225
MoDOT TE 5092 .M8A3 no .90-4 RI COOPERATIVE HIGHWAY RESEARCH PROGRAM FINAL REPORT DETERMINATION OF AASHTO DRAINAGE COEFFICIENTS MISSOURI HIGHWAY AND TRANSPORTATION DEPARTMENT FEDERAL HIGHWAY ADMINISTRATION 90-4 Property of MoDOT TRANSPORTATION LIBRARY

Determination of AASHTO Drainage Coefficients

Embed Size (px)

Citation preview

MoDOT

TE 5092 .M8A3 no.90-4

RI COOPERATIVE HIGHWAY RESEARCH PROGRAM

FINAL REPORT

DETERMINATION OF AASHTO DRAINAGE COEFFICIENTS

MISSOURI HIGHWAY AND TRANSPORTATION DEPARTMENT FEDERAL HIGHWAY ADMINISTRATION

90-4

Property of

MoDOT TRANSPORTATION LIBRARY

DETERMINATION OF AASHTO DRAINAGE COEFFICIENTS

STUDY 90-4

Prepared for

MISSOURI HIGHWAY AND TRANSPORTATION DEPARTMENT

by DAVID N. RICHARDSON WILLIAM J. MORRISON

PAUL A. KREMER KEVIN M. HUBBARD

DEPARTMENT OF CIVIL ENGINEERING UNIVERSITY OF MISSOURI - ROLLA

ROLLA, MISSOURI

in cooperation with U.S. DEPARTMENT OF TRANSPORTATION

FEDERAL HIGHWAY ADMINISTRATION

June 1996

The opinions, findings and conclusions expressed in this publication are not necessarily those of the Federal Highway Administration.

I IIIIIIII IIII Ill/ ll]ij1~li1]~~f 11111 1111111111111 RD0009047

I I I I I I I I I I I I I I I I I I I

ii

EXECUTIVE SUMMARY

This study was conducted to determine the drainage (m) coefficients of

granular bases and subbases for use in the 1986 AASHTO Guide pavement design

method. The project entailed a review and compilation of published literature,

laboratory testing, analysis of results, and preparation of this report.

One existing method which is used to determine drainage coefficients was

examined, and two potential strategies for development into a new method were

explored.

Use of the AASHTO Guide method necessitates the determination of: 1) the

Percent Time of Saturation of the pavement structure, and 2) the Quality of

Drainage of the base and subbase. With these, the m-coefficients are found from

a table. Unfortunately, little direction was given in regard to the determination of

the necessary input data that is necessary in order to use the table.

Carpenter developed a method to determine m-coefficients which can be

easily implemented by use of software called DAMP. Carpenter provided a method

to determine the Percent Time of Saturation for the pavement structure, and the

Quality of Drainage of the combined base and subgrade. DAMP does not lead to

reasonable results if one subscribes to the theory that the Road Test pavement

drainage was "Fair." The whole idea of the use of m-coefficients is to rate any

pavement's drainage relative to that at the Road Test . Better drainage should be

"Good" or "Excellent," worse drainage should be "Poor" or "Very Poor." Also, the

effect of freeze-thaw cycles and frost heave is not emphasized. The extrapolation

of the Thornthwaite method of regional moisture available to the conditions in the

iii

pavement structure is of concern. Also, the manner in which the time of

saturation for various environmental conditions is calculated is arbitrary. However,

it is recognized that at the present time there are no practical working solutions to

this dilemna, and that the time of saturation procedure in DAMP is a significant

step forward, and it or some modification should be used until a more

fundamentally sound, user-friendly method can be developed.

The TTI Integrated Model of the Climatic Effects on Pavements was

evaluated with the idea that it could be used to determine environmental effects on

granular base/subbase materials, which would lead to the calculation of m­

coefficients. Unfortunately, the scheme· was unsuccessful because of limitations in

the TTI model. First, the model requires a minimum of 100 input variables, many

of which are not easily obtained and must be assumed. Model output is sensitive

to the magnitude of some of these input values. Secondly, the model has a low

sensitivity to variables that are thought to be important to the derivation of m­

coefficients. Third, the program is somewhat cumbersome. And most

importantly, the output parrots the input base course modulus. This is a fatal flaw

and rendered the program unusable for purposes of drainage coefficient

determination as envisioned in this study.

The materials under study included two sources of crushed stone and two

gravels. All materials were selected, sampled, and delivered to UMR by MHTD

personnel. The primary tests performed were: 1) resilient modulus testing at a

low and high degree of saturation to assess the moisture sensitivity of the

materials, and 2) permeability and effective porosity to assess the drainage

I

I I I I I I I

I

I

I I I I I I I I I I I I I I I I I I

iv

characteristics of the materials.

Two gradations of granular material were used in the resilient modulus

testing: one followed the midpoint of the MHTD Type 1 gradation (MHTD Middle)

acceptance band, and the other was the so-called New Jersey (NJ) open-graded

gradation. An additional gradation (OGS) was used in the permeability portion of

the study, along with the MHTD Middle and the NJ. The aggregates were also

tested for specific gravity, plasticity index (Pl), moisture-density relationships, and

relative density.

Particle shape/surface texture tests were performed on the four aggregates.

The measured angularities of the two stones were about the same, and were more

angular than the two gravels, which were about equal. The difference in

angularity/texture was not great between the crushed stones and the gravels.

Resilient modulus (Eul test results were required for use in the TTI method

and in the new method developed in this study. The tests were run on all four

aggregates using two gradations, two compactive efforts, and two degrees of

saturation, with replications. Fourteen combinations of confining pressure and

cyclic applied deviator stress were used for each specimen in the test sequence.

The results of the testing indicated the Eu increases with a lower

degree of saturation. The average percent loss in k1 (intercept of the Eu - bulk

stress plot) due to increased saturation was 31 %. This information was used in

the development of the m-coefficients. The data showed that the drained open­

graded moduli were greater than the undrained dense-graded moduli . An increase

in degree of saturation acted to lower k1 and raise k2 (slope of the Eu - bulk stress

V

plot) of the granular material, and to lower subgrade support, all of which acted to

lower the E9

of the granular material.

Permeability of base material is required input for DAMP, TTI Integrated

model, and the method developed in this study. Permeability data are necessary in

order to calculate the time-to-drain for base layers. The rigid wall constant head

test procedure was used for the NJ and OGS open-graded materials, while the

dense-graded specimens were tested in a triaxial compression chamber (flexible

wall, constant head).

Permeabilities estimated from the Moulton equation significantly

underestimated the observed values by an average of seven times . A review of

the data on which the Moulton equation is based reveals potential problems with

air blockage, effect of end conditions, and possibly incorrect use of specific gravity

data, all of which would lead to falsely low values.

I

I

The gravels exhibited slightly greater permeabilities than the crushed stones, I but not statistically so at the 0.05 level.

On the average, the effective porosities of the dense-graded and open-

graded materials were about 27% and 68% of the total porosities, respectively.

The dense-graded effective porosity is considerably smaller than the open-graded

value, which is to be expected because of the finer pore sizes in the dense-graded

material.

Overall, the permeabilities of the dense-graded materials were significantly

lower by several orders of magnitude than the open-graded materials (average of

0.8 vs 1014 ft/day).

I

I

I

I I

I

I I I I I I I I I I I

vi

A regression equation to estimate permeability was developed by combining

the results from several studies found in the literature with the results of this

study. The equation had an adjusted-R2 = 0.900. The equation is considered

accurate in the range of 0. 1 to 1000 ft/day.

Although the Moulton equation significantly underpredicts permeability, it

may be the equation of choice because field conditions may render the base layer

less permeable than what would be predicted with good quality laboratory testing.

A new method of calculation of drainage coefficients was developed. In

essence, m-coefficients were calculated as a ratio of the layer coefficient of Road

Test granular base material under a given drainage and climate condition to the

layer coefficient under Road Test site conditions. The layer coefficients were

calculated from resilient moduli. The resilient moduli were calculated with the

program KEN LA YER under varying conditions. By changing subgrade and base

moisture conditions for a given time of year, the moduli were varied. The base

material moisture sensitivity (effect on k1 and k2 ) was determined in part by the

resilient modulus laboratory testing of granular materials. The result of the above

analysis was the creation of a Quality of Base Drainability table (based on time-to­

drain to 85 percent saturation), a Quality of Subgrade Drainability table (based on

subgrade permeability, position of water table, flooding potential, presence of

impermeable layers, potential for water seepage, and so forth), a Quality of

Pavement Drainage table (based on the previous two tables), a Climate Condition

table (based on estimated season lengths), and finally, an m-coefficient table

(based on Quality of Pavement drainage and Climate Condition).

vii

A regression equation was developed to assist in the estimation of

compacted dry density in order to estimate permeability with the Moulton equation

and the UMR equation. The equation had an adjusted-R2 = 0. 729.

A sensitivity analysis was performed. The most important variables in

regard to m-coefficient calculation were climate condition, base drainability, and

subgrade drainability. These, in turn, affected base thickness calculation

significantly.

In comparison of Missouri sites to the Road Test site, in a regional sense

actual data indicates that most of Missouri is in a climatic zone that is rated as

having a greater time of saturation, so a given paveme.nt in Missouri should fare

worse (from moisture-related problems) and therefore should have m-coefficients

less than 1 .0, unless something is done to improve the pavement drainage.

Conversely, for a situation where any water that enters the base is quickly

removed laterally and where the soil drains well and does not supply water from

the surrounding soil or side hill wet weather springs and so forth, then an

expectation of a 10 to 20% improvement in pavement performance would be

reasonable. For a somewhat lesser quality of subgrade drainage with a highly

drainable base, a 10% credit may be more realistic. And, going with the belt-and­

suspenders approach, there is the option of supplying a drainable section with no

reduction in pavement thickness.

I

I I I I I I I

I I I I I I I I I I I

viii

TABLE OF CONTENTS PAGE

EXECUTIVE SUMMARY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ii

TABLE OF CONTENTS ....................................... viii

LIST OF FIGURES ........................................... xii

LIST OF TABLES I • • • I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I xiv

INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 GENERAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1 OBJECTIVES AND SCOPE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

DATA PROCUREMENT .............. . ....................... 5 GENERAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5

. SOIL SURVEYS ........ . · . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 MATERIAL PHYSICAL, THERMAL, AND MOISTURE

PROPERTIES/DAT A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 CLIMATOLOGICAL DAT A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 UNBOUND GRANULAR MATERIAL DRAINABILITY PROPERTIES . . . . . 6 PAVEMENT STRUCTURAL PROPERTIES . . . . . . . . . . . . . . . . . . . . . . 6 PAVEMENT SECTION GEOMETRY . . . . . . . . . . . . . . . . . . . . . . . . . . 7

ALTERNATIVE METHODS FOR DETERMINATION OF DRAINAGE COEFFICIENTS ...................................... . GENERAL .......................................... . 1986 AASHTO METHOD ............................... .

8 8 9

General . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9 Standards for Quality of Drainage . . . . . . . . . . . . . . . . . . . . . . 9 Time-to-Drain Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 Flexible Pavement Drainage Coefficients (m) .............. .

Relation to Quality of Drainage .................. . Effects of Varying Moisture Levels ................ .

DAMP ............................................ .

11 11 12 15 15 16 16 21 23 25 25 26 26

General ....................................... . Drainage Coefficient Determination .................... .

Drainage Layer Characteristics and Base Drainage Times .. Percent Time of Saturation ..................... . Subgrade Drainage ........................... . Quality of Drainage .......................... . AASHTO Drainage Coefficient Selection ............ .

TII I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I

General I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I

ix

Precipitation Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28 Infiltration and Drainage Model . . . . . . . . . . . . . . . . . . . . . . . . 28 Climatic-Materials-Structural Model . . . . . . . . . . . . . . . . . . . . . 30 CRREL Frost Heave and Thaw Settlement Model . . . . . . . . . . . . 30

MATERIAL TYPES AND SOURCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32

LABORATORY INVESTIGATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 GENERAL ........................................... 33 EXPERIMENT AL GRADATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 GRADATION CURVE SHAPE/POSITION . . . . . . . . . . . . . . . . . . . . . . 34 PARTICLE SHAPE/TEXTURE . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34 SPECIFIC GRAVITY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 SCREENING . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 SPECIMEN FABRICATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37 MOISTURE - DENSITY RELATIONSHIP . . . . . . . . . . . . . . . . . . . . . . . 38 RESILIENT MODULUS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

General ........................................ 38 Equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 Test Variables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

Stress State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 Degree of Saturation . . . . . . . . . . . . . . . . . . . . . . . . . . 42 Degree of Compaction . . . . . . . . . . . . . . . . . . . . . . . . . 42 Particle Shape/Surface Texture . . . . . . . . . . . . . . . . . . . 44

Testing Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 Test Procedure . . . . . . . . . . . . . . . .. . . . . . . . . . . . . . . . . .. . . 44

PERMEABILITY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46 General ........................................ 46 Testing Concerns . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

Air Blockage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49 Movement of Fines . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 Excessive Gradients . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 Direction of Flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51 Off-Target Density . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

Rigid Wall Permeameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 Equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52 Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

Flexible Wall Permeameter . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 Equipment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 Procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61

POROSITY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62 EFFECTIVE POROSITY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

RES UL TS OF THE LABORATORY INVESTIGATION . . . . . . . . . . . . . . . . . . . 68 AS-RECEIVED GRADATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

I

I

I

I

I

I

I

I I

X

EXPERIMENTAL GRADATIONS ............................ 68 GRADATION CURVE SHAPE/POSITION . . . . . . . . . . . . . . . . . . . . . . . 69 MOISTURE-DENSITY RELATIONSHIPS AND SPECIFIC GRAVITIES ..... 74 PARTICLE SHAPE AND SURFACE TEXTURE . . . . . . . . . . . . . . . . . . . 75 PLASTICITY OF FINES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 RESILIENT MODULUS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76 ST A TISTICAL ANALYSIS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 PERMEABILITY, POROSITY, AND EFFECTIVE POROSITY ........... 87

Open-Graded Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87 Dense-Graded Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

ESTIMATION OF PERMEABILITY . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

RESULTS OF MODELS EVALUATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 TTI INTEGRATED MODEL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114 DAMP ............................................ 116 CONTRAST BETWEEN THE INTEGRATED PROGRAM AND DAMP . . . . 120

DRAINAGE COEFFICIENT DETERMINATION . . . . . . . . . . . . . . . . . . . . . . . 122 GENERAL . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 AASHO ROAD TEST . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122 DETERMINATION OF AASHTO DRAINAGE COEFFICIENTS-UMR

METHOD ...................................... 124 General Methodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 124 Quality of Base Drainage . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 Subgrade Quality of Drainage . . . . . . . . . . . . . . . . . . . . . . . . 134 Pavement Structure Quality of Drainage . . . . . . . . . . . . . . . . . 134

DEVELOPMENT OF M-COEFFICIENT TABLE . . . . . . . . . . . . . . . . . . . 136 Reasonableness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 146

MO DAMP SENSITIVITY ANALYSIS - . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148 M-Coefficient Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 148

SUMMARY AND CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152

RECOMMENDATIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 163

FUTURE RESEARCH NEEDS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 164

ACKNOWLEDGEMENT ..................................... 165

REFERENCES 166

APPENDIX A . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175 REQUIRED INPUT FOR TII MODEL . . . . . . . . . . . . . . . . . . . . . . . . . 175

xi

APPENDIX 8 ............................................. 180

USE OF DAMP MANUAL ..................................... 181 GENERAL ........................................... 181 USE OF DAMP DATA FILES .............................. 181 INTERACTIVE DAMP INPUT SELECTION ...................... 182 SUMMARY REQUIREMENTS FOR DAMP INPUT ................. 182

Pavement Geometry ............................... 182 Layer Thicknesses ................................ 183 Layer Densities ................................... 183 Base Course Gradation and Specific Gravity ............... 183 Subgrade Drainability/Permeability ...................... 183 Other Input Variables .............................. 184

DRAINAGE COEFFICIENT OUTPUT SCREEN ................... 185

APPENDIX C ............................................. 186 MODAMP MANUAL .................................... 186 GENERAL ........................................... 187 LOADING MODAMP AND CLIMATOLOGICAL DATA SPREADSHEETS . 187

Software and Hardware ............................. 187 Loading ........................................ 187

MO DAMP DISPLAY .................................... 188 Permeability of the Drainage Layer ..................... 188 Slope Factor .................................... 189 Thickness of Base ................................. 189 Subgrade Drainability .............................. 189 Time-to-Drain/Quality of Base Drainage .................. 191 Quality of Pavement Drainage ........................ 192 Time of Saturation ................................ 192 Climate Condition ................................. 193 Drainage Coefficients .............................. 193

APPENDIX D ............................................. 194 I

I I

I I I I I I I I

I

I I

xii

LIST OF FIGURES 1. Recommended m-values as a function of the Quality of Drainage and

the Exposure to Saturation (after Seeds and Hicks (4)) . . . . . . . . . . . . 14 2. Moulton Nomograph for Estimation of Base Course Permeability . . . . . 17 3. Integrated Program Flowchart . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 4. Semilog Plot of Three Experimental Gradations . . . . . . . . . . . . . . . . . 35 5. Resilient Modulus Testing Equipment . . . . . . . . . . . . . . . . . . . . . . . . 40 6. Rigid Wall Permeameter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53 7. Schematic of Rigid Wall Permeameter Test Station . . . . . . . . . . . . . . . 55

. 8. Field Gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60 9. Flexible Wall Permeameter Test Station . . . . . . . . . . . . . . . . . . . . . . 63 10. Schematic of Flexible Wall Permeameter Test Station . . . . . . . . . . . . . 64 11. Gradations Used in the Industry-wide Permeability Algorithm . . . . . . . . 72 12. Typical Vibratory Table Test Result . . . . . . . . . . . . . . . . . . . . . . . . . 77 13. Compaction Curves for MHTD Base Rock . . . . . . . . . . . . . . . . . . . . . 78 14. Typical Bulk Stress - Resilient Modulus Relationship . . . . . . . . . . . . . . 79 15. Relationship Between Experimentally Derived Factors (k1 and k2 ) for

the Theta Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84 16. . Effect of Degree of Saturation and Aggregate Source on Resilient

Modulus ............................................ 85 17. Effect of Gradation, Degree of Saturation, and Compactive Effort on

Resilient Modulus . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86 18. Typical Constant Head Rigid Wall Permeameter Test Result . . . . . . . . . 89 19. Relationship of Porosity and Effective Porosity . . . . . . . . . . . . . . . . . . 93 20. FHWA 0.45 Power Paper Plot of Experimental Gradations . . . . . . . . . . 94 21. Plot of Individual Percent Retained for NJ and OGS Gradations . . . . . . 95 22. Relationship of Porosity and Permeability . . . . . . . . . . . . . . . . . . . . . 97 23. Relationship of Effective Porosity and Permeability . . . . . . . . . . . . . . . 98 24. Relationship of Observed Permeability and Estimated Permeability for

Open-Graded Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99 25. Relationship of Observed Permeability and Estimated Permeability for

Dense-Graded Materials .................................. 103 26. Relationship of Observed Permeability and Estimated Permeability for

Several Studies ....................................... 111 27. Relationship of Observed Permeability and Permeability Estimated by

Moulton Equation ..................................... 112 28. Relationship of Observed Permeability and Permeability Estimated by

the UMR Equation ..................................... 113 29. Variation of Subgrade Resilient Modulus Through the Year ......... 127 30. Six Climate Zones in the United States ....................... 129 31. Average AASHO Road Test Cross-Section .................... 138 32. Relationship of Road Test Resilient Modulus and Deviator Stress for

Three States of Moisture Content .......................... 142 33. Resilient Modulus Seasonal Variation With Variations in Base and

Subgrade Drainability ........ . .............. . ........... 145

xiii

C1. Typical MODAMP Screen ................................ 190

I I I I

I

1.

I 2.

3.

4. 5. 6. 7. 8. 9. 10. 11 . 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23.

24. 25. 26. 27. 28. 29. 30. 31. 32. 81. C1. C2.

xiv

LIST OF TABLES Recommended mi Values for Modifying Structural Layer Coefficients of Untreated Base and Sub-base Materials in Flexible Pavements . . . . . 13 AASHTO Quality of Drainage Using Calculated Base and Subgrade Drainability Values . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25 Percentage Index (PD) of Free Draining Water for Different Types of Base Course . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29 Material Types and Sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 Test Sequence for Granular Specimens of Base/Subbase Material . . . . . 43 Testing Variable Scheme . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44 As-Received Gradations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68

· Experimental Gradations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 Gradation Shape Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70 Experimental Gradation Slopes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 Usefulness of Individual Particle Sizes in Prediction of Permeability . . . . 73 Specific Gravity and Moisture Density Data . . . . . . . . . . . . . . . . . . . . 74 Particle Shape/Texture Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 Atterberg _Limits of the Base Materials .................. · ..... 76 Resilient Modulus Test Data ................... ·. . . . . . . . . . . 82 Statistical Significance of Testing Variables to Resilient Modulus . . . . . 87 Results of Rigid Wall Permeameter Permeability Testing . . . . . . . . . . . 90 Typical Set of Data for a Rigid Wall Permeameter Test . . . . . . . . . . . . 91 Results of Flexible Wall Permeameter Permeability Testing . . . . . . . . . 100 Effect of Material Variables on Permeability . . . . . . . . . . . . . . . . . . . 104 Data Used in the Permeability Predictive Equation . . . . . . . . . . . . . . 107 Drainage Coefficient Sensitivity Anaysis for DAMP . . . . . . . . . . . . . . 119 Recommended Drainage Coefficients for Flexible Pavements for Untreated Base and Subbase Materials . . . . . . . . . . . . . . . . . . . . . . 125 MODAMP Quality of Drainage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 125 Climate Condition Season Lengths . . . . . . . . . . . . . . . . . . . . . . . . . 128 Zone - Climate Condition Relationships . . . . . . . . . . . . . . . . . . . . . . 130 Determination of Climate Condition for Several Missouri Sites . . . . . . 131 Required Permeabilities for Quality of Drainage Levels . . . . . . . . . . . . 132 Quality of Subgrade Drainage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 135 Input Values for KENLA YER Analysis . . . . . . . . . . . . . . . . . . . . . . . 140 Drainage Coefficient Sensitivity Analysis for MODAMP . . . . . . . . . . . 149 Thickness Sensitivity Analysis for MODAMP . . . . . . . . . . . . . . . . . . 150 Subgrade Drainability Input Information . . . . . . . . . . . . . . . . . . . . . . 184 Quality of Subgrade Drainage . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191 Required Drainage Times for Quality of Drainage Levels . . . . . . . . . . 192

I I

I

I

I

GENERAL

1

INTRODUCTION

The 1986 AASHT0 1 Guide (1) recommends that consideration be given to

the inclusion of the concept of pavement drainage into the design of pavement

structures. The benefits of positive drainage of pavements is well documented in

the literature and there seems to be an increasing trend toward the use of

internally drained pavements.

Seeds and Hicks (2) and the AASHTO Guide list the moisture-induced

pavement problems associated with lack of pavement structure drainage. These

include asphalt stripping, loss of asphalt stiffness, unbound granular base strength

and stiffness loss, erosion of cement-treated base, subgrade strength and stiffness

loss, and subgrade distress induced by volumetric change.

Mathis (3) and Mann (4) have summarized the trends in the use of various

types of pavement drainage designs. Both stabilized and unstabilized drainable

bases are increasingly being used. The non-stabilized materials tend to be 4 to 6 in

thick with a dense-graded subbase underneath, which acts as a filter to prevent

contamination of the open-graded base by the subgrade. The drainage base

aggregates are usually crushed and include some finer fractions for stability under

construction traffic.

This report involves the determination of AASHTO pavement design method

drainage coefficients for several highway materials commonly specified by the

Missouri Highway and Transportation Department (MHTD). The study was made

1 American Association of State Highway and Transportation Officials

2

at the request of the MHTD Research Advisory Committee. The project was

executed by personnel from the University of Missouri-Rolla (UMR) Department of

Civil Engineering.

Based on the results of the AASH02 Road Test, a pavement design method

has been developed. This is commonly known as the AASHTO method(1 ). In the

application of the method, the highway designer determines a "structural number"

(SN) by knowledge of such factors as designed traffic level, subgrade support,

desired reliability, and desired terminal serviceability. The magnitude of the SN

reflects the degree to which the subgrade must be protected from the effects of

traffic. For example, a relatively high SN would indicate that a thick or stiff

pavement structure would be necessary to protect the subgrade from failing or

causing pavement structure failure. Once the SN is calculated, it becomes

necessary to determine the manner in which the SN will be achieved, i.e., what the

required thicknesses and quality of each pavement layer should be. This is done

by solving the following equation:

where:

SN = structural number

a1,a2,a3 = layer coefficients for the surface, base, and subbase layers,

respectively

drainage coefficients of the base and subbase, respectively.

2American Association of State Highway Officials

I

3

D1 ,D2,D3 = thickness of surface, base, and subbase layers, respectively.

Drainage coefficients are essentially modifiers of the layer coefficients, and

take into account the relative effects of pavement structure internal drainage on

performance of the pavement.

Determination of the layer coefficients is addressed under a separate

contract in a second report submitted concurrently with this study (5).

Examination of Eq. 1 indicates that the thickness of any particular layer is,

to a significant extent, dependent upon the layer drainage coefficient. Hence, an

accurate determination of drainage coefficients can have a significant economic

impact in regard to the design of the pavement structure .

As originally used in the AASHO Road Test results, layer coefficients were

actually regression coefficients which were the result of relating layer thicknesses

to road performance under the conditions of the Road Test. The problem is to

translate the Road Test findings to other geographic areas where the construction

materials and climate are different. Drainage coefficients must be determined in

order to use Eq. 1 for design purposes.

It should be noted that by definition the m2 and m3 coefficients only address

the effects of drainage in the unbound granular base and subbase, respectively .

They do not address the effects of moisture in the asphalt-bound layer(s), other

stabilized layers, or the subgrade. The effects of moisture in the subgrade should

be addressed during calculation of the effective subgrade modulus in the design

phase of a given project.

4

OBJECTIVES AND SCOPE

This study entailed the determination of flexible pavement drainage

coefficients (m-coefficients) for MHTD materials. This included procurement of

existing soil, pavement material, and climatological data, the performance of

drainability and moduli testing for four unbound granular base materials, and

analysis of an existing method of m-coefficient determination to evaluate the effect

of the above factors on pavement performance. The method developed by

Carpenter (6) was evaluated and is termed herein the "DAMP" method. Also, a

study performed at the Texas Transportation Institute (7) in regard to moisture and

temperature effects beneath pavements looked promising in regard to being

adaptable to the determination of drainage coefficients. This method is referred to

as the TTI method. A third method was also developed as a part of this study. A

recommendation was to be made as to the choice of method. The report includes

a method suitable for use in routine design which will enable the designer to solve

Eq. 1 and hence obtain the desired layer thicknesses.

I I

I I

DATA PROCUREMENT

GENERAL

Five types of existing data were necessary to evaluate the m-coefficient

methods of determination. These were 1) routine soil properties and location, 2)

soil and material thermal/moisture-related properties, 3) climatological data, 4)

pavement material structural properties, 5) unbound granular material drainability

properties, and 6) MHTD typical pavement geometry information.

SOIL SURVEYS

5

Routine soil properties suitable for classification purposes were obtained

from USDA county soil maps. In the future, this material can be supplemented

with data from MHTD construction projects as necessary and if available. These

data were used in the estimation of subgrade drainability for any particular site and

were used in both the DAMP and TTI methods.

MATERIAL PHYSICAL, THERMAL, AND MOISTURE PROPERTIES/DATA

Certain material physical, thermal, and moisture properties and data were

necessary as input for the TTI method. These values were located in the literature

rather than actually obtained from testing the materials. From climatological data,

I moisture available to the soil in any given area was calculated.

I

CLIMATOLOGICAL DATA

Climate moisture availability was necessary for use in the DAMP method.

These data were in the form of mean monthly temperatures, mean monthly rainfall

data, and latitudes for various areas across the state. The TTI model required

additional data: mean monthly wind speed, averages of monthly maximum and

minimum air temperatures, number of wet days per month, number of

thunderstorms, and percentage of sunshine. U.S. Weather Bureau Data Summary

Sheets were the source of such information.

UNBOUND GRANULAR MATERIAL DRAINABILITY PROPERTIES

In the development of the m-coefficients, it was necessary to compute

drainage times for various unbound base materials in given situations. The data

required for calculation of drainage times includes two laboratory-derived

properties: permeability and effective porosity.

6

Four sources of aggregate were tested for their permeability and effective

porosity characteristics and various index properties. The sources of aggregate

included two crushed stones and two gravels representing various particle shapes.

Each of the four aggregate types had one gradation prepared with an amount of

fines as allowed in the MHTD section 208 specifications, and two open-graded

gradations for a total of three gradations per source. For each gradation, the dry

unit weight was determined. The specific gravity, liquid limit, and plasticity index

of each of the four aggregate types was also determined.

PAVEMENT STRUCTURAL PROPERTIES

Resilient modulus data for each of the pavement layer materials were

necessary as input for the TTI method. This information for asphalt surface and

bituminous base mixes and for the dense-graded unbound base were obtained from

the companion project that was executed by UMR for the MHTD, which deals with

the determination of AASHTO layer coefficients. The resi lient modulus (E.gl values

for all possible subgrade soils came from estimates based on group index

I I

I

I

7

classifications derived from county soil maps or other sources.

The resilient modulus (E5g) of unbound granular base materials was

necessary in computing m-coefficients. In conjunction with the study that was a

companion to this report, one open and one dense gradation were tested using

each of the four aggregate types as mentioned in the previous section. Testing at

two degrees of saturation was valuable in assessing the moisture sensitivity of

these materials. This information was helpful in the determination of the m­

coefficients developed in this study.

PAVEMENT SECTION GEOMETRY

To assess the hydraulic capacity of the drainage layers, it was necessary to

have information regarding typical cross grades, ranges of longitudinal grades, layer

thicknesses, and pavement widths. This information was obtained from the

MHTD.

I

I

8

ALTERNATIVE METHODS FOR DETERMINATION OF DRAINAGE COEFFICIENTS

GENERAL

Two methods were explored for possible use in the determination of m­

coefficients. The DAMP method is an adaptation of the Moisture Accelerated

Distress (MAD) identification system which was published by FHWA (8). It is also

a mod_ification of the basic method in the AASHTO Guide (1 ). Carpenter

postulated that the MAD system could be adapted for use in determining AASHTO

drainage coefficients. The other method that the UMR project team evaluated was

an integrated computer model for the estimation of moisture and temperature

effects on pavements. This program was developed by the Texas Transportation

Institute. It was hypothesized at UMR that there might have been some promise in

adapting this integrated model to the problem of determination of drainage

coefficients. Because layer (a) coefficients can be related to resilient modulus

values, and because drainage (m) coefficients are modifiers of a-coefficients, then

m-coefficients could simply become a ratio of base material modulus (adjusted for

differences in response to environmental effects) to a normal unadjusted base

modulus. These environmental effects could even be site-specific because of such

localized effects as rainfall, temperature, solar radiation, soil type, topography, and

so forth . Both methods required laboratory testing of granular base materials and

the location of soil survey information and certain climatological data. The DAMP

system has the advantage of simplicity and requires less input when using it. The

TTI method offers the possibility of being more accurate because it considers

actual changes in the pavement subgrade and structure .

9

1986 AASHTO METHOD

General

Appendix DD of the 1986 AASHTO Guide ( 1) describes the development of

the drainage coefficients to be used in the 1986 flexible and rigid pavement design

procedures. Seeds and Hicks (2) also described the development of the drainage

coefficients, couched in the same words. The presumption is that Seeds and Hicks

were the authors of Appendix DD of the 1986 Guide.

Standards for Quality of Drainage

Seeds and Hicks discussed the standards for quality of drainage and suggest

the following time-to-drain to a degree of drainage of 50% (they incorrectly termed

this as time to reach 50% saturation).

Quality of Drainage

Excellent

Good

Fair

Poor

Very Poor

Recommended (hrs)

2

24

168

720

Does not drain

No discussion was provided in the paper that might allow the reader to evaluate

the basis for the recommended times-to-drain. Also, the categorization of the

Quality of Drainage ("Excellent", "Good", "Fair", "Poor", "Very Poor") was not

discussed, except that calculations for the Road Test indicated a time-to-drain to

50% degree of drainage as about 120 to 240 hours. A reference was made to the

FHWA Highway Subdrainage Design manual by Moulton (9), but Moulton is silent

I

I

I I

I I

10

on the topic. Others, such as Carpenter (8) have suggested the following time-to-

drain to 85% saturation for heavy pavement structures (fL_g., interstate) with

significant truck traffic:

Quality of Drainage

Satisfactory

Marginal

Unacceptable

Time-to-Drain Calculations

Recommended (hrs)

<5

5-10

>10

Seeds and Hicks present a table (Table DD.1 in the 1986 Guide) that

summarizes the results of calculations of time-to-drain a base layer to a purported

50% saturation. The material properties of the base are said to be taken from the

AASHO Road Test materials. The FHWA Highway Subdrainage Design manual by

Moulton is referenced as the method used to calculate the values in the table. It is

clear, however, that Seeds and Hicks have calculated the time-to-drain to a degree

of drainage of 0. 50 rather than 50% saturation. Also, the table values for porosity

are obviously effective porosity, not total porosity.

Table DD.1 lists 10 days to drain a 12 in. thick base 12 feet wide with a

porosity of 0.015 to 50% saturation. Actually, this base would take approximately

255 hours to drain to 96% saturation when calculated using Moulton's procedures.

This includes the assumption that 0.015 was effective porosity, not porosity. It

never would reach 50% saturation or even 85 % saturation unless subjected to a

prolonged time where air drying could occur. However, a degree of drainage of

0.5 would be obtained in 255 hours or approximately 10 days. Also, for the given

11

permeabilities, the "porosity" values correspond to effective porosities in Moulton's

Fig. 30, which shows the permeability--effective porosity relationship. Thus, the

conclusion is that the table is for 0.5 degree of drainage, not 50% saturation, and

that the "porosity" values are actually effective porosity.

Flexible Pavement Drainage Coefficients (m)

Introduction. The development of drainage coefficients for flexible pavement (m­

coefficient) is discussed in Appendix DD of the 1986 AASHTO Guide. If the base

course layer coefficient (a) is multiplied by some factor (called drainage

coefficient), there is a corresponding increase or decrease in the thickness of the

pavement layer while the structural number (SN) is maintained as a constant (Eq.

1). Seeds and Hicks plotted the change in surface thickness so obtained vs the

assumed drainage coefficients. This plot can be called the "SN constant plot."

Relation to Quality of Drainage. Seeds and Hicks then looked at a method to relate

drainage coefficients to quality of drainage of the granular base course. They

selected a theoretical mechanistic analysis. AASHO Road Test material data was

used to establish the asphalt concrete modulus = 500,000 psi, the aggregate base

modulus = 30,000 psi, and the roadbed soil modulus = 3,000 psi. These

conditions were said to correspond to a drainage coefficient of 1.0. These

modulus values were entered into a public domain multilayered elastic system

analysis computer program called ELSYM5 ( 10). Surface deflections were

calculated by ELSYM5.

The base modulus was then set at 10,000 psi, 20,000 psi, and 40,000 psi

and the surface thickness was adjusted to maintain the same surface deflection .

I

12

This yielded a set of incremental changes in surface thickness associated with

each base modulus. These constant surface deflection incremental changes were

used to enter the SN constant plot to obtain drainage coefficients associated with

the different base modulus values.

A graph of the base moduli vs their associated drainage coefficients

demonstrated a nearly straight line. This line was extrapolated to 50,000 psi to

obtain the following values:

Base Modulus

50,000 psi

40,000 psi

30,000 psi

20,000 psi

10,000 psi

Drainage Coefficient

1.4

1.2

1.0

0.7

0.4

Effects of Varying Moisture Levels. An attempt was made to quantify the effects

of varying moisture levels that may occur over the course of a calendar year. The

discussion on this topic was brief and is quoted in its entirety. "However, it is

recognized that these values would vary also with the percent of time the

pavement structure is exposed to moisture levels approaching saturation. Fig. 8

summarizes the approach for considering the variation in m-value with percent of

time the structure is in or near a saturated condition." Their Fig. 8 is included in

I this report as Fig. 1 .

I

This figure was stated to be the source of the m-value table presented in the

1986 AASHTO guide presented herein as Table 1. However, interpretation of the

table as presented could yield a different m-value than one obtained from Fig. 1.

Table 1 . Recommended mi Values for Modifying Structural Layer Coefficients of Untreated Base and Sub-base Materials in Flexible Pavements.

Quality of Percent of Time Pavement Structure is Exposed to Moisture Drainage Levels Approaching Saturation

13

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

One is left to speculate on the source of Fig. 1 or the reasons for the

categories chosen to differentiate the percent of time the pavement structure is

exposed to moisture levels approaching saturation.

In the development of the m-coefficients, a drainage time of 255 hours

corresponds to a quality of drainage between "Fair" and "Poor" on page 00-2 of

Appendix DD. Yet on page 00-12, for a modulus of 30,000 psi (assumed value of

Road Test base material), the quality of drainage is listed as "Fair". Further, in

Table 1, for m = 1.0 and "Fair" drainage, the percent time of saturation would have

to be between 1-5% and 5-25%. As will be seen in the next section, Table 4.1 in

the AASHTO Guide puts the Road Test time of saturation in the "over 25%"

catagory. Here, for m = 1.0, drainage is shown to be "Good". This lack of

agreement presents a problem when trying to compute the m-coefficients in a

given locale in comparison to the m-coefficient (1.0) at the Road Test. Was the

D') -+' C a,

E C ., 0 >

C +'

Cl. C L.

L. :,

0 +' C - (/'J

""C a, .c

I +' °' 0 ·-a, :r: L. L. 0 0 ... u

CD L. :, OJ ., 0 :, a. - X C

> Lu I E

I

2.0

Quality of

1.5 Drainage

Excellent

Good

1.0 Fair

Poor 0.5

Very Poor

0.0 1 1-5 5-25 >25

Percent of Time Structure is Near Saturation

Fig.1. Recommended m-Values cs a Function of

the Quality of Drainage and Exposure to

Sctu ration.

14

Road Test drainage Good? Fair? Poor? This is discussed further in the next

section.

DAMP

General

15

Carpenter has written an interactive computer program (Drainage Analysis

and Modelling Program (DAMP)) (11) designed to perform a drainage analysis of

pavement structures. DAMP basically takes the FHWA Highway Subdrainage

Design Manual (9) written by Moulton and computerizes the calculations.

Carpenter added sections on recent geocomposite fin drains and filter fabrics and

procedures for selection of m-coefficients. For m-coefficient determination, DAMP

considers base drainage capacity, subgrade drainage capacity, and climatological

data. The data necessary for these considerations include: subgrade soil drainage

characteristics, granular base width/thickness/cross-slope/longitudinal

grade/density/effective grain size/percent passing the #200 sieve, average monthly

precipitation and temperature, and latitude. Additionally, DAMP considers surface

infiltration, meltwater, roadway geometrical inflow/outflows, edge drain capacities,

and filtration criteria. The data necessary for these calculations include: weather

data, pavement type/thickness, transverse joint or crack spacing, number of

longitudinal joints, transverse joint or crack length, number of layers in the

pavement, thickness and density of each layer, permeability of subgrade, heave

rate of subgrade, and cross sections of the roadway right-of-way.

DAMP has the capability of performing many sorts of drainage-related

activities. However, in terms of the selection of drainage (m) coefficients, only the

I

I

I I

I

I

16

topic titled "Drainage Coefficient Determination" is presented.

Drainage Coefficient Determination

Drainage Layer Characteristics and Base Drainage Times. The permeability of the

drainage layer is computed by DAMP using the relationship from Moulton which

was based upon permeability tests reported in the literature ( 12-17). This

relationship is:

6.214x105(D101·478)(n 6·654)

kd = . . . . . . . . . . . . . . (2) p2000.597

where:

kd = permeability of the drainage layer, ft/day

D10 = drainage layer's effective (10 percent passing) particle size, mm

P 200 = amount of the drainage layer material passing the #200 sieve, %

n = porosity of the drainage layer material.

where:

n = 1 Yd - -- ......... . .......... (3)

YwG•

where:

yd = compacted dry unit weight, pcf

G. = apparent specific gravity

Yw = unit weight of water.

Fig. 2 depicts the nomograph that Moulton provided in his manual.

DAMP calculates time for drainage using the relationships developed by

., > ., ii

) 8 •

N 0 z bO

C

: ,

'iii ... Ill

0..

-I

C: ~

10

., 0.. I

'° 8 N

0.

.

1.4

78

6

.654

6.2

14

xl0

5(0

10)

(n)

(ft,

/day

) k

• p

)0.5

97

(

200

yd

n

• P

oro

alt

y

• (l

-

62

.4G

)

G •

S

pecif

ic

Gra

vit

y

(Aae

umed

2, 7

0)

-----

-----... '·"

" O

OI

E

E

ooi

- ., N

O 1

0 V

) .c: Ill

0~

t3 ., _

> --

·-e -

-_

_ ,_

to

~

w

,., 0

,~,

.....

a

10.

0

.. IO

100

'10

u... :,

u ' vi .D

-_

_ .j.,

;;-.~

----

"' C: C1I

a ----

------

Exa

mpl

e:

p200

2X

010

0,6

rra

n

yd

117

lb./

cu

.ft.

Rco

d:

k •

65

ft./

day

\JO

~

uo

a ·, '-J

Fl g

. 2.

Mo

ult

on

N

om

ag

rap

h

for

Es

tim

ati

on

o

f B

as

e

Co

urs

e

Pe

rme

ab

ilit

y

(aft

er

Mo

ult

on

).

' ,, ' 10

, IO

.., " .,

:,... "' 'O

':-

:,...

.D "' ., E

~ ., C

L 0 C: ., '=

I(•

CV

0 u I

,~· l

.:J

C

.J

"' ... "

.... -.J

I I I I I I

I I I

I

18

Casagrande and Shannon ( 18). The assumptions made for this analysis included

1) the subgrade is impermeable and 2) the base course is saturated when drainage

begins . To facilitate the solution, they defined three dimensionless quantities, U,

T, and S. The degree of drainage, U is:

U = Drained cross sectional area ( of drainable voids) Total cross sectional area (of drainable voids)

. . . . . . (4)

The time factor, T is:

tkdH T = -- ......••.•..•..•...•. (5)

n L2 B

where:

t = time for drainage for U to be reached, days

kd = permeability of the base course, ft/day

H = thickness of the drainage layer, ft

L = length of the drainage path, ft

= w/1 +(gfsJ2

where:

w = width of drained area on same cross slope, ft

g = longitudinal grade

sc = cross slope

c = geometrical constant (defined later).

and from Strohm, et fil. ( 17):

19

n,, = v.VWD = 1 - [ yd (1 +GsWJ] . . . . . . . . . . . . (6) T G8 *62.4

where:

n8

= effective porosity

Vwo = volume of water drained from the sample

. Vr = total volume of the sample

yd = dry density, pcf

G5 = apparent specific gravity

W8

= water content of the sample after 24 hours of drainage, %.

DAMP, however, determines effective porosity (n8 ) by means of a statistical

correlation with measured permeabilities (kd) published by Moulton which was

based upon work reported by Barber ( 13) and Strohm et al. ( 17):

n,, = 0.027 kJ·234

The slope factor S is:

(7)

H H 5 = Ltancx = LS · · · · · · · · · · · · · · · · · · (B) d

where:

H = thickness of the base course, ft

L = length of the drainage path, ft

Sd = slope = (Sc 2 + g2)0.5

a = angle of the drainage path with horizontal, degrees.

Using these three dimensionless coefficients, DAMP computes the time

I

I I I I

factors with the following Casagrande-Shannon relationships and geometry.

for U > 0.5:

T = c[S + S In 2S-2US+1 _ s2 In S+1] 2 (2-2~(S+1) S

and for U ::5 0.5:

20

(9)

c 2 S+2U T= -[2US - S In ] .............. (10) 2 S

I where all notation has been previously defined except c. From model tests,

I I I I I

I I I

Casagrande and Shannon found c to be:

C = 2.4 - 0.8

TS ( 11)

Time-to-drain (t in days) is found from Eq. 5 knowing kd, H, ne, L, and using

a T (time factor)based upon the U (degree of drainage) necessary to achieve the

design degree of saturation.

The relationship between degree of drainage and percent saturation used in

DAMP is:

u = v. (1-°7o;at) ................ 1121

ne

and

V -(n ~ Deg Sat = v 8 x 100 . . . . . . . . . . . . . . ( 13)

Vv

where:

21

V v = volume of voids in the drainage layer

Deg Sat = percent of saturation selected for the design, %.

DAMP calculates time-to-drain to both 85% saturation and 0.5 degree of drainage

(U).

There is a significant difference in the philosophy of acceptable behavior

between the AASHTO Guide and DAMP. The drainability acceptance criteria in the

Guide is based on t 50, the time to drain to 50 percent of the drainable water. The

acceptance criteria in DAMP is based on the time to reach 85% saturation. For a

dense graded base, of all the void space, the amount of drainable water may be

very small. So, 50 percent drainage of this may still result in a very high percent

of saturation. Percent saturation is based on the percent water remaining in the

total voids. It would seem that, because the behavior of pavement structures

depends on the percent of saturation rather than degree of drainage, the criteria in

DAMP is more realistic.

Percent Time of Saturation. DAMP accounts for moisture in the pavement

structure using a concept published by Thornthwaite (19) to classify climatic

regions in a rational manner. The indices in his system are calculated from:

monthly average temperatures (°C)

monthly average rainfall (cm)

North latitude (degrees)

The index of interest here is potential evapo-transpiration (PET) which

quantifies the amount of moisture that could be given up by the soil in the selected

time period. This value is compared to the amount of rainfall and change in soil

11 I

I I I

22

moisture storage during the same time period. This presents a measure of time

that excess moisture conditions are present and leads to a percent saturation

estimate for use in the selection of drainage coefficients.

Thornthwaite's procedure as utilized by DAMP begins with calculation of the

I annual heat index (I) using local weather data:

I '= ~2 (;>1.514 . . . . . . . . . . . . . . . . . . (14)

I I I

where:

T = monthly average temperature (°C).

Next, DAMP calculates the constant "a":

a = 0.000000675(~ 3 - o.oooon1 (~ 2 + o.01792(~ + 0.49239 (15)

I Then the normalized monthly (30 days of 12 hours of sunlight) PET is found:

I I I I I I I I I

T " . PET= 1.6 (10-) . . . . . . . . . . . . . . . . . (16) I

This normalized value of PET is corrected to reflect the actual number of

days and number of hours of sunlight per day for each month in the year. The

difference between the latitude - corrected PET value and rainfall data is used by

Thornthwaite to identify months with surplus moisture conditions. Thornthwaite

accounted for moisture stored in the soil by assuming the soil could accept a

maximum of 10 cm of rainfall and hold this moisture in storage until needed to

support evapotranspiration. Thus monthly water surpluses or deficits are not

generated until the storage is satisfied.

23

DAMP approximates percent saturation time from the monthly water surplus

and storage data using the following scheme:

1 . Frozen period - When the average monthly temperature is below freezing, no contribution to saturation can be made regardless of moisture conditions. 2. Surplus period following the winter, Zone A (northeast portion of USA) - The soil will be continually saturated during this period, with a contribution from frost heave. In Zone A all months with a surplus following a winter period will contribute to the saturation time. 3. Surplus period following the winter, Zone B (zone below A) - Here the spring thaw phenomenon is not critical. There may be periods where the soil is not totally saturated, and these may accurately correspond to dry days in the "rain and dry" day sequences. In Zone B, include the first month and one-fourth of remaining months having surplus as contributing to the saturation time . . 4. Surplus following a recharge which does not follow a frozen period - Here one-fourth of the months in the surplus period should contribute to the saturation time. 5. Utilization period following a surplus - During this period, the evaporation potential exceeds rainfall, and the storage moisture is being depleted. The soil is going from saturated to a dry condition over the period. The initial month may have a portion during which it is close to saturation. Include one-fourth of all months during this period which have a storage exceeding 7.5 cm, representing wet months with rain. 6. Utilization after a recharge which did not lead to a surplus - None of the time in this moisture state contributes to saturation time. 7. Recharge leading to a surplus - The final months when the storage moisture is close to full (saturation) may contribute to saturation. Include one-fourth of all months which have storage values above 7.5 cm. 8. Recharge leading to no surplus - This period will contribute little moisture. Include none of the time during this period toward saturation time. 9. Deficit - During this time, there is no water available at all, and if there were to be rainfall, it could evaporate before entering the soil. Include none of this period in the saturation time.

The sum of the months that DAMP includes for saturation purposes is

divided by 12 to determine a "Percent Time of Saturation" for the given climate.

Subgrade Drainage. The source for soil drainage classification of subgrades is the

I

I 24

County Soil Survey produced for most counties under the direction of the Soil

Conservation Service. The county soil surveys contain several categories of data

related to soil drainage (20) such as runoff, internal soil drainage, and soil

permeability. This information is combined with knowledge of the underlying

geological formation, slope of the land, and location of the soil with regard to

elevation/position in the topography to determine the soil drainage classification in

the county soil survey that is used by DAMP. The soil drainage classification is

I given in each soil description (rather than in the many tables) in the county soil

I I I

I I I I I I

survey. Soil drainage classifications include very poorly drained, poorly drained,

somewhat poorly drained, moderately well drained, well drained, somewhat

excessively drained, and excessively drained.

DAMP accounts for the contribution that the subgrade makes to drainage of

the pavement structure by employing a concept introduced by Hole (21) termed

Natural Drainage Indices of Soil. Bodies (NDI). The NDI arbitrarily assigns the value

of + 1 to well drained soils, -10 to excessively drained soils, and + 2.5 to + 10 to

the range of soils classed as moderately well drained to very poorly drained.

DAMP classifies the NDI numbers as follows:

NDI Classification

-10 to <-2 Good

-2 to 2.5

> 2.5 to 10

Fair

Poor

These descriptive classifications are used to enter the quality of drainage table

discussed next. Quite simply, one uses the terms "Good, Fair, or Poor" to describe

25

the contribution of the soil to the Quality of Drainage. A soil classified as "Good"

will actually improve the performance of the granular layer, a "Fair" soil will not

augment the base layer drainage, but will not detract from performance, while a

"Poor" soil will actually provide a source of water to the structure.

Quality of Drainage. The drainage time of the granular base (time to 85 %

saturation) and the subgrade drainage (good, fair, poor) discussed above are used

to enter Table 2 to determine "Quality of Drainage".

Table 2.

Subgrade Drainability (NOi)

Good -10to-2

Fair -2 to 2.5

Poor> 2.5

AASHTO Quality of Drainage Using Calculated Base and Subgrade Drainability Values.

Base Drainability (to 85 % saturation)

Excellent Good Fair Poor Very Poor s. 5 hrs 5-30 hrs 30-100 hrs 100-200 hrs 200+ hrs

Excellent Excellent Good Fair Very Poor

Excellent Good Fair Poor Very Poor

Fair Fair Poor Very Poor Very Poor

AASHTO Drainage Coefficient Selection. The "Quality of Drainage" and "Percent

Time of Saturation" are the inputs to the 1986 AASHTO drainage (m) coefficient

table (Table 1 ). These drainage coefficients are then used in the AASHTO

structural number equation (Eq. 1 ).

Thus, the column is selected by going through the Thornthwaite PET

analysis, and the row is selected by combining the effects of base drainability (time

to 85% saturation) and subgrade contribution (good, fair, poor drainage).

I I I I I I I I I I I I I I I I I I

26

Carpenter also recommends that the drainage coefficient should be adjusted

up or down within the range in each cell of Table 1 depending on certain features

of the pavement structure. Increased m-values are allowed by the presence of

edge drains and a working drainage layer. Decreased m-values result from a

bathtub-type structure.

Further evaluation of DAMP is given in the section "Results of Models

Evaluation."

TTI

General

Titled the "Integrated Model of the Climatic Effects on Pavements" (7), this

method produced by the Texas Transportation Institute combines into a single

program several component models that had been developed independently. The

component models include the Precipitation Model, the Infiltration and Drainage

Model, the Climatic-Materials-Structural Model, and the CAREL Frost Heave-Thaw

Settlement Model. A flow chart of the integrated model is included in Fig. 3. The

component models were developed over a period of several years for use on main­

frame computers. In combining these models, the authors eliminated portions of

the original programs, substituted some methods of computation, and suppressed

some of the orginal output. The program was provided in a compiled form and as

such was not available for examination. It is not clear from the documentation

which links exist between the various programs and exactly what data is passed

from program to program.

PRECIP MODEL

Input 2 Pavement Geometry Physical and Thermal

Material Properties Initial Soil Suction Profile Initial Soil Temp. Profile Heat Transfer Coeff. Rainfall Intensity Coeff. Pavement Infiltration

Parameters

ID Model

CRREL MODEL

Output

Soil Temp. Profile with Time Soil Suction Profile with Time Frost Penetration with Time Thaw Depth with Time Surface Heave with Time Degree of Drainage with Tim Dry & Wet Probabilities of

Base Course . Adequacy of Base Course

Design

Output

CMS MODEL

Asphalt Stiffness with Time Base & Subbase Mod. with Time Subgrade Mod. with Time Climatic Data

27

Fig.3. Integrated Program Flowchart (after Lytton, et al.).

I I I I

I I

I I I I I I I I I I I

28

Precipitation Model

The Precipitation Model (PRECIP) was described by Liang and Lytton (22) as

a deterministic algorithm that uses recorded data to simulate rainfall patterns that

are used for infiltration and drainage calculations. Stochastic processes and

random methods are employed to analyze past climatological data, and to estimate

and predict the effects of the environment on the performance of pavement with

specified confidence levels. This description by the authors includes the statement

"the effects of the environment on the performance of pavement." However, the

model only produces a simulation of rainfall patterns to a given confidence level.

The model considers both convective and frontal types of rainfall. Convective

rainfall generally occurs in a brief intense thunderstorm. Frontal rainfall may be

steady and of longer durations. Long duration rainfall is associated with more

water entering the pavement structure. See Fig. 3 for the relationship of the

PRECIP program to the whole.

Infiltration and Drainage Model

The Infiltration and Drainage Model (ID) was written by Lytton and Liu (23)

and evaluates the pavement base course drainage, the probabilities associated with

the rainfall data, the infiltration analysis, and the resulting probabilities of having

either a wet or dry base course. That is, each day of the month is declared either

a day when the base is saturated to greater than 85% or a day when it is drier.

The drainage output describes the degree of drainage and corresponding times

using a model developed by Liu, Jeyapalan, and Lytton (24). This model is very

similiar to the Casagrande-Shannon model with the exception that the phreatic

surface is parabolic rather than linear. The Liu model also allows for subgrade

drainage. The pavement base course evaluation uses an empirical procedure to

include the percentage of gravel, sand, and fines and the type of fines in

determining the acceptability of the drainage design. This empirical procedure is

represented by the equation:

Sa = 1 - ( PD) * U . . . . . . . . . . . . . . . ( 17)

where:

Sa = degree of saturation

29

PD = percentage index which represents the drainability of the base course

material

U = degree of drainage.

PD is found from a table published by Carpenter in the MAD Index (8). This table is

shown below:

Table 3. Percentage Index (PD) of Free Draining Water for Different Types of Base Course.

Amt of Fines <2.5% Fines 5% Fines 10% Fines

Type Fines Inert Silt Clay Inert Silt Clay Inert Silt Clay

Gravel* 70 60 40 60 40 20 40 30 10

Sand** 57 50 35 50 35 15 25 18 8

* Gravel, 0% fines, 75 % > #4: 80% water loss * * Sand, 0% fines, well graded: 65 % water loss Gap graded material will follow the predominant size.

The probabilities associated with the rainfall data and the infiltration of that

I

I

I I I

I I I I I I I I I

30

rainfall into the pavement structure are combined with the drainage times to

produce an estimate of the probibility of the amount of time the base course is

saturated. The method for establishing the probability of having a wet or dry base

course involves several steps in which the probability of a number of events and/or

conditions are established and then are combined by multiplication, addition, and/or

subtraction to obtain the final result.

Climatic-Materials-Structural Model

The Climatic-Materials-Structural Model (CMS) was written by Dempsey, .et

.a.l. (25) and uses sunshine percentage, wind speed, air temperature, and solar

radiation to find the temperature profile in the pavement structure.

CAREL Frost Heave and Thaw Settlement Model

The CAREL Frost Heave and Thaw Settlement Model (CAREL) was written

by Berg, Guymon, and Johnston (26) and provides a measure of frost heave using

a coupled heat and moisture flow mathematical model. The CAREL model uses the

temperature profile found by the CMS model.

The Integrated Model requires a minimum of 100 input variables to operate.

These are listed in Appendix A.

The exact number of variables is dependent upon the number of layers of

the pavement system modeled. There is great flexibility in the Integrated Model

that allows accurate models of many different pavement structures.

The source of the input data depends upon the location and design of the

pavement structure. Given this information, the user can construct the necessary

finite element sketch to define the pavement layers. The weather data published

31

under the title of "Local Climatological Data, Monthly Summary" by NOAA is

sufficient to complete the weather data input. The users manual for the Integrated

program offers suggestions for much of the more obscure input requirements.

I I I I I I I I I I I I I I I I I I I

32

MATERIAL TYPES AND SOURCES

All unbound aggregates in the study were MHTD approved materials. The

materials were selected and sampled by MHTD personnel. Two Type 1 crushed

stone base aggregates were studied, and were selected by MHTD personnel to

give a wide range of particle shape and texture. Additionally, in a companion

project (29), two Type 2 gravel materials (re-graded to Type 1 specifications) were

tested for resilient modulus. Test results for these two materials are also included

in this report. The materials, sources, and identification are shown in Table 4.

Table 4. Material Types and Sources.

Nomenclature Material Sources Location

DR-12 Type 1 crushed Burlington Mertens Quarry Millersburg limestone

DR-13 Type 1 crushed Jefferson City Smith Quarry Rolla dolomite

DR-14 Type 2 Crowley Ridge gravel Delta Dexter base Aggregates

DR-15 Type 2 Black River gravel base Williamsville Poplar Stone Co. Bluff

Note: All sources are located in Missouri

GENERAL

33

LABORATORY INVESTIGATION

The principal properties to be determined for unbound granular base

materials were the 1) resilient modulus at a low and high degree of saturation to

assess moisture sensitivity, 2) permeability, and 3) effective porosity to assess

drainability. However, performance of other tests and procedures were necessary

in order to conduct the primary tests and to analyze the results. These other

procedures included sieve analyses, gradation formulation, specific gravity

determination, moisture-density relationship testing, particle shape/texture testing,

and plasticity of fines determination. These operations are outlined below.

EXPERIMENTAL GRADATIONS

The experimental gradations utilized in this study were: 1) a curve situated

midway between the upper and lower limits of the allowable gradation

specification band for MHTD Type 1 unbound base material ("MHTD Middle"); 2)

the New Jersey (NJ) open gradation; and 3) the PennDOT OGS open gradation.

All three gradations were used in the permeability portion of the study, while the

MHTD Middle and the New Jersey were used in the resilient modulus (Egl part of

the study. These gradations were used for both the two crushed stones and the

two gravels. At the finer size end, the MHTD Middle gradation was extended to

include a controlled amount passing the #200 sieve, which was 8%. This value

was chosen because: 1) it matched one of the gradations used in the layer

coefficient study which is the companion project to the present study; and 2) this

was approximately the same percentage as both as-delivered gradations of the

I I I I I I I I I I I I I I I I I I I

34

Type 1 aggregates supplied by MHTD for this project. The MHTD Middle, New

Jersey, and OGS gradations are shown in Fig. 4.

GRADATION CURVE SHAPE/POSITION

An analysis was performed to determine the effect of gradation upon

permeability and the effect of the interaction of gradation and degree of saturation

on resilient modulus. The most promising methods were later tried in the

development of the predictive permeability multiple regression equations. To

accomplish this, there was a need to characterize the gradations so that a single

value of gradation "modulus" would represent the shape and position of the

gradation curves. Nine different methods were tried and are described in detail in

Volume I of the companion study to this report (27).

PARTICLE SHAPE/TEXTURE

Numerous test methods have been devised to quantify particle shape and/or

texture. These can be divided into direct methods (those that result in

measurement or aspects of individual particle shape or texture) and indirect

methods (those that measure some sort of bulk aggregate property, such as void

content, which is related to particle shape/texture). Recent evaluations of these

methods were reported by Kandhal fil al, (28) at NCAT (National Center for

Asphalt Technology). There are several methods available which can be used in

lieu of the standard test, ASTM D 3398 (29), which is somewhat cumbersome to

perform. Kandhal fil al. recommended the National Aggregate Association's (NAA)

proposed method (A or B) for fine aggregate (30). Both of these are indirect

methods of particle shape determination.

35

I

I

100

90

80 CJ)

C 70 Cl)

Cl)

0 60 Q_

+' C

50 (l)

u L (l)

40 Q_

0 +' 30 0 I-

20

1 0

j /;, r

) w // I

/// /) '/

~ // V / ~ //

/ y (b"' ~

I

0 200 40 16 4 1/2 II 1 II 1 1 /2 11 .3 II

I 0 Middle Sieve Size • New Jersey

V OGS

Fig. 4. Semilog Plot of Three Experimental Gradations.

I

I I I I I I I I I I I I I I I I I I I

36

In this study, the (-) #8 to ( +) #100 sieve size material of each gradation

was tested using NAA Method A. The method is given in Appendix A of Volume I

of the layer coefficient study. For the ( +) #4 size, the aggregates were tested in

accordance with ASTM D 3398. This method is also given in the previously

mentioned Volume I. The results of both methods were used in developing the

permeability regression equations discussed later in the "Results" section of this

report. Photographs of the NAA test device and the D 3398 equipment are shown

in Figs. 9 and 10 of Volume I of the layer coefficient study.

SPECIFIC GRAVITY

Aggregate fractions of each of the three gradations were separated at the

#4 and #100 sieve sizes and tested in accordance with AASHTO T85-88 (31) and

T84-88 (32) for the ( + )#4 material and the (-)#4 to ( + )#100 material. These data

were necessary for use in the degree of saturation and porosity calculations.

Weighing was performed on a scale readable to the nearest 0.1 g. Weighted

averages of apparent specific gravities were used to calculate the specific gravity

for each gradation of each of the four aggregates as follows:

G=------10_0 ____ ~ % Passing #4 + % Retained #4 . . . . . . . . . . (18)

ASG ASG

where:

G = apparent specific gravity, weighted average

ASG = apparent specific gravity of each fraction.

SCREENING

All aggregates were shaken in an air dry state through the appropriate

screens in a Gilson shaker. A dust baffle/cover was designed to restrict the

movement of particles in order to minimize problems with incorrect sizes of

material being retained on any given sieve.

Upon shaking, the split material was stored in 20 gal plastic cans with lids

until the aggregate was needed for specimen fabrication.

SPECIMEN FABRICATION

37

Specimens for unbound granular base were fabricated as follows: Each

specimen was produced by taking the indicated amount of material for each sieve

size in accordance with the experimental gradation previously discussed. Each

layer was proportioned separately. The largest particle size in the gradation was

approximately 5/8 in. Thus, the specimen diameter was greater than six times the

maximum particle size, in accordance with AASHTO T-XXXC91 (33) (flexible wall

resilient modulus), and permeability method ASTM D 5084. For the triaxial-type

specimens (resilient modulus/flexible wall permeability), the 4 in diameter 8 in high

specimens were compacted in 1 in lifts with a Dayton air hammer in a split

aluminum mold lined with a nitrile rubber membrane. Various membrane materials

and thicknesses were tried, including 0.012 and 0.025 in latex and nitrile rubber.

It was found that during compaction the thinner membranes would tear, especially

the latex, even if two membranes were used. A 0.06 in thick nitrile rubber

membrane was the minimum that was sufficiently rugged . The AASHTO

specification limits membrane thickness to 0.08 in. A vacuum of approximately 20

I

I I I I I I

I

I I

I I I I I I I I I I I I I I I I I I I

38

in was applied to the specimen prior to removing the split mold. The specimen

was compacted directly on the triaxial cell pedestal. A more complete description

of the specimen fabrication process is included in Ref. 27.

The 1 O in diameter rigid wall permeameter specimens were also compacted

in 1 in lifts with the air hammer.

MOISTURE - DENSITY RELATIONSHIP

In order to choose target densities for compaction of resilient modulus and

permeability specimens, standard and modified proctor tests were performed in

accordance with AASHTO T-99 (34) and T-180 (35). Additionally, the maximum

density of the open-graded gradations were determined via the vibratory table

method, ASTM D 4253 (36). For each of the four aggregates, a double amplitude

~ dry density curve was obtained in accordance with the dry method to obtain the

optimum power setting. This power setting was then used for the determination

of the density utilizing the wet method. The vibratory table is shown in Fig. 3 of

Volume II of the layer coefficient study where a more complete description of the

test method and equipment is given.

RESILIENT MODULUS

General

The relationship between repeated applied stress and the resulting strain of

unbound granular base materials is most commonly defined by the resilient

modulus test. This test was performed by subjecting a compacted specimen to an

all-around confining pressure and then applying a vertical cyclic load. Total applied

load (a1 ), displacement resulting from the load, and confining pressure (a3 ) were

39

monitored. The applied load and confining pressure were varied to achieve a range

of stress states which should represent the expected stress states in actual

pavement structures. The specimen was encased in a flexible membrane and

tested in a triaxial cell. Fifteen combinations of confining (cell) pressure and cyclic

applied (deviator) stress were used for each specimen .

. The procedure that was followed in this study is essentially in conformance

with the 1991 Interim AASHTO method of test (33). The test procedure is also

essentially in conformance with SHRP Protocol P46 (37). One notable exception is

that the AASHTO stress state sequence (not the SHRP) was followed. However,

as per Claros fil fil. (38), a1/a3 ratios were not allowed to exceed three in order to

prevent possible excessive specimen straining.

A more complete discussion of the test equipment and procedure is given in

Ref. 27.

Equipment

The testing equipment setup is shown in Fig. 5. The equipment consisted of

an MTS electrohydraulic load system, a triaxial chamber capable of housing a 4 in

diameter specimen while subjected to cyclic loads, and a data acquisition system.

Load was measured with an internal 1000 lb capacity load cell and deformation

was measured with two L VDT'S mounted externally to the cell. This type of

measuring system is allowed in the AASHTO method and is recommended in the

SHRP method. Minimum resolution of the vertical LVDT's and the load cell met

the AASHTO standard. Actual minimum deformations and loads during the testing

were kept at least ten times the minimum resolutions to assure confidence in the

I I I

I I

I I

I I

I

I I

I I I

I I I I I I I I

40

Fig.5. Resilient Modulus Testing Equipment.

41

test results. Air was used as the confining fluid instead of water in order to

protect the internal load cell. Triaxial cell pressure and back pressure were

controlled via a Geotest control panel. The Research Engineering triaxial cell that

was used had several advantages. First, the chamber cylinder wall could be placed

after the loading piston is brought into contact with the specimen. Also, end caps I could be purged of air very easily by the unique design of the caps.

Test Variables

Although the effect of the degree of saturation was of primary interest, four

test parameters were controlled as independent variables.

Stress State. As previously mentioned, several variables affect the modulus of

granular materials. Stress state is considered to be the most important. As in

shear strength, the more confined a granular material is, the higher will be the

modulus. In the field, confinement is supplied by the layer underneath the granular

material, the granular material itself in the lateral (tangential and radial) (u2 and u3)

direction, the overburden above the point of interest, and the momentary load from

a vehicle. In a triaxial test, the difference between total vertical stress (u,) and u 3

is called the deviator stress or stress difference (ud). Cell pressure supplies the

lateral confinement to the specimen (u2 and u 3). A small static load (0.1 ud)

supplies the "overburden" pressure, and cyclic deviator stress (0.9 ud) supplies the

"vehicle" momentary stress. All of the stresses combined are known as the bulk

stress:

I

I I

I

I I I

42

= Ud + 3u3 •••••••••••••••••••••••••••••••• (19)

For each specimen, resilient modulus was determined at 14 stress states in which

effective confining pressure ranged from 2 to 20 psi and ud varied from 2 to 40

psi. This resulted in a range of bulk stress from 8 to 100 psi. This was considered

I adequate to cover the range of stress states likely to be encountered in practice.

I

I I I I I I

I I I I

The testing sequence and stress state schedule is shown in Table 5. Thus, E0 =

Degree of Saturation. In general, an increased water content will cause modulus

to decrease. Several reasons are given for explanation of this behavior. These are:

1) decrease in modulus of subgrade, thus a decrease in granular layer bulk stress;

2) reduction of apparent cohesion; 3) reduction of effective overburden pressure if

the base is below the water table; and 4) increase in positive pore pressure under

quick loads. Several degrees of saturation (0 S) have been put forth as break­

points in behavior. Base materials are considered to be relatively "dry" at degrees

of saturation 60 percent and less (39). AASHO Road Test granular base materials

suffered a marked increase in distress above 85 percent saturation. In the present

study, each material was tested at two degrees of saturation: approximately 60%

and 100%. This variable was explored with the idea that a change in modulus due

to changes in saturation could lead to the development of m-coefficients. Resilient

behavior has been shown to deteriorate above 80 to 90 percent saturation (40).

Degree of Compaction. As previously discussed, modulus generally increases with

higher levels of compaction. Two levels of compactive effort were evaluated for

each material and gradation. For the dense gradation, specimens were compacted

Table 5 . Test Sequence for Granular Specimens of Base/Subbase Material.

Phase Sequence Deviator u, Confining CT1/CT3 0 No. of No. Stress Pressure Repetitions

(c,d)(psi) * (psi) ** Specimen 1 15 35 20 1.75 75 1000 Conditioning

2 10 30 20 1.5 70 50

3 20 40 20 2.0 80 50

4 30 50 20 2.5 90 50

5 40 60 20 3.0 100 50

6 10 25 15 1.67 55 50

7 20 35 15 2.33 65 50

8 30 45 15 3.0 75 50

Testing 9 5 15 10 1.5 35 50

10 10 20 10 2.0 40 50

11 20 30 10 3.0 50 50

12 5 10 5 2.0 20 50

13 10 15 5 3.0 25 50

14 5 8 3 2.67 14 50

15 2 4 2 2.0 8 50

Note: 1psi = 6.9kPa * Cyclic loads = 0.9 c,d; constant contact loads = 0. 1 c,d ** For all stress states the minimum number of repetitions necessary is 50. The

maximum is determined as per the AASHTO procedure and was redetermined for each confining pressure.

to 100% standard and 100% modified proctor densities. For the New Jersey

gradations, in most cases, one level of compaction corresponded to the maximum

index density via vibratory compaction (wet method), while the second level of

density usually corresponded to an impact-type of compaction, such as 100%

43

I

standard proctor.

Particle Shape/Surface Texture. As stated earlier, the effect of particle

shape/surface texture is not well-defined. Two crushed stones and two gravels

were chosen to delineate the effect of particle shape/surface texture.

44

I Testing Scheme

The testing scheme involved the following variables: four sources of

aggregate, two compactive efforts, two gradations, and two degrees of saturation

I for a total of 32 "tests". Each test was run with duplicate specimens. The testing

scheme is shown in Table 6.

I I I

Table 6. Testing Variable Scheme.

Crushed Stone Gravel

DR-12 DR-13 DR-14 DR-15

Mid. NJ Mid. NJ Mid. NJ Mid. NJ

CEL 0 S = 60 X X X X X X X X

0 S = 100 X X X X X X X X

CEH 0 S = 60 X X X X X X X X

05 = 100 X X X X X X X X

Note: Mid. = middle of MHTD Type 1 gradation band NJ = New Jersey gradation CEL = lower compactive effort CEH = higher compactive effort 0 S = 60 or 100% saturated X = combination of variables was utilized

Test Procedure

The resilient modulus testing procedure involved the following steps:

45

specimen compaction; assembly of the triaxial cell; consolidation; specimen

conditioning at a given stress state; load application through 14 additional stress

states at 60% saturation; backpressure saturation to 100% saturation;

consolidation; and load application through 14 stress states at 100% saturation .

After the load application at 100% saturation step, the dense-graded specimens

were tested for permeability. As a final step, the specimens were allowed to drain

overnight in order to calculate their effective porosities.

The specimens were compacted in eight layers of equal thickness with a

hand-held air hammer. The material was compacted at the optimum moisture

content (which was about at 60% saturation) into a split mold. After cell

assembly and consolidation , the specimen was conditioned with 1000 repetitions.

The various stress states and loads were then applied as per Table 4. The number

of load applications varied from 50 to 200, depending on the number of

applications necessary to meet the AASHTO modulus repeatability requirements .

Load and deformation data were taken for every load application over the

entire sequence, but only the last five repetitions were used for calculation of

resilient modulus.

The load duration for each repetition was 0.1 sec followed by 0.9 sec rest.

The stress pulse shape was haversine in nature. Repeated load equipment

deflection was determined on an aluminum dummy specimen and was subtracted

from total deflections for each stress state. Initially, calibration of the load cell and

LVDT's was performed before each test, but the interval was increased upon

determining that the drift in calibration was insignificant. The change in specimen

I

I

I I

height was constantly monitored. None of the specimens approached the

maximum allowable permanent strain of 5%.

In an effort to determine the effect of drainability on pavement bases, the

tests at 60% saturation were performed in a drained condition while the 100%

saturation tests were run in an undrained state.

PERMEABILITY

General

46

The ability of a pavement structure to drain water rapidly is partly a function

of the permeability of the various layer materials of its structure. Thus, for proper

design, knowledge of the permeability of the materials, especially of the granular

base layer, is necessary.

Even though permeability is expressed in, say, ft/day (the same as velocity),

the two are not necessarily equal. The permeability "ft/day" is a contraction of ft3

per day/ft2 as derived from Darcy's Law:

0 = ki A ................................. . ........ . (20)

where:

0 = discharge, ft3

k = permeability, ft/day

i = hydraulic gradient, ft/ft

A = cross-sectional area of discharge, ft 2•

Rearranging, 0/ A = ki

and V = 0/A (V = velocity, ft/day)

thus V = ki

47

So, velocity is equal to the product of permeability and gradient. If the gradient is

unity, then V = k. But, at any other gradient, they are not equal. For many

granular pavement base situations, gradients are 0.2 to 1.0, so velocity is usually a

fraction of permeability.

Traditionally, the determination of permeability of granular base materials

involves the assumption that Darcy's Law is in effect. For this to be true, the

underlying assumptions are that the material is 100% saturated, the flow is

laminar, and the discharge is proportional to the hydraulic gradient. Additionally, in

laboratory testing, for proper flow conditions to exist, the largest material particle

should not be excessively large in relation to the total cross-sectional diameter of

the permeameter. Typically, testing procedures impose limits so that the maximum

particle size does not exceed 10 times the specimen diameter for rigid wall

permeameters, or 6 times the specimen diameter for flexible wall permeameters.

The permeability of granular materials is a function of pore size distribution,

pore continuity, and pore shape. These are affected by grain size distribution,

particle shape, and relative density. Permeability is also a function of the degree of

saturation, specimen mineralogical composition, and nature of the permeant. This

last factor is to a certain extent affected by the viscosity, unit weight, and

chemical composition of the permeant. For coarse grained materials under normal

circumstances, due to the large volumes of water involved, tapwater is used as the

permeant, and the interaction of particle mineralogical composition and permeant

chemical nature is considered negligible. Thus, in a practical sense, the significant

variables in permeability testing of granular materials are particle size gradation,

I I

I

I

48

particle shape, degree of saturation, relative density, mineralogical composition of

the fines (~, plasticity index) and permeant temperature (which affects the

viscosity and unit weight of the permeating water). Porosity (the ratio of volume

of voids to total volume) is related roughly to permeability, and is used in various

drainability algorithms. Porosity is a function of relative density, specific gravity

and , indirectly, particle shape. In general, permeability increases with a more open

gradation (less fine material, described variously as minus #4, minus #16, or minus

#200 sieve sizes), a more angular particle shape (although, at least one study has

shown the opposite to be true (41 )), a higher degree of saturation, a lower relative

density, a higher temperature (lower viscosity), a lower mineralogical activity (~.

a lower Pl) and, in a crude way, a higher porosity.

One of two types of testing methods are usually employed in the

determination of permeability in the laboratory: one in which the flow driving head

is constant, and, alternately, one in which it is variable (decreasing, or falling). The

constant head method is usually applicable to materials with permeabilities greater

than 2 to 3 ft/day (about 10 x 10·4 cm/sec). In constant head tests, the flow is

measured. For low permeability materials, the flow may be too small to measure

accurately, and thus the falling head method is employed. The dimensions of the

apparatus can be adjusted so the measurement of head and time can be carried out

over a range of permeabilities. However, the falling head test is more sensitive to

errors (such as small leaks) because of the small amounts of permeant involved.

Thus, the constant head method is preferred.

Two different permeameters are available for testing the permeability of base

49

material: rigid wall and flexible wall. Rigid wall permeameters are generally less

costly and less complex in operation, can handle relatively large flow rates, and

probably are the more common type. The major disadvantages are potential

leakage along the permeameter wall/specimen interface, the large sample size

required, the potential for difficulties in successful specimen saturation and the

limitation in available head that can be applied. The major advantages of flexible

wall permeameters are the ability to seal the permeameter wall/specimen interface,

the ability to back pressure saturate, and the ability to apply larger heads.

Disadvantages of flexible wall permeameters are complex operation, cost of large

(4 to 6 in) diameter triaxial chamber equipment, and difficulty in specimen

compaction. In general, the constant head test in a rigid wall permeameter can be

used successfully with open-graded specimens, while the constant head test in a

flexible permeameter is more applicable to dense-graded specimens.

Testing Concerns

Air Blockage. In some studies ( 15-17, 41), submergence of the specimen has

been the method of bringing the specimen to a saturated condition. Unfortunately,

for some gradations, it is nearly impossible to fill all voids using this method. Non­

water filled voids in the granular material are air filled. These bubbles tend to block

the flow of water, reducing the measured permeability. The air can be present in

an initially non-saturated specimen, or can be carried into the specimen by the

permeant, either as air bubbles or by air coming out of solution. Air can also

accumulate in the testing apparatus plumbing. Air accumulation can be detected

by plotting several successive permeability tests under identical conditions against

I

50

time. A drop in permeability indicates clogging of some sort, usually air. Solutions

to the problem include the use of de-aired water (42-43), initial vacuum saturation

(42-43), back pressure saturation (42-43), using warm water (15), and removal of

air in the specimen with CO2 prior to saturation ( 1 5).

Movement of Fines. Specimen particle segregation can occur during the

comp_action step, during vacuum saturation, or during testing. A segregated or

altered specimen may result in a significantly different measured permeability. To

prevent this from occurring, the specimen should be compacted at less than

optimum moisture content, the vacuum applied during initial saturation stages

should not be excessive, and the hydraulic gradient during testing should be kept

low enough so turbulent flow is avoided. Additionally, head loss through the

specimen should be measured by use of manometer tubes attached to ports which

are positioned to avoid the end portions of the specimen which may have a

disproportionate amount of fines, or some other end condition which renders that

area non-representative. Use of a flexible wall permeater will also reduce the

possibility of fines being washed up the side of the permeameter and out of the

specimen.

Excessive Gradients. Excessive gradients lead to turbulent flow, a condition which

should be avoided for several reasons. First, Darcy's Law no longer applies under

turbulent conditions. Second, high seepage pressures could lead to consolidation

of the specimen. And, turbulent flow may induce fines movement, as previously

discussed. For rigid wall permeameters, AASHTO T215 suggests allowable upper

gradient limits of 0.2 to 0.5, depending on gradation. For flexible wall

51

permeameters, ASTM D 5084 recommends using the gradient expected in the

field, which may range from less than 1 and up to 5. Recognizing that low

gradients will lead to very long testing times, D 5084 ties recommended gradient

to permeability: the lower the permeability, the higher the allowable gradient. For

materials with permeabilities of 0.3 to 3 ft/day (1 x 1 o·4 to 1 x 1 o·3 cm/sec), the

maximum recommended gradient is 5. Moulton says that, in a practical sense,

even at relatively low gradients, coarse graded materials may exhibit turbulent

behavior at gradients similar to those in the field. Thus, the coefficient of

permeability would not be a true Darcy coefficient, but would represent the desired

design situation for estimation of seepage flow. Excessive hydraulic gradients can

be detected by plotting discharge vs. gradient. Darcy's Law says that these two

variables are directly proportional and that permeability is the coefficient of

proportionality, or slope of the line. If at some point the slope begins to decrease

with increasing gradient, somet_hing is inhibiting flow. This could be the onset of

turbulent flow. Alternately, permeability could be plotted against gradient. A drop

in permeability would indicate a problem.

Direction of Flow. There are several concerns in regard to the direction of flow of

the permeant relative to the specimen. First, in the field, flow is horizontal and

thus runs parallel to the planes of compaction. In the laboratory, specimens are

usually tested with flow running perpendicular to the planes of compaction. Thus,

all things being equal, laboratory measured permeability should give conservative

(lower) values. One study (44) has shown permeability values for the horizontal

direction to be 1 .1 to 1 .8 times greater than in the vertica l direction for dense-

I

I

52

graded materials. The more fines, the more stratification, and the larger the

difference between horizontal and vertical permeabilities. Thus, in practice an

open-graded material will be less affected by stratification. However, the whole

concept may be moot because in a pavement structure a 4 to 6 in drainage layer

most likely will be placed in one lift.

Secondly, in laboratory testing, flow can be vertically upward or downward.

Downward would help prevent specimen expansion, but may even density the

material. Upward flow will assist in washing air bubbles out of the specimen, but

will tend to decompact the specimen. Upward expansion can be prevented by a

spring and perforated plate holding down the specimen. Thus, upward flow is the

preferred direction.

Off-Target Density. In a range of possible densities, the permeability of a given

material can vary from one to 20 times. This tendency is reduced as the range of

particle sizes narrows. Specimens can be inadvertently tested at densities that are

significantly different than what is intended. This can happen by incorrectly

fabricating the specimen, or through expansion or densification during the test, as

discussed earlier. Strohm m ru. ( 17) have shown that being off the target density

by 1 % can lead to a difference of measured permeability of 32% at a gradient of

0.2. In a 10 in (25.4 cm) diameter specimen, a height differential of 1 /8 in will

render a difference in density of about 2%. A change in height as a consequence

of testing can be detected by measuring the initial and final heights.

Rigid Wall Permeameter

Equipment. In this study, the open-graded materials were tested in the rigid wall

53

Fig.6. Rigid Wall Permeameter.

54

permeameter which is shown in Fig. 6. This included the NJ and OGS gradations

for the DR-12 through DR-15 aggregates. Tests were attempted on the MHTD

Middle gradation, but this proved to be impractical because of the low permeability

of this material. One problem that was noted early in the program happened

during the vacuum saturation phase. It was observed that the water tended to

rush up the voids between the permeameter wall and the compacted material,

washing fines from the aggregate. Thus, if tested, the permeating water would

tend to follow the path of least resistance and bypass the specimen, rendering

falsely high permeability test results. Sherard (45) has noted that for open-graded

materials, the voids along the walls are larger than the interior voids. In the

I present study, a series of experiments was conducted to find a way to reduce the

void space between the permeameter wall and the particles of the specimen.

Several other researchers have used various materials, such as rubber (46) and

sand (45, 47). In this study, a material was sought which would be a balance

between good flow inhibition and durability. The material that was finally decided

upon was an open-cell neoprene sponge rubber. A liner of this material was

permanently affixed to the lucite permeameter interior wall surface and sealed

against short-circuiting.

I

The entire rigid wall permeability test station is shown in Fig. 7. The

equipment included a variable height deaeration sand tank, an inlet tank, the lucite

permeameter with manometer ports, a variable · height outlet tank, and two sets of

manometers--one set attached to the permeameter and the other set attached to

the inlet and outlet tanks. In operation, tapwater was fed into the deaeration tank

VA

RIA

BL

E

HEI

GH

T O

UTL

ET

TAN

K

'7

"

\.I A

ST

E_

.._

__

_,

BLE

EDER

V

AL

VE

TD

BLE

EDER

V

AL

VE

INLE

T

TA

NK

,.

Fig

. 7

. S

chet

rto.

tic

o

f R

igic

l \J

a.ll

P

erM

ea

Me

ter

Te

st

Sta

. tio

n.

VA

RIA

BL

E

HEI

GH

T D

EAER

AT!

ON

TA

NK

. s·~ND

:,

-......

.:-

---B

AC

KF

LO

\./

LIN

E

<J1

<J1

I 56

which functioned in two ways. First, the water was introduced into the freeboard

above the sand bed. Here, upon depressurization, air was allowed to come out of

solution and rise to the top of the body of water and escape. The water itself

permeated downward through the sandbed. Most of the remaining air bubbles

I were trapped in the top portion of the sand bed. Periodically, the sand bed was

backflushed and stirred to remove bubbles. The gradient in the permeameter was

varied by raising or lowering the outlet tank; the driving head was the difference in

I I I

I

I I

elevation between the inlet and outlet tanks. To get a truer measurement of

permeability, the difference in readings in the two permeameter manometer tubes

were used as a measure of the driving head. This removed the problems

associated with effects at the top and bottom surfaces of the specimen, where

disturbance, smearing, and fines collection can alter the measured specimen

permeability. Flow was collected from the outlet tank and measured with a

graduated cylinder. Flow rate was calculated from time interval readings taken

with a stopwatch. Temperature was measured at the inlet and outlet tanks; the

average was used for temperature corrections to permeability calculations.

The permeameter was a 10 in diameter lucite cylinder, with manometer

ports approximately 10 in apart. The compacted specimen was 12 in in height. At

either end of the specimen was a perforated aluminum plate with an adjacent #200

screen to limit fines loss. The ability to see the specimen during saturation and

permeability testing proved to be very helpful in determining specimen behavior.

The effects of vacuum application and release, various flow rates, trapped air,

buildup of air, specimen disturbance, and so forth were observed. This ability is

57

not available with most permeameters which are made of opaque materials.

However, this ability was lost when the use of the inside liner was initiated.

Various air bleeder valves were installed in the system because the

entrapment of air was a major obstacle that had to be overcome. The first line of

defense was a good deaeration system. Various schemes were tried, including a

vacuu_m tank with the water being sprayed into the tank. The system with the

sand bed as previously described proved to be the most effective and simplest to

operate. Secondly, the specimen needed to be subjected to an initial vacuum to

remove as much air as possible, and to draw water up into the specimen. The

vacuum should not be excessive because the resulting gradient would introduce

turbulent flow which could disturb the sample and wash fines toward one end of

the specimen. Thus, only 1.5 psi vacuum was applied during the saturation phase.

The flow of water through the specimen was vertically toward the top in order to

wash air bubbles upward in their natural direction of flow, thus preventing them

from being entrapped. Specimen expansion was not a problem, partly because of

the spring which held the top plate (and specimen) in place, and partly because of

the low gradients involved. Air tended to collect in the permeameter inlet line, at

the bottom manifold, at the top of the permeameter, and in the plumbing between

the permeameter and the outlet tank. Where possible, the materials used for

construction of the device were transparent, which allowed visual observation of

air bubble buildup. Periodically, the lines and permeameter were tapped by hand or

with a rubber mallet to entice the bubbles to a point where they could be bled off,

thus preventing the specimen from clogging or vapor lock from occurring in the

I

I

I

58

plumbing.

Procedure. Each specimen was compacted into the permeameter at a slightly drier­

than-optimum moisture content. This moisture content was found to be necessary

in order to prevent fines migration during compaction. The specimens were

compacted in 12 layers of one in thickness using an air hammer and steel plate.

Final compacted height was measured; it was measured again after completion of

the permeability testing to calculate any specimen expansion or densification. The

I specimen was subjected to 1.5 psi vacuum for 15 minutes, then deaired water

was introduced at the base. After water was observed to flow into the outlet

tank, testing was initiated. Each specimen was subjected to five increasingly

higher test gradients. The flows and times were repeated five times at each

gradient. The gradients ranged from 0.02 to 0.6, which covered the expected

range of gradients in the field up to turbulent flow. The expected gradient in the

field was calculated as follows:

I

D1 +D2 +(S • \IV) = .•............... (21)

w

where:

0 1 = maximum anticipated surface layer thickness

0 2 = base layer thickness

S = slope of base

W = base width.

This assumes that water has completely filled the pavement up to the pavement

surface at the centerline joint. For a 1 5 in asphalt layer over 6 in granular subbase

59

at a cross-slope of 3/16 in per ft and a 14 ft lane width, the maximum expected

gradient at total saturation of a very thick pavement would be 0. 14. Average

gradients would be considerably less than this. The permeabilities calculated in

this study were limited to gradients of less than 0. 1 to assure that laminar flow

conditions existed. The concept is shown in Fig. 8.

Hydraulic gradient Y.S. flow rate was plotted to check for turbulent flow.

During calculation of permeability, the permeability values for all gradients up to

0.1 were averaged; however, any value of permeability which was suspected of

being under turbulent conditions was not used in order to assure compliance with

Darcy's Law.

Temperature readings at the inlet and outlet tanks were averaged.

Calculated permeabilities were corrected to 20°C.

Permeability was calculated as follows:

k = Ql R/Ath ................... (22)

where:

k = coefficient of permeability at 20°C, ft/day

Q = volume of flow, cu ft

L = distance between manometer ports, ft

Rt = temperature correction factor to 20°C

A = cross-sectional area of the specimen, sq ft

t = time interval, days

h = head loss through specimen, ft.

60

I

I I

1· t

D1

I +- -D .

f-S'v/

<J

t Ll

I ·o I F. 8 Field Gradient. 19. .

I

I

61

Gradient was kept constant by keeping the difference in elevation between the

outlet and inlet tank water levels constant. Gradients were adjusted by raising or

lowering the outlet tank. Calculations of permeability utilized the change in

permeameter manometer levels. The inlet and outlet tank manometer readings

were used as a check to see if the permeameter manometer readings were

behaving normally.

Flexible Wall Permeameter

Equipment. The permeability of the dense-graded materials was too low for these

materials to be used in the rigid wall permeameter because of available applied

head limitations of the equipment. Thus, a flexible wall permeameter was used for

testing the MHTD Middle gradation. Actually, the equipment used was the cyclic

triaxial apparatus used for resilient modulus testing. After the cyclic loading used

in the modulus testing was completed, the specimen was subjected to permeability

testing. The equipment and specimen fabrication has been discussed in previous

sections. Special perforated aluminum manifolds were machined to replace the

more traditional porous stones at each end of the specimen in order to assure that

sufficient flow would be applied to the specimen. Also, the porous stones tended

to break during the vibratory compaction of the specimen during fabrication.

Number 200 screen was placed at the specimen ends to reduce the chance of

fines migrating out of the specimen during fabrication and testing.

Procedure. Upon completion of the resilient modulus testing, the specimen was in

a saturated and consolidated condition. The procedure found in ASTM D 5084-90

was then followed for permeability testing. Tap water was used as the permeant.

I

62

For materials with permeabilities of 0.28 to 0.028 ft/day, hydraulic gradients of

five or less are recommended. This was achieved in the testing program by

limiting the cell pressure to 3 psi above back pressure, and the head pressure to

1.5 psi above back pressure for the 8 in high specimens. And, effective confining

pressure never exceeded 3 psi at one end of the specimen and 1.5 psi at the other

end.

The test equipment is shown in Fig. 9. A schematic of the system is

depicted in Fig. 10. As can be seen, back pressure is applied to the top of the

specimen and a higher pressure (head pressure) is applied to the specimen bottom,

forcing water flow upwards. Back pressure and head pressure are achieved

principally by applying air pressure to water-filled burettes. Flow through the

specimen is measured in both the inflow (head pressure) and outflow (back

pressure) burettes. These were essentially equal in the testing program. The time

interval over which the flow takes place is also recorded. Eq. 22 was used to

calculate permeability. The head loss (h) was the average of the initial and final

inflow burette plus head pressure readings minus the average of the initial and final

outflow burette plus backpressure readings. The final calculated permeability was

the average of five test runs.

POROSITY

In the estimation of permeability, it is useful to use porosity as a predictor,

although a high porosity does not necessarily mean a correspondingly higher

permeability. Porosity is calculated in accordance with Eq. 3, rep.eated here:

63

Fig.9. Flexible Wall Permeability Test Station.

I I

I

I

I I

I

Fig.10.

Pr e 1 .aurt. S up p l y

C t 11

F;_ e""'!. • f VO Ir

Tollwoler

R e , ervo l r

Permeo b lll t y

C C 1 1

V e n 1 -----<<>----~

L I n c >--«>------'

Heod w oler

Rfl s e.r volr

Inf I u • n t

L i n t

Schematic of Flexible Wall Permeameter Test

Station (after ASTM 05084-90).

64

65

fJ = 1 _ Yd G• Yw

The question arises as to whether to use bulk or apparent specific gravity. Both

values are readily available from the same test procedure. A case can be made for

using bulk specific gravity as follows. When water is added to granular material,

water penetrates voids in the particles, filling them to the surface. Drainability

studies have shown that this void space is not available to conduct water through

the compacted granular material. Only the voids between the particles have a

possibility of being available. Thus calculation of porosity, in a practical sense,

should be based on the particles having no particle voids communicating to the

particle surface -- a condition described by bulk specific gravity, which is defined

as dry weight divided by bulk particle volume. Examination of Eq. 6 leads to the

conclusion that porosities calculated using bulk specific gravity (BSG) will be

smaller than porosities calculated with apparent specific gravity (ASG) which uses

the non-water penetrating particle volume in its definition. Unfortunately, many

studies that are reported in the literature do not specify which type of specific

gravity was used in the porosity calculations. It was assumed that apparent

specific gravity is used in the Moulton equation (Eq. 2). Examination of Eq. 2

indicates that the use of BSG would lead to lower predicted permeabilities. The

concept of effective porosity, which takes all this a step further, is explained in a

later section.

The value of specific gravity used in the porosity calculation can significantly

alter the results. Values of specific gravity seen in practice range from BSG's of

66

2.45 to ASG's of 2.80. Thus, calculated porosities can vary from 11. 7 to 27. 7%

I for a compacted unit weight of 135 pcf. It follows that an accurate value for

specific gravity is necessary for an accurate calculation of porosity.

I I

EFFECTIVE POROSITY

As previously discussed, time-to-drain calculations require data for the

effective porosity data for the material that is draining. Effective porosity is the

ratio of the volume of voids that can be drained under gravity to the total volume

of base material. It has been shown ( 16) that for open-graded materials, the

effective porosity can be close to the calculated porosity, but for dense-graded

base materials, the effective porosity can be quite small. The water that is

essentially nondrainable is water that is held in the pores by capillary action or

water films on the aggregate particle surfaces. The equation (Eq. 6) for effective

porosity is essentially the traditional porosity equation, with the term that

represents the solid volume being increased for the nondrainable water. The

equation is repeated below:

After permeability testing is complete, the water content remaining after 16 hr

drainage under gravity conditions is determined. This water content is w.. Here,

the apparent specific gravity should be used, not the bulk specific gravity, because

I w. reflects all the water remaining including the water in the particle pores. Thus,

the smaller particle volume (as calculated by use of apparent specific gravity)

should be used in the above equation. Even after 24 hr drainage, for dense-

I

67

graded, highly compacted base material with a fines content as low as 5%,

effective porosity can approach zero (17). A low TJ. directly affects drainage times.

Additionally, the mineralogical type of fines affects effective porosity much as it

does permeability (13). The more active the fines, the lower the effective porosity.

68

RESULTS OF THE LABORATORY INVESTIGATION

I AS-RECEIVED GRADATIONS

I

I

The as-received gradations of the four granular materials are shown in Table

7.

Table 7. As-Received Gradations.

Sieve Percent Passing Size

DR-12 DR-13 DR-14 DR-15

1 in 100 100 100 100

1/2 in 96 83 72 83

#4 68 50 46 50

16 -- 26 37 26

40 24 18 17 12

100 12 13 2 5

200 8 7 1 4

EXPERIMENTAL GRADATIONS

The three experimental gradations were the MHTD Middle, the New Jersey

(NJ), and the OGS. These are shown in Table 8. The NJ and the OGS were used

in the rigid wall permeameter permeability testing, while the Middle and the NJ

were used in the resilient modulus portion of the study. Only the Middle was

tested for permeability with the flexible wall permeability procedure because of

equipment limitations.

69

Table 8. Experimental Gradations.

Sieve Size % Passing

Middle NJ OGS

3 in. 100 100 100

1 1/2 100 100 100

1 100 100 100

3/4 (95) (95) (91)

1/2 75 68 60

3/8 (63) (58) (48)

#4 50 47 30

8 (40) (20) (16)

16 33 5 7

30 (28) (4) (6)

40 25 3 5

50 (22) (2.5) (4)

100 (16) (2.5) (3)

200 8 2 2

( ) = estimated from semilog plot

GRADATION CURVE SHAPE/POSITION

In an attempt to determine the effect of gradation on permeability, values for

gradation curve shape/position were required. Nine different methods were tried.

Sieve size data from all three experimental gradation curves were used for

calculation of various parameters; the parameters were then used in the

development of the multiple regression models for permeability to see which

method increased the accuracy of the model the most. The nine methods or

parameters were as follows: fineness modulus (FM), coefficient of uniformity (Cul,

I I I I I I I I I I I I I I I I I I I

70

coefficient of skew (Cz), surface fineness (SF), specific surface factor (SSF),

-(SF/SSF), Hudson's A, slopes-of-gradation-curve (mn-nl, and the percent passing

individual sieves. The results are shown in Table 9. These parameters are

discussed more fully in Volume I of this report. The slopes-of-gradation-curve

method was altered from that described in Volume I to better match the natural

break points of the experimental gradation curves. Thus the slopes of each curve

were determined between the 1 in and #4 sieve, the #4 and #16 sieves, and the

#16 and #200 sieves. The results of these calculations are shown in Table 10.

Table 9. Gradation Shape Results.

Middle NJ OGS

FM 4.53 5.66 5.95

cu 82.6 5.29 8.47

CZ 1.19 0.21 0.25

SF 1588 1938 1929

SSF 294.4 65.8 76.3

SF/SSF 5.40 29.4 25.3

A 4.55 3.36 3.07

Mn-n 94.3, 78.0, 100.0, 132.1, 138.9 192. 7, 105.5,

16.7 27.8

3/4", #4, 16, 200 95, 50, 33, 8 95, 47, 91, 30, 5, 2 7, 2

71

Table 10. Experimental Gradation Slopes.

Gradation M,.4 M4.1e M,e-200

Middle 94.3 78.0 138.9

NJ 100.0 192.7 16.7

OGS 132.1 105.5 27.8

For the permeability testing, none of the single curve shape/position

parameters was significant to the model. However, percent passing certain

individual sieves was significant. In terms of individual particle sizes, the effect of

gradation on permeability was ascertained as follows. As will be discussed more

fully later, a model was developed to represent the permeability results of several

other studies in addition to the results of the present study to give a more

generally applicable equation to predict permeability. All gradations used in the

model are shown in Fig. 11. Because grain size distribution is linked to the pore

volume available to transmit water, a method was developed to relate the two.

This was done by determining how close the percent passing on each sieve size

was to the percent passing for the densest possible gradation. The maximum

density gradation was estimated by use of 0.45 power FHWA paper, commonly

used in asphalt gradation work. For each gradation, the maximum density line

(MDL) was drawn in accordance with the method given in Volume I of the

companion report to this study (27). Next, the vertical distance (in percent

passing) between the MDL and the percent passing gradation line on each sieve

was found and converted to a percent of total possible distance. This

measurement was determined for each of the base rock gradations and was

I

I I I

I I I I I I I I I

I

I I

bl)

t:: •r-4 rt.I l1l IC

11.. ,+,)

~ Q) t.) J-4 Q)

11.. -IC ,+,)

0 ~

100

90

80

70

60

50

40

30

20

10

0

72

200 100 50 30 16 B 4 3/B" 3/4" 1 1/2"

Sieve Size

Fig. 11. Gradations of Materials Used in the Permeability Model Development.

73

correlated with permeability. Pearson's correlation coefficients (R) were

calculated. The results are shown in Table 11 . The particle sizes that were most

effective in predicting permeability were (individually) the #4, #8, and 3/8 in

sieves, as well as the D10 size. This information proved useful in the development

of the permeability model as discussed later in this report. Although the #200

sieve did not indicate a high individual correlation, it was found to be helpful in the

model development.

Table 11. Usefulness of Individual Particle Sizes in Prediction of Permeability.

Material k Percent Difference Between Maximum Density Line and Gradation Line (ft/day)

3/8 in #4 #8 #16 #30 #50 #100 #200 D,o (mm)

MHTD 0.8 8.7 0.2 8.1 25.9 47.4 57.1 60.0 14.3 0.092 Middle

S&G 1.2 0.2 5.1 0.4 19.0 53.3 63.6 1.2 33.3 0.16

Cr. 372 29.0 28.0 24.3 18.5 10.5 14.3 30.0 28.6 0.20 Gravel

MHTD 820 15.9 6.0 45.9 81.5 78.9 85.7 80.0 71.4 1.7 NJ

Cr. 1073 38.9 41.0 35.7 47.6 46.7 45.4 37.5 50.0 0.94 Stone

MHTD 1158 27.3 37.5 54.3 72.0 66.7 69.2 70.0 71.4 1.5 OGS

S&G 18,144 38.9 69.2 71.4 66.7 60.0 54.5 50.0 33.3 3.9

S&G 21,546 63.0 89.7 92.9 100 100 100 100 100 7.0

R -- 0.773 0.893 0.839 0.644 0.572 0.435 0.487 0.444 0.944

S&G = Sand and Gravel (ref. 41) Cr. Gravel = Crushed gravel (ref. 17) Cr. Stone = crushed stone (ref. 17)

I I I I I I I I I I I I

I I I I I I I I I I I I I I I I I

I

MOISTURE-DENSITY RELATIONSHIPS AND SPECIFIC GRAVITIES

Moisture-density relationship and specific gravity information were

determined in regard to the three test gradations for each of the four granular

materials. The data is shown in Table 12.

Table 12. Specific Gravity and Moisture Density Data.

74

T-99 T-180 D4253 Material Gradation Sp. Gravity

ASG BSG MOD OMC MOD OMC MOD OMC (pcf) (%) (pcf) (%) (pcf) (%)

DR-12 Middle 2.69 2.53 136.5 7.0 137.6 7.4 -- --OGS 2.69 2.53 -- -- -- -- 127.6 10.3

NJ 2.69 2.53 -- -- 131.2 9.0 121.1 12.2

DR-13 Middle 2.78 2.55 138.3 7.7 141.0 5.9 -- --OGS 2.78 2.55 -- -- -- -- 124.6 12.2

NJ 2.78 2.55 -- -- 135.2 8.3 124.1 13.0

DR-14 Middle 2.65 2.52 132.5 7.8 134.5 6.7 -- --OGS 2.65 2.52 -- -- -- -- 123.0 10.8

NJ 2.65 2.52 125.8* 7.3 122.8 10.3 121.3 12.0

DR-15 Middle 2.65 2.41 134.4 7.6 136.9 6.1 -- --OGS 2.65 2.41 -- -- -- -- 112.2 13.8

NJ 2.65 2.41 109.5* * 8.9 -- -- 114.3 14.0

Note: T-99 = Standard proctor T-180 = Modified proctor D4253 = Vibratory table *E0 specimen compacted with vibratory hammer to refusal. This value was used as the maximum target value for the High Compactive effort specimens. * * T-99, T-180, and D4253 densities were very close. To get a wider difference in values a peak vibratory density achieved with a different surcharge weight was used.

75

Some difficulty was experienced in performing the impact-type of moisture-density

tests for certain materials when graded with the New Jersey and OGS gradations.

So, the vibratory table density (wet) method was also performed. It is anticipated

that this gives a more realistic density value similar to that which will be achieved

in the field. The tests were performed in a dry state at different power settings to

determine the optimum setting which would result in the highest density. Then

the test was run with the material in a moisture state which is more in line with

field compaction conditions. Fig. 12 is a typical vibratory table test result. Fig. 13

shows the moisture-density relationships for each of the four granular materials.

· PARTICLE SHAPE AND SURFACE TEXTURE

Particle shape/texture characteristics were quantified by use of ASTM D

3398 for the ( +) #4 sieve material and by NAA Method A for the (-) #8 through

( +) #100 material for each aggregate source. Both are measures of void content

of bulk aggregate; void content has been shown to be related to shape/texture.

03398 results in a "Particle Index" (IP); NAA Method A gives an "Uncompacted

Voids Percent" (U). The results are shown in Table 13.

Table 13. Particle Shape/Texture Results.

Aggregate Particle Index (IP) Uncompacted Voids (U) %

DR-12 12.5 43.6

DR-13 11.9 45.3

DR-14 10.0 41.2

DR-15 10.4 40.6

I I

I I I I I

76

Round, smooth particles give IP's of 6 or 7, while angular, rough particles result in

values of more than 15. The range of IP's of the aggregates in this study was

10.0 to 12.5. The Particle Index was determined for the coarse aggregate fraction

of each gradation and the Uncompacted Voids content was determined for the fine

aggregate fraction.

Looking at Particle Index and especially the Uncompacted Voids values, the

crushed aggregates were somewhat more angular than the gravels, as expected,

but the ranges were limited.

PLASTICITY OF FINES

The results of the Atterberg Limits testing are shown in Table 14. All four

I aggregates were essentially non-plastic in nature.

I I I I I I I I

1•

Table 14. Atterberg Limits of the Base Materials.

Aggregate Liquid Limit Plasticity Index

DR-12 16 NP

DR-13 18 NP

DR-14 22 NP

DR-15 19 NP

RESILIENT MODULUS

Resilient modulus tests were performed on two crushed stone aggregates

(DR-12 and DR-13) and two gravels (DR-14 and DR-15) at two degrees of

saturation (aproximately 60 and 100%), two compactive efforts (low and high),

two gradations (MHTD Middle and New Jersey), and 14 stress states, with

77

123

122

,........_ 1 21 rr)

-+-J ~ Wet

.......__ ~

..0 120 ....._,,

+' ...c 1 1 9 O"l

(l)

3: +' 1 1 8 C:

::J

>- 1 1 7 I...

0

1 1 6 Sample: DR-12 New Jersey

1 1 5 0.008 0 .010 0.012 0.014 0.016 0.018

Double Amplitude of Vibration (in.)

Fi g . 1 2 . Ty pi ca I Vi b rat o ry Ta b I e Test Res u It.

I I I I I I I I I I I I I I I I I I I

145

140

-'-.. ..... :3 1 35

; 130

C ::::J

~ 125 0

120 5

145

,......_ 140

+' -'-.. ..... :3 135

+' ..c Cl'

; 130

C: ::::J

~ 125 .... 0

120 0.0

T180

'

Semple: DR-12 Middle Optimum H20 Content : 7 .OX (T99) Optimum H20 Content: 7 .4!1: (T\80) Mo x. Dry Unit Wt. : 136.5 Lbf/ft

3 (T99)

Mo x. Dry Unit Wt.: 137 .6 Lbf/ft (T180)

6 7 8 Water Content (%)

Sample : DR-14 Middle Optimum H

20 Content: 7 .8 (T99)

Optimum H20 Content: 6.7 (T180)

T180

TIIII

Mo x. Dry Unit Wt. : Mo x. Dry Unit Wt. :

J 132.5 Lbf/ft

3 (T99)

134 .5 Lbf/ft (T180)

9

,-.. I")

+' .... ......... ....

145

140

:3 135

+' .r. Cl

.iii 130 3;

C ::::J

>- 125 .... 0

120

,...... 140 I")

+' .... ......... .... ..0

.:::, 1 35 +' .r. C7' Q)

3; 130

C ::::J

~ 125 0

4

Sample: DR-13 Middle Optimum H

10 Content: 7 .7",(, (T99)

Optimum H20 Content : 5.9% (T\80) t.Aox . Dry Unit Wt.: 138.3 Lbf/ft3 (T99) t.Aox . Dry Unit Wt. : 141.0 Lbf/fl (T1BO)

5 6 7 8 9 Water Content (%)

Sample: DR-15 Middle Optimum H20 Content: 7 .6 (T99) Optimum H20 Content : 6.1 (T18jl) Mox. Dry Unit Wt. : 134 .'4 Lbf/ft3 (T99) Mox. Dry Unit Wt. : 136.9 Lbf/ft (T1 80)

78

10

120 '-~~ ...... ~~ ...... ~~ ...... ~~---~~ ...... 0 . 1 0.2 4 5 6 7

Water Content (%) Water Content (%)

Fig. 13. Moisture - Density Relationships for

MHTD Middle Unbound Granular Materials.

8 9

~

en a.

'-"

en :::,

:::, "'O 0 ~

~

C: Q)

en Q)

0:::

1 0 100

Bulk Stress (psi)

Fig . 14. Typical Resilient Modulus Test Results.

79

I

I I

I

I

I

I I I I I I I I I I I I I I I I I I I

80

duplicate samples, for a total of 896 tests. Because the same 14 stress states and

the two degrees of saturation were used for each specimen, there were 32

specimens. The testing sequence for each specimen is shown in Table 5. Fig. 14

is a typical plot of bulk stress ~. resilient modulus. Each data point is

representative of one stress state. As can be seen, modulus increases with an

increase in stress state. The classic equation of the line (known as the "theta

model") is:

or

where:

log E(l = log k 1 + k2 log 0 ............... (23)

E(l = k1

e*• . . . . . . . . . . . . . . . . . . . . (24)

k1 = intercept of Eu at 0 = 1 psi, log-log plot

k2 = slope of line, log-log plot.

The results of all resilient modulus testing are tabulated in Table 15.

Fig. 15 shows the relationship of coefficients k1 and k2 as reported by Rada

and Witczak (40), with the results of the present study also plotted. As can be

seen, the results of the present study data fall in the range that has been reported

elsewhere.

The effect of increased degree of saturation is shown in Fig. 16. The

general trend is a loss of Eu as the degree of saturation increases from a moist

state to a saturated state. This is similar to the trend reported by others

(2,40,59,60). As will be discussed later, the loss of k1 due to an increase in

81

moisture was required to develop m-coefficients in this study. The overall average

k1 at 60% saturation was 4797 psi, which dropped to 3314 psi, for a loss of 31 %.

Other reported losses have ranged from 8 to 94% (40, 54 and 56).

The interaction between gradation, compactive effort, and degree of

saturation is shown in Fig. 17. It appears that open-graded material suffers more

of a loss in Eg than dense-graded material as indicated by steeper curve slopes.

One possible explanation is that dense bases may not suffer so much because they

may tend to dilate under load, thus positive pore pressures may not be so much in

evidence as in a less dense-graded material. The percent loss for the dense graded

materials with low and high compactive efforts was 9.8 and 11.0, while the loss

for the open graded materials was 17. 7 and 19.5, respectively. However, the

effect on Eg of providing a drained base can be seen from Fig. 17 by looking at the

dashed lines. For both the low and high compactive effort cases, there is a benefit

to changing from a dense-graded material which will remain saturated for extended

periods of time to an open-graded material which will remain in a drained state

most of the time.

STATISTICAL ANALYSIS

A statistical analysis was performed to determine the effect of several of the

independent variables on Eg. A more complete analysis is included in Ref. 5.

Paired-t tests were performed to see if there was a significant difference between

the mean of all Eg data of a low degree Y..S., a high degree of saturation.

Additionally, a Tukey HSD analysis was performed to determine if aggregate

source made a significant difference in Eg results and if so, which source(s) were

I

I I

I

I I I I I

I

----

----

----

----

---

Ta

ble

15

. R

esi

lien

t M

od

ulu

s T

est

Dat

a.

Mat

eria

l G

rada

tion/

CE

D

ry D

ensi

ty (

pcf)

M

AD

D

Tar

get

As-

test

ed

(pcf

) (%

)

DR

-12

Mid

(lo

w)

13

6.5

1

36

.4

13

8.9

9

8.2

DR

-12

Mid

(lo

w)

13

6.5

1

36

.4

13

8.9

9

8.2

DR

-12

Mid

(hi

gh)

13

7.6

1

38

.4

13

8.9

9

9.6

DR

-12

Mid

(hi

gh)

13

7.6

1

38

.4

13

8.9

9

9.6

DR

-12

NJ

(low

) 12

1.1

12

0.5

1

31

.2

91

.8

DR

-12

NJ

(low

) 12

1.1

12

0.5

1

31

.2

91

.8

DR

-12

NJ

(hig

h)

13

1.2

1

27

.7

13

1.2

9

7.3

DR

-12

NJ

(hig

h)

13

1.2

1

27

.7

13

1.2

9

7.3

DR

-13

Mid

(lo

w)

13

8.3

1

38

.2

14

1.0

9

8.0

DR

-13

Mid

(lo

w)

13

8.3

1

38

.2

14

1.0

9

8.0

DR

-13

Mid

(hi

gh)

14

1.0

1

39

.5

14

1.0

9

8.9

DR

-13

Mid

(hi

gh)

14

1.0

1

39

.5

14

1.0

9

8.9

DR

-13

NJ

(low

) 12

4.1

12

5.9

1

35

.2

93.1

DR

-13

NJ

(low

) 12

4.1

12

5.9

1

35

.2

93.1

DR

-13

NJ

(hig

h)

13

5.2

13

4.1

13

5.2

9

9.2

DR

-13

NJ

(hig

h)

13

5.2

13

4.1

13

5.2

9

9.2

DR

-14

Mid

(lo

w)

13

2.5

13

1.1

13

5.4

9

6.8

DR

-14

Mid

(lo

w)

13

2.5

13

1 .1

1

35

.4

96

.8

DR

-14

Mid

(hi

gh)

13

4.5

1

34

.4

13

5.4

9

9.3

DR

-14

Mid

(hi

gh)

13

4.5

1

34

.4

13

5.4

9

9.3

DR

-14

NJ

(low

) 1

22

.8

121.

1 1

25

.8

96

.2

DR

-14

NJ

(low

) 1

22

.8

121.

1 1

25

.8

96

.2

DR

-14

NJ

(hig

h)

12

5.3

12

4.1

12

5.8

9

8.6

DR

-14

NJ

(hig

h)

12

5.3

1

25

.0

12

5.8

9

9.3

DR

0S

at.

k,

(%)

(psi

)

92.1

6

1.4

3

04

0

92.1

1

00

2

95

8

98

.4

62

.5

38

28

98

.4

10

0

37

58

69

.3

59

.6

43

07

69

.3

10

0

19

40

90

.5

53

.1

51

34

90

.5

10

0

27

06

91

.3

63

.8

42

12

91

.3

10

0

36

06

95

.2

59

.8

83

12

95

.2

10

0

59

18

78

.3

63

.8

34

70

78

.3

10

0

28

24

97

.5

58

.8

31

64

97

.5

10

0

29

97

85

.9

58

.6

44

43

85

.9

10

0

3401

96

.8

67

.2

54

68

96

.8

10

0

57

93

84

.7

58

.2

66

18

84

.7

10

0

45

04

94

.7

55.1

7

63

9

97

.4

10

0

35

69

k2

0.8

5

0.8

6

0.8

2

0.8

0

0.7

2

0.9

0

0.7

4

0.9

2

0.7

6

0.8

0

0.5

8

0.6

2

0.81

0.8

2

0.9

0

0.8

6

0.6

0

0.6

8

0.6

4

0.6

7

0.5

6

0.7

2

0.5

3

0.7

4

Eg(

psi)*

8=

10

21

,52

2

21

,42

8

25

,29

5

23

,43

7

22

,86

5

15

,58

4

27

,89

0

22

,50

8

24

,23

8

22

,75

2

31

,96

7

24

,39

0

22

,40

7

18

,65

8

25

,42

8

21

,96

3

17

,68

8

16

,27

8

23

,86

9

27

,09

6

24

,30

9

23

,36

9

25

,88

4

19

,84

0

CX)

I\.)

DR

-15

Mid

(lo

w)

13

4.4

1

31

.5

13

6.9

DR

-15

Mid

(lo

w)

13

4.4

1

31

.5

13

6.9

DR

-15

Mid

(hi

gh)

13

6.9

1

35

.4

13

6.9

DR

-15

Mid

(hi

gh)

13

6.9

1

35

.4

13

6.9

DR

-15

NJ

(low

) 1

09

.5

11

0.4

1

15

.8

DR

-15

NJ

(low

) 1

09

.5

11

0.4

1

15

.8

DR

-15

NJ

(hig

h)

11

4.3

1

15

.8

11

5.8

DR

-15

NJ

(hig

h)

11

4.3

1

15

.8

11

5.8

CE

=

Com

pact

ive

eff

ort

M

AD

D =

M

axim

um A

ttain

able

Dry

Den

sity

DR

=

R

elat

ive

Den

sity

* E

g =

k 1

91<2.

---

96

.0

83.1

5

1.4

96

.0

83.1

1

00

98

.9

95

.2

59

.7

98

.7

94

.5

10

0

95

.3

76.1

5

8.9

95

.3

76.1

1

00

10

0

10

0

63

.6

10

0

10

0

10

0

--

50

58

0

.55

26

45

0

.69

44

98

0

.65

27

02

0

.75

45

54

0

.57

23

58

0

.76

30

12

0

.72

13

38

0

.96

17

,94

6

12

,95

5

20

,09

0

15

,19

4

16

,92

0

13

,56

9

15

,80

7

12

,20

3 --

00

(.,.)

I I I I I I I I I I I I I I I I I I I

,.--....

VJ 0...

'--""

N 0 ,..-

X

~

~

1000

100

1 0

1 0.0 0.2

Fig. 15:

0 .4

Lim its of Rado and

Witczok 's Data

• ., .• . - .. .. . \ . .. ~:J. ....

0.6 0.8

k2

•• • •

1 .0

Relationship Between Experimentally Derived Factors (k 1 and k2 ) for the Theta Model.

84

1 .2

85

I

30

I 25 ,,....._ 13---- •

I")

0 ..--

X I Ul 20 0.. ...__,,

Ul :::::,

:::::, 1 5

.... DR12 ""'O 0 6 DR12 NJ I ~

DR13 M ... ......, C V DR13 NJ Q)

10 • DR14 M Ul

DR14 NJ Q) C 0:: • DR15 M

I 0 DR15 NJ 5 30 40 50 60 70 80 90 100 11 0 1 20

Degree of Saturation (%)

Fig. 1 6. Effect of Degree of Saturation and Aggregate Source on Resilient Modulus.

I

I I I I I I I I I I I I I I I I I I

1•

......... ~

0 -X

U)

0. -U) ::,

::, -0 0 ~

4,J

C: Q)

U) Q)

a::

86

30 ~-------------------------------------------,.------

25

20

• Mid

0 NJ 15

Mid ~

V NJ

- High CE

- High CE

- Low CE

- Low CE e - 10 psi

Comparison ·of Mid to Mid or NJ to NJ Comparison of NJ to Mid

10 1.-----........ ----------------------------------l.-----~ Low High

Degree of Saturation (%)

Fig. 17. Effect of Gradation, Degree of Saturation, and Compactive Effort on Resilient Modulus.

87

significantly different. The results are shown in Table 16. As can be seen, degree

of saturation was significant to differences in Eg at the 0.05 level, and the

interaction of gradation and saturation was significant at the 0.088 level.

Reduction in Eg came from increasing the saturation from 60 to 100%, and having

a saturated, dense-graded material as opposed to a drained, open-graded material.

Table 16. Statistical Significance of Testing Variables to Resilient Modulus.

Eg at 8 = 1 O psi (psi)

Condition Maximum Minimum Difference Significance at 0.05 level

All Mixture~:

0 Saturation, low~. 22,706 19,452 3164 yes high

Gradation and 22,704 20,442 2262 yes at 0.088 saturation: open level graded drained ~. dense graded undrained

PERMEABILITY, POROSITY, AND EFFECTIVE POROSITY

Open-Graded Materials

Permeability tests were performed on specimens of the four aggregates (DR-

12 through DR-15) which were graded in accordance to the two open gradations:

NJ and OGS. Duplicate specimens were tested for each gradation in the rigid wall

permeameter. The target test density was the result of the lower of the two

compactive efforts which corresponded to approximately 100% T-99 maximum

density. Each specimen was tested three to five times each at up to eight

I

I

I 88

successively higher hydraulic gradients. However, only the gradients at 0.1 and

lower were averaged to produce the reported permeability because this range

represented the gradients that are expected in the field. This gradient is well

below the AASHTO recommended maximum of 0.2 to 0.5 for the rigid wall

permeameter method. The rigid wall permeameter constant head tests were

applicable to these gradations because the permeabilities were considerably in

excess of the recommended minimum of 2 to 3 ft/day. Plots of flow rate Y.§.

hydraulic gradient indicated that laminar flow was in effect over the range that the

permeabilities were averaged, as shown in Fig. 18. At the end of each

permeability test, the specimens were drained for 16 hours under gravity, then

tested for drained moisture content for use in calculation of effective porosity.

Upon dissassembly of the permeameter, some fines were observed to be

accumulated at the upper perforated plate.

The summary of the results are shown in Table 17. A typical set of data is

I shown in Table 18. As shown in Table 17, average compacted densities ranged

from 98. 7 to 99.8 percent of target densities. Average porosities ranged from

0.245 to 0.324. Effective porosities were approximately 68 percent of the

"standard" porosities, as shown in Fig. 19.

I I

All permeabilities determined by test were considerably greater than those

estimated by the Moulton equation. This was somewhat expected after a careful

reading of the research studies upon which the Moulton equation is based. First,

much of the reported data is from tests where the specimens were soaked by mere

submergence. Thus the specimens were in all liklihood unsaturated, a condition

,,.-..... u Q) (/)

N

E u

" -E ......__,,

>-_...., u 0 Q)

>

0.20

0.15

0.1 0

0 .05

0.00 --------------------0.0 0. 1 0.2 0.3 0 .4 0.5

Hydraulic Gradient

Fig. 18 . Typical Constant Head Rigid Wall Permeameter Test Result.

0.6

89

-I ll

I

I I I I

I

I I I I

Table 17. Results of Rigid Wall Permeameter Permeability Testing.

Material Gradation Dry Density (pcf) Target ,, * * "· * *

Hydraulic % Gradient*

Target As-Tested

DR-12 NJ 121.1 119.0 98.3 0.291 0.209 0.025-0.1

NJ 121.1 120.9 99.8 0.280 0.215 0.021-0.1

avg 99.0 0.286 0.212

DR-12 OGS 127.6 126.1 98.8 0.249 0.173 0.023-0.1

OGS 127.6 127.5 99.9 0.240 0.164 0.027-0.1

avg 99.4 0.245 0.169

DR-13 NJ 124.1 122.4 98.6 0.294 0.180 0.023-0.1

NJ 124.1 123.8 99.8 0.286 0.198 0.025-0.1

avg 99.0 0.290 0.189

DR-13 OGS 124.6 123.0 98.7 0.291 0.190 0.027-0.1

OGS 124.6 123.8 99.4 0.286 0.185 0.021-0.1

avg 99.0 0.289 0.188

DR-14 NJ 122.8 121.1 98.6 0.268 0.161 0.022-0.1

NJ 122.8 122.4 99.6 0.260 0.173 0.029-0.1

avg 99.1 0.264 0.167

DR-14 OGS 123.0 121.5 98.8 0.265 0.161 0.023-0.1

OGS 123.0 122.9 99.9 0.257 0.163 0.021-0.1

avg 98.9 0.261 0.162

DR-15 NJ 114.3 112.7 98.6 0.318 0.230 0.022-0.1

NJ 114.3 112.9 98.8 0.317 0.234 0.022-0.1

avg 98.7 0.318 0.232

DR-15 OGS 112.2 110.9 98.9 0.329 0.242 0.029-0.1

OGS 112.2 112.8 100.6 0.318 0.238 0.017-0.1

avg 99.8 0.324 0.240

* Additional tests were conducted at higher gradients. * * Estimated permeability based on ASG using Moulton Equation

90

k (Hiday)

Est.** Test

244 402

188 1178

216 790

72 552

57 783

64 668

264 402

219 773

242 588

204 1048

181 1027

192 1038

140 228

115 889

128 559

109 948

88 938

98 943

444 955

434 1730

439 1342

461 1613

362 2746

412 2180

91 -

Table 18. Typical Set of Data for a Rigid Wall Permeamet er Test.

Per11,eabJl11u Test for 00-12 NJ (06/11 / 93 Qun 3)

Je s: t nano11ete<s Heao T.int: . I Pore ( ! •psed Mveraqe "'-/df..UI IC lh,er~qe N...,.ber HI H2 h He>d a lvoluaes l 1 , .. vetoc, HJ veloct tt.J qr~d1en1 qr~tent

<Oesc '1 o 1 1 on> <c•) <c• > <c• > <c ,.> <•I> u•ula11ve <sec> (a1n l 0.11'1 0 .1A1 h,t h,L

I 69. 90 69. 35 0. 55 0. 90 239 0.00 52.09 3.00 0 . 0112 0. 0229 69 . 90 69. 10 0 . 50 0. 90 238 0 . 00 50 . 81 5.00 0.0111 0. 020S 69. 90 69. 10 0. 50 0 . 90 2,0 0.00 50. 21 7.00 0.0 11 6

.. 0. 0208

69. ':a 69. "!O 0 . 50 a. 3a 1" 0 .00 50. 90 2. 00 0 .011 7 0 . 0116 0.0108 0. 0208

2 69 €G 68. 55 1. 25 I 1.70 138 0 . 00 21. 0, 11. 00 0 . 02)] 0.0511 o9. eo 66. S5 '- ,5 I I. 70 111 0. 00 15. so ! ., . 00 0.0233 0. C52 I 69. 90 ,6. 55 I. "5 I. 70 ] ... ] 0.00 ?5.cO !5 co C. 0230 0.0?32 0.050 0 . 05}5

3 l9 . 50 57.&0 I. 70 2. 70 :i,o 0.00 ,O. 90 ~ 1. 00 0 . 0?80 0 . 070il 69 . 50 67. 75 ·- ."S 1. 70 2,2 0.00 10. 93 13. 00 0. 02E2 0. Cl?~

69 . 50 67. 70 I. BG 1. 70 136 0.00 20. 93 ] ~. GO 0.0775 0 . 0279 0 . 075G 0. 0729 1 69. ,o 07. IC l . JO 3. 60 2« 0 .00 I< . 99 2':!.00 0 . OJS? 0 . 095'!

69. 30 67. 10 I 2. ?tJ ) . 60 238 0. 00 11 . 71 JO.OD 0 . 0,91 0 . 0317 69. JO 67 . 10 : . 20 3. 60 2,0 0.00 15 . 21 31. 00 0 . o,s, 0.0392 0. 0917 0 .0931

5 67. 80 63 . 20 <. 60 8. 90 2'7 0.00 J. 15 35.00 o . oa1! 0. IS !7

67.80 63.10 1 . 60 8. 90 111 0 .00 7. 21 36.00 0. 0225 0. 1917 67 . 80 6). 20 ~. 60 8 . 90 239 0 .00 6. 97 37. 00 0 . 0835 0.0831 0 . 1917 0 . 1917

6 65. 50 58. 55 6. S5 16 . 00 230 0.00 • . 28 11.00 0. 1)09 0 . ?896

65. 50 58. 55 6. 95 16 . 00 251 0 . 00 1. 90 12 . 00 0. 1218 0. 2896

65 . 50 SB . 55 6. ~:, 16.00 226 0.00 1. 12 11. 00 0 . I ])6 0. 1298 0 . 1896 0. 2696

J 50 . 10 13 . JO I~. SO 39 . 50 770 0.00 l. 87 <8. 00 0 .2 291 0. 5615

5; . 10 '3.60 11 . 50 J~ . ~a 139 0 . 00 2. 16 1,.00 0.1)57 0. 5625

57 . 00 13 . 50 !'J. so 39. 50 710 0.00 1. 8 I 50 . 00 0 . 1311 0. 2330 0 . 5615 0. 5625

8 69 . 90 69. 30 ,) . 60 0. 30 218 0. 00 '18. 28 57. 00 0 . 0125 0 . 0150

69. 90 69 . 30 0. 60 0. '30 230 0.00 11 . 60 59.00 0.0126 0. 0250

69. 90 69. 30 o . 60 0. 90 136 0.00 16 . 03 61.00 0 .01 25 0. 0125 0. 0250 0. 0250

a:lhe nuaoer of pore volu111es "" 6. 00 •TN' 1n le1 u.a1er scheae u.iS ch.inqed lor 1h1 s 1es1. M plc1s11c C,irbo~ ullh c1 sM'\d bed in 11 uc1s otc1ced 19 1tne t>elor• the, u,iter re.aches 1he conslc1n1 head 1nle1 1c1n~. The Scln.d bed u.as cre c1ted b4 rc11n1nq doun 1"e

Sclnd 1n the c.arb<X.I into c1bou1 cl 1001 of s1c1nd1nQ u.:ner, .and 1hen cl v.acuu• uu pulled on 1he 1c1n~ to reaove cltr 1n 1he sc1nd.

Tes, Teiaper.at ur e VISCOS I I~ fwer.aqe Averc1qe Aver.aqe Aver .aqe

N1.1t1ber (deo CI correct 10 l l l Qn 0< K

<Deserio, 10n) In lei Out let Ava. Q, (( • l' Se>C) (f ll'dcl!J) Cl 1.1dau >

I 23.0 21 . 0 13. s 0 . 921 0. '193 12 73. SJ

23 . 0 21 . 0 13. 5 o. 921 0.5012 1129.32

23. 0 21. 0 23. 5 0. 921 0 . 5115 1158. 51

23. 0 21. 0 23 . S 0. 92 1 0. 5163 1163 . 62 1150.19 0.6815 3. 0522 7030.15

2 22.5 21. 0 23. J 0. 927 0 . 1155 11 77 . 78

22. 5 21. 0 13. 3 0 . 917 0.1119 1176.11

22. 5 21 . 0 2) . 3 0. 927 0 . 3796 1075. 91 1113 . 31 I. 3590 6. 1256 '111 . 29

3 22.0 21 :; 1). 3 0. n1 0. 3661 1037 . 92

22. 0 11 5 1]. 3 0. 927 0. 3581 1015.11

22. 0 11 . 5 23. 3 0. 927 0. 3396 961. 51 1005 . 22 I. 6307 7. )502 121 2. 61

1 n. o 11 0 23. 0 0. 931 0. 3855 1092 . 87

21. 0 21 0 ?) 0 0. 931 0 . 1001 I 131 . 90 21. 0 ! ~. 0 :J. 0 0. 331 0. 3905 11 06. S1 1111. 53 1.2827 10 . )315 2110. 12

5 11.0 12. 5 '?7.] 0. 919 0. 1167 11 81. 32

22. 0 n 5 ~? . ) 0. 91~ 0. 1083 1157 .16 12. 0 22. 5 ?2.] 0 . 919 0. 1137 11 72.SS 1170. 36 i . l66'5 21. 9891 1209. 12

6 ,____ 21. 0 ?2. D n .o a . 951 0. '309 11? I. 11

21.0 11. 0 ?'J. 0 0. ~53 a. 1107 1161. 27 22. a 22 . 0 ;? 0 0. ~5) 0 . 1398 1216. 77 1210. e2 7 . 3863 31. 2261 751.07

7 11. 5 i I. 'i !I ., 0. 965 0. 3532 1111 1,6

I! . 5 ii 5 !I 1 IJ. '3~5 0 . 1011 111<\ . J I

ii. 5 21. ' .. 'i !J. ~f.'5 0.1016 I 1'8. 16 1133. 15 !]. 1003 61. 111 2 117. 01

a 21. 5 ? I 5 11. 5 0.%5 0 . 1831 13-S9 . i3

11 5 11. 5 : l. 'i 0. ~65 0 . 1850 1)71 . 79

? l. ~ 21 . 5 ?! 5 0. 965 0 . 181? 1366 . &J 1370. )1 0. 70 11 J . 3025 6137.'. 7') -

I

92

which can severely limit the measured permeability . Second, much of the data

was taken without benefit of specimen manometer ports. Thus, any contaminated

or smeared end conditions would render low results. The divergance of estimated

vs. observed data is even wider if bulk specific gravities are used in the Moulton

equation rather than apparent specific gravities.

Table 17 also reveals that of three of the four aggregate sources, the OGS

gradation exhibited higher permeability than the NJ. The exception was the DR-12

material. The DR-12 OGS density was significantly higher and the resulting

porosity was lower. This may explain why the DR-1 2 NJ specimens were more

permeable than the OGS specimens. For the other three aggregates, the porosities

I resulting from the two gradations were about equal. Moulton's equation

incorrectly predicts that the NJ would render higher permeabilities in all cases, due

to a slightly larger 0 10 size. Thus, it appears that some other sieve size(s) may

also be significant to permeability. It is noted that a plot on 0.45 power FHWA

I I

paper (Fig. 20) shows the NJ gradation much closer to the maximum density line

for the ( +) #4 sieve material. It also shows the gap-graded nature of the NJ. Fig .

21 shows this feature more clearly. Here the major difference between the two

gradations is the lack of #16 sieve size material in the OGS.

Table 17 also reveals that for both the NJ and OGS gradations, in five cases

out of eight, both gravels (DR-14 and 15) had greater permeabilities than the

crushed stone materials. This was somewhat unexpected, as the opposite is

generally held as true. However, at least one study (41) has shown results that

concur with those of the present study.

>-+'

U)

0 I...

0 Cl..

Q)

> +' u Q)

'+-'+-w

0.25

0.24

0.23

0.22

0.21

0.20

0.19

0.18

0.17

0.16

0.15 0.24 0.26 0.28 0.30 0.32

Porosity

Fig. 19. Relationship of Effective Porosity and Porosity.

93

I

I

/1 94

I

100 ,, . . --,, /

,I /

90 -LJ

I 80

Tl I Fl/ I

d I I , J I

I~ ., -IVI u A. 11 II U I 11 -, I I I ~ "1 n lC I IJ

I I

' _,._ -- . / I I \ - - - ._,._ . - -

70 '-' .... 'y L.. " .,

I I C, .::, LY ~ " . , I ~ .., r IP ,, -,, I .lou, lo..- kc- 0\/

,.--._ , / I - - - - -J

~

I ....__,, 60

CJ)

C:

~/ / I U/

u7 / ,, / I

" / , . / , Cf)

50 / 1,

Cf) / , ,

0 /I ,

0... 40 I/ I ,

/ I ,

/ I ,

0 / I ,

~

0 30 I-

/ J

/ I J , • , I I

20 ,,, , , ,

I I I/ , I/

10 " • I -

I 0 ,

200 50 JO 18 II 4 J/11 In J/4 In 1 In 1.50 In J In 100 1/2 In

Sieve Sizes • Middle

(0.45 Power) • New Jersey

I A. OGS

Fig. 20. FHWA 0.45 Power Paper Plot of Experimental Gradations.

I I I

50

40

-0 Cl) C ·-C +' 30 Cl)

a::: +' C Cl)

0 lo.. 20 Cl)

a..

0 New Jersey

• OGS

1.5 1 3/4 1/2 3/8 4 8 16 30 40 50 100 200 325 Pan

Sieve

Fig. 21. Plot of Individual Percent Retained for NJ and OGS Gradations.

95

I

96

Fig. 22 shows the relationship between porosity and permeability. The fact

I that there is a general trend of increasing permeability with increasing porosity is

supported in the literature. The same could be said of the effect of effective

porosity on permeability, as shown in Fig. 23. However, the effect is small over

I

I

I

the greater part of the range of porosities used.

A multiple regression equation was fit to the rigid wall open-graded

permeability data. Many combinations of variables were analyzed. These variables

included porosity, percent of maximum density, particle shape indicators U and IP,

effective porosity, various parameters representing gradation, and logs thereof.

The statistical criteria for final selection have been presented previously (29). The

model which most successfully predicted permeability had an R2 = 0.823, an

adjusted-R2 = 0. 779, and a standard error of estimate SEE = 287.9:

k=-40,962+15,284(n8,J +205.89(P16) +380.70(PDens) . . . . . . (25)

where:

k = permeability at 20°C, ft/day

/Jett = effective porosity

P16 = percent passing #16 sieve

Pd ens = percent of maximum achievable density.

Fig. 24 shows the relationship of permeability and estimated permeability for

the open-graded materials included in the present study.

Dense-Graded Materials

The gradation of all four aggregates (DR-12 through DR-15) were built

~

>. 0

"U ........... +-' '+-'--"

>. +-'

..0 0 Q)

E I... Q)

Cl..

• 2500

2000

• 1500

• 1000 •• • •

• 500

•• •

0 0.24 0 .26 0.28 0.30 0.32

Porosity

Fig. 22 . Relationship of Permeability and Porosity.

97

98

I 3000

I • 2500 ->.

. a, 'C

........... 2000 +' 111-1 ......... >. +' •.-1 1500 -•.-1 ,c cd Q)

I s 1000 J-4 Q) • ll.! •

I 500 • •

I 0 0.16 0.18 0.20 0.22 0.24

Effective Porosity

Fig. 23. Relationship of Permeability and Effective Porosity.

.a C

2500

2000

CD 1500 E L. CD

Q.

"D 1 000 CD > L. CD m .a 500 0

• •

•• .. o---~~__._~~~--~~~"'--~~__,_~~~...i

0 500 1000 1500 2000 2500

Estimated Permeability (ft/day)

Fig.24. Relationship of Observed Permeability and Estimated Permeability for Open- Graded Materials.

99

100

Table 19. Results of Flexible Wall Permeameter Permeability Testing.

Material Grad- Dry Density (pcf) Target MADD nASG n. i k (ft/day)

ation % pcf % Est. Test Target As-

Tested

DR-12 136.5 137.3 100.6 137.6 99.8 0.182 0.054 5.5 0.06 0.29

Middle 136.5 135.6 99.3 137.6 98.5 0.192 0.047 5.6 0.09 0.17

avg 100 99.2 0.08 0.23

137.6 137.9 100.2 137.6 100.2 0.178 0.042 6.4 0.06 0.15

137.6 138.9 101.0 137.6 100.9 0.172 0.033 6.1 0.04 0.59

avg 100.6 100.6 0.05 0.37

DR-13 138.3 137.1 99.1 141.0 97.2 0.210 0.063 6.3 0.16 1.20 Middle

138.3 138.2 . 99.9 141.0 98.0 0.203 0.032 5.8 0.13 2.07

avg 99.5 97.6 0.14 1.63

141.0 139.1 98.6 141.0 98.6 0.198 0.042 6.2 0.11 0.43 ,

141.0 139.6 99.0 141.0 99.1 0.194 0.033 5.9 0.10 1.44

avg 98.8 98.8 0.10 0.94

DR-14 132.5 131.1 98.9 134.5 97.5 0.207 0.061 6.0 0.15 1.36 Middle

132.5 133.9 101.0 134.5 94.6 0.190 0.042 6.1 0.08 0.78

avg 100.0 98.5 0.12 1.07

134.5 133.9 99.6 134.5 99.6 0.190 0.032 5.9 0.08 1.00

134.5 135.4 100.7 134.5 100.7 0.181 0.033 6.0 0.06 1.14

avg 100.2 100.2 0.07 1.07

DR-15 134.4 131.5 97.8 136.9 96.4 0.205 0.056 5.5 0.14 1.46 Middle

134.4 134.6 100.1 136.9 98.3 0.186 0.051 6.0 0.07 0.08

avg 99.0 97.2 0.10 0.77

136.9 134.1 98.0 136.9 98.0 0.189 0.050 5.9 0.08 0.84

136.9 136.6 99.8 136.9 99.8 0.174 0.041 6.0 0.05 0.40

avg 98.9 98.9 0.06 0.62

101

according to the MHTD Middle dense gradation and tested for permeability in the

flexible wall permeameter. Duplicate specimens were used. The target densities

corresponded to the low and high compactive efforts discussed previously. Each

specimen was tested five times at the recommended gradient equal to five.

Successive replications of test runs resulted in permeabilities that agreed well

within the recommended 25 percent. At the end of each suite of permeability

tests, the specimen was allowed to drain by gravity for 16 hours; then the drained

moisture content was determined and the effective porosity was calculated.

A summary of the test results is shown in Table 19. Average compacted

densities ranged from 97 .8 to 101.0 percent of target densities. Average

porosities ranged from 0.172 to 0.210. Effective porosities were approximately

27 percent of the standard porosities. The relationship between the two is

included in Table 19.

As with the rigid wall permeability results, all permeabilities determined by

test were considerably greater than those estimated by the Moulton equation. As

with the rigid wall results, there is a rough trend of increasing permeability with

increasing porosity and effective porosity.

A multiple regression equation was fit to the flexible wall dense-graded data.

Various combination of variables were analyzed. These included porosity

(calculated with either BSG or ASG values), effective porosity, particle shape

parameters U and IP, drained degree of saturation, and logs thereof. Effective

porosity is determined by draining the permeability specimen and measuring final

drained moisture content. From this is calculated the final drained degree of

I I I

I

102

saturation. The less permeable the material, the greater the drained moisture

content. Only the MHTD Middle gradation was tested, so gradation parameters

could not be used in the model. The most successful model had an R2 = 0.624,

an adjusted-R2 = 0.566, and an SEE = 0.381:

k=-8.457 +35.753(11)+0.033(satfin~ ........ . ... (26)

where:

k = permeability at 20°C, ft/day

fl = porosity (apparent specific gravity)

I satfinal = final degree of saturation, % .

I I I

Fig. 25 shows the relationship of permeability and estimated permeability for

the dense-graded materials included in this study.

In Table 20 is shown the relative effect of material property variables on

permeability. As can be seen, gradation change from open-graded to dense-graded

resulted in statistically different permeabilities at the 0.05 significance level. For

the open-graded gradations, the OGS was not significantly different from the NJ.

Secondly, a more highly compacted dense-graded material exhibited a lower

permeability than a material of lesser density, but not statistically so.

Unfortunately, the difference in percent maximum dry density was only on the

order of 1 to 1.5%. This may not have been a large enough difference to be

significant. In regard to particle shape, the flexible wall testing showed that the

gravels were not significantly different from the crushed stones. However, the

I rigid wall testing resulted in the DR-15 gravel being statistically more permeable

I I

-~ 2.5

C "tJ 2.0

' -,.;. -..c C I)

E ...

1.5

~ 1.0

"tJ I)

> ... I)

IJ ..c 0

0.5

• • • • 0.0

0.0 0.5 1.0 1.5 2.0

Estimated Permeability (ft/day)

Fig.25. Relationship of Observed Permeability and Estimated Permeability for Dense-Graded Materials.

103

I

I I

104

than both stones. But, it was also shown to be more permeable than the other

gravel, despite the fact that their particle shapes were not much different. Thus,

the effect of particle shape on permeability did not seem to be a factor for the four

aggregates tested. It should be noted that these aggregates did not exhibit a wide

range of shapes.

Table 20. Effect of Material Variables on Permeability.

Permeability (ft/day)

Condition Maximum Minimum Difference Significance

All Data at 0.05 level

Gradation OGS (avg) to 1158 0.925 1157 yes Middle (avg)

Open-Graded

Gradation OGS (avg) to 1158 820 338 no NJ (avg)

Particle Gravel to 1782 812 970 yes shape stone

Dense-Graded

Particle Gravel to 1.286 0.697 0.589 no shape Stone

% MADD Low CE (avg) 0.927 0 .750 0. 177 no to High CE

(avg)

Note: % MADD = % maximum achieveable density CE = Compactive effort

I I

105

ESTIMATION OF PERMEABILITY

It is sometimes useful to estimate permeability from readily available

information about a material rather than perform the permeability testing. As

previously mentioned, Moulton (9) has produced a widely-used algorithm based on

data from the literature ( 12-17). For the most part, these tests were of the rigid

wall low head variety, using either constant head or falling head procedures. There

was no provision for prevention of water short circuiting along the permeameter

walls. "Saturation" was usually achieved either by mere submergence of the

specimen or by applying a vacuum to the specimen and permeameter in a tub of

water. Thus, it is doubtful that saturation even approached 100% for the more

densely-graded materials. Additionally, manometer ports were missing in many of

the studies' permeameters.

Moulton's equation is based on porosity, effective aggregate size (010), and

the percent passing the #200 sieve. Three of the six references cited for the data

base lacked porosity data or the means to calculate porosity (specific gravity).

However, because compacted density data was available, apparently specific

gravity values were assumed. As previously shown, this can lead to a wide range

of calculated porosities.

In the present study, a regression equation was developed using data from

the literature and data from this study. Initially, the material parameters of interest

were gradation, porosity (or specific gravity and compacted density), and test

temperature (so that all data could be put on a common temperature basis). Very

few studies in the literature met these criteria. The acceptable data set is shown

106

in Table 21. The materials included dense (well) graded granular base material,

sand, moderately open-graded material, and very open (uniformly) graded granular

base material. Natural sands, gravels and crushed stones were represented.

Maximum aggregate sizes ranged from #8 sieve to 2 in, percent passing the #200

sieve from Oto 15%, and 0 10 sizes from 0.032 to 7 mm. The gradations are

shown in Fig. 11. The model would have been more accurate by the inclusion of a

measure of the activity of the fines, such as plasticity index, but unfortunately very

few studies report this sort of data. The possible variables available for analysis

were measures of overall gradation curve shape/position, individual sieve sizes, and

porosity.

As previously discussed, the particle sizes that correlated best with

permeability were the #4, #8, and 3/8 in sizes. Gradation descriptors that were

-analyzed were D10 size and Hudson's A. For porosity, despite the previously

mentioned idea that BSG would be more appropriate than ASG in calculation of

porosity, it was determined that ASG correlated better with permeability. Thus,

ASG-based porosities were considered.

The criteria for model selection have been delineated in the companion study

of this report (27). The model selected had an R2 = 0.906, an adjusted-R2 =

0.900, and an SEE = 0.433:

logk = 17 .358 +8.9951ogl]+0.5911ogD10-5.5111ogP3/8-0.3491ogP200 (27

) or

k = (2.28x1017){1J)8.995(D,o)0.591 ~(P3/8)5.511 (P200)0.349]

where:

- -- -

- -- -

Ta

ble

21

.Da

ta U

sed

in

th

e P

erm

ea

bili

ty P

red

icti

ve E

qu

ati

on

.

RE

F.

MA

TE

RIA

L

PE

RC

EN

T P

AS

SIN

G

1.5

3

/4

3/8

#

4

#8

#

16

#

30

17

C

LS

1

00

6

3

33

2

3

16

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

16

11

8

CL

S

10

0

63

3

3

23

16

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

16

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CL

S

10

0

63

3

3

23

1

6

11

8

CG

1

00

1

00

4

9

36

2

8

22

1

7

CG

1

00

1

00

4

9

36

2

8

22

1

7

CG

1

00

1

00

4

9

36

2

8

22

1

7

CG

1

00

1

00

4

9

36

2

8

22

1

7

CG

1

00

1

00

4

9

36

2

8

22

1

7

CG

1

00

1

00

4

9

36

2

8

22

1

7

CG

1

00

, 1

00

4

9

36

2

8

22

1

7

CG

1

00

1

00

4

9

36

2

8

22

17

CG

1

00

1

00

4

9

36

2

8

22

1

7

CG

1

00

1

00

4

9

36

2

8

22

1

7

15

S

AN

D

10

0

10

0

10

0

10

0

10

0

99

9

5

SA

ND

1

00

1

00

1

00

1

00

1

00

9

9

95

SA

ND

1

00

1

00

1

00

1

00

1

00

9

9

95

SA

~'"

1

00

1

00

. 1

00

1

00

1

00

9

9

95

SA

~

10

0

10

0

10

0

10

0

10

0

99

9

5

SA

i\ 1

00

1

00

1

00

1

00

1

00

9

9

95

SA

N:J

1

00

1

00

1

00

1

00

1

00

9

7

80

SA

NO

1

00

1

00

1

00

1

00

1

00

9

7

80

SA

ND

1

00

1

00

1

00

1

00

1

00

9

7

80

SA

ND

1

00

1

00

1

00

1

00

1

00

9

7

80

SA

ND

1

00

1

00

1

00

1

00

1

00

9

9

98

SA

ND

1

00

1

00

1

00

1

00

1

00

9

9

98

SA

ND

1

00

1

00

1

00

1

00

1

00

9

9

98

SA

ND

1

00

1

00

1

00

1

00

1

00

9

9

98

41

S&

G

10

0

67

1

9

4 2

0 0

S&

G

98

7

2

32

1

2

8 7

6

S&

G

98

7

7

54

3

7

28

2

5

23

-- -

-0

10

A

#5

0

#1

00

#

20

0

mm

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

5 3

0 0

.94

2

.62

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

12

8

5 0

.2

3.7

7

75

3

2

12

0

.06

2

8.1

3

75

3

2

12

0

.06

2

8.1

3

75

3

2

12

0

.06

2

8.1

3

55

2

0

15

0

.03

2

7.8

4

55

2

0

15

0

.03

2

7.8

4

55

2

0

15

0

.03

2

7.8

4

20

3

1 0

.25

7.0

1

20

3

1 0

.25

7

.01

20

3

1 0

.25

7

.01

20

3

1 0

.25

7

.01

70

1

5

6 0

.12

7

.88

70

1

5

6 0

.12

7.8

8

70

15

6

0.1

2

7.8

8

70

1

5

6 0

.12

7

.88

0 0

0 7

1.9

2

5 4

3.5

3

.9

2.4

75

18

8

4 0

.16

3

.72

- n

0.2

61

76

0.2

61

76

0.2

34

06

9

0.2

31

71

2

0.2

10

50

2

0.2

08

73

4

0.1

91

64

8

0.1

89

88

0.1

89

88

0.1

75

15

1

0.1

73

97

2

0.1

46

28

1

0.1

46

0.2

75

54

5

0.2

77

94

6

0.2

48

53

5

0.2

16

12

4

0.2

15

52

4

0.2

11

92

3

0.2

10

12

2

0.2

04

12

0.2

05

32

0.2

01

11

9

0.3

96

0.3

66

0.3

45

0.3

56

0.3

15

0.2

78

0.3

81

0.3

75

0.3

55

0.3

4

0.4

2

0. 4

08

0.3

83

0.3

57

0.3

7

0.3

2

0.2

3

-k fp

d

26

67

23

14

15

99

21

50

911

12

77

73

0

65

7

10

40

29

2

21

6

52

44

12

99

10

14

73

4

181

16

8

13

4

88

50

34

22

39

.5

24

.5

18

.4

6.9

2.4

1.4

25

8.5

23

7.5

19

9.4

17

5.4

77

.1

66

.3

54

.5

34.1

21

54

6

18

14

4

1.2

.... 0 ....

J

-

K)G

S

GR

AV

EL

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.25

7

93

8

GR

AV

EL

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.26

5

94

8

GR

AV

EL

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

. 07

0

.32

9

16

13

GR

AV

EL

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.30

8

12

92

GR

AV

EL

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.31

8

27

46

CL

S

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.28

6

10

27

CL

S

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.291

1

04

8

CL

S

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.24

7

83

CL

S

10

0

91

48

3

0

16

7

6 4

3 2

1.5

3

.07

0

.24

9

55

2

NJ

GR

AV

EL

10

0

95

5

8

47

2

0

5 4

2.5

2

.5

2 1

.7

3.3

6

0.2

68

2

28

GR

AV

EL

10

0

95

5

8

47

2

0

5 4

2.5

2

.5

2 1

.7

3.3

6

0.2

6

88

9

GR

AV

EL

10

0

95

5

8

47

2

0

5 4

2.5

2

.5

2 1

.7

3.3

6

0.3

17

1

73

0

GR

AV

EL

10

0

95

5

8

47

2

0

5 4

2.5

2

.5

2 1

.7

3.3

6

0.3

18

9

55

CLS

1

00

9

5

58

4

7

20

5

4 2

.5

2.5

2

1.7

3

.36

0

.29

45

4

02

CL

S

10

0

95

5

8

47

2

0

5 4

2.5

2

.5

2 1

.7

3.3

6

0.2

86

7

73

CLS

1

00

9

5

58

4

7

20

5

4 2.

5 2

.5

2 1

.7

3.3

6

0.2

91

40

2

CL

S

10

0

95

5

8

47

2

0

5 4

2.5

2

.5

2 1

.7

3.3

6

0.2

8

11

78

MID

DLE

C

LS

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.1

92

0

.16

54

CLS

1

00

9

5

63

5

0

40

3

3

28

22

1

6

8 0

.09

2

4.5

5

0.1

82

0

.28

95

CLS

1

00

9

5

63

5

0

40

3

3

28

2

2

16

8

0.0

92

4

.55

0

.17

8

0.1

52

5

CLS

1

00

9

5

63

5

0

40

3

3

28

22

1

6

8 0

.09

2

4.5

5

0.1

72

5

0.5

94

8

CLS

1

00

9

5

63

5

0

40

3

3

28

22

1

6

8 0

.09

2

4.5

5

0.2

1 1

.19

62

CLS

1

00

9

5

63

5

0

40

3

3

28

22

1

6

8 0

.09

2

4.5

5

0.2

03

2

.07

25

CL

S

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.1

98

0

.43

2

CL

S

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.1

94

1

.44

25

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.2

07

1

.36

58

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.1

9

0.7

80

2

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.1

9

1.0

04

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

16

8

0.0

92

4

.55

0.

181

1.1

36

9

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.2

05

1

.46

27

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

1

6

8 0

.09

2

4.5

5

0.1

86

0

.08

47

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

16

8

0. 0

92

4

.55

0

.18

9

0.8

37

9

GR

AV

EL

10

0

95

6

3

50

4

0

33

2

8

22

16

8

0. 0

92

4

.55

0

.17

4

0.4

03

2

No

te:

CLS

= c

rush

ed

lim

est

on

e

CG

, ~

rush

ed g

rave

l

S&C

sa

nd a

nd g

rave

l

I

I

I

I I I I I

I I

109

k = coefficient of permeability at 20°C, ft/day

fJ = porosity = 1 - (density, pcf/(apparent sp. grav. * 62.4))

D10 = size that represents 10% passing, mm

P 3/8 = percent passing 3/8 in sieve

P200 = percent passing #200 sieve.

The relationship of observed permeability with estimated permeability is shown in

Fig. 26.

As mentioned previously, the Moulton equation underpredicts observed

permeability by almost an order of magnitude. This is shown in Fig. 27; the wide

scatter should be noted. It is postulated that the above inaccuracies are

associated with the manner in which the permeability testing was performed

(complete saturation unlikely) and inaccuracies in porosity determination. The

algorithm developed in the present study is relatively accurate up to permeabilities

of around 1000 ft/day. Beyond that, the model overpredicts permeability

significantly. This is shown in Fig. 28 (arithmetic-scale version of Fig. 26).

Despite all efforts of arriving at good estimates of field permeability, it must

be remembered that there are several factors that will render field permeabilities

different from laboratory derived values. First, it is unlikely that the pavement base

layer will be close to 100% saturated--even in the laboratory, this is very difficult

to achieve under conditions of vacuum. Entrapped air can lower the expected

permeability significantly. Additionally, in pavement sections, the permeant

temperature will be higher or lower than the standard 20°C -- thus raising or

lowering actual permeability. And, the aggregate can become segregated and the

110

density may be off-target, all resulting in a permeability different than expected.

So, the question becomes, which predictive equation should be used if

actual laboratory permeability testing is not possible 7 The equation developed in

this study appears to be a truer representation of permeability in the range of 0. 1

to 1000 ft/day. The Moulton equation underpredicts permeability at all levels.

However, in the field, several factors are at work which result in achieved

permeabilities lower than that which are expected. In this regard, a conservative

estimate may be in order, and the use of the Moulton equation may be justified.

This decision falls in the area of design judgement.

1, I I I I I I I I I I I I

I I I I I

->. C

"tJ

' .... -.._,, CD -C u

(/l

Cl 0

_J

->. .... .c C CIJ

E L. a,

a.. "O a, > L. CIJ rn .c 0

5

4

J •

2 ••

1

0 •

-i •

-2 -1 0 1 2 J 4 5

Estimated Permeability, Log Sec le (ft/dcy)

Fig.26. Relationship of Observed Permeability end Estimated Permeability for Several Studies.

1 1 1

->,

~ 2500

' -.... ->. 2000 .... .c C ID

E ... cu

a.. 'lJ

CD > ... CD rn .c 0

1500

1000

500

0

• •

• •

• • •

• • • • • • •

• • • • •••

0 100 200 300 400 500

Moulton Estimated Permeability {ft/day)

Fig.27. Relationship of Observed Permeability and Permeability Estimated by Moulton Equation.

112

I I

I I

,.... ~ C

"'C

' ~ --~ ~

.a C G

E t.... G a.. "C I)

> t.... G ., .a 0

3000

• • 2500

2000

1500

1000

500

o-------------...... ---------0 2500 5000 7500 10000 12500

Estimated Permeability (ft/day)

Fig.28. Relationship of Observed Permeability and Permeability Estimated by UMR Equation.

113

I

I

I

I

I I

I I I I

I I I I I I

RESULTS OF MODELS EVALUATION

TTI INTEGRATED MODEL

114

This model does not appear suitable for use in the determination of drainage

coefficients. First, it does not have drainage coefficient determination as a goal of

the model, and consequently does not generate drainage coefficients or have

drainage coefficients as an output of the program. However, since some of the

output from the program included the modulus of the pavement layers, a scheme

was devised to use the base course modulus generated by the TTI program in the

calculation of drainage coefficients. Unfortunately, this proved to be impossible

because of limitations in the TTI model output as discussed below.

Review of the program documentation (7) leads one to believe that base and

subbase modulus vs time is an output of the CMS model module. This is shown in

Fig . 1 on page 3 of the documentation. Also, on page 75 of the documentation

under the discussion of the CMS model module, the following statement is made,

"A subroutine to compute accompanying changes in material moduli is

incorporated in the model." After several unsussessfull attempts to obtain base

modulus output that differed from the initial input, the following comment was

discovered on page 82 of the documentation: "The CMS Model, therefore, ignores

the effect of moisture on the resilient modulus of these granular materials when

they are unfrozen." This in fact is correct. The output parrots the input base

course modulus for the frozen and unfrozen condition. This is a fatal flaw in any

program that purports to model the environmental effects on pavement behavior.

This rendered the program unusable for purposes of m-coefficient determination.

115

Other considerations of the TTI model include the sensitivity of results to

changes in inputs. The authors identified the following inputs as highly sensitive:

length of elements used by the CMS program to model the soil pavement column,

temperature at which the soil is considered frozen, the liquid limit for fine grained

soils, material type, plastic limit of the non-asphalt layers, stiffness relationship of

the asphalt bitumen and temperature, saturated water content, water table depth,

maximum daily air temperature, minimum daily air temperature, temperatures for

the bitumen stiffness-temperature relationship, coefficient of variation for

unsaturated permeability, temperature at the upper boundary, modifier of

unsaturated permeability in the frozen zone per layer, saturated permeability of the

soil per layer, exponent of pore pressure for Gardner's unsaturated permeability

function per layer, base course hydraulic permeability, index of base course

material, and rainfall occurrences. According to the authors, the model displayed

low sensitivity to the following inputs {a partial list only): base depth, porosity of

the subgrade, amount of sand and gravel in the base course, and linear length of

cracks and joints and cracks. In the opinion of this reviewer, these low sensitivity

items should be important players in a model that is dealing with drainage of base

courses. Further, it was discovered by trial and error that the prediction of

temperatures in the pavement layers was driven by the input selected for thermal

capacity and resistance to heat flow in each layer. The results generally obtained

were counter-intuitive and did not agree with those obtained by other methods.

An attempt was made to determine if FHWA had evaluated the Integrated

Program separately from the Texas evaluation. ·Persons contacted included Byron

I I

I

I

I I

I

116

Lord, Roger Larson, Jim Sherwood, and Bill Dearasaugh of FHWA in July, 1992.

These gentlemen had not used the program and did not know of any serious users.

Mr. Larson relayed that he had the understanding that the program was

cumbersome and did not handle moisture well. Mr. Larson suggested that Dr.

Barry Dempsey of the University of Illinois might have had an opinion on the

program. Contact with Dr. Dempsey revealed that he thought the program was

not sensitive to moisture.

The TTI program is cumbersome, requires a great deal of input, and outputs

the frozen and unfrozen base course modulus that is input by the users regardless

of the other input variables. For the user interested in base course moduli adjusted

for climatic conditions, this program is not useful.

DAMP

The parts of DAMP that are based upon the work of Moulton as published in

the FHWA Highway Subdrainage Design Manual seem well founded in verifiable

research efforts published in the literature. The parts of DAMP added to the

original to extend the work to drainage coefficient determination are less well

supported.

The percent time the pavement structure is saturated is determined using a

method based upon Thornthwaite's classification of climatic regions. This method

is rational on a regional basis. The leap from Thornthwaites's determination of

monthly water surplus conditions to percent of time saturated is discussed earlier

in this paper. Here, Carpenter simply states that some part of a month the base

course shall be considered saturated because of some conditions being met. For

117

instance, "Surplus following a recharge which does not follow a frozen period

here one-fourth of the months in the surplus period should contribute to the

saturation time." No rationale or references are available to determine the basis for

the allocation of time saturated. Also, the relationship between general soil

moisture content and base course moisture content is assumed to be equal.

Another problem with DAMP is the calculated percent time of saturation.

DAMP was used to analyze data for several sites per each of the six AASHTO

climate zones . First, DAMP classified all six zones into the 5-25 percent saturation

time catagory. Worse yet, when counting months as frozen, thawed, wet, and

dry, problems arose in distinguishing zones. For instance, the Road Test area,

normally thought of as being a moist area, became very similar to Phoenix. One

explanation may be as follows. It was noted that in all the dry (west of the

Mississippi River) states, DAMP puts a 10 cm storage (maximum) in January, no

matter what the rainfall and PET were in previous months. It takes until about July

for this large amount of storage to be diminished to a level that is more in keeping

with the percent-of-time-of-saturation rules stated in DAMP. Included in the output

of DAMP is the Thornthwaite Moisture Index:

TM/ = (100 S-600,/PET . . . . . . . . . . . . . . . (28)

where:

TMI = Thornthwaite Moisture Index,

S = surplus of moisture during the year,

D = deficit of moisture during the year,

I I I

I I I I I I I I I I I I I I

118

PET = potential evapotranpiration during the year.

It appears that when a threshold value of TMI between 5.8 and 18 is calculated,

10 cm storage is automatically invoked in January. If the TMI is greater than the

threshold, the output of DAMP storage seems to follow the stated rules.

The Quality of Drainage determination presented by Carpenter (Table 2) is

entered with information on the subgrade drainability and the base drainability.

These values are determined in a rational manner and are logical. The matrix of the

table is filled in with descriptive terms for the Quality of Drainage such as

Excellent, Good, Fair, Poor, and Very Poor. No rational or documentation is

presented for the position of these descriptors in the table. Cursory review reveals

a base that drains in less than five hours that is placed on a typical low

permeability subgrade earns the Quality of Drainage description of Fair. This

"Fair", when used in the Drainage Coefficient selection (Table 1) for typical

Missouri times of saturation yields an m-coefficient of, say, 0.80. If this is tied

back to the AASHO road test, the base that drains in less than five hours is less

desirable than the AASHO base that had difficulty in draining below 85%

saturation during spring periods. This does not make sense and does not correlate

well with the general assumption in the pavement design community that well­

drained bases are more desirable.

A sensitivity analysis was conducted on DAMP by establishing the most

likely value for each variable, a high value and a low value. Several runs were

made with each successive run changing one variable from low to median to high

while holding all other variables constant at the median value. Base drainability

119

was the variable with the greatest effect upon the drainage coefficient. With a

Fair-draining subgrade, changing from Very Poor to Excellent base drainage

changed the m-coefficient from 0.4 to 1.2, although the improvement was only

from 0.40 to 0.80 for a poor-draining subgrade. The permeability of the subgrade

was the next most important variable in the program. For an excellent-draining

base, the m-coefficient changed from 0.80 to 1.20 as subgrade drainage varied

from Poor to Good, although for a very poorly-draining base, the effect of subgrade

drainability was negligible. Table 22 shows the results of the sensitivity analysis.

Unless noted, the standard situation was: W = 13 ft, H = 6 in, g = 0.02,

subgrade drainability = Fair, base drainability = Very Poor. Most other variables

had no discernable effect upon the drainage coefficients.

Table 22. Drainage Coefficient Sensitivity Anaysis for DAMP.

Variable Value/m-coefficent Range in Change

W, ft 13/0.40 25/0.40 37/0.40 0.0

H, in 6/0.40 -- 12/0.40 0.0

g, ft/ft 0.0/0.40 0.03/0.40 0 .06/0.40 0.0

Subgrade Drainability

kh = V. Poor poor/0.40 Fair/0.40 Good/0.40 0.0

kh = Fair poor/0.60 Fair/0.80 Good/1.00 0 .4

kh = Excellent poor/0.80 Fair/1.20 Good/1.20 0.4

Base Drainability

SG = Poor V. Poor/0.40 Fair/0.60 Ex./0.80 0.4

SG = Fair V. Poor/0.40 Fair/0.80 Ex./1.20 0.6

I I I I I I I I I I I I I I I I I I I

120

CONTRAST BETWEEN THE INTEGRATED PROGRAM AND DAMP

Infiltration of surface water into the base course is modeled using

Ridgeway's or Carpenter's methods in DAMP and Ridgeway or Dempsey-Robnett's

methods in the TTI program. There is a choice in each program.

Meltwater contribution to the base course is modeled in the DAMP program

using Moulton's approach of heave rate based upon Unified Soil Classification

system correlations. The TTI program does not add melt water to the water

contained in the base course for saturation purposes.

DAMP provides routines for groundwater and artesian inflow estimation.

TTl's program does not offer this feature .

DAMP calculates time for drainage using Moulton's regression equation for

permeability and Casagrande-Shannon's relationships for time-to-drain. The TTI

model uses a model by Liu fil .fil. that assumes a parabolic phreatic surface and

permeable subgrade for time-to-drain. At a degree of drainage of 0.5, the

Casagrande model better agrees with tests conducted by Casagrande. Liu's model

better predicts at degrees of drainage at either end of the scale. In general, the

region of interest for base course is around 0.5 degree of drainage. The TTI

method uses permeability and porosity inputs for the drainage time calculations

and further modifies those by use of the PD table presented earlier. Thus, if the

input permeabilities are based upon data from material that included plastic fines,

the TTI program will adjust for this fact a second time in the PD calculation .

Saturation exposure has been discussed earlier. DAMP uses the

Thornthwaite method to estimate the saturation exposure. TTI uses the

121

probabilities of wet and dry base course on a monthly basis.

Subgrade drainage is included by DAMP using the NOi concept discussed

earlier. TTI includes this in the drainage analysis when calculating time to drain.

The quality of drainage determination in DAMP is based upon the

combination of time-to-drain and the subgrade drainage. In TTI, both time-to-drain

and subgrade permeability are considered in the infiltration and drainage model to

produce a satisfactory or unsatisfactory drainage categorization. The DAMP

descripton of "excellent" generally corresponds to TTl's "satisfactory".

DAMP has a procedure to select drainage coefficients. TTI does not select

drainage coefficients.

I I I I I I I I I I I I.

I I I I I I I

GENERAL

122

DRAINAGE COEFFICIENT DETERMINATION

Drainage (m) coefficients may be determined by using DAMP. The detailed

procedure is included in APPENDIX 8, "Use of DAMP Manual," and has been

discussed to a certain extent in previous sections.

An alternative method was developed as a part of this study. This new

method is similar in nature to the DAMP method, but the manner in which the

drainage coefficients were calculated for the m-coefficient table is different. Also,

the "Quality of Drainage" table has been modified. In practice, the user can

determine m-coefficients manually, as explained in the next section, or by use of ·

the computer spreadsheet MODAMP, which was also developed as a part of this

study. MODAMP is a single-screen spreadsheet (using spreadsheet program

QUATIRO PRO or LOTUS 123) that incorporates much of that which is in DAMP,

plus the new m-coefficient table. Additionally, the Missouri weather station files

incl.uded with this report are easily imported into MO DAMP to aid · in the time-to­

drain calculations. A user's manual, "MO DAMP Manual," is included in Appendix

C.

AASHO ROAD TEST

One cannot discuss m-coefficients without mentioning the drainage

condition at the Road Test. By definition, the m-coefficient was 1 .0. Looking at

Table 1, an analysis by DAMP indicates that Ottawa, Illinois is in the 5-25%

saturation column. Hence, for m = 1 .0, the Quality of Drainage must have been

"Good" or "Fair." The granular base material had 11 % minus #200. For this

material, Moulton's equation predicts a permeability of 0.03 ft/day. The soil at the

123

site was an A-6. This combination of base and subgrade does not appear to be of

a good drainage quality. Indeed, Carpenter's Table 2 rates the drainage as very

poor, thus Table 1 would indicate m = 0.40 to 0. 75.

In setting up Table 1, the authors of Appendix DD in the AASHTO Guide

based a "Good" rating on a base modulus of 30,000 psi, which is the value that

was assigned to the Road Test material. However, in the same document, the

permeability of the base was put at 0.1 ft/day which renders time-to-drain as 1 O

days, which in turn was rated as between "Fair" and "Poor" drainage. The

discrepancy of drainability description was not reconciled. However, in a study by

Haynes and Yoder (53), permeability test results on Road Test crushed stone

granular material were reported as 7.5 ft/day. Furthermore, the laboratory

specimens were so permeable that a uniform moisture content could not be

achieved because the water kept draining down to the bottom portion of the

specimen, indicating a well draining material. Additionally, trench studies at the

Road Test indicated that at the times that the base was tested, the degrees of

saturation were 45 to 60%. Only during the short spring breakup period was

saturation higher. And, the Road Test cross section featured the granular base and

subbase layers extending to the ditches. The ditch bottoms were well below the

grade, and the topography was relatively level. Considering all this, perhaps the

subgrade did behave in a "Fair" manner (not contributing water) and perhaps the

base did manage to be (barely) rated as Fair because it apparently drained in less

than a month (720 hr), which would put the permeability at about 1 ft/day. And,

if one does not count the frozen months as saturated months, then the time of

I I I I

I I I I I I I I I I

I I I I I I I I I I I I I I I I I I I

124

saturation would compute close to 25%. So, if the site was in the 5 to 25%

column, it is conceivable that the table could be reconciled to an m = 1.0.

Certainly, it is conceivable that some subgrade drainage situations could be

considerably worse, and some bases could be significantly less permeable than 1

ft/day . So, a "Fair" rating is reasonable in this context. Finally, it makes sense

that if the Road Test conditions are to be considered as a baseline to which all

other conditions are compared, than a middle value or "Fair" rating is sensible.

DETERMINATION OF AASHTO DRAINAGE COEFFICIENTS-UMR METHOD

Perhaps the best way to describe the manner in which the m-coefficients

were developed in this study is to look at how the designer will use the design

tables, then discuss how the tables were developed.

General Methodology

The AASHTO Guide provides the framework for choosing an m-coefficient

through use of a table (Table 1 ). This has been replaced by Table 23. Two sorts

of information are required: · 1) "Climate Condition" which deals with the climate in

which the project site is located, and 2) "Quality of Drainage" which relates to

how well the pavement structure sheds the water that enters it and the amount of

water supplied by the various components of the highway cross-section. Included

in the following sections is a method to determine the Climate Condition (column

choice in Table 23) and a method to determine the Quality of Drainage (row

choice). The Quality of Drainage is found by use of Table 24 by knowing: 1) the

Granular Layer Quality of Drainage (Base or Subbase), and 2) the Quality of

Subgrade Drainage. The purpose of this section is to provide information to assist

I 125 I

the Missouri designer in determining the above three inputs.

Table 23.

Quality of

Drainage

Excellent

Good

Fair

Poor

Very Poor

Table 24.

Quality of Subgrade Drainage

Good

Fair

Poor

Very Poor

Recommended Drainage Coefficients for Flexible Pavements for Untreated Base and Subbase Materials.

Climate Condition

A B C D E F

1.25- 1.25- 1.25- 1.25- 1.20- 1.20-1.20 1.20 1.20 1.20 1.15 1.15

1.25- 1.20- 1.20- 1.20- 1.20- 1.20-1.20 1.15 1.15 1.15 1.15 1.15

1.20- 1.15- 1.05- 1.05- 1.05- 1.15-1.15 1.05 0.85 0.85 0.85 1.05

1.15- 1.15- 1.05- 1.05- 1.05- 0.85-1.05 1.05 0.85 0.85 0.85 0.70

1.05- 1.05- 0.85- 0.70- 0.85- 0.70-0.85 0.85 0.70 0.60 0.70 0.60

MODAMP Quality of Drainage.

Granular Layer Quality of Drainage (Base or Subbase)

Excellent Good Fair Poor Very Poor < 2 hr* 2 to 24 hr 24 to 168 hr 168 to 720 hr > 720 hr

Excellent Good Fair Fair Fair-Poor

Good Good-Fair Fair Fair Poor

Fair Fair Fair Poor Very Poor

Poor Very Poor Very Poor Very Poor Very Poor

* drainage time to 85% saturation

One problem with the original AASHTO Table (Table 1) is the method of choice of

column. The choice is dependent on the percent of time that the soil in the general

area of the project site is close to saturation. This does not deal well with a soil

I I I I I I I

I I I I I I I I

I I I I I I I I I I I I I I I I I I I

126

that undergoes a seasonal change in resilient modulus as shown in Fig. 29, nor

with the conditions of drainage within the pavement structure. To address this

issue, it is necessary to determine the number of months in the year that the

subgrade is frozen, thawed, wet (recovering from thaw condition) and "dry"

(relative to wet). One way to do this is to import site-specific weather station data

into MODAMP and delineate the number of dry and wet months. From local

experience and MODAMP output, the number of months frozen can be estimated.

With this information, the user should enter Table 25 to find the Climate Condition

description closest to that of the project site. With the Climate Condition defined,

the column choice in Table 23 can be made. It is recommended that for sites in

Zones fl and V, a minimum interval of 0.5 month for frozen and thawed periods

should be used to account for freeze-thaw periods. This will help distinguish

between zones with some of this kind of activity from zones with none. The

season lengths were based on the results of an analysis of weather data by use of

MODAMP for several sites from each of the six AASHTO climate zones. The basic

methods for calculation of water surplus and deficit are identical to those of

DAMP. However, the rules for identifying a particular month as to its providing a

certain saturated pavement time interval were modified to render a better

differentiation between climate zones. In essence, the only rule that was changed

was that if the subgrade was storing the maximum amount of water possible ( 10

cm) and there was a surplus, then the full month was considered to be in a

saturated state. Under certain conditions, DAMP would have considered this as a

one-fourth month saturated interval, namely if there was not a previous frozen

X

(/)

Cl.

(/)

::::,

::::, "'O 0 ~

(/)

Cl> 0::::

22

20

1 8

1 6

1 4

1 2

1 0

8

6

4

2

0 J F M A M J J A s 0 N

Fig. 29. Variation of Subgrade Resilient Modulus Through the Year.

127

D

I I I I I I I I

I I I I I I

I I

I I I I I I I I I

I I I I I I I I I

128

month. This change allowed areas such as the mid west to show saturated

condition intervals greater than arid regions, which seems to make sense.

Unfortunately, the total percent time of saturation tends to appear excessive.

However, it was believed that this was the lesser of the two evils, and if in error,

was on the conservative side.

Table 25. Climate Condition Season Lengths.

Climate Season (Months) Condition

Roadbed Roadbed Roadbed Roadbed Frozen Thawing Wet Dry

A 0.0 0.0 2.0 10.0

B 0.0 0.0 5.5 6.5

C 3.0 1.5 1.0 6.5

D 0.5 0.5 1.5 9.5

E 3.0 1.5 2.0 5.5

F 0.5 0.5 5.0 6.0

If local weather data are not available, a second, more generic, way is to

utilize the zone method. The AASHTO Guide divides the USA into six zones in

regard to climate, as shown in Fig. 30. Examination of weather data indicates that

the following relationships can be used to choose the proper column in Table 23:

I

REGION CHARACTERISTICS

I . Wet, no freeze

JI Wet, freeze - thaw cycling m Wet, hard-freeze, spring thaw .nz: Dry, no freeze Jz: Dry, freeze - thaw cycling E: Dry, hard freeze , spring thaw

Fig.JO. Six Ciimate Zones in the United States (after 1986

AASHTO Guide for Design of Pavement Structures).

129

I I I I I I I I I I I I I I I I I I I

130

Table 26. Zone - Climate Condition Relationships.

ZONE Table 23 Column

I 8

II F

111 E

IV A

V D

VI C

Note that even though Zone V is quite dry, it is recommended that the designer

consider downgrading the column choice from "less than 1 % to the "5 to 25%"

column if freeze/thaw conditions exist at the particular project site. Choice of row

will be presented next.

It should be noted that for Zones I, II, and Ill (high TMI areas), the storage

calculated by DAMP and MODAMP are essentially identical. Below some TMI

threshold value (somewhere between 5.8 and 18), usually in Zones IV, V, and VI,

DAMP appears to automatically place a 10 cm (maximum) storage in January,

MODAMP does not. In Zones IV through VI output, DAMP and MODAMP output

agree from July through December. It is believed that MODAMP is functioning

correctly because the output agrees well with the various examples in

Thornthwaite's paper (19).

Several Missouri sites were examined. By looking at: 1) Appendix D

weather files for temperature, 2) the output of MODAMP for number of months

saturated (but not frozen or thawing), and 3) Table 25 for Climate Condition

131

(tempered with the suggested number of months thawed), the proper Climate

Condition was found and is shown in Table 27. The Climate Condition was

compared to the corresponding AASHTO Zone by use of Table 26. The two types

of descriptions were found to agree well.

Table 27. Determination of Climate Condition for Several Missouri Sites.

Location Season (Months) Climate Condit-

Roadbed Roadbed Roadbed Roadbed ion Frozen Thawing Wet Dry

Amity 3 1.5 3.5 4.0 E

Rolla 1 0.5 5.5 5.0 F

Springfield 1 0.5 6.5 4.0 F

Sikeston 0.5 0.5 7.0 5.0 F

Obviously, the months that the subgrade is frozen and thawing is only

approximate. Local experience will give a much more realistic estimate.

Quality of Base Drainage

AASHTO Zone

Ill

II

II

II

The Quality of Base Drainage is a function of the time to drain to 85%

saturation. For any pavement cross section and granular material permeability,

MODAMP quickly computes the drainage time. Built into MODAMP is the Moulton

equation for prediction of permeability. MODAMP allows the designer to adjust the

material gradation and dry density to achieve the drainage time that is sought.

Once the time of drainage is determined, the Quality of Base Drainage is

determined in accordance with Table 27. The Quality of Base Drainage criteria

and approximate permeabilities required to meet the criteria are shown in the table.

I

I

I 132

These were determined by making numerous runs of MODAMP with various

I material characteristics. Note that the criteria are in accordance with the AASHTO

I

I I I

I I

Guide, as opposed to those of Carpenter.

The reason that the Appendix DD criteria were adopted rather than those of

Carpenter (DAMP) is that the less stringent criteria of Appendix DD fit better into

the concept that the Road Test drainage condition should be rated as fair. If one

looks at Carpenter's criteria in Table 2, it is seen that the Quality of Drainage at

the Road Test would be rated "Very Poor," and by use of Table 1 the combined

drainage condition plus environment would be considered close to the worst in the

country. This is not reasonable. Plus, the resulting "m" would be 0.4 to 0. 75,

and that cannot be. As discussed previously, use of Appendix DD drainage criteria

leads to a more reasonable assessment of "m" at the Road Test.

Table 28. Required Permeabilities for Quality of Drainage Levels.

Quality of Base or Time to Drain (hr) Required Permeability* Subbase Drainage (ft/day)

Excellent < 2 > 700

Good 2-24 40-699

Fair 24-168 5-39 ~

Poor 168-720 1-4

Very Poor >720 <1

* Based on drainage path = two 12 ft lanes plus 6 ft shoulder; 4 in. thick layer; cross slope = 0.0156, grade = 2%.

From Table 28, it can be seen that most of the open-graded materials tested in this

study would fall into the "Excellent" category when using actual permeability

values or "Good" when using Moulton estimates. Conversely, the MHTD Middle

133

gradation with 8% minus #200 sieve (0.1 ft/day per Moulton or 1 ft/day per test)

would be regarded as "Very Poor" (via Moulton) or "Poor" (via test). These

outcomes fit well with the experience gained from observing these materials both

in the laboratory and in the field.

Typically, an open-graded base will be underlain by a dense-graded subbase,

which serves to protect the base from soil intrusion. Thus, the base Quality of

Drainage may be "Excellent" while the subbase may be "Very Poor".

To assist the designer in achieving a given level of drainage by adjustment of

gradation, a regression equation has been developed to estimate compacted dry

density from gradation and specific gravity data. The adjusted-R2 = 0. 729 with a

S.E.E. = 5.22:

-DryDensity = -87.146+9.522A+66.785ASG · · · · · ·,, (29)

where:

Dry Density = 100 + ·% of T-99 maximum dry density, pcf (the data base of

the equation had only materials tested at 100% T-99 or

greater)

-A = sum of percent passing of sieve series: 1.5 in, 3/4, 3/8, #4, 8, 16,

30, 50, 100, 200, divided by 100

ASG = apparent specific gravity.

Thus, for a given gradation and specific gravity, the dry density can be found.

Then, in MODAMP, knowing the dry density, specific gravity, D10, and percent

minus #200 sieve material, the permeability can be estimated. Finally, by use of

I I

I I I

I I

I

134

Table 28, the Quality of Base Drainage can be found for use in determination of

the Pavement Quality of Drainage.

Subgrade Quality of Drainage

To use the Pavement Quality of Drainage table (Table 24), the Subgrade

Quality of Drainage is required in addition to the Granular Layer Quality of

Drainage. This is found by assessing the subgrade's contribution to the pavement

drainage: 1) a positive drainage of the pavement is "Good", 2) neither draining nor

supplying moisture is "Fair"; and 3) actually supplying water to the pavement is

"Poor" or "Very Poor". Use of Table 29 is recommended for determination of

Quality of Subgrade Drainage.

Pavement Structure Quality of Drainage

By knowledge of both the Granular Layer and Subgrade Quality of Drainage,

the overall Pavement Quality of Drainage can be determined from Table 24.

For instance, an open-graded base on a well-drained subgrade ought to be

rated considerably better than the Road Test, and thus is rated "Excellent." A

pavement whose base has a permeability of about 1 ft/day on a subgrade that

does not help drainage but is not a source of water should be rated about the same

as the Road Test ("Fair"). And, a base that traps water in a worse manner than

that at the Road Test should have a lesser rating ("Poor", "Very Poor") no matter

what the subgrade is doing.

More specifically, a moderately drained soil not in a wet-weather spring

situation combined with an MHTD middle gradation with fines adjusted 3.5%

(k = 13) would have a "Fair" Quality of Drainage. Using this example,

135 I Table 29. Quality of Subgrade Drainage.

Rating Soil Drainage Additional Moisture Contribution

Good • Relatively high permeability Low Moisture Contribution: (predominantly granular • deep water table soils) • absence of wet-weather

springs

• at-grade or on fill

• flooding potential: none or rare

Fair • moderate permeability (fine Moderate to none: to moderately fine soil • deep water table texture) • absence of wet-weather

• may have layer that springs impedes downward • at-grade or on fill drainage • flooding potential: none to

occasional

Poor • low permeability (fill: silty Positive moisture contribution: clays) • shallow water table

• may have layer that • absence of wet-weather impedes downward springs drainage • at-grade or in fill

• flooding potential: occasional to frequent

Very • very low permeability Positive moisture contribution: Poor (heavy clays) • shallow water table

• contains layer that impedes • in area of wet-weather drainage springs

• sidehill cut or cut section

• flooding potential: frequent or common

• marshy area

entering Table 23 finding m-coeffient on the "Fair" row and in the Condition F

column would result in an m of about 1. 1. Thus, a pavement in this situation

would be expected to perform somewhat better in regard to drainage than the

Road Test pavement.

I I I I I I I I I I I I I I I

I I

136

At this point, the user can adjust within the range given based on 1) the

goodness of match between the number of months per season at the project site

and the Climate Catagory chosen, 2) presence or absence of edge drains, 3)

presence or absence of any type of outlet for the draining water, and 4) quality of

the filter/separator. For instance, if the project site's number of wet months is

actually somewhat less than the value shown for the Climate Condition, then the

designer could increase the m-value towards its upper limit (still staying within the

cell). If a drainable base is provided, but the layer is extended under the shoulder

and daylighted at the ditch rather than incorporating edge drains, then the m-value

should be downgraded. If no outlet whatsoever is provided, then the m-value

should be less than 1.0. And, if a drained base is provided, but the filter/separator

layer gradation is not designed to be compatible with both the subgrade and

drainage layer gradations, the m-value should be downgraded. These are examples

of some considerations in the choice of m-coefficient. Edge drains must be

provided in order for a base layer to be catagorized as "Excellent."

DEVELOPMENT OF M-COEFFICIENT TABLE

In the following paragraphs, the development of the drainage coefficient

table will be explained. It will be shown that the m-coefficients are a function of

granular material resilient moduli, and these moduli were calculated from numerous

runs of KEN LA YER for different moisture and temperature conditions.

Drainage coefficients are in essence modifiers of layer coefficients that

adjust for conditions of drainage of a given project relative to the conditions of

drainage at the Road Test. Each drainage coefficient developed in the present

st udy is basically a ratio of the layer coefficient of Road Test granular material

adjusted for specific site conditions to the layer coefficient of the Road Test

material for Road Test conditions:

137

a site m = ---- .................. . (30)

aRoadTest

The determination of each layer coefficient is as outlined in the companion

study of this report (29). The layer coefficient for bases and subbases are

functions of resilient moduli (Egl, and can be calculated by use of the equations

given in the AASHTO Guide:

a2 = 0.249 log Eg - 0.977 . . . . . . . . . . . . . . (31)

a3 = 0 .227 log Eg - 0.839 . . . . . . . . . . . . . . (32)

In this portion of the study, the moduli of the Road Test base and subbase

were determined by use of the program KENLAYER. The average Road Test

pavement cross-section was used for each condition, as shown in Fig. 31. ·

The moduli of any granular base and subbase are functions of the stress

state of the material, represented by the bulk stress (8), and of material

characteristics, represented by k1 and k2 , as shown in Fig. 14. The bulk stress is

a function of the asphalt layer thickness, asphalt layer resilient modulus, granular

layer thickness, and subgrade resilient modulus.

The effect of drainage on the support capacity of a granular layer can be

determined by 1) ascertaining the effect on base modulus by changing subgrade

modulus (as a consequence of it getting wetter or staying wet longer) and 2)

I I

DI

D 2

D 3

= 4 .00

5 .39

= 8.75

138

4500 Lbf 4500 Lbf

p 70 ps1

In u = 0.40 I

In u 0 .35 2

1n u = 0.35 3

u - 0 .45 4

Fig. 31. Average AASHO Road Test Cross-Section .

139

determining the effect of the degree of saturation on the granular material

constants k1 and k2 .

Numerous runs of KENLAYER were required to give a range of modulus

values covering the spectrum of moisture and temperature conditions across the

country. The values for the KEN LAYER input are listed in Table 30. The value

reported by Traylor for the base material k1 ,wet was thought to be excessive. This

necessitated the estimation of k1 ,wet· In order to estimate a more reasonable value

for k1 ,wet' a k 1 reduction factor was required which would allow for a decrease in

k1 as the moisture content changed from a moist state to a wet state. The

reduction factor was determined by examining data from several sources (57, 59)

as well as the data generated in this study. This analysis led to the use of a value

of 20% reduction, which resu lted in a k1 ,wet = 8300 psi.

The subgrade Emin• Emax• and K1 values for various moisture states were

determined as follows. Emin and Emax refer to the minimum and maximum values

that the subgrade resilient modulus is expected to reach throughout the year. K1

is the modulus at a deviator stress of 6.2 psi, and is at the knee of the E59-ad

curve. This is shown in Fig . 32. A more thorough explanation of the derivation

and use of this type of curve is found in reference 29. Briefly, for the moist

condition, K1 moist was calculated by use of the Robnett-Thompson equation (61 ),

while Emin,moist and Emax,moist were found by knowing the slopes of the curves.

K1,thawed was estimated as 55% of K1moist · This value is recommended by

Witczak (56) for the Road Test conditions. The value of 55% was verified by the

Robnett-Thompson K 1-correction relationship for changes in moisture content.

I

I

I I

Table 30. Input Values for KENLAYER Analysis.

Parameter Source

thickness of asphalt Road Test average (54) layer

thickness of granular Road Test average (54) base

thickness of granular Road Test average (54) subbase

load, ESAL Van Til fil gf. (55), Witczak (56)

number of tires Van Tit fil gf. (55), Witczak (56)

tire spacing Van Til fil gf. (55), Witczak (56)

tire pressure Van Til fil gf. (55), Witczak (56)

Poisson's ratio, asphalt AASHTO Guide, App. 00(1)

Poisson's ratio, base AASHTO Guide, App. 00(1)

Poisson's ratio, subbase AASHTO Guide, App. 00(1)

Poisson's ratio, subgrade AASHTO Guide, App. 00(1)

Resilient Modulus, Richardson, fil gf. (30) asphalt

unit wt. of asphalt Road Test average (54)

unit wt. of base Road Test average (54)

unit wt. of subbase Road Test average (54)

unit wt. of subgrade Road Test average (54)

k, ,moist of base Traylor (57)

kl.wet of base Traylor (57)

140

Value

4.0 in

5.39 in

8.75 in

18 k

2

13.57 in

70 psi

0.40

0.35

0;35

0.45

656,800 psi

0.086 pci

0.082 pci

0.081 pci

0.063 pci

10,360 psi

8300 psi

141

k, ,thawed of base Traylor (57) 2850 psi

k, ,moiat of subbase Traylor (57) 6840 psi

k,,wet of subbase Traylor (57) 6270 psi

k, ,thawed of sub base Traylor (57) 4075 psi

k2,moist of base Traylor (57) 0.35

k2,wet of base Traylor (57) 0.34

k2,thawed of base Traylor (57) 0.62

k2,moiat of subbase Traylor (57) 0.32

k2,wet of subbase Traylor (57) 0.30

k2,thawed of subbase Traylor (57) 0.40

k0 of base and subbase Huang (58) 0.6

ko of subgrade Huang (58) 0.8

K2 of subgrade, moist Huang 6.2 psi

K3 of subgrade, moist Huang 1110 psi

K4 of subgrade, moist Huang 178 psi

Emin of subgrade, moist from Fig. 32 5389 psi

Emax of subgrade, moist from Fig. 32 13,420 psi

K, of subgrade, moist Richardson, fil al. (29) 8759 psi

Emin of subgrade, wet from Fig. 32 3515 psi

EmllX of subgrade, wet from Fig. 32 10,402 psi

K, of subgrade, wet calculated 5740 psi

Emin of subgrade, thawed from Fig. 32 1705 psi

Emax of subgrade, thawed from Fig. 32 7382 psi

K, of subgrade, thawed calculated 2720 psi

I I I

I I

I I

I

I I

.,.-,,. l'1

0 ,--

X

Cl)

a. '-"

Cl ID

w -Cl)

:J

:J -0 0 ~

-1--'

C Q.)

Cl) Q.)

n::: Q.)

-0 0 "-C)'l '

..c :J

(/)

1 8

1 6

14

12

10,402 1 0

8 7,382

6

4

2

0 0

142

13,421

- 111 0

Dry 5389

Wet 3515

Thawed 1705

5 10 15 20 25 30 35

Repeated Deviator Stress, CT d (psi)

Fig. 32. Relationship of Road Test Subgrade Resilient Modulus and Deviator Stress for Three States of Moisture Content.

143

Again, Emin,thawed and Emax,thawed were determined using geometry. Finally, K 1 wet

was calculated as midway between K1moist and K1thawed because the wet

recovery period E59-time curve is assumed to be a straight line (see Fig. 31 ). And,

Emin,wet and Emax,wet were again calculated by geometric means.

The analysis standard year was divided into 24 one-half month periods, and

the k1 (base), k 1 (subbase), and K1, Emin' Emax (subgrade) were varied for each

period, depending on prevailing environmental conditions. The environmental

conditions were varied according to the particular zone of interest. The six zones

in the U.S. are shown in Fig. 30. Table 25 is adapted from the AASHTO Guide

and tabulates the number of months a subgrade can be expected to be frozen,

thawing, wet, or "dry". These season lengths were modified depending on the

drainage conditions of the subgrade and the base course as shown below.

KENLAYER does not allow a variation in k2 (granular) values, so an average k2 was

input which was weighted for the number of months moist, wet, and thawed for

the particular case being analyzed.

As discussed earlier, soil drainage was considered to be good, fair, poor, or

very poor. These conditions are described in Table 29. Base drainability is tied to

permeability, and is described as excellent, good, fair, poor, and very poor -- these

states are shown in Table 28.

The manner in which the benefit or lack of drainage asserted itself was

taken care of by varying incrementally the season length. For a plot of E59 vs.

time, the dry period was divided into four parts and the thawed period was divided

into five parts. Although quite arbitrary, this was done as an expedient to allow

I

144

for the gradual drying of subgrades through the year coupled with the effect of the

quality of drainage. As base layer drainability decreased from excellent to very

poor (five stages), at each stage one part of the dry period was changed to a wet

period, and the thawed period was increased by one fifth. Thus base drainability

was addressed. Better soil drainage was handled by increasing the dry period and

I decreasing the thawed period one part for all base drainage cases. This is shown

in Fig. 33. There are five cases (E.g plots) for each of the good, fair, and poor soil

drainage situations, and one case for very poor soil drainage. These 16 cases

were applied to each of the six zones. Thus 96 runs would have been required of

KENLAYER. However, not that many were performed because of some replication

of case/zone situations.

I

Each run of KENLAYER utilized a unique Eg-time curve, and thus produced a

unique set of Eg,b•• and Eg,aubb•• values. Each of these was converted to an a­

coefficient, which was then divided by the a-coefficient which represented the

Road Test conditions. The resulting ratio was an m-coefficient.

The delineation of layer coefficients into the ranges shown in Table 23 was

arrived at by plotting a frequency diagram of calculated m-coefficient occurrence.

The separation points of m-values were at natural breaks in the frequency

I histogram.

I I

The analysis for the subbase m3 coefficients was identical, and the results

were similar to the m2 values. For the sake of expediency, Table 23 represents

both m2 and m3 values.

The Quality of Drainage table (Table 24) is basically the same as put forth by

"' " w

Soil - Good

Very Poor "' " w

Soil Fair 145

Base = Very Poor

"' " w

"' " w

"' " w

"' " w

Time

Time

Time

Base Excellent

w "' "

~ . w

"' ., w

"' " w

Time

Time

Time

Base Excellent

Time Time

Note: Seasonal Lengths Vary in Accordance with Climate Zone.

Fig. 33. Resilient Modulus Seasonal Variation With Variations in Base and Subgrade Drainability.

I

I

I I I I

I I I I I I I I I

w

w

'" ..

'" "

'" w

.,

'" ., w

'" .. w

Soil Poor

Bose Very Poor

'" " w

Soil = Very Poor

Very Poor

Time Ti me

Time

Time

Time

Time

Note: Seasonal Lengths Vary in Accord a nee With C Ii mate Zone. Fig. 33. Continued.

Carpenter, although it was adjusted to match the outcome of the m-coefficient

frequency plot.

Reasonableness

147

In comparison of Missouri sites to the Road Test site, in a regional sense

actual data indicates that most of Missouri is in a climatic zone that is rated as

having a greater time of saturation, so a given pavement in Missouri should fare

worse (from moisture-related problems) and therefore should have m-coefficients

less than 1.0, unless something is done to improve the pavement drainage.

Conversely, for a situation where any water that enters the base is quickly

removed laterally and where the soil drains well and does not supply water from

the surrounding soil or side hill wet weather springs and so forth, then an

expectation of a 10 to 20% improvement in pavement performance would be

reasonable. For a somewhat lesser quality of subgrade drainage with a highly

drainable base, a 10% credit may be more realistic. And, going with the belt-and­

suspenders approach, there is the option of supplying a drainable section with no

reduction in pavement thickness, a conservative approach to be sure, but not

w ithout merit for as the above discussion reveals, there is an element of conjecture

in this whole business.

II

I I

MODAMP SENSITIVITY ANALYSIS

A sensitivity analysis was performed to observe the effect of the major

drainage variables on design. This information should be helpful for future or

routine work when one is trying to decide which variable can be estimated and

148

I which must be tested. Also, at any time during an actual project, from pavement

I

I I

I

I

design and material selection to construction and inspection, it can be very useful

to have insight into which deviations from certain criteria will significantly affect

the performance of the pavement.

The analysis was divided into two parts: 1) an m-coefficient analysis, and 2)

a pavement thickness analysis.

M-Coefficient Analysis

Drainage coefficients are a function of : 1) pavement drainability, and 2)

climate. Within a given climate, pavement drainability is a function of : 1) width,

2) longitudinal grade 3) subgrade drainability, 4) base thickness, and 5) base

permeability. Cross-slope is usually not a variable that can be changed by the

highway designer. The sensitivity analysis was conducted by establishing three

values for each variable: the most likely value for each variable, a high value, and a

low value. Several runs with MODAMP were made with each successive run

changing one variable from low to median to high while holding all other variables

constant at the median value. Base drainability was the variable with the greatest

effect upon the drainage coefficient. With a Fair-draining subgrade, changing from

Very Poor to Excellent base drainage changed the m-coefficient by 0.30 to 0.45.

The permeability of the subgrade was also an important variable in the program.

149

For a very poorly draining base, the m-coefficient changed by 0.1 to 0.45 as

subgrade drainage varied from Poor to Good. Table 31 shows the results of the

sensitivity analysis. Unless noted, the standard situation was: W = 13 ft, H = 6

in, g = 0.02, subgrade drainability = Fair, base drainability = Very Poor, climatic

condition = F. Most other variables had no discernable effect upon the drainage

coefficients.

Table 31. Drainage Coefficient Sensitivity Analysis for MODAMP.

Variable Variable Magnitude/(m-coefficient Range of Change

W, ft: 13,25,37

kb = Very Poor 13/(0. 70-0.85) 25/(0. 70-0.85) 37 /(0. 70-0.85) 0.0

kb = Excellent 13/(1.15-1.20) 25/(1.15-1.20) 37/(1.15-1.20) 0.0

Hb, in: 4, 6, 12

kb = Very Poor 4/(0.70-0.85) 6/(0. 70-0.85) 12/(0.70-0.85) 0.0

kb = Excellent 4/(1.15-1.20) 6/(1.15-1.20) 12/(1.15-1.20) 0.0

g, ft/ft: 0.0, 0.03, 0.1

kb = Very Poor 0.0/(0. 70-0.85) 0.03/(0. 70-0.85) 0.1 /(0.70-0.85) 0.0

~ = Excellent 0.0/(1.15-1.20) 0.03/( 1 .1 5-1 .20) 0.1 /(1.15-1.20) 0.0

Subgrade Drainability: V. Poor, Fair, Good

kb = Very Poor V. Poor/(0.60-0.70) Fair/(0. 70-0.85) Good/(0.70-1.15) 0.1-0.45

kb = Fair V. Poor/(0.60-0.70) Fair/(1.05-1.15 Good/(1.05-1.15) 0.45

kb = Excellent V. Poor/(0.70-1.151 Fair/(1.15-1.201 Good/(1.15-1.20) 0.35-0.45

Base Drainability: V. Poor, Fair, Ex.

SG = V. Poor V. Poor/(0.60-0.70) Fair/(0.60-0. 701 Ex./(1.05-1.051 0.45

SG = Fair V. Poor/(0.70-0.85) Fair/(1.05-1.151 Ex./(1.15-1.201 0.30-0.45

SG = V. Poor MHTD Fine (Very Poor): 0.60-0.70 --SG = V. Poor MHTD Coarse (Poor): 0.70-0.85 --

SG = Poor NJ, 2% minus #200 (Good): 1.05-1.15 --SG = Poor NJ, clean (Excellent): 1.05-1.15 0.45

I I I

I

I I

I I

I I

Hb and P200 : 2 in and P200 = 0 or 2%

SG = Fair Hb = 2 in; NJ, clean (Excellent): 1.15-1 .20 0.0

SG = Fair Hb = 2 in; NJ, 2% minus #200 (Good): 1.05-1 .15 0.05-0.10

The gradation of the base material had a pronounced effect on m-coefficient.

A gradation at the fine side of the Type 1 limit with 15 % fines is rated Very Poor

and consequently has low m-values. A Type 1 on the coarse side of the limits

with 2% fines is an improvement, but still a poorly-draining material. A NJ

gradation, somewhat dirty, and a clean NJ render m-values greater than 1.0.

Table 32. Thickness Sensitivity Analysis for MODAMP.

Variable m2 m~ D2, in

Climate:

Worst (Zone F) 1.10 0.78 12.0

Best (Zone A) 1.18 1.10 10.2

(Base, Subbase, and Subgrade all = Fair)

Drainability:

Base: Very Poor

Subgrade: Very Poor 0.65 0.65 21.0

Base: Very Poor

Subgrade: Good 0.95 0.95 13.2

Base: Excellent

Subgrade: Very Poor 0 .78 0.65 17.5

Base: Excellent

Subgrade: Good 1.18 0.95 10.7

All cases o, = 4 in , E1 = 450,000 psi, a1 = 0.44, 82 = 0.10, 83 = 0.09, D3 = 4 in, SN = 3.36

Finally, the effect of accidentally pinching the base layer down to 1 in and

fouling the NJ gradation lowers the m-values somewhat.

150

151

The second portion of the analysis involved an examination of how the major

variables affected base thickness design. Using an asphalt layer thickness (D 1 ) of

4 in and E1 = 450,000 psi; a base layer thickness (D2 ) of 12 in, a subbase layer

thickness (D3 ) of 4 in, and a somewhat soft subgrade (K 1 = 6000 psi), the elastic

layer analysis program KENLAYER was used to calculate the moduli in the base

and subbase layers. These were converted to layer (a2, a3) coefficients by the use

of the AASHTO nomograph equations (see Ref. 5 for further discussion). By use

of Eq. 1, the SN was calculated, and was considered constant for the rest of the

analysis.

First, the effect of climate was considered. Using the above parameters, for

a constant pavement drainability rating, there was a difference in D2 (base

thickness) of 15 % between the best environmental zone and the worst. Second,

for constant base drainability, the required D2 decreased about 38% when the

subgrade drainability improved from Very Poor to Good. Finally, looking at the

variable over which the designer has some control, by improving base drainability,

the required base thickness decreased by about 18%, depending on subgrade

conditions.

I

I I

I I I

I

I I I I I

I I I

152

SUMMARY AND CONCLUSIONS

The purpose of this study was to determine the drainage (m) coefficients of

granular bases and subbases for use in the 1986 AASHTO Guide pavement design

method for flexible pavements. The project entailed a review and compilation of

published literature, laboratory testing, analysis of results, and preparation of this

report.

One existing method with which to determine drainage coefficients was

examined, and two potential strategies for development into a new method were

explored.

1. Seeds and Hicks developed the basic method for determination of drainage

coefficients, which is presented in the 1986 AASHTO Guide. The method

necessitates the determination of: 1) the Percent Time of Saturation of the

pavement structure, and 2) the Quality of Drainage of the base and subbase.

With these, the m-coefficients are found from a table. Unfortunately, little

direction was given in regard to the determination of the necessary input

data that is required for using the m-coefficient table.

2. Carpenter developed a method for determining m-coefficients which can be

easily implemented by use of software called DAMP. DAMP is based on the

method given in the 1986 AASHTO Guide. Carpenter's main contribution

was to provide a method with which to determine the Percent Time of

Saturation for the pavement structure, and the Quality of Drainage of the

combined base and subgrade. Required input for DAMP includes base

material permeability, or dry density and gradation information, certain

weather data, base layer thickness and density, and subgrade drainability

153

based on soil characteristics. DAMP calculates time-to-drain to 85%

saturation, and rates the Quality of Base Drainage. With this, and the

Quality of Subgrade Drainage, the overall Pavement Quality of Drainage is

found. At this point, the AASHTO Guide m-coefficient table is used for

determination of m-coefficients.

3 . A sensitivity analysis was performed using DAMP. Base drainability had the

greatest effect on the magnitude of the coefficients, with subgrade

drainability the second most important. All other variables had a minor or

negligible effect.

4. The major concerns with DAMP were:

a. DAMP's Table 2 (Quality of Drainage) does not lead to reasonable results

if one subscribes to the theory that the Road Test pavement drainage

was "Fair." The whole idea of the use of m-coefficients is to rate any

pavement's drainage relative to that at the Road Test. Better drainage

should be "Good" or "Excellent," worse drainage should be "Poor" or

"Very Poor."

b. The effect of freeze-thaw cycles and frost heave is not addressed well.

The thaw/melt period will typically result in a worse condition than

merely getting the soil wet. This effect is lost in the Time of Saturation

calculations, and the result is that AASHTO Zones I, II, and Ill are usually

lumped together, which could lead to erroneous results.

c. DAMP appears to arbitrarily increase the January storage to the

maximum for Thornthwaite Moisture Indices below a certain threshold

I I I I I I I I I I I I

I I I I

11

I

5.

154

value, regardless of previous month's history of moisture changes.

d. The extrapolation of the Thornthwaite method of regional moisture

available to the conditions in the pavement structure is of concern.

Also, the manner in which the time of saturation for various

environmental conditions is calculated is arbitrary. However, it is

recognized that at the present time there are no practical working

solutions to this dilemna, and that the time of saturation procedure in

DAMP is a significant step forward, and should be used until a more

fundamentally sound, user-friendly method can be developed.

The TTI Integrated Model of the Climatic Effects on Pavements was

evaluated with the idea that it could determine environmental effects on

granular base/subbase materials, which would lead to the calculation of m­

coefficients. The TTI model consists of four independent models linked

together: the Precipitation Model, the Infiltration/Drainage Model, the

Climatic-Materials-Structural Model, and the CRREL Frost Heave-Settlement

Model. The Integrated Model examines many variables and processes in

very fundamental terms and thus has the potential for rather detailed and

accurate analysis.

Unfortunately, the scheme was unsuccessful because of limitations in

the TTI model. First, the model requires a minimum of 100 input variables.

Many of these are not easily obtained and must be assumed. Model output

is sensitive to the magnitude of some of these input values. Secondly, the

model has a low sensitivity to variables that are thought to be important to

155

the derivation of m-coefficients. Third, the program is cumbersome. And

most importantly, the output parrots the input base course modulus. This is

a fatal flaw and rendered the program unusable for purposes of drainage

coefficient determination as envisioned in this study.

6 . The materials under study included two sources of crushed stone and two

. gravels. All materials were selected, sampled, and delivered to UMR by

MHTD personnel. The primary tests performed were: 1) resilient modulus

testing at a low and high degree of saturation to assess the moisture

sensitivity of the materials, and 2) permeability and effective porosity to

assess the drainage characteristics of the materials.

7. Two gradations of granular material were used in the resilient modulus

testing: one followed the midpoint of the MHTD Type 1 gradation (MHTD

Middle) acceptance band, and the other was the so-called New Jersey (NJ)

open-graded gradation. An additional gradation (OGS) was used in the

permeability portion of the study, along with the MHTD Middle and the NJ.

8. The aggregates were separated into the appropriate size fractions,

recombined, and tested for specific gravity, plasticity index (Pl) moisture­

density relationships, and relative density. Standard and modified proctor­

type tests (T-99, T-180) were performed for the dense gradation, while

vibratory table densification was used for the open-graded material. The

specific gravity information was required for calculation of porosity, effective

porosity, and permeability estimation. The Pl data were useful in evaluation

of permeability and effective porosity results. The finer fraction of all four

I I I I

I I I I I I I I I I I I I

9.

156

aggregate types was found to be non-plastic. The density results were

needed in order to determine the compaction target densities and moisture

contents for the resilient modulus and permeability specimens.

Nine different methods were used to characterize the gradation curve shapes

and positions. These methods were fineness modulus, coefficient of

-. uniformity, coefficient of skew, Hudson's A, surface fineness (SF), specific

surface factor (SSF), SF/SSF, slopes-of-gradation curve, and percent passing

or retained on individual sieves. None of the single parameters were superior

in the prediction of k1 or E9 . The percent passing the 3/8 in, #4, #8, ano

#200 sieves proved to be useful in prediction of permeability.

10. Particle shape/surface texture tests were performed on the four aggregates.

The ( +) #4 sieve material was tested in accordance with ASTM D3398,

while the (-) #8 to ( +) #100 fraction was tested using the NAA method.

The measured angularities of the two stones were about the same, and were

more angular than the two gravels, which were about equal. The difference

in angularity/texture was not great between the crushed stones and the

gravels.

11 . Resilient modulus test results were required for use in the TTI method and in

the new method developed in this study. The tests were run on all four

aggregates using two gradations, two compactive efforts, and two degrees

of saturation, with replications. Fourteen combinations of confining pressure

and cyclic applied deviator stress were used for each specimen in the test

sequence. Effective confining pressures ranged from 2 to 20 psi and cyclic

157

deviator stress ranged from 2 to 40 psi. Thirty-two specimens were

fabricated. The total number of tests run was 896.

The results of the testing indicated the Eg increases with a lower degree

of saturation. The average percent loss in k1 due to increased saturation

was 31 %. This information was used in the development of the m­

coefficients.

12. An analysis of the interaction of gradation and degree of saturation indicated

that dense-graded material suffered less loss of modulus than open-graded

material. However, in practice the open-graded material would not be in a

saturated state nearly as much (if ever) as a dense-graded state. The data

showed that the drained open-graded moduli were greater than the

undrained dense-graded moduli.

13. A statistical analysis was performed to determine the significance of the

variables. Paired-t tests indicated that change in degree of saturation gave

significantly different results at the 0.05 level. In comparing saturated

dense-graded material with drained open-graded material, there was a

significant (0.088 level) increase in Eg with superior drainage.

An increase in degree of saturation acts to lower k1 and raise k2 of the

granular material, and to lower subgrade support, all of which act to lower

the Eg of the granular material.

14. Permeability of base material is required input for DAMP, TTI Integrated

model, and MODAMP. It is necessary in order to calculate the time-to-drain

for base layers. The rigid wall constant head test procedure was used for

I I I I I I I I I I I I I I I

I I I

158

the NJ and OGS open-graded materials, while the dense-graded MHTD

Middle specimens were tested in a triaxial compression chamber following

the resil ient modulus testing.

15. Measured permeability values have been shown to be sensitive to: 1) air

blockage, 2) specimen sergregation, 3) relative density, 4) gradient, and 5)

short-circuiting. Considerable effort was exerted to eliminate or at least

minimize the above problems, either through the development of the test

procedure or the design of the equipment.

16. The open-graded specimens were tested at a minimum of five gradients with

five repetitions. The permeability coefficients were calculated as the

average of the values which were obtained at or less than the expected field

gradient (approximately 0.1 ). The testing variables included four aggreate

sources and two gradations (NJ and OGS). Duplicate specimens were

tested, for a total of 400 tests; this represented data from 16 total

specimens.

17. The dense-graded specimens were tested at the recommended gradient of

five with five repetitions . Two compactive efforts were used on the four

aggregate sources. Duplicate specimens were used. This resulted in a total

of 80 tests, representing data from 16 specimens.

18. The testing program results for the open-graded materials revealed the

following :

a. Permeabilities estimated from the Moulton equation significantly

underestimated the observed values by an order of magnitude or more in

159

most cases. A review of the data on which the Moulton equation is

based reveals potential problems with air blockage, effect of end

conditions, and possibly incorrect use of specific gravity data, all of

which would lead to falsely low values.

b. The OGS gradation appeared to be more permeable than the NJ, but was

not so at the 0.05 significance level.

c . The #16 sieve size was important to the predicition of permeability of

the open-graded materials.

d. The DR-15 gravel was more permeable in a statistical sense than both of

the crushed stones, but was also more permeable than the DR-14 gravel.

Because their particle shapes were about the same, a strong statement

cannot be made as to the effect of particle shape.

e. A regression equation was developed for the open-graded materials . It

had an adjusted-R2 = 0. 779:

k = -40,962 + 15,284 (/Jett) + 205.89 (P16) + 380. 70 (PDENS)

where:

/Jett = effective porosity

P16 = percent passing #16 sieve

PDENS = percent of maximum achievable density.

f. The average effective porosity was about 68% of the average porosity.

19. The results of the dense-graded permeability testing were:

a. Again , permeabilities estimated with the Moulton equation were

significantly lower than observed values.

I I I I I I I I I I I I I I I I I I I

160

b. Gravels exhibited slightly greater permeabilities than crushed stones, but

not statistically so at the. 0.05 level.

c. A regression equation was fit to the data which resulted in an adjusted­

R2 = 0.624:

k = -8.457 + 35.753 (17) + 0.033 (satfinal)

where:

17 = porosity (calculated with apparent specific gravities)

satfinal = final degree of saturation, %.

d. On the average, the effective porosities were about 27% of the total

porosities. This is considerably smaller than the open-graded value,

which is to be expected because of the finer pore sizes in the dense­

graded material.

e. Overall, the permeabilities of the dense-graded materials were

significantly lower by several orders of magnitude than those of the

open-graded materials (average of 0.8 ~ 1014 ft/day).

20. A regression equation to estimate permeability was developed by combining

the results from several studies found in the literature with the results of this

study. The equation had an adjusted-R2 = 0.900:

k = (2.28 x 1017)(l7)8.995 (D,o)0.591 /[(P3/8)5.511 (P200)0.349]

where:

fJ = porosity (calculated with apparent specific gravity data)

D10 = size that represents 10% passing, mm

P3/8 = percent passing 3/8 in sieve

161

P200 = percent passing #200 sieve.

The equation is considered accurate in the range of 0.1 to 1000 ft/day.

21. Although the Moulton equation significantly underpredicts permeability, it

may be the equation of choice because field conditions may render the base

layer less permeable than what would be predicted with good quality

laboratory testing.

22. A new method of calculation of drainage coefficients was developed. In

essence, m-coefficients were calculated as a ratio of the layer coefficient of

Road Test granular base material under a given drainage and climate

condition to the layer coefficient under Road Test site conditions. The layer

coefficients were calculated from resilient moduli. The resilient moduli were

calculated with KENLAYER under varying conditions. By changing subgrade

and base moisture conditions for a given time of year, the moduli were

varied. The base material moisture sensitivity (effect on k1 and k2) was

determined in part by the resilient modulus laboratory testing of granular

materials.

23. The result of the above analysis was the creation of a Quality of Base

Drainability table (based on time-to-drain to 85% saturation), a Quality of

Subgrade Drainability table (based on subgrade permeability, position of

water table, flooding potential, presence of impermeable layers, potential for

water seepage, and so forth), a Quality of Pavement Drainage table(based

on the previous two tables), a Climate Condition table (based on estimated

season lengths), and finally, an m-coefficient table (based on Quality of

I

I

I

I

I

I I I I I

I I I I I I I I I I I I I

24.

25.

162

Pavement drainage and Climate Condition).

A regression equation was developed to assist in the estimation of

compacted dry density in order to estimate permeability with the Moulton

equation and the UMR equation. The equation had an adjusted-R2 = 0. 729

and a S.E.E. = 5.22. The equation is:

yd = - 87.146 + 9.522 A + 66.785 ASG

where: yd

A

= 100+ % of T-99 maximum dry density, pcf

= sum of percent passing of sieve series from 1. 5 in through

#200, divided by 100

ASG = apparent specific gravity.

A sensitivity analysis was performed. The most important variables in

regard to m-coefficient calculation were climate condition, base drainability,

and subgrade drainability. These, in turn, affected base thickness

calculation significantly.

I

I I I I I I I I I

I

I

I I I I

I I I I I I I I I I I I I

1 .

2 .

3.

163

RECOMMENDATIONS

A choice is presented as to the procedure for estimation of permeability.

Moulton's equation underpredicts permeability significantly. The UMR

equation is more accurate. However, due to uncertainties of the impact of

field construction practices, permeabilities estimated by the Moulton method

would err on the conservative side. It is recommended that field

permeabilities should be run to determine which equation should be used .

Until this is done, the Moulton equation is recommended.

In regard to determination of drainage coefficients, it is recommended that

for a given project site, the time-of-saturation should be calculated by

MODAMP to assist in choosing the proper column to use in the m-table.

Local experience should be used to estimate the season length of the project

site. The Climate Condition can then be estimated from season lengths.

Knowing site conditions, Table 28 can be used to assess the Quality of

Subgrade Drainage . Next, using MODAMP, the base time-to-drain can be

calculated. Using the proper table, MODAMP renders the Quality of Base

Drainage, the Quality of Pavement Drainage, and initial value for the m­

oefficient. The designer should then adjust the m-value to account for

specific site conditions.

It is recommended that the drainage coefficients developed in this study

should be evaluated for reasonableness by MHTD personnel experienced

with pavement drainability and performance.

I I

I

I I I

I

I

I

I

I I I

I I I

I I I I I

1.

2 .

164

FUTURE RESEARCH NEEDS

One of the major weaknesses of DAMP and MODAMP is the determination

of the time of pavement saturation. As it is now, this value is tied to

monthly precipitation data. Even the TTI method is limited to daily data.

Since drainage times are discussed in terms of hours, a method needs to be

. developed, if possible, that is tied to hourly precipitation.

In order to decide which permeability predictive equation to use (Moulton ~.

UMR), field permeability tests should be performed and correlated with

laboratory test data.

I

I I

I I I I I

I I I

I I I I

I I I I

165

ACKNOWLEDGEMENT

The authors wish to thank the MHTD for its sponsorship and support of this

research project. They also thank the UMR Department of Civil Engineering for its

support. Special thanks go to Mr. Kevin Hubbard and Mr. Aswath V. Rao for their

assistance in the figure preparation portion of the study.

------------------------------------ - --

I

I I

I I

I

I

I I I I

1.

2.

REFERENCES

AASHTO 1986 Guide for Design of Pavement Structures, AASHTO,

Washington, D.C., 1986.

166

Seeds, S.B. and R.G. Hicks, "Development of Drainage Coefficients for the

1986 AASHTO Guide for Design of Pavement Structures," TRB 68th Annual

Meeting, Washington, D.C., 36 p.

3. Mathis, D.M., "Permeable Base Design and Construction," Stone Review,

Vol. 5, No. 4, 1989, pp. 12-14.

4.

5.

6.

7.

8.

9.

Mann, W.D., "Behind the Shift to Permeable Bases," Hwy. and Heavy

Construction, Vol. , No. , 1990, pp. 38-41.

Richardson, D.N. and P.A. Kremer, "Determination of AASHTO Layer

Coefficients, Vol. II: Unbound Granular Bases and Cement Treated Bases,"

MCHRP Report, Study 90-5, University of Missouri-Rolla, Rolla, Missouri,

1994, 169 p.

Carpenter, S.H., "Selecting AASHTO Drainage Coefficients," TRB 68th

Annual Meeting, Washington, D.C., 1989, 28p.

Lytton, R.L., D.E. Pufahl, C.H. Michalak, H.S. Liang and B. Dempsey, An

Integrated Model of the Climatic Effects on Pavements, Rpt. No. FHWA-RD-

90-033, Fed . Hwy. Admin., McLean, Virginia, 1989, 289 p.

Carpenter, S.H., M.I. Darter, B.J. Dempsey, and S. Herrin, A Pavement

Moisture Accelerated Distress (MAD) Identification System - Volume 1. Rpt.

No. FHWA-RD-81-079, Fed. Hwy. Admin., McLean, Virginia, 1981, 136 p.

Moulton, L.K., Highway Subdrainage Design Manual, Rpt. No. FHWA-TS-80-

167

224, Fed. Hwy. Admin., Mclean, Virginia, 1980, 162 p.

10. Kopperman, S., G. Tiller, and M. Tseng, "ELSYM5, Interactive

Microcomputer Version," FHWA Rot. No. FHWA-TS-8-206, 1986, 33 p.

11. Carpenter, S.H., Highway Subdrainage Design by Microcomputer: DAMP,

Rpt. No. FHWA-IP-90-012, Fed. Hwy. Admin., McLean, Virginia, 1990, 118

p.

12. Lane, K.S., and D.E. Washburn, "Capillarity Tests by Capillarimeter and by

Soil Filled Tubes,"~. Hwy. Res. Bd., 1946, pp. 460-473.

13. Barber, E.S. and C.L. Sawyer, "Highway Subdrainage," fm.c..., Hwy. Res.

Bd., 1952, pp. 643-666.

14. Yemington, E.G., "A Low-Head Permeameter for Testing Granular Materials,"

Permeability of Soils, ASTM Spec. Tech. Pub. No. 163, American Society

for Testing Materials, Philadelphia, Penn., 1955, pp. 37-42.

15. Chu T.Y., D.T. Davidson,. and A.E. Wickstrom, "Permeability Tests for

Sands," Permeability of Soils, ASTM Spec. Tech. Pub. No. 163, American

Society for Testing Materials, Philadelphia, Penn., 1955, pp. 43-55.

16. Smith, T.W., H.R. Cedergren, and C.A. Reyner, "Permeable Materials for

Highway Drainage," Hwy. Res. Rec. No. 68, Hwy. Res. Bd., Washington,

D.C., 1964, pp. 1-16.

17. Strohm, W .E., E.H. Nettles and C.C. Calhoun, "Study of Drainage

Characteristics of Base Course Materials," Hwy. Res. Rec. No. 203. Hwy.

Res. Bd., Washington, D.C., 1967, pp. 8-28

18. Casagrande, A., and W.L. Shannon, "Base Course Drainage for Airport

I 168

Pavements," Proc. of the American Society of Civil Engineers, Vol. 77, June

1951, pp. 792-820.

19. Thornthwaite, C.W. "An Approach Toward a Rational Classification of

Climate," Geographical Review, Vol. 38, 1, 1948, pp. 55-74.

20. Soil Survey Manual, Handbook No. 18, Soil Conservation Service, U.S. Dept.

Ag., 1951, -p.

21. Hole, F.D., "Suggested Terminology for Describing Soils as Three

I Dimensional Bodies," Soil Science Proceedings, Vol. 17(2), 1953, pp. 131-

135.

22. Liang H.S. and R.L. Lytton, "Rainfall Estimation for Pavement Analysis and

I Design," TRB 68th Annual Meeting, Washington, D.C., 19 p.

I

23. Lytton, R.L. and S.J. Liu, Environmental Effects on Pavements - Drainage,

Rpt. No. FHWA/RD-84/116, Fed. Hwy. Admin., McLean, Virginia, 1983,

174 p.

24. Liu, S.J., J.K. Jeyapalan, and R.L. Lytton, "Characteristics of Base and

Subgrade Drainage of Pavements," Trans. Res. Rec. 945, Trans. Res. Bd.,

19-, pp. 1-9.

25. Dempsey, B.J., W.A. Herlache, and A.J. Patel, Environmental Effects on

Pavements - Theory Manual, Rpt. No. FHWA/RD-84/115, Fed. Hwy. Admin.,

McLean, Virginia, 1983, 134 p.

26. Berg, R.L., G.L. Guymon, and T.C. Johnson, Mathematical Model of Frost

Heave and Thaw Settlement in Pavements, Rpt. of the U.S. Army Material

Command, Cold Regions Research and Engineering Laboratory, Hanover,

169

New Hampshire, 1986, 49 p.

27. Richardson, D.N., J.K. Lambert, and P.A. Kremer, "Determination of

AASHTO Layer Coefficients, Vol I: Bituminous Materials," MCHRP Final

Rpt., Study 90-5, Univ. of Missouri-Rolla, Rolla, Missouri, 1994, 237 p.

28. Kandhal, P.S. J.B. Motter, and M.A. Khatri, "Evaluation of Particle Shape

and Texture: Manufactured ~. Natural Sands, "NCAT Rpt. No. 91-3,

NCAT, Auburn, Alabama, 1991, 23 p.

29. "Standard Test Method for Particle Shape, Texture, and Uncompacted Void

Content of Fine Aggregate," Draft, National Aggregate Assn., Silver Spring,

Maryland, 1991, 12 p.

30. "Test Method for Index of Aggregate Particle Shape and Texture," ASTM

03398-87, Annual Book of ASTM Standards, Vol. 05.03 ASTM,

Philadelphia, Penn., 1992, pp. 393-396.

31. "Standard Method of Test for Specific Gravity and Absorption of Coarse

Aggregate, T85-88," Standard Specifications for Transportation Materials

and Methods of Sampling and Testing, 15th Ed., Part II, Tests, AASHTO,

Washington, D.C., 1990, pp. 183-186.

32. "Standard Method of Test for Specific Gravity and Absorption of Coarse

Aggregate, T84-88," Standard Specifications for Transportation Materials

and Methods of Sampling and Testing, 15th Ed., Part 11, Tests, AASHTO,

Washington, D.C., 1990, pp. 179-182.

33. "Interim Method of Test for Resilient Modulus of Subgrade Soils and

Untreated Base/Subbase Materials," AASHTO T-XXXC-91, AASHTO, I

I

I

I

I

34.

170

Washington, D.C., 1991, p. 1-35.

"Standard Method of Test for the Moisture-Density Relations of Soils Using

a 5.5 lb Rammer and 12 in Drop, T99-90," Standard Specifications for

Transportation Materials and Methods of Sampling and Testing, 15th Ed.,

Part II, Tests, AASHTO, Washington, D.C., 1990, pp. 226-230.

35. "Standard Method of Test for Moisture-Density Relations of Soils Using a

1 O lb Rammer and an 18 in Drop, T 180-90," Strandard Specifications for

Transportation Materials and Methods of Sampling and Testing, 15th Ed.,

Part II, Tests, AASHTO, Washington, D.C., 1990, pp. 455-459.

36. "Standard Test Methods for Maximum Index Density of Soils Using a

Vibratory Table, 04253," Annual Book of ASTM Standards, Vol. 04.08,

ASTM, Philadelphia, Penn, 1990, pp. 572-583.

37. "Resilient Modulus of Unbound Granular Base/Subbase Materials and

Subgrade Soils," SHRP Protocol P46, S.H.R.P., Washington, D.C., 1992, 43

p.

38. Claros, G., W.R. Hudson, and K.H. Stokoe, II, "Modifications to the Resilient

Modulus Testing Procedure and the Use of Synthetic Samples for Equipment

Calibration," TRB 69th Annual Meeting, Trans. Res. Bd., Washington, D.C.,

1990, 27 p.

39. Rada, G. and M.W. Witczak, "Material Layer Coefficients of Unbound

Granular Materials from Resilient Modulus," Trans. Res. Rec. 852. 1982. pp.

15-21.

40. Rada. G. and M.W. Witczak. "Comprehensive Evaluation of Laboratory

171

Resilient Moduli Results for Granular Material," Trans. Res . Rec. 81 O. Trans.

Res. Bd., 1981, pp. 23-33.

41. Highlands, K.L. and G.L. Hoffman, "Subbase Permeability and Pavement

Performance," Trans. Res. Rec. 1159. Trans. Res. Bd., 1988, pp. 7-20.

42. "Standard Test Method for Measurement of Hydraulic Conductivity of

Saturated Porous Materials Using a Flexible Wall Permeameter," ASTM

05084-90 Annual Book of ASTM Standards, Vol. 04.08, ASTM,

Philadelphia, Penn, 1990, pp. 1070-1077.

43. "Standard Method of Test for Permeability of Granular Soils (Constant

Head),_ T 215-90," Standard Specifications for Transportation Materials and

Methods of Sampling and Testing, 15th Ed .• Part II, Tests, AASHTO,

Washington, D.C., 1990, pp. 554-558.

44. Moynahan, Jr., T.J. and Y. Sternberg, "Effects on Highway Subdrainage of

Gradation and Direction of Flow within a Densely Graded Base Course

Material," pp. 50-59.

45. Sherard, J.L., L.P. Dunningan, and J.R. Talbot, "Basic Properties of Standard

Gravel Filters," ASCE Geotech. J., Vol. 110, No. 6, 1984, pp. 684-700.

46. Jones, C.W., "The Permeability and Settlement of Laboratory Specimens of

Sand and Sand-Gravel Mixtures," Symp. on Permeability of Soils, ASTM

Special Tech. Pub. No. 163, 1955, pp. 68-79.

47. Karlsdotter, J., "Optimization of Coarse Particles in a Filter Soil," .M....S....

Thesis, University of Missouri-Rolla, Rolla, Missouri, 1991, 96 p.

48. Allen, W .L., Subsurface Drainage of Pavement Structures: Current Corps of

I

I

172

Engineers and Industry Practice, DOT/FAA/RD-91 /24, US DOT, FAA,

Springfield, Virginia, 1991, 31 p.

49. OUATTROPRO, Borland International, Inc.

50. Haynes, J.H. and E.J. Yoder, "Effects of Repeated Loading on Gravel and

Crushed Stone Base Course Materials Used in the AASHO Road Test," ~

Res. Rec. 39, Hwy. Res. Bd., Washington, D.C., pp. 82-96.

51. "The AASHO Road Test, Rpt. 2-Materials and Construction" Hwy. Res. Bd.

Spec. Rot. 61 B, Hwy. Res. Bd., 1962, 173 p.

52. Van Til, C.J., B.F. McCullough, B.A. Vallerga, and R.G. Hicks, "Evaluation of

AASHO Interim Guides for Design of Pavement Structures, "NCHRP Rot.

~ Hwy. Res. Bd., Washington, DC, 1972, 111 p.

53. Witczak, M.W., Development of Regression Model for Asphalt Concrete

Modulus for Use in MS-1 Study, Asphalt Institute, 1978, 39 p.

54. Traylor, M.L., "Characterization of Flexible Pavements by Non-Destructive

Testing," PhD Dissertation, Univ. of Illinois-Urbana-Champaign, Illinois,

1978, 213 p.

55. Huang, Y.H., Pavement Analysis and Design, Prentice Hall, Englewood

Cliffs, NJ, 1993, 805 p.

56. Hicks, R.G., and F.N. Finn, "Prediction of Pavement Performance from

Calculated Stresses and Strains at the San Diego Test Road," Proc. of Assn.

of Asphalt Paving Tech., Vol. 43, 1974, pp. 1-40.

57. Woolstrum, G., "Dynamic Testing of Nebraska Soils and Aggregates," TRB

69th Annual Meeting, Washington, D.C., 1990, 24 p.

173

58. Thompson, M.R. and O.L. Robnett, "Resilient Properties of Subgrade Soils,"

ASCE Trans. Journal, TE1, 1979, pp. 71-89.

I

I

I

I

I

I I

174

APPENDICES

I 175

APPENDIX A

REQUIRED INPUT FOR TTI MODEL

I

I

I

I

I

I

I I

Required Input for TTI Model.

climatic region for the analysis

city or weather station used in the analysis

maximum allowable convection coefficient

total number of layers in the pavement system

total number of finite elements in the pavement system

number of asphalt layers in the pavement system

coefficient of variation for unsaturated permeability

nodes for which the output is to be printed

name or the AASHTO soil classification of each layer in the pavement system

thickness of each layer

spacing between two nodes

last node in each pavement layer

176

thermal conductivity of the asphalt or stabilized layer under unfrozen conditions

heat capacity of the unfrozen asphalt or stabilized layer

total unit weight of the unfrozen asphalt or stabilized layer

gravimetric water content of the asphalt or stabilized layer

thermal conductivity of the asphalt or stabilized layer under freezing conditions

heat capacity of the freezing asphalt or stabilized layer

thermal conductivity of the asphalt or stabilized layer under frozen conditions

heat capacity of the frozen asphalt or stabilized layer

air content of each asphalt layer

coarse aggregate content of each asphalt layer

number of points in the temperature-stiffness relationship for each asphalt layer

temperature values for an asphalt layer in the bitumen temperature-stiffness relationship

porosity of a soil layer

dry unit weight of a soil layer

thermal conductivity of the dry soil particles of a soil layer

specific heat of a dry soil layer

frozen resilient modulus of a soil layer

unfrozen resilient modulus of a soil layer

Poisson's ratio of a frozen soil layer

Poisson's ratio of an unfrozen soil layer

coefficient of volume compressibility of a soil layer

saturated permeability of a soil layer

177

multiplier of pore pressure for Gardner's unsaturated permeability function for a soil layer

exponent of pore pressure for Gardener's unsaturated permeability function for a soil layer

multiplier of pore pressure for Gardner's moisture content function for a soil layer

exponent of pore pressure for Gardner's moisture content function for a soil layer

length of the recovery period in days

factor of resilient modulus reduction due to thawing

emissivity factor

surface short wave absorptivity

constant deep ground temperature

modifier of overburden pressure during thaw

Geiger long wave back radiation Equation Factor A

Geiger long wave back radiation Equation Factor 8

vapor pressure near the ground surface

cloud base factor for back radiation

time of day at which the minimum air temperature occurs

time of day at which the maximum air temperature occurs

upper temperature limit of the freezing range

lower temperature limit of the freezing range

number of times each day that the temperature profile is recorded an indicator for which temperature profile to print

times at which the temperature profile is recorded

I

I I

I I

year in which the first consideration period is started

starting date of a given consideration period

number of days in the consideration period

initial soil temperature profile

initial soil pore pressure profile

coefficient of pore pressure transfer

number of the day for the minimum and maximum temperature

daily minimum temperature

daily maximum temperature

average monthly wind speed

average monthly sunshine percentage

standard deviation of the monthly sunshine percentage

time at which the sun rises each day on a 24 hour clock

time at which the sun sets each day on a 24 hour clock

daily extraterrestrial radiation

the number of lower boundary points for pore pressure; pore pressure at the low boundary

time to begin with the corresponding low boundary pore pressure value

one side width of base

slope ratio or value of tangent alpha of the base

indicator for lower boundary condition

indicator for type of fines added

indicator for amount of fines added

percentage of gravel in the sample

percentage of sand in the sample

pavement type

linear length of cracks and joints of one side of the pavement

total length surveyed for cracks and joints

recurrence period for rainfall

constant K for the intensity-duration-recurrence equation

178

power for the recurrence interval term

power for the rainfall duration term

constant due to the curve shape of the rainfall intensity vs. rainfall period

number of years of rainfall data

monthly rainfall amount

number of wet days in the month

number of thunderstorms in the month

average monthly air temperature

confidence interval to use for the rainfall calculations

179

I

I

I I I

180

APPENDIX B

USE OF DAMP

I

I

I I I I

GENERAL

181

USE OF DAMP MANUAL

The DAMP Manual is self explanatory and can be relied upon by the first

time user of the program. The DAMP version tested at UMR was not able to print

the output reports mentioned in the manual. This may have been due to some

hardware or software problem with our systems that we were not able to detect.

In any case, output was produced by using the printscreen feature of DOS. If you

have not yet installed DAMP on your machine, do so at this time following the

procedures in the DAMP Manual. It must be remembered that the m-coefficients

obtained from DAMP are not in agreement with the recommendations of this

report.

USE OF DAMP DATA FILES

The DAMP data notebook (Appendix D) and accompanying floppy disks

contain the weather data necessary to use the DAMP program for all areas of

Missouri. The user need only copy the weather data files to the DAMP directory.

These files are in the format, "TOWN NAME" .dat. Once the data files are on the

DAMP directory, they can be read while in the DAMP program from the "Main

Selection Menu" by selecting "#4. Read Input Data from Disk File". This loads the

data into the DAMP program for the user to edit as desired. From this point on the

DAMP program can be used as described in the DAMP Manual.

The data files were created at UMR using historical climatological data

published by the National Climatic Data Center.

182

INTERACTIVE DAMP INPUT SELECTION

The DAMP Manual and program are well documented. The user will be able

to proceed through the program provided the subgrade soil type is known or can

be estimated, the gradation of the base/subbase is known, the geometry of the

sections is known.

SUMMARY REQUIREMENTS FOR DAMP INPUT

This section briefly summarizes the data that should be assembled by the

designer before starting either program. The requirements are very simple and fall

into these general categories: pavement geometry, layer thicknesses, layer

densities, base course gradation/specific gravity, and subgrade

drainability/permeability.

Pavement Geometry

This includes the pavement grade (gGrade), the cross slope (Sc), the width

of the drainable base (W), and the number of longitudinal joints in the pavement

surface that water flows over (Ne). Include the joint at the crown if present. The

length of the transverse cracks or joints in the pavement is (We). In the length

measurement, include the portion of the cracks or joints that extend through the

shoulders. The average spacing of these transverse cracks is (Cs).

A typical two lane highway with 12 foot wide lanes that had transverse

cracks every 25 ft would have the following numerical values for input: W = 12

ft, Ne = 3 (one at the crown and one at each shoulder), W c = 12 ft (assume no

cracking in the shoulder), c. = 25 ft. Only W is used for determination of

drainage coefficient. The other parameters are for calculation of crack infiltration,

if so desired.

Layer Thicknesses

183

This is the thickness of the combined asphalt pavement layers above the

base course, the thickness of the base course, and the thickness of the subbase.

If there is no subbase, enter zero. Only the thickness of the base is used for

calculation of m-coefficients.

Layer Densities

This is the expected in-place density of the combined asphalt pavement

layers, the basecourse layer, and the subbase layer. Typical values might be 150

(pcf) for asphalt pavement, 138 (pcf) for base course, and 138 (pcf) for subbase.

I If there is no subbase layer, enter zero. Only the density of the base is used for

calculation of m-coefficient.

I

Base Course Gradation and Specific Gravity

The apparent specific gravity (Gs) of the base course aggregate is needed.

The gradation of the base course aggregate needs to be plotted so the effective

size (010) in millimeters can be delineated. This is the grain size that 10 percent of

the aggregate passes. Also the P200 is required. This is the percent of the base

course aggregate that passes the #200 sieve. Typical numbers for an open

gradation might be 0 10 = 2mm and P200 = 2 percent.

Subgrade Drainability/Permeability

Knowledge of the subgrade soil type will allow approximation of the

subgrade permeability (km) with sufficient accuracy to use these programs. The

soil drainability from the general Missouri Soil Survey is used to enter the SG

184

Drainage input into the programs.

The frost heave rate of the subgrade (h2) in mm/day is a required input. An

acceptable number can be determined from the FHWA Highway Subdrainage

Design Manual (9), Table 4, page 72 or the DAMP Manual, Table 1, page 64 ( 11).

The necessary information to complete this input to DAMP can be found in

the County Soil Survey or in the Missouri General Soil Survey's soil drainage

classification. DAMP asks for either "Good", "Fair", or "Poor" as an input. The

following table summarizes the drainage classifications and the proper relation to

DAMP input.

Table B1. Subgrade Drainability Input Information.

Damp Input Soil Survey Drainability Natural Drainage Index

1. Good Excessively Drained -10 to -2 Somewhat Excessively Drained

2. Fair Well Drained -2 to 2.5 Moderately Well Drained

3. Poor Somewhat Poorly Drained > 2.5 Poorly Drained Very Poorly Drained

This Good, Fair, or Poor rating is all that is used in the m-coefficient determination.

Other Input Variables

The permeability of the surface of the pavement layer (kp) can be input by

the user. It is recommended, however, that the value of zero be used unless

unusual circumstances are present.

The crack infiltration rate is a user input, but the value of 2.4 cfd/linear foot

is the number provided as a default. The originator of the crack infiltration method

I I

I

I

185

determined 2.4 cfd/lft experimentally by testing a small number of pavements.

These variables are not used for calculation of m-coefficients.

DRAINAGE COEFFICIENT OUTPUT SCREEN

In execution of DAMP the screen that comes up immediately after the

subgrade drainability input discussed above is the final summary of the program's

efforts. If a hard copy of the results of the analysis is desired, it is best to

'printscreen' before entering any other keystrokes.

I

I I

I I

I I

I I

I I I

186

APPENDIX C

MODAMP MANUAL

I I I

I I I I

I I

I

I I I I

187

GENERAL

The UMR version of DAMP has been developed into a spreadsheet format.

The calculation procedure for permeability, time-to-drain, and time-of-saturation is

identical to the rules stated in the DAMP documentation. However, for drier

climates, DAMP increases storage for the first several months of the year.

LOADING MODAMP AND CLIMATOLOGICAL DATA SPREADSHEETS

Software and Hardware

The spreadsheet has been created using Quattro Pro from Borland

International (49)and was rewritten to operate in Lotus 123. It has successfully

operated in Quattro Pro versions 3 through 5.0. The spreadsheet has been

operated on IBM PC XT, 386 and 486 machines.

Loading

Running Lotus 123 or Quattro Pro from the c: drive is necessary in order for

the macros in the spreadsheets to operate properly. The enclosed floppy disks

have the MODAMP spreadsheet in file MODAMP.WK1. The Quattro Pro version

uses MODAMP.WQ1. The enclosed DAMP data notebook (Appendix D) and floppy

disks have the weather data necessary to use the MODAMP spreadsheet for all

areas of Missouri. The user need only copy the weather data files to the c:\ 123\

directory. These files are in the format, "TOWNNAME".WK1. The MODAMP.WK1

file needs to be copied to the same drive. Quattro Pro versions of data files are

also available, with .WQ1 extensions.

The user needs to get to the c:\ 123\ > prompt. At this prompt Lotus 123

can be loaded and MODAMP can be retrieved by typing 'modamp' followed by

enter. If the user is already in 123, the same result can be obtained by closing all

188

open spreadsheets and retrieving the MODAMP.WK1 spreadsheet. Once the user

is in the MODAMP spreadsheet, the climatic conditions for the location of interest

can be automatically entered in the proper place in MODAMP by use of a macro.

This macro asks the user what weather file to use. Upon receiving the requested

input, the macro opens the specified file, loads the required weather data into

MODAMP and closes the weather file. The specific keystrokes for the macro are

as follows: "Alt-w" followed by "enter."

The user will be prompted for the name of the weather file to use. Refer to

the first few pages of the DAMP Data Notebook for the proper filename. In

general, the file names are simply the city name or the first eight letters of the city

name where the weather station is located. At the prompt, type in the filename

followed by "enter." The word "macro" will appear in the bottom right corner of

the screen. Wait until the word "macro" disappears from the screen. The end of

the macro will be signalled by a beep. If a file name is entered that is not in the

directory, an error message will be displayed. Simply hit the escape key and rerun

the macro entering the proper filename when prompted.

MODAMP DISPLAY

The display that appears should look like the sample shown in Fig. C1 if the

Quattro Pro version is used. The Lotus version will not show the desktop settings.

The name of the data file that you have imported will appear in the cell next

to the cell labeled "Location."

Permeability of the Drainage Layer

The fourth row in the table solves Moulton's relationship for permeability

(kd) of the drainage layer (base course) material (Eq. 2). The dry density of the

I 189

drainage layer material (dd), apparent specific gravity (G.L drainage layer material

I effective size (D10), and percent drainage layer material passing the #200 sieve

(P200 ) are material properties the user needs to input on line five. Eq. 29 can be

used to estimate "dd" from gradation data. The gradation of the basecourse

I aggregate needs to be plotted so the effective size (D10) in millimeters can be

picked off. This is the grain size that 10 percent of the aggregate passes. Also

the P 200 is required. This is the percent of the base course aggregate that passes

I the #200 sieve. Typical numbers for an open gradation might be D10 = 2mm and

P200 = 2 percent. A nomograph of Moulton's relationship can be found in the

FHWA Highway Subdrainage Design Manual (9), Fig. 28, page 51. Permeability

I (kd), porosity (n) and estimated effective porosity (n.) are calculated and displayed.

The kd and n. values are used in the time-to-drain calculations.

Slope Factor

I The calculations for the slope (slope) and length (L) of the drainage path are

I I I I

I I

based upon the user's input of longitudinal grade (gGrade), cross slope (Sc) and

width of drainage layer drainin_g to the collector (w). Both gGrade and Sc are

expressed in foot/foot or inch/inch, not percent. These inputs are used to calculate

a time-to-drain using the Casagrande-Shannon method as described in this paper.

Thickness of Base

The thickness of the layer in question is input assist in the time-to-drain

calculations.

Subgrade Drainability

This user input is entered in the block just below the label "SG Drainage."

MODAMP 16-fet>-915 190

" . 120 26844 0.1

0.0201 0.018 6

0.11999119 1.038422 1.6099909 10.624437 114.53 0.11 1.038422 1.6099909 1.1682393 1249 0.8 1.038422 1 . IIOIIIIIIOII 0.711481M .67 0.7 1.038422 1.eo99909 0.4869115 6.25 76 0.6 1.038422 1.l!OIIIIIIOII 0.3471681 a14 7'11

1.038422 1.IIO!IIIIIOII 0.2604628 270 83 0.4 1.038422 1.IIOIIIIIIOII 0.172II092 1.116 116 0.3 1.038422 1.l!Clll9SIOII 0.1067002 1.14 IIO 0.2 1.IXJIM22 1.809990II 0.0515231 0.66 113 0.1 1.038422 1.l!08IIII09 0.01421116 0.16 117 0 1.038422 1.6099909 0 0.00 100

8::a . eQu::.aiDrctn:im.:11

222 7.11 1a83 18.311 22.111 26.33 24.311 20.44 14.22 7.33 217 6.38 8.215 11.12 11 .IM 10.01 11.70 8.315 7.lll 7.116 11.88 6.IM

0.21129616 1.7044832 4.667111111 .11128438 10.006326 11 .666742 11 .014511 8.432774 4.868018 1.786771 0.281936 61.903 0.36 1.96 6.20 7.90 10.89 1263 11 .915 11.22 6.42 205 0.34 0.31 202 6.61 11.08 1262 14.78 1a311 11.41 6.36 1.87 0.31 0.31 202 6.67 11.16 1263 14.111 1a60 11.41 6.31 1.116 0.31 0.31 202 6.62 11.32 1274 16.16 1a62 11.60 6.31 1.83 0.30

202 6.62 11.40 12915 16.251 13.74 11.60 6.31 1.81 0.30 202 6.67 11.48 1a0& 16.41 1a86 11.60 6.215 1.78 0.30 202 6.72 11.66 13.28 16.66 1a86 11.50 6.26 1.76 0.251 202 11.63 1a311 1 1 118 11.60 6.215 1. 4 0.28 202 6.72 11.71 1a60 16.7'11 1all8 II.fill 6.20 1.72 0.28 202 6.n 11.71 13.60 16.112 14.10 II.fill 6.20 1.72 0.28 202 6.n 11.7'11 13.61 111.04 14.10 II.fill 6.20 1.70 0.28 202 6.83 11.915 1a83 111.1 14.22 II.fill 16 1.f 18 0.27 200 6.118 10.03 14.04 111.42 14.34 II.fill 6.16 UM 0.215 2.00 6.118 10.111 14.215 16.68 14.68 11.69 6.0II 1.62 0.25 202 6.72 11.71 1a60 16.7'11 13.118 11.69 6.20 1.72 0.28 11.24 a.co 222 -3.49 ~OIi ~62 -1.lll 244 6.66 0.00 0.00 ~OIi -0.42 242

10.00 10.00 0.42 0.00 10.00 tl.24 2.22 0.00 0.00 3.24 0.00 0.00 0.00 6.20 00 2.02 16.7'11 1.78 0.28

1.05-0.85 0.85-0. 70 0.85-0. 70 0. 70-0.60

Fig. C1. Typical MODAMP Screen (OuattroPro Version).

I

I I I I

I I I

I

I I I I I I

191

The program is looking for inputs of "Good", "Fair", "Poor" or "V. Poor".

Guidance for selection can be found in Table C1. Knowledge of the subgrade soil

type will allow approximation of the subgrade drainability with sufficient accuracy

to use these programs. The soil drainability from the general Missouri Soil Survey

(20) is used to enter the SG Drainage input into the programs.

Table C1. Quality of Subgrade Drainage.

Rating Soil Drainage Additional Moisture Contribution

Good • Relatively high permeability Low Moisture Contribution: (predominantly granular • deep water table soils) • absence of wet-weather springs

• at-grade or on fill

• flooding potential: none or rare

Fair • moderate permeability (fine Moderate to none: to moderately fine soil • deep water table texture) • absence of wet-weather springs

• may have layer that • at-grade or on fill impedes downward • flooding potential: none to occasional drainage

Poor • low permeability (.eg: silty Positive moisture contribution: clays) • shallow water table

• may have layer that • absence of wet-weather springs impedes downward • at-grade or in fill drainage • flooding potential: occasional to frequent

Very • very low permeability Positive moisture contribution: Poor (heavy clays) • shallow water table

• contains layer that impedes • in area of wet-weather springs drainage • sidehill cut or cut section

• flooding potential: frequent or common

• marshy area

Time-to-Drain/Quality of Base Drainage

The second large box displays the results of the time-to-drain calculations.

192

For degrees of drainage (U) from zero to 0.99999, the times-to-drain and the

corresponding saturation percentages are shown. The time-to-drain to 85%

saturation is displayed, as is the rating (Very Poor, Poor, Fair, Good, or Excellent)

at the bottom of the box. Numerical ratings (1-5) are also given. Criteria for

ratings are given in Table 28, repeated here as Table C2.

Table C2. Required Drainage Times for Quality of Drainage Levels.

Quality of Base or Time to Dra in (hr) Subbase Drainage

Excellent < 2

Good 2-24

Fair 24-168

Poor 168-720

Very Poor >720

Quality of Pavement Drainage

The Quality of Pavement Drainage is determined based on the input Quality

of Subgrade Drainage and the above-determined Quality of Base Drainage (line 22).

The Pavement Quality of Drainage table is identical to Table 24 in the MCHRP

report 90-4. The rating for Pave OD is shown at the bottom of the table,

expressed both as a description (Very Poor, Poor, Fair, Good, or Excellent) and

numerically (1-5).

Time of Saturation

The fourth large block in MODAMP is where the Percent Time of Saturation

is calculated for a typical year. Input includes average monthly temperatures in

I I I I I I I I I

I I I I I I I I I

193

degrees Celsius and average monthly precipitation in cm. At the bottom of the

block is displayed the saturation condition of each month: 11 1.0 11 if saturated the

entire time, 11 0.25 11 if part of the month is considered saturated.

Climate Condition

The Climate Condition block is identical to Table 25 in the MCHRP report,

except zones are delineated as 1 through 6 instead of A through F. Here the user

chooses the Climate Condition ( 1-6) that is most similar to the project site

condition based on the monthly time of saturation and frozen conditions from the

previous block. The choice is input by the user at the bottom of the next block on

line 73 for "ClimCond. 11

Drainage Coefficients

The last block displays m-coefficients. Input is the Climate Condition (1-6)

based on the information from the previous block, and Pavement Quality of

Drainage which MODAMP brings down from line 31. The resulting m-value is

displayed at the bottom of the block.

I I I

I

I

I I I I I I I I I I I I I I I I I

APPENDIX D

MISSOURI CLIMATOLOGICAL DAT A

(Bound Separately)

194

I

I