some important key conditions. In addition, it was shown that

SEEP/W could also be used to accurately model the infiltration process that pavement

structures may go through during rainfall events. Finally, it was shown that SEEP/W could

also be used to model more complicated structures consisting of layers of soil.

The verification of results to field results for cells 33, 34, and 35 (Chapter 5) illustrated how it

is possible to match well with SEEP/W measured field volumetric moisture contents, through

only minor adjustments in the unsaturated soil hydraulic properties. The results obtained show

that SEEP/W is a valuable tool for modeling the unsaturated flow of water through pavement

systems, and can be used to predict results that compare favorably to measured field results.

It is not surprising that some adjustment in unsaturated hydraulic soil properties was required to

obtain a match between predicted and measured TDR volumetric moisture contents. Some fac-

tors that may influence the results include the possibility of slight differences in the soil

packing arrangement between the field and the laboratory due to differences in field and

laboratory compaction processes, and the potential presence of residual stresses and anisotropic

stress conditions in the field. However, due to lack of data at this point in time, it is premature

to speculate about the presence and effects of these various factors on measured unsaturated

hydraulic properties of granular base material.

Page 138: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

120

In addition, the numerical results presented in Chapter 5 show the importance of considering

that only a part of the water coming from precipitation events is going to infiltrate the

pavement. Hence, an adjustment is needed for the infiltration function that represents these

events in the finite element model.

The parametric study presented in Chapter 6 showed the sensitivity of predicted volumetric

moisture contents due to variations in key unsaturated soil hydraulic properties and SEEP/W

input parameters was evaluated. The most sensitive parameters were: 1) the air entry value of

the granular base course material, which if changed from 4 kPa to 5 kPa, changed the predicted

volumetric moisture content at TDR location 101 from 12.20 percent to 23.41 percent, and 2)

water table location, and 3) gradation and type of Mn/DOT Class aggreagate base course

material. Similarly, moderate sensitivity of the predicted volumetric moisture content at TDR

location 101 was observed due to changes in: 1) the saturated hydraulic conductivity of the

subgrade, and 2) time history of infiltration. Finally, the predicted volumetric moisture content

at TDR location 101 was found to be relatively insensitive to changes in: 1) the initial slope of

the soil water characteristic curve, 2) the saturated hydraulic conductivity of the granular base

material, and 3) the air entry value for the subgrade soil.

The results of the sensitivity study show that it is important to establish good estimates of the

air entry value for granular base course materials, as the water table location throughout the

cross section to be analyzed. The sensitivity to material type (gradation) was expected. The

results imply the importance of obtaining actual measured soil water characteristic curves.

Similarly, the other effects that should be considered important in unsaturated finite element

modeling of unsaturated flow through pavements include establishing the saturated hydraulic

conductivity of the subgrade, as well as a detailed time history of infiltration that not only

includes, but starts well before and extends beyond the rain event of interest.

Page 139: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

121

After the evaluation of the effect of edge and under drains on water flow through flexible pave-

ments (Chapter 7), the results show that under drains by themselves do not significantly

improve the drainability of dense graded bases, whereas the introduction of either collector

pipes or edge drains in combination with under drains is very effective in reducing the amount

of moisture in a Mn/DOT Class 6 Special crushed granite base course material. These benefits

are likely due to the effects of zero head boundary conditions around the collector pipes and

edge drains, which in combination with the shortened drainage path due to the under drains

affect the distribution of suctions significantly throughout the base material, thus promoting

better drainage.

The results in Appendix B show that saturated flow assumptions do not take into consideration

the variation of the hydraulic conductivity with volumetric water content or suction. This

means that the time to drain evaluated on the basis of traditional “time to drain” equations

(Equation 3.2) results in an unrealistically short time to drain compared to the more realistic

unsaturated conditions. The more realistic unsaturated flow theory considers the hydraulic

conductivity as a function of the matric suction experienced in the material with drainage.

Hence, the time to drain calculated based on unsaturated flow theory will generally be longer

that that one evaluated under saturated flow assumptions for dense graded granular base

materials.

In summary, the results obtained in this study show that pavement drainage should generally be

modeled using unsaturated flow theory. The quality of the predictions is heavily dependent

upon the accuracy of both the soil water characteristic curve and the hydraulic conductivity

curve obtained in the laboratory, as well as on having a detailed knowledge of field conditions.

In addition to having detailed water table measurements throughout the cross section to be

analyzed, it is also important to have knowledge of the variation of moisture contents in both

the vertical and horizontal plane. In particular, horizontal arrays of TDRs around key

interfaces could play a major role in furthering the understanding of unsaturated flow through

flexible pavements obtained in this project.

Page 140: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

122

LIST OF REFERENCES 1. McAdam, J.L., Report to the London Board of Agriculture, 1820.

2. American Association of State Highway and Transportation Officials, AASHTO Guide

for Design of Pavement Structures. Washington, D.C., 1998.

3. Forsyth, R.A., The Economic Impact of the Pavement Subsurface Drainage. Transporta-

tion Research Record 1121, Transportation Research Board, National Research Council.

Washington, D.C., 1987.

4. Cedergren, H.R., Drainage of Highways and Airfield Pavements. Wiley Interscience

Publication. New York, 1974.

5. Cedergren, H.R., Why All Important Pavements Should be Well Drained. Transportation

Research Record 1188, Transportation Research Board, National Research Council.

Washington, D.C., 1988.

6. US Army Corps of Engineers, Engineering and Design Drainage Layers for Pavements,

Engineer Technical Letter 1110-3-435, Department of the Army, Washington DC 1992.

7. Philip, J.R., Theory of Infiltration. Advances in Hydroscience, Vol. 5, Academic Press.

New York, 1969.

8. Roberson, R. and B. Birgisson, “Evaluation of Water Flow Through Pavement Systems.”

Proceedings. International Symposium on Subdrainage in Roadway Pavements and

Subgrades, pp. 295-302, 1998.

9. Birgisson, B. and R. Roberson, Drainage of Pavement Base Material: Design and Con-

struction Issues. Transportation Research Record 938, Transportation Research Board,

National Research Council. Washington, D.C., 2000.

10. Lambe, R. and R. Whitman, Soil Mechanics, Wiley, New York, 1969.

11. Tindall, J.A. and J.R. Kunkel, Unsaturated Zone Hydrology for Scientists and Engineers.

Prentice Hall. New Jersey, 1999.

12. FHWA. Drainable Pavement Systems. Participant Notebook, Demonstration Project 87.

Washington, D.C., 1992.

13. Casagrande, A. and W.L. Shannon, “Base Course Drainage for Airport Pavements.”

Proceedings of the American Society of Civil Engineers, Vol. 77, pp 792-814, 1952.

Page 141: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

123

14. Cedergren, H.R., Seepage Requirements of Filters and Pervious Bases. Soil Mechanics

and Foundation Division. ASCE, SM5, pp. 15-23, 1956.

15. Cedergren, H.R, K.H. O’Brien, and J.A. Arman, “Guidelines for the Design of

Subsurface Drainage Systems for Highway Structural Sections.” Report No. FHWA-RD-

72-30. Washington, D.C., 1972.

16. Liu, S.J., J.K. Jeyapalan, and R.L. Lytton, Characteristics of Base and Subgrade Drain-

age of Pavements, Transportation Research Record 945, Transportation Research Board,

National Research Council. Washington, D.C., 1983.

17. Carpenter, S.H, “Highway Subdrainage Design by Microcomputer (DAMP). Drainage

Analysis and Modeling Programs.” Report No. FHWA-IP-90-012. Washington, D.C.,

1990.

18. Pufahl, D.E., R.L. Lytton, and H.S. Liang, An Integrated Computer Model to Estimate

Moisture and Temperature Effects beneath Pavements. Presented at the Annual Meeting

of the Transportation Research Record. Washington D.C.,1990.

19. Huang, Y.H., Pavement Analysis and Design. Prentice-Hall. New York, 1993.

20. Betram, G.E, An Experimental Investigation of Protective Filters. Publication No. 267,

Graduate School of Engineering, Harvard University. Cambridge, 1940.

21. Christopher, B.R. and V.C. McGuffey, Pavement Subsurface Drainage Systems.

National Cooperative Highway Research Program, Synthesis of Highway Practice 239,

National Academy Press, Washington, D.C., 1997.

22. American Society for Testing Materials (ASTM), Annual Book of ASTM Standards,

Concrete and Aggregates, Vol. 04.02. Philadelphia, 1989.

23. FHWA Geotextile Engineering Manual, Course Text. Report No. FHWA-HI-89-050,

Federal Highway Administration, 1989.

24. Fredlund, D.G. and H. Raharjo, Soil Mechanics for Unsaturated Soils. 1993 pp 30-31.

25. Hillel, D., Introduction to Soil Physics. 1980.

26. Freeze, R.A. and J.A. Cherry, Groundwater. Prentice-Hall. Englewood Cliffs, 1979.

27. Gupta, S.C. and D. Wang (2001). Soil Water Retention. In: R. Lal. Encyclopedia of Soil

Science. Marcel-Dekker (In print).

28. Fredlund, D.G. and A. Xing, Equations for the Soil-Water Characteristic Curve. Can.

Geotech J., Vol 31, pp. 521-532, 1994.

Page 142: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

124

29. Espinoza, R.D., P.L. Bourdeau, and T.D. White, Pavement Drainage and Pavement

Shoulder Joint Evaluation and Rehabilitation. Numerical Analysis of Infiltration and

Drainage in Pavement Systems. Report. Purdue University, West Lafayette, 1993.

30. Brooks, R.H. and A.T. Corey, Hydraulic Properties of Porous Media. Hydrology Paper

No. 3, Civil Engineering Dep., Colorado State Univ. Fort Collins, 1966.

31. Bear, J., Dynamics of Porous Media. Dover Publications, Inc. New York, 1972.

32. Van Genuchten, M.A., Closed-form equation for predicting the hydraulic conductivity of

unsaturated soils. Soil Sci. Soc. Amer. Proc., Vol. 44, pp. 892-898, 1980.

33. Wallace, K.B. and F. Leonardi, The influence of soil properties on the wetting-up of

earth structures. ARRB Proceedings, Volume 8, session 18. 1976.

34. Brutsaert, W, Probability Laws for Pore-size Distribution. Soil Sci., 101, pp. 85-92,

1966.

35. Vauclin, M., D. Khanji, and G. Vachaud, Experimental and Numerical Study of a

Transient Two Dimensional Unsaturated-Saturated Water Table Recharge Problem.

Water Resources Research, Vol. 15, No. 5, 1089-1101, 1979.

36. Bear, J. and A.M. Verruijt, Modeling Groundwater Flow and Pollution. D. Reidel Pub-

lishing Company. Boston, 1990.

37. Gray, W.G. and S.M. Hassanizadeh, Unsaturated Flow Theory Including Interfacial

Phenomena. Water Resources Research, Vol. 27, No. 8, 1855-1863, 1991.

38. Mualen, Y., Modified Approach to Capillary Hysteresis Based on a Similarity Hypoth-

esis. Water Resources Research, Vol. 9, No. 5, 1324-1331, 1973.

39. Brutsaert, W., The Permeability of a Porous Medium Determined from Certain Proba-

bility Lays for Pore Size Distribution. Water Resources Research, Vol. 4, No. 2, 425-

434, 1968.

40. Burdine, N.T., Relative Permeability Calculations from Pore Size Distribution Data.

Trans. AIME, Vol. 198, pp. 71-78, 1953.

41. Mualen, Y., A New Model for Predicting the Hydraulic Conductivity of Unsaturated

Porous Media. Water Resources Research, Vol. 12, No. 3, 503-522, 1976.

42. Richards, B.G., Behavior of Unsaturated Soils. In Soil Mechanics-New Horizons, Ch. 4.

American Elsevier Publishing Company Inc. New York, 1974.

Page 143: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

125

43. Gardner, W.R., Calculation of Capillary Conductivity from Pressure Plate Outflow

Data. Soil Sci. Soc. Am.J. 3, pp. 317-320, 1956.

44. Rawls, W. J., Infiltration and Soil Water Movement. Handbook of Hydrology, Editor in

Chief D. R. Maidment. McGraw Hill. New York, 1992.

45. Green, R.E. and J.C. Corey, Calculation of Hydraulic Conductivity: A Further Evalua-

tion of Some Predictive Methods. Soil Sci. Soc. Amer. Proc., Vol. 35, pp. 3-8, 1971.

46. Childs, E.C. and N. Collis-George, The Permeability of Porous Materials. Proc. Roy.

Soc. London Vol. 201 A, pp. 392-405, 1950.

47. Fredlund, D.G., A. Xing, and S. Huang, Predicting the Permeability Function for

Unsaturated Soils Using the Soil-Water Characteristic Curve. Can. Geotech. J., Vol. 31,

pp. 533-546, 1994.

48. SEEP /W Users Manual, Version 4.24. GEO-SLOPE International Ltd. Calgary, 2001.

49. NHI Course No. 130126 Pavement Subsurface Drainage Design, National Highway

Institute, Arlington, VA, April, 1999.

50. DRIP (Drainage Requirements in Pavements), Version 1.00. Federal Highway Adminis-

tration. Washington, D.C., 1996.

51. Barber, E.S. and C.L. Sawyer, Highway Subdrainage, Proceedings, Highway Research

Board, pp. 643-666, 1952.

52. Roberson, R., Personal Communication, 2002.

Page 144: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 1

APPENDIX A

EVALUATION OF TIME TO DRAIN

CALCULATIONS FOR PAVEMENTS

Page 145: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 2

APPENDIX A EVALUATION OF TIME TO DRAIN CALCULATIONS FOR PAVEMENTS

Drainage performance of base materials is often measured in terms of the time it takes to drain

a certain amount of water out of the paper. The two drainage levels that are most often used

are the time to drain either 50 percent or 90 percent of the water out of the pavement.

Traditional time to drain calculations are performed using: 1) traditional formulations by

Casagrande and Shannon (13), and Barber and Sawyer (51), discussed in Chapter 3, and 2)

from unsaturated flow theory. The differences in predicted time to drain for these two cases are

compared and discussed. The saturated time to drain calculations were performed using the

program DRIP (50), whereas the unsaturated flow calculations were obtained with SEEP/W

(48).

A.1 Time to Drain from Unsaturated Flow Theory

A one-dimensional finite element-based flow model was run under unsaturated conditions,

using SEEP/W. The finite element model consisted of a one m tall column that had a 0.4 m by

0.4 m cross section, with 0.05 x 0.05m quadrilateral elements (Figure A.1). The column was

setup so that the material was fully saturated initially; therefore no water table was set. Lateral

sides were considered impervious (q = 0 m/s per square meter). Subsequently, the bottom of

the column was subjected to atmospheric conditions, and the column was allowed to drain

freely.

Figure A.1 Finite element model – Geometry and boundary conditions

Page 146: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 3

The purpose of this finite element model is to observe how the flow behaves under unsaturated

conditions, with different material characterizations. Four different materials were used: Class

3 Special, Class 4 Special, Class 5 Special, and Class 6 Special. Figures A.2 to A.5 present the

different soil-water characteristic curves for these materials.

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 0.10 1.00 10.00 100.00

Suction (kPa)

Figure A.2 Class 3 special – Soil water characteristic curve

0.0

5.0

10.0

15.0

20.0

0.01 0.10 1.00 10.00 100.00 1000.00

Suction (kPa)

Figure A.3 Class 4 special – Soil water characteristic curve

Page 147: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 4

0.05.0

10.015.020.025.030.035.040.045.0

0.01 0.10 1.00 10.00 100.00 1000.00

Suction (kPa)

Figure A.4 Class 5 special – Soil water characteristic curve

0.0

10.0

20.0

30.0

40.0

50.0

60.0

0.01 0.10 1.00 10.00 100.00

Suction (kPa)

Figure A.5 Class 6 special – Soil water characteristic curve

For each material, five hydraulic conductivity models were applied. The hydraulic

conductivity models used include models by Gardner (43), Brooks and Corey (30), Green and

Corey (45), Van Genuchten (32), and Fredlund and Xing (28). Figures A.6 to A.9 present the

hydraulic conductivity for each material, based on the different models.

Page 148: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 5

1.0E-37

1.0E-33

1.0E-29

1.0E-25

1.0E-21

1.0E-17

1.0E-13

1.0E-09

1.0E-05

0.01 0.10 1.00 10.00 100.00Suction (kPa)

Brooks & CoreyVanGenuchtenGardnerGreen & CoreyFredlund and Xing

Figure A.6 Comparison of hydraulic conductivity models in Class 3 special

1.0E-341.0E-311.0E-281.0E-251.0E-221.0E-191.0E-161.0E-131.0E-101.0E-071.0E-04

0.01 0.10 1.00 10.00 100.00 1000.00Suction (kPa)

Brooks & CoreyVanGenuchtenGardnerGreen & CoreyFredlund and Xing

Figure A.7 Comparison of hydraulic conductivity models in Class 4 special

Page 149: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 6

1.0E-351.0E-321.0E-291.0E-261.0E-231.0E-201.0E-171.0E-141.0E-111.0E-081.0E-05

0.00 0.01 0.10 1.00 10.00 100.00 1000.00Suction (kPa)

Brooks & CoreyVanGenuchtenGardnerGreen & CoreyFredlund and Xing

Figure A.8 Comparison of hydraulic conductivity models in Class 5 special

1.0E-31

1.0E-27

1.0E-23

1.0E-19

1.0E-15

1.0E-11

1.0E-07

0.00 0.01 0.10 1.00 10.00 100.00Suction (kPa)

Brooks & CoreyVanGenuchtenGardnerGreen & CoreyFredlund and Xing

Figure A.9 Comparison of hydraulic conductivity models in Class 6 special

The evolution of the volumetric water content at the top of the soil column at selected times

during the experiment is shown in Figures A.10 to A.13.

Page 150: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 7

5.0

10.0

15.0

20.0

25.0

30.0

35.0

1.0E-05 1.0E-02 1.0E+01 1.0E+04 1.0E+07 1.0E+10

Time (days)

Vol

umet

ric

Wat

er C

onte

nt (

%)

s

Brooks & CoreyVan GenuchtenGardnerGreen & CoreyFredlund & Xing

Figure A.10 Class 3 special – Evolution of volumetric water content with time at the top of the soil column

4.0

6.0

8.0

10.0

12.0

14.0

16.0

18.0

20.0

1.0E-05 1.0E-03 1.0E-01 1.0E+01 1.0E+03Time (days)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

Brooks & CoreyVan GenuchtenGardnerGreen & CoreyFredlund & Xing

Figure A.11 Class 4 special – Evolution of volumetric water content with time at the top of the soil column

Page 151: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 8

15.0

20.0

25.0

30.0

35.0

40.0

45.0

1.0E-01 1.0E+01 1.0E+03 1.0E+05 1.0E+07 1.0E+09

Time (days)

Vol

umet

ric

Wat

er C

onte

nt (%

) s

Brooks & CoreyVan GenuchtenGardnerGreen & CoreyFredlund & Xing

Figure A.12 Class 5 special – Evolution of volumetric water content with time

at the top of the soil column

0.05.0

10.015.020.025.030.035.040.045.050.055.0

1.0E-05 1.0E-02 1.0E+01 1.0E+04 1.0E+07 1.0E+10

Time (days)

Brooks & CoreyVan GenuchtenGardnerGreen & CoreyFredlund & Xing

Figure A.13 Class 6 special – Evolution of volumetric water content with time

at the top of the soil column

Based on the change in volumetric water content, the time to drain was determined (Table A.1)

for each combination of hydraulic conductivity model and material.

Page 152: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 9

Table A.1 Comparison of drainage times under unsaturated flow conditions

Material Hydraulic Conductivity

Model

Time for 50%

Drainage (days)

Time for 90%

Drainage (days)

Brooks & Corey (30) 2.08 3.54E+04

Van Genuchten (32) 7.64 1.04E+05

Gardner (43) 0.42 5.21

Class 3

Special

Green & Corey (45) 7.64E+02 7.64E+04

Brooks & Corey (30) Not reached Not reached

Van Genuchten (32) Not reached Not reached

Gardner (43) Not reached Not reached

Class 4

Special

Green & Corey (45) Not reached Not reached

Brooks & Corey (30) 1.46 Not reached

Van Genuchten (32) 17.36 Not reached

Gardner (43) 0.42 Not reached

Class 5

Special

Green & Corey (45) 2.64E+03 Not reached

Brooks & Corey (30) 2.22 1.94E+06

Van Genuchten (32) 14.58 7.64E+06

Gardner (43) 0.28 3.47

Class 6

Special

Green & Corey (45) 4.17E+03 2.78E+04

Due to the geometry of the finite element model (Figure A.1), the maximum suction that can be

achieved is 9.8 kPa. The total height of the mesh is 1 m, therefore, if the water table reaches

the bottom of the mesh, the pressure head created above the water table is equal to 1m. This

suction head corresponds to a suction pressure of 9.8 kPa. The materials will only show

significant drainage if the air entry potential is considerably less than the maximum suction

sustained by the height of the soil column. Hence, the results from the finite element analyses

for all four materials showed a negligible to small change in volumetric water content, meaning

that the times to drain to 50 and 90 percent were also affected greatly. For example, according

to the Class 4 Special soil water characteristic curve, a pressure higher than 10 kPa is needed to

have a reduction of volumetric water content. Because the system maximum developed suction

Page 153: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 10

is 9.8 kPa, this material never achieves 50 percent drainage, let alone 90 percent drainage. This

phenomenon was also observed with the Class 5 Special at about 90% drainage, meaning that

the Class 4 and Class 5 base materials never drain fully.

The results from the numerical simulations using SEEP/W and full-unsaturated conditions

show that the different hydraulic conductivity models all result in very different times to drain.

The time difference varies from a few hours to many days. Although all of the hydraulic

conductivity models follow the same general trend, the materials drain at dissimilar times. For

all the materials, the fastest time to drain is obtained when using Gardner’s model, for 50 and

90%. If this model is compared against the others (Figures 5.6 through 5.9), it is possible to

see that the slope of the hydraulic conductivity curve is more gradual than for the other models.

Therefore, the reduction in hydraulic conductivity is slower and the material could be drained

faster. On the other hand, when using Green and Corey’s model, the time to drain is really

long, due to the steep slope in the hydraulic conductivity curve, as well as the continuous

change in hydraulic conductivity starting at zero (0) kPa.

A.2 Time to Drain from Saturated Flow Theory

The parameters used to perform the calculations in DRIP are summarized in Table A.2. Due to

the saturated condition assumption, the hydraulic conductivity does not change. Therefore, a

ksat value is needed for each material. This value is the same for all the hydraulic conductivity

models used previously. From each soil water characteristic curve, the effective porosity (ne) is

determined because is the difference between θ sat and θ R.

The height (H) of the model is 1 m, the width of the drainage path (W) is 0.4 m, and the

resultant length of the drainage path (LR) is 0.4 m.

Page 154: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 11

Table A.2 Saturated case – Model parameters

Material H

(m)

k

(m/s)

W

(m)

LR

(m)

θ sat

(%)

θ R

(%)

ne

(%)

Class 3 Special 4.0 2.20E-07 0.4 0.40 32.01 16.97 15.04

Class 4 Special 4.0 4.20E-06 0.4 0.40 14.35 11.28 3.08

Class 5 Special 4.0 3.80E-06 0.4 0.40 40.70 16.09 24.61

Class 6 Special 4.0 2.00E-06 0.4 0.40 52.08 5.92 46.16

The calculations in DRIP of time-to-drain are based on two methods: the Barber and Sawyer

method (51) and the Casagrande and Shannon method (13). These two procedures were

applied to find time to drain at 50% and 90% of drainage. Table A.3 shows the results.

Table A.3 Comparison of drainage times under saturated flow conditions

Method

Barber and Sawyer (51) Casagrande and Shannon (13) Material

Time for 50%

Drainage (min)

Time for 90%

Drainage (min)

Time for 50%

Drainage (min)

Time for 90%

Drainage (min)

Class 3 Special 2.22 38.40 4.41 39.56

Class 4 Special 0.02 0.41 0.05 0.42

Class 5 Special 0.21 3.64 0.42 3.75

Class 6 Special 0.75 12.94 1.48 13.33

Interestingly, the time for 50% drainage based on the Barber and Sawyer (51) method (Equa-

tion 3.2) is about 50 percent of the time to drain obtained with Casagrande and Shannon’s (13)

approach (Equation 3.2).

Page 155: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 12

In stark contrast to the time to drain based on unsaturated conditions, the Class 4 Special material showed the quickest drainage time for 50% and 90%. This soil has the highest Ksat of all the materials used for the modeling, and subsequently shows faster drainage using saturated flow theory that ignores suction effects. Although saturated flow theory considers the material characterization and geometry of the sys-tem, it does not take into account the variation of the hydraulic conductivity with volumetric water content or suction that represents the true behavior of the material under field conditions. Therefore, by taking just the ksat value and using that throughout to characterize the drainage of base materials the time to drain will obviously be significantly shorter than for the more realistic unsaturated conditions, in which the hydraulic conductivity is a function of the suction experienced in the material with drainage. A comparison of Tables 5.1 and 5.3 shows that most materials would drain under saturated conditions in just a few minutes, while for unsaturated conditions it would take days or not be achieved.

A.3 Comparison Applied to a Real Example Based on the previous results, a more detailed comparison between saturated and unsaturated conditions was performed, using a SEEP/W finite element model based on a case proposed by Huang (19). Geometry attributes of a drainage base layer were used to generate a Finite Ele-ment Model system, 0.5 m tall and 6.5 m long, with a slope of 2%, and quadrilateral elements. Figure A.14 presents a section of the Finite Element Model left side for the unsaturated case. The layer was setup so that the material was fully saturated initially, the bottom was subjected to atmospheric conditions, and the layer was allowed to drain freely. The layer was considered impervious on right and left sides. Class 3 Special, Class 4 Special, Class 5 Special, and Class 6 Special were used as base mate-rials. The soil water characteristics curves were described at the beginning of the chapter. The hydraulic conductivity curves corresponded to Fredlund and Xing (28) model. The time to drain for the saturated case was obtained by applying Casagrande and Shannon method (Equation 2.1). Table A.4 summarizes the parameters used for this case.

Page 156: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 13

Figure A.14 Example – Geometry and boundary conditions

Table A.4 Geometry and material parameters for saturated case

SOIL Class 3 Special Class 4 Special Class 5 Special Class 6 Specialk (m/s) 2.20E-07 4.20E-06 3.80E-06 2.00E-06 θmax (%) 32.01 14.35 40.70 52.08

θR (%) 16.97 11.28 16.09 5.92 ne (%) 15.04 3.08 24.61 46.16 L (m) 6.5 6.5 6.5 6.5 H (m) 0.15 0.15 0.15 0.15 S (%) 2 2 2 2

As it can be seen from Table A.5, the saturated conditions approach (Casagrande and Shannon

method) results in a quicker time for 50% drainage. On the other hand, the more realistic

unsaturated conditions show that 50% drainage is never reached.

Table A.5 Time to drain – Saturated vs. unsaturated conditions for a geometry closer to

pavement conditions

SOIL Class 3 Special

Class 4 Special Class 5 Special Class 6 Special

t50 (days) Saturated conditions

597.03 6.39 56.56 201.53

t50 (days) Unsaturated conditions

Not reached Not reached Not reached Not reached

Page 157: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 14

Due to the geometry of the layer under unsaturated conditions, the maximum suction pressure

that can be obtained is approximately 1.5 kPa. When the water table reaches the layer bottom,

the maximum pressure head developed is 0.15 m. This is equivalent to 1.47 kPa of suction.

Therefore, only a small reduction in the volumetric water content will occur according to the

characterization of the base materials.

In conclusion, saturated flow conditions will result in a quicker drainage of soil materials.

However, in most cases, this does not represent real conditions. Unsaturated conditions result

in a more realistic behavior of the drainage flow, showing that the drainage depends on the

suction pressure, and therefore on the variability of the volumetric water content and hydraulic

conductivity.

Within unsaturated conditions is also important to take into account, the air entry potential for

the material. This value determines the suction at which the material starts to drain. For ex-

ample, by looking at the soil water characteristic curve for Class 4 Special (Figure A.15), this

soil is saturated before 10 kPa are reached. This suction value is its air entry potential. This

value can also be observed at the hydraulic conductivity curve (Figure A.16).

0.0

5.0

10.0

15.0

20.0

0.01 0.10 1.00 10.00 100.00 1000.00

Suction (kPa)

Air entry = 10 kPa

Saturated condition

Unsaturated condition

Figure A.15 Class 4 special – Air entry potential at soil water characteristic curve

Page 158: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 15

1.0E-10

1.0E-08

1.0E-06

1.0E-04

0.01 0.10 1.00 10.00 100.00 1000.00Suction (kPa)

Air entry = 10 kPa

Saturated conditionUnsaturated condition

Figure A.16 Class 4 special – Air entry potential at hydraulic conductivity curve

A.4 Sensitivity to Gradation

As a complementary component to the evaluation of the time to drain, the effects of variations

in material gradation on time to drain were estimated using SEEP/W. The one-dimensional

finite element-based flow model shown previously in Figure A.1 was used to evaluate the

sensitivity to gradation. Same Finite Element Model, geometry and boundary conditions were

applied.

Table A.6 describes the limits used for Class 3 Special, Class 4 Special, Class 5 Special, and

Class 6 Special. These materials were used as base layers. The upper and lower limits in Table

A.6 refer to the gradation specification limits used by Mn/DOT for these materials.

The soil water characteristic curves used to characterize the base materials are the same as

presented in Figures A.2 to A.5. Due to the fact that Gardner’s model (43) and Green and

Corey’s model (45) do not depend on the gradation of the material, the Finite Element Model

was only run for Brooks and Corey’s model (30) and Van Genuchten’s model (32), which

depend on the pore size index that is function of the percentage of sand. Table A.7 summarizes

the percentage of sand used for the different gradation limits.

Page 159: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 16

Table A.6 Gradation limits for different base materials

Material Class 3 Special Class 4 Special Class 5 Special Class 6 SpecialLimit Lower Upper Lower Upper Lower Upper Lower Upper

Passing 1" (25mm) 100 % 95 % 100 % 100 % 100 %

Passing 3/4" (19mm) - 90 % 100 % 90 % 100 % 85 % 100 %

Passing 3/8" (9.5mm) 95 % 100 % 80 % 95 % 70 % 85 % 50 % 70 %

Passing #4 (4.75mm) 85 % 100 % 70 % 85 % 55 % 70 % 30 % 50 %

Passing #10 (2mm) 65 % 90 % 55 % 70 % 35 % 55 % 15 % 30 %

Passing #20 (0.85mm) - - - -

Passing #40 (0.425mm) 30 % 50 % 15 % 30 % 15 % 30 % 5 % 15 %

Passing #60 (0.225mm) - - - -

Passing #100 (0.15mm) - - - -

Passing #200 (0.075mm) 8 % 15 % 5 % 10 % 3 % 8 % 0 % 5 %

Table A.7 Percentages of sand for different base materials

Limit Class3

Special

Class 4

Special

Class 5

Special

Class 6

Special

Lower 77 65 52 30

Upper 85 75 65 45

Figures A.17 to A.20 show predicted hydraulic conductivity with suction, as determined by

Brooks and Corey’s model (30). When the upper limit of the specifications is used, the

hydraulic conductivity is greater than in the lower limit, as expected. Also, when the sand

percentage is increased, the Brooks and Corey’s parameter n (n=3+2/l) decreases, resulting in

an increase in hydraulic conductivity.

Page 160: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 17

1.0E-341.0E-311.0E-281.0E-251.0E-221.0E-191.0E-161.0E-131.0E-101.0E-071.0E-04

0.01 1 100Suction (kPa)

Lower LimitUpper Limit

Figure A.17 Class 3 special – Comparison upper and lower limit –

Brooks and Corey (30) model

1.0E-341.0E-311.0E-281.0E-251.0E-221.0E-191.0E-161.0E-131.0E-101.0E-071.0E-04

0.01 1 100 10000Suction (kPa)

Lower Limit

Upper Limit

Figure A.18 Class 4 special – Comparison upper and lower limit –

Brooks and Corey (30) model

Page 161: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 18

1.0E-361.0E-331.0E-301.0E-271.0E-241.0E-211.0E-181.0E-151.0E-121.0E-091.0E-061.0E-03

0.01 1 100 10000

Suction (kPa)

Lower LimitUpper Limit

Figure A.19 Class 5 special – Comparison upper and lower limit –

Brooks and Corey (30) model

1.0E-31

1.0E-28

1.0E-25

1.0E-22

1.0E-19

1.0E-16

1.0E-13

1.0E-10

1.0E-07

1.0E-04

0.01 1 100

Suction (kPa)

Lower Limit

Upper Limit

Figure A.20 Class 6 special – Comparison upper and lower limit –

Brooks and Corey (30) model

Figures A.21 to A.24 show similar results for the Van Genuchten (32) model. Again, when the

sand content is increased (Upper limit applied), Van Genuchten’s parameter m (m=λ / (λ + 1))

also increases, resulting in an increase in hydraulic conductivity.

Page 162: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 19

1.0E-32

1.0E-29

1.0E-26

1.0E-23

1.0E-20

1.0E-17

1.0E-14

1.0E-11

1.0E-08

1.0E-05

0.01 1 100

Suction (kPa)

k (m

/s)

Lower LimitUpper Limit

Figure A.21 Class 3 special – Comparison upper and lower limit –

Van Genuchten (32) model

1.0E-32

1.0E-29

1.0E-26

1.0E-23

1.0E-20

1.0E-17

1.0E-14

1.0E-11

1.0E-08

1.0E-05

0.01 1 100 10000Suction (kPa)

Lower LimitUpper Limit

Figure A.22 Class 4 special – Comparison upper and lower limit –

Van Genuchten (32) model

Page 163: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 20

1.0E-28

1.0E-25

1.0E-22

1.0E-19

1.0E-16

1.0E-13

1.0E-10

1.0E-07

1.0E-04

0.01 1 100 10000

Suction (kPa)

k (m

/s)

Lower LimitUpper Limit

Figure A.23 Class 5 special – Comparison upper and lower limit – Van Genuchten (32) model

1.0E-31

1.0E-28

1.0E-25

1.0E-22

1.0E-19

1.0E-16

1.0E-13

1.0E-10

1.0E-07

1.0E-04

0.01 1 100

Suction (kPa)

k (m

/s)

Lower LimitUpper Limit

Figure A.24 Class 6 special – Comparison upper and lower limit – Van Genuchten (32) model

Page 164: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 21

Using the Brooks and Corey (30) and Van Genuchten (32) predictions, along with the Upper

and Lower limits on hydraulic conductivity, the time to drain to 50 and 90 percent was

obtained. Tables A.8 and A.9 summarize the results for Class 3, 4, 5, and 6 Special, in terms of

a decrease in volumetric water content. For the lower gradation limit, the results show that

more time is needed to drain 50% as well as 90% as compared to the results obtained for the

Upper limit.

Table A.8 Time to drain for lower and upper limit – Brooks and Corey (30) model

Time (days) Limit Material

50% Drainage 90% Drainage Lower Class 3 Special 2.20 4.79E+04 Upper Class 3 Special 2.08 4.10E+04 Lower Class 4 Special Not reached Not reached Upper Class 4 Special Not reached Not reached Lower Class 5 Special 1.53 Not reached Upper Class 5 Special 1.39 Not reached Lower Class 6 Special 2.01 2.43E+06 Upper Class 6 Special 1.74 1.46E+06

Table A.9 Time to drain for lower and upper limit – Van Genuchten (32) model

Time (days) Limit Material

50% Drainage 90% Drainage Lower Class 3 Special 7.90 1.32E+05 Upper Class 3 Special 7.64 1.04E+05 Lower Class 4 Special Not reached Not reached Upper Class 4 Special Not reached Not reached Lower Class 5 Special 18.06 Not reached Upper Class 5 Special 14.58 Not reached Lower Class 6 Special 13.89 1.25E+07 Upper Class 6 Special 13.19 6.60E+06

Page 165: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

A - 22

In conclusion, the results clearly show that under unsaturated conditions the gradation of a

material has a strong effect on the time to drain. However, this is not the only factor to take

into account. The soil water characteristic curve of the material is also important, as well as the

data that is measured in field because it provides initial parameters as the hydraulic

conductivity when is saturated.

A.5 Summary

Time to drain is a variable that describes the drainage performance of a soil layer (i.e. base,

subbase, subgrade). It helps to understand how long takes the water to leave a specific system.

Evaluation of the reduction in water content of a material, under saturated or unsaturated con-

ditions, can be used to determine this parameter.

Due to the fact that saturated flow assumptions do not take into consideration the variation of

the hydraulic conductivity with volumetric water content or suction, the drainage performance

of a simulated system will result in a short time to drain compared to unsaturated conditions.

Therefore, true behavior of the material under field conditions are not being reproduced in a

realistic way. In contrast unsaturated flow theory considers the hydraulic conductivity as a

function of the suction experienced in the material with drainage. Hence, the time to drain the

system will be longer that that one evaluated under saturated flow assumptions.

As mentioned before, the material characterization is an important input to simulate the

drainage performance of a system. soil water characteristic curve and hydraulic conductivity

curve permit to evaluate when and how fast a material can drain under certain conditions.

Therefore, the trend and slope of the curves is really significant. Besides, knowing the point at

which the material will start to drain (i.e. air entry potential) helps to have a more complete

perspective of what it is taking place in the drainage of a material.

Page 166: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

B - 1

APPENDIX B

DETERMINATION OF AIR ENTRY VALUE

FOR CLASS 6 SPECIAL

Page 167: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

B - 2

APPENDIX B

DETERMINATION OF AIR ENTRY VALUE FOR CLASS 6 SPECIAL The air entry value for the Mn/DOT Class 6 Special aggregate course material was determined

by using two different methods. The first method consisted on plotting its soil water character-

istic curve (Figure 5.16) in a semi log scale and finding the air entry value as the intercept

between the tangents to the slopes of this curve. Figure B.1 presents this procedure.

According to this graph, the air entry value for the class 6 special material is 3 kPa.

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

0.01 0.10 1.00 10.00 100.00

Suction (kPa)

Air Entry = 3 kPa

Figure B.1 Air entry value determination for Class 6 special material – Semi log scale

In order to check the air entry value determined through the first method, a second method was

applied. This method consisted of plotting the soil water characteristic curve of the material in

a log-log scale and determining the air entry value from the intercept of the tangents to the

slopes of this curve. Figure B. 2 shows that the air entry value determined confirms the first

founded value.

Page 168: Final 2002-30 Research - Minnesota Department of Transportationdotapp7.dot.state.mn.us/research/pdf/200230.pdf · 2006-09-25 · Paola Ariza, Bjorn Birgisson, Ph.D., P.E. UNIVERSITY

B - 3

1.0

10.0

100.0

0.01 0.10 1.00 10.00 100.00Suction (kPa)

Air entry = 3 kPa

Figure B.2 Air entry value determination for Class 6 special material – Log-log scale