You are on page 1of 612

U.S.

Department of Transportation
Federal Highway Administration

Publication No. FHWA NHI-07-092


August 2008

NHI Course No. 132013____________________________

Geosynthetic Design & Construction Guidelines


Reference Manual

National Highway Institute

NOTICE
The contents of this report reflect the views of the authors, who are responsible for the
facts and the accuracy of the data presented herein. The contents do not necessarily
reflect policy of the Department of Transportation. This report does not constitute a
standard, specification, or regulation. The United States Government does not endorse
products or manufacturers. Trade or manufacturer's names appear herein only
because they are considered essential to the objective of this document.

Technical Report Documentation Page


1. REPORT NO.

FHWA-NHI-07-092

2. GOVERNMENT
ACCESSION NO.

4. TITLE AND SUBTITLE

3. RECIPIENT'S CATALOG NO.

5. REPORT DATE

Geosynthetic Design and


Construction Guidelines
7. AUTHOR(S)

August 2008
6. PERFORMING ORGANIZATION CODE
8. PERFORMING ORGANIZATION REPORT NO.

Robert D. Holtz, Ph.D., P.E., Barry R.


Christopher, Ph.D., P.E. and Ryan R. Berg, P.E.
9. PERFORMING ORGANIZATION NAME AND ADDRESS

Ryan R. Berg & Associates, Inc.


2190 Leyland Alcove
Woodbury, MN 55125
12. SPONSORING AGENCY NAME AND ADDRESS

National Highway Institute


Federal Highway Administration
U.S. Department of Transportation
Washington, D.C.

10. WORK UNIT NO.


11. CONTRACT OR GRANT NO.

DTFH61-02-T-63036
13. TYPE OF REPORT & PERIOD COVERED
14. SPONSORING AGENCY CODE

15. SUPPLEMENTARY NOTES

FHWA COTR Larry Jones


FHWA Technical Consultants: Jerry A. DiMaggio, P.E. and Daniel Alzamora, P.E.
This manual is the updated version of FHWA HI-95-038 (updated 1998) prepared by Ryan R. Berg & Associates,
Inc.; authored by R.D. Holtz, B.R. Christopher and R.R. Berg.

This manual is an updated version of the FHWA Reference Manual for the National
Highway Institutes training courses on geosynthetic design and construction. The update was
performed to reflect current practice and codes for geosynthetics in highway works. The manual was
prepared to enable the Highway Engineer to correctly identify and evaluate potential applications of
geosynthetics as alternatives to other construction methods and as a means to solve construction
problems. With the aid of this text, the Highway Engineer should be able to properly design, select,
test, specify, and construct with geotextiles, geocomposite drains, geogrids and related materials in
drainage, sediment control, erosion control, roadway, and embankment of soft soil applications.
Steepened reinforced soil slopes and MSE retaining wall applications are also addressed within, but
designers are referred to the more detailed FHWA NHI-00-043 reference manual on these subjects.
This manual is directed toward geotechnical, hydraulic, pavement, bridge and structures, construction,
maintenance, and route layout highway engineers, and construction inspectors and technicians
involved with design and/or construction and/or maintenance of transportation facilities that
incorporate earthwork.

16. ABSTRACT

geosynthetics, geotextiles,
geogrids, geomembranes, geocomposites,
roadway design, filters, drains, erosion control,
sediment control, separation, embankments, soil
reinforcement

17. KEY WORDS

19. SECURITY CLASSIF.

20. SECURITY CLASSIF.

Unclassified

Unclassified

18. DISTRIBUTION STATEMENT

No restrictions.

21. NO. OF PAGES

592

22. PRICE

Symbol

SI CONVERSION FACTORS
APPROXIMATE CONVERSIONS FROM SI UNITS
When You
Multiply By
To Find
Know

Symbol

LENGTH
mm
m
m
km

millimeters
meters
meters
kilometers

0.039
3.28
1.09
0.621
AREA

inches
feet
yards
miles

in
ft
yd
mi

mm2
m2
m2
ha
km2

square millimeters
square meters
square meters
hectares
square kilometers

0.0016
10.764
1.195
2.47
0.386
VOLUME

square inches
square feet
square yards
acres
square miles

in2
ft2
yd2
ac
mi2

ml
l
m3
m3

millimeters
liters
cubic meters
cubic meters

0.034
0.264
35.71
1.307
MASS

fluid ounces
gallons
cubic feet
cubic yards

fl oz
gal
ft3
yd3

g
kg
tonnes

grams
kilograms
tonnes

ounces
pounds
tons

oz
lb
tons

EC

Celsius

0.035
2.202
1.103
TEMPERATURE
1.8 C + 32
WEIGHT DENSITY

Fahrenheit

EF

kN/m3

kilonewton / cubic
meter

poundforce / cubic foot

pcf

6.36

FORCE and PRESSURE or STRESS


N
kN
kPa
kPa

newtons
kilonewtons
kilopascals
kilopascals

0.225
225
0.145
20.9

poundforce
poundforce
poundforce / square inch
poundforce / square foot

lbf
lbf
psi
psf

PREFACE

The 2007 update to the Geosynthetic Design & Construction Guidelines manual was initiated
to reflect the following recent publications:
AASHTO Standard Specifications for Geotextiles M 288; AASHTO, Standard
Specifications for Geotextiles - M 288, Standard Specifications for Transportation
Materials and Methods of Sampling and Testing, 26th Edition, American Association
of State Transportation and Highway Officials, Washington, D.C., 2006.
AASHTO, Geosynthetic Reinforcement of the Aggregate Base Course of Flexible
Pavement Structures PP 46-0, Standard Specifications for Transportation Materials
and Methods of Sampling and Testing, 26th Edition, and Provisional Standards,
American Association of State Transportation and Highway Officials, Washington,
D.C., 2006.
Ground Improvement Methods, FHWA NHI-06-019 Volume I and FHWA NHI-06020 Volume II, 2006;
Geotechnical Aspects of Pavements, FHWA-NHI-05-037, 2006;
Development of Design Methods for Geosynthetic Reinforced Flexible Pavements,
FHWA DTFH61-01-X-00068, May 2004, 263p.;
Available at: http://www.coe.montana.edu/wti/wti/pdf/426202_Final_Report.pdf
NCHRP 1-37A Design Guide (2002). 2002 Design Guide Design of New and
Rehabilitated Pavement Structures, Draft Final Report, Part 1 Introduction and Part
2 Design Inputs, Prepared for the National Cooperative Highway Research Program
by ERES Division of ARA.
AASHTO Standard Specifications for Highway Bridges, Seventeenth Edition, 2002;
Mechanically Stabilized Earth Walls and Reinforced Soil Slopes Design and
Construction Guidelines, FHWA NHI-00-043, March 2001;
Corrosion/Degradation of Soil Reinforcements for Mechanically Stabilized Earth
Walls and Reinforced Soil Slopes, FHWA NHI-00-044, March 2001;
Geosynthetic Reinforcement of the Aggregate Base/Subbase Courses of Pavement
Structures B GMA White Paper II, Geosynthetic Materials Association, Roseville,
MN, 2000, 176 p.; and
Geosynthetics in Pavement Systems Applications, Section One: Geogrids, Section
Two: Geotextiles, prepared for AASHTO, Geosynthetics Materials Association,
Roseville, MN, 1999, 46 p.

FHWA NHI-07-092
Geosynthetics Engineering

i-1

August 2008

The 2007 revised Geosynthetic Design & Construction Guidelines manual evolved from the
following FHWA manuals:
Geosynthetic Design & Construction Guidelines by Robert D. Holtz, Barry R.
Christopher, and Ryan R. Berg; Ryan R. Berg & Associates, Inc., FHWA HI-95-038;
1995 and updated in 1998; 460 p.
Geotextile Design & Construction Guidelines - Participant Notebook by Barry R.
Christopher and Robert D. Holtz; STS Consultants, Northbrook, Illinois, and
GeoServices, Inc., Boca Raton, Florida; October 1988 and selectively updated to
April 1992.
Geotextile Engineering Manual by Barry R. Christopher and Robert D. Holtz; STS
Consultants, Northbrook, Illinois; March, 1985; 917 p.
Use of Engineering Fabrics in Transportation Type Related Applications by T. Allan
Haliburton, J.D. Lawmaster, and Verne C. McGuffey; 1981.
Guidelines for Design, Specification, and Contracting of Geosynthetic Mechanically
Stabilized Earth Slopes on Firm Foundations; by Ryan R. Berg; Ryan R. Berg &
Associates, St. Paul, Minnesota; January, 1993; 88p.
Reinforced Soil Structures - Volume I, Design and Construction Guidelines, and
Volume II Summary of Research and Systems Information; by B.R. Christopher, S.A.
Gill, J.P. Giroud, J.K. Mitchell, F. Schlosser, and J. Dunnicliff; STS Consultants,
Northbrook, Illinois, November 1990.

Special Acknowledgement
Jerry A. DiMaggio, P.E. is the FHWA Technical Consultant for this work, and served
in the same capacity for most of the above referenced publications. Mr. DiMaggio's
guidance and input to this and the previous works was invaluable.
The Geosynthetics Materials Association (GMA), the North American Geosynthetics
Society (NAGS), and the International Geosynthetics Society (IGS) provided support for this
revision. Their support to help initiate and to review this update is gratefully appreciated.

FHWA NHI-07-092
Geosynthetics Engineering

i- 2

August 2008

TABLE OF CONTENTS

1.0 INTRODUCTION............................................................................................................... 1-1


1.1 BACKGROUND ..................................................................................................... 1-1
1.2 DEFINITIONS, MANUFACTURING PROCESSES,
AND IDENTIFICATION.................................................................................... 1-2
1.3 FUNCTIONS AND APPLICATIONS .................................................................... 1-5
1.4 DESIGN APPROACH ............................................................................................ 1-9
1.5 EVALUATION OF PROPERTIES......................................................................... 1-9
1.6 SPECIFICATIONS................................................................................................ 1-21
1.7 FIELD INSPECTION............................................................................................ 1-24
1.8 FIELD SEAMING ................................................................................................. 1-24
1.9 REFERENCES ...................................................................................................... 1-29

2.0 GEOSYNTHETICS IN SUBSURFACE DRAINAGE SYSTEMS ................................ 2-1


2.1 BACKGROUND ..................................................................................................... 2-1
2.2 APPLICATIONS ..................................................................................................... 2-2
2.3 GEOTEXILE FILTER DESIGN - PRINCIPLES AND CONCEPTS .................... 2-4
2.4 FHWA FILTER DESIGN PROCEDURE............................................................... 2-8
2.4-1 Retention Criteria ..................................................................................... 2-8
2.4-2 Permeability and Permittivity Criteria.................................................... 2-10
2.4-3 Clogging Resistance ............................................................................... 2-12
2.4-4 Survivability and Durability Criteria...................................................... 2-15
2.4-5 Additional Filter Selection Considerations and Summary ..................... 2-16
2.5 DRAINAGE SYSTEM DESIGN GUIDELINES ................................................. 2-19
2.6 DESIGN EXAMPLE ............................................................................................. 2-25
2.7 COST CONSIDERATIONS.................................................................................. 2-30
2.8 SPECIFICATIONS ............................................................................................... 2-31
2.9 INSTALLATION PROCEDURES ....................................................................... 2-36
2.10 FIELD INSPECTION........................................................................................... 2-37
2.11 IN-PLANE DRAINAGE; PREFABRICATED
GEOCOMPOSITE DRAINS............................................................................. 2-40
2.11-1 Design and Selection Criteria ............................................................... 2-41
2.11-2 Construction Considerations................................................................. 2-44
2.12 REFERENCES .................................................................................................... 2-45

FHWA NHI-07-092
Geosynthetics Engineering

i-3

August 2008

3.0 GEOTEXTILES IN RIPRAP REVETMENTS AND


OTHER PERMANENT EROSION CONTROL SYSTEMS .................................... 3-1
3.1 BACKGROUND ..................................................................................................... 3-1
3.2 APPLICATIONS ..................................................................................................... 3-2
3.3 DESIGN OF GEOTEXTILES BENEATH HARD ARMOR AND DESIGN
CONCEPTS ......................................................................................................... 3-4
3.3-1 Retention Criteria for Cyclic or Dynamic Flow ....................................... 3-4
3.3-2 Permeability and Effective Flow Capacity Requirements
for Erosion Control .................................................................................. 3-4
3.3-3 Clogging Resistance for Cyclic or Dynamic Flow and for Problematic
Soils.......................................................................................................... 3-5
3.3-4 Survivability Criteria for Erosion Control................................................ 3-5
3.3-5 Additional Filter Selection Considerations and Summary ....................... 3-6
3.4 GEOTEXTILE DESIGN GUIDELINES ................................................................ 3-9
3.5 GEOTEXTILE DESIGN EXAMPLE ................................................................... 3-15
3.6 GEOTEXTILE COST CONSIDERATIONS ........................................................ 3-20
3.7 GEOTEXTILE SPECIFICATIONS ...................................................................... 3-21
3.8 GEOTEXTILE INSTALLATION PROCEDURES.............................................. 3-27
3.8-1 General Construction Considerations..................................................... 3-27
3.8-2 Cut and Fill Slope Protection.................................................................. 3-30
3.8-3 Streambank Protection............................................................................ 3-34
3.8-4 Precipitation Runoff Collection and Diversion Ditches ......................... 3-35
3.8-5 Wave Protection Revetments.................................................................. 3-36
3.8-6 Scour Protection ..................................................................................... 3-37
3.9 GEOTEXTILE FIELD INSPECTION .................................................................. 3-38
3.10 GEOCELLS ......................................................................................................... 3-38
3.11 EROSION CONTROL MATS ............................................................................ 3-39
3.11-1 Summary of Planning, Design, and Installation ................................... 3-41
3.11-2 Specification .......................................................................................... 3-43
3.12 REFERENCES .................................................................................................... 3-44

4.0 TEMPORARY RUNOFF AND SEDIMENT CONTROL ............................................. 4-1


4.1 INTRODUCTION ................................................................................................... 4-1
4.2 FUNCTION OF SILT FENCES.............................................................................. 4-3
4.3 DESIGN OF SILT FENCES ................................................................................... 4-4
4.3-1 Estimates of Runoff and Sediment Volumes............................................ 4-5
4.3-2 Hydraulic Design of the Geotextile .......................................................... 4-6
4.3-3 Physical and Mechanical Properties; Constructability Requirements...... 4-7
FHWA NHI-07-092
Geosynthetics Engineering

i- 4

August 2008

4.4
4.5
4.6
4.7
4.8
4.9

SPECIFICATIONS.................................................................................................. 4-9
INSTALLATION PROCEDURES ....................................................................... 4-13
INSPECTION AND MAINTENANCE ................................................................ 4-13
SILT AND TURBIDITY CURTAINS.................................................................. 4-15
EROSION CONTROL BLANKETS .................................................................... 4-18
REFERENCES ...................................................................................................... 4-20

5.0 GEOSYNTHETICS IN ROADWAYS AND PAVEMENTS.......................................... 5-1


5.1 INTRODUCTION ................................................................................................... 5-1
5.2 APPLICABILITY AND BENEFITS OF GEOSYNTHETICS IN ROADWAYS . 5-2
5.2-1 Temporary Roads and Working Platforms............................................... 5-2
5.2-2 Permanent Paved and Unpaved Roads .................................................... 5-3
5.2-3 Subgrade Conditions in which Geosynthetics are Useful ........................ 5-4
5.2-4 Benefits..................................................................................................... 5-6
5.3 ROADWAY DESIGN USING GEOSYNTHETICS .............................................. 5-8
5.3-1 Functions of Geosynthetics in Roadways and Pavements........................ 5-8
5.3-2 Possible Failure Modes of Permanent Roads ......................................... 5-11
5.3-3 Design for Separation ............................................................................. 5-12
5.3-4 Design for Stabilization .......................................................................... 5-13
5.3-5 Reinforced Base/Subbase Design........................................................... 5-15
5.3-6 Material Properties used in Design......................................................... 5-16
5.4 DESIGN GUIDELINES FOR USE OF GEOTEXTILES IN TEMPORARY
AND UNPAVED ROADS ................................................................................ 5-21
5.4-1 Temporary Road Design Example ....................................................... 5-27
5.5 DESIGN GUIDELINES FOR USE OF GEOGRIDS IN TEMPORARY AND
UNPAVED ROADS.......................................................................................... 5-30
5.5-1 Empirical Design Method: Modified Steward et al............................. 5-30
5.5-2 Empirical Design Method of Giroud and Han...................................... 5-31
5.5-3 Design Example for Geogrid Reinforced Unpaved Road .................... 5-35
5.6 DESIGN GUIDELINES FOR USE OF GEOTEXTILES IN PERMANENT
PAVED ROADWAYS ...................................................................................... 5-42
5.6-1 Separation ............................................................................................. 5-42
5.6-2 Stabilization .......................................................................................... 5-42
5.6-3 Permanent Road Subgrade Stabilization Design Example................... 5-45
5.6-4 Improved Drainage ............................................................................... 5-49
5.7 DESIGN GUIDELINES FOR USE OF GEOGRIDS IN PERMANENT PAVED
ROADWAYS .................................................................................................... 5-52
5.7-1 Empirical Design Method from AASHTO PP46-01............................ 5-53
5.7-2 Mechanistic-Empirical Approach for Pavement Design ...................... 5-56
FHWA NHI-07-092
Geosynthetics Engineering

i-5

August 2008

5.7-3 Design Example for Geogrid Reinforced Paved Roadway .................. 5-58
5.8 INSTALLATION PROCEDURES ....................................................................... 5-64
5.8-1 Roll Placement...................................................................................... 5-64
5.8-2a Geotextile Overlaps .............................................................................. 5-67
5.8-2b Geogrid Overlaps.................................................................................. 5-69
5.8-3 Seams.................................................................................................... 5-70
5.8-4 Field Inspection .................................................................................... 5-70
5.9 SPECIFICATIONS................................................................................................ 5-70
5.9-1 Geotextile for Separation and Stabilization Applications ...................... 5-70
5.9-2 Geogrids for Subgrade Stabilization....................................................... 5-76
5.9-3 Geosynthetics for Base Reinforcement of Pavement Structures ............ 5-80
5.10 COST CONSIDERATIONS................................................................................ 5-85
5.11 REFERENCES .................................................................................................... 5-86
6.0 PAVEMENT OVERLAYS ................................................................................................ 6-1
6.1 BACKGROUND ..................................................................................................... 6-1
6.2 PAVEMENT OVERLAYS AND REFLECTIVE CRACKING............................. 6-1
6.3 GEOTEXTILES....................................................................................................... 6-4
6.3-1 Functions .................................................................................................. 6-4
6.3-2 Asphalt Concrete (AC) Pavement Applications....................................... 6-5
6.3-3 Portland Cement Concrete Pavement Applications.................................. 6-6
6.3-4 HMAC-Overlaid PCC Pavements ............................................................ 6-7
6.3-5 Chip Seals for Unpaved Roads and AC Pavements ................................. 6-8
6.3-6 Advantages and Potential Disadvantages ................................................. 6-8
6.3-7 Design.................................................................................................... 6-10
6.3-8 Geotextile Selection................................................................................ 6-13
6.3-9 Cost Considerations................................................................................ 6-13
6.3-10 Specifications.......................................................................................... 6-16
6.3-11 Field Inspection ...................................................................................... 6-23
6.3-12 Recycling ................................................................................................ 6-23
6.4 GEOGRIDS ........................................................................................................... 6-24
6.4-1 Geogrid Functions .................................................................................. 6-24
6.4-2 Applications............................................................................................ 6-24
6.4-3 Design.................................................................................................... 6-25
6.4-4 Installation ............................................................................................. 6-26
6.4-5 Cost Considerations................................................................................ 6-26
6.4-6 Specifications.......................................................................................... 6-27
6.5 GEOCOMPOSITES .............................................................................................. 6-29
6.5-1 Membrane and Composite Strips ........................................................... 6-29
FHWA NHI-07-092
Geosynthetics Engineering

i- 6

August 2008

6.5-2 Specifications.......................................................................................... 6-30


6.6 REFERENCES ...................................................................................................... 6-31

7.0 REINFORCED EMBANKMENTS ON SOFT FOUNDATIONS ................................. 7-1


7.1 BACKGROUND ..................................................................................................... 7-1
7.2 APPLICATIONS ..................................................................................................... 7-2
7.3
DESIGN GUIDELINES FOR REINFORCED EMBANKMENTS
ON SOFT SOILS ................................................................................................ 7-3
7.3-1 Design Considerations.............................................................................. 7-3
7.3-2 Design Steps ............................................................................................. 7-5
7.3-3 Comments on the Design Procedure ...................................................... 7-13
7.4 SELECTION OF GEOSYNTHETIC AND FILL PROPERTIES......................... 7-25
7.4-1 Geotextile and Geogrid Strength Requirements..................................... 7-26
7.4-2 Drainage Requirements .......................................................................... 7-28
7.4-3 Environmental Considerations ............................................................... 7-28
7.4-4 Constructability (Survivability) Requirements....................................... 7-28
7.4-5 Stiffness and Workability ....................................................................... 7-31
7.4-6 Fill Considerations.................................................................................. 7-33
7.5 DESIGN EXAMPLE ............................................................................................. 7-33
7.6 SPECIFICATIONS................................................................................................ 7-40
7.7 COST CONSIDERATIONS.................................................................................. 7-44
7.8 CONSTRUCTION PROCEDURES...................................................................... 7-45
7.9 INSPECTION ........................................................................................................ 7-52
7.10 REINFORCED EMBANKMENTS FOR ROADWAY WIDENING ................ 7-52
7.11 REINFORCEMENT OF EMBANKMENTS COVERING LARGE AREAS.... 7-54
7.12 COLUMN SUPPORTED EMBANKMENTS .................................................... 7-54
7.13 REFERENCES .................................................................................................... 7-57

8.0 REINFORCED SLOPES ................................................................................................... 8-1


8.1 BACKGROUND ..................................................................................................... 8-1
8.2 APPLICATIONS ..................................................................................................... 8-1
8.3 DESIGN GUIDELINES FOR REINFORCED SLOPES........................................ 8-4
8.3-1 Design Concepts ....................................................................................... 8-4
8.3-2 Design of Reinforced Slopes .................................................................... 8-5
8.3-3 Reinforced Slope Design Guidelines........................................................ 8-7
8.3-4 Computer Assisted Design ..................................................................... 8-27
8.4 MATERIAL PROPERTIES .................................................................................. 8-28
FHWA NHI-07-092
Geosynthetics Engineering

i-7

August 2008

8.4-1 Reinforced Slope Systems ...................................................................... 8-28


8.4-2 Soils ........................................................................................................ 8-28
8.4-3 Geosynthetic Reinforcement .................................................................. 8-30
8.5 TREATMENT OF OUTER FACE........................................................................ 8-33
8.6 PRELIMINARY DESIGN AND COST EXAMPLE............................................ 8-37
8.7 COST CONSIDERATIONS.................................................................................. 8-43
8.8 IMPLEMENTATION............................................................................................ 8-44
8.9 SPECIFICATIONS AND CONTRACTING APPROACH .................................. 8-46
8.9-1 Specification for Geosynthetic Soil Reinforcement ............................... 8-47
8.9-2 Specification for Geosynthetic Reinforced Soil Slope System .............. 8-53
8.10 INSTALLATION PROCEDURES ..................................................................... 8-55
8.11 FIELD INSPECTION.......................................................................................... 8-59
8.12 STANDARD DESIGNS...................................................................................... 8-59
8.13 REFERENCES .................................................................................................... 8-62
9.0 MECHANICALLY STABILIZED EARTH RETAINING
WALLS AND ABUTMENTS ....................................................................................... 9-1
9.1 BACKGROUND ..................................................................................................... 9-1
9.2 APPLICATIONS ..................................................................................................... 9-3
9.3 DESCRIPTION OF MSE WALLS ......................................................................... 9-5
9.3-1 Soil Reinforcements ................................................................................. 9-5
9.3-2 Facings...................................................................................................... 9-5
9.4 DESIGN GUIDELINES FOR MSE WALLS ....................................................... 9-10
9.4-1 Approaches and Models ......................................................................... 9-10
9.4-2 Design Steps .......................................................................................... 9-12
9.4-3 Comments on the Design Procedure ...................................................... 9-18
9.4-4 Drainage.................................................................................................. 9-30
9.4-5 Seismic Design ....................................................................................... 9-32
9.5 LATERAL DISPLACEMENT.............................................................................. 9-34
9.6 MATERIAL PROPERTIES .................................................................................. 9-34
9.6-1 Reinforced Wall Fill Soil........................................................................ 9-34
9.6-2 Geosynthetic Reinforcement .................................................................. 9-36
9.7 COST CONSIDERATIONS................................................................................. 9-43
9.8 COST ESTIMATE EXAMPLES ......................................................................... 9-44
9.8-1 Geogrid, MBW Unit-Faced Wall ........................................................... 9-44
9.8-2 Geotextile Wrap Wall............................................................................. 9-48
9.9 SPECIFICATIONS................................................................................................ 9-51
9.9-1 Geosynthetic, MBW Unit-Faced Wall ................................................... 9-51
9.9-2 Modular Block Wall Unit ....................................................................... 9-59
FHWA NHI-07-092
Geosynthetics Engineering

i- 8

August 2008

9.10

9.11
9.12

9.13

9.14

9.9-3 Geosynthetic Wrap Around Wall ........................................................... 9-66


CONSTRUCTION PROCEDURES.................................................................... 9-74
9.10-1 Concrete Faced Wall ............................................................................ 9-74
9.10-2 Geotextile Wrap-Around Wall ............................................................. 9-76
INSPECTION ...................................................................................................... 9-78
IMPLEMENTATION.......................................................................................... 9-80
9.12-1 Design Responsibility........................................................................... 9-81
9.12-2 Standardized Designs ........................................................................... 9-81
9.12-1 Geosynthetic Design Strength .............................................................. 9-85
SUMMARY OF LOAD RESISTANCE FACTOR DESIGN ............................. 9-85
9.13-1 Introduction .......................................................................................... 9-85
9.13-2 Background........................................................................................... 9-86
9.13-3 MSE Wall Design................................................................................. 9-87
9.13-4 External Stability .................................................................................. 9-87
9.13-5 Internal Stability ................................................................................... 9-88
REFERENCES .................................................................................................... 9-89

10.0 GEOMEMBRANES AND OTHER GEOSYNTHETIC BARRIERS ....................... 10-1


10.1 BACKGROUND ................................................................................................. 10-1
10.2 GEOSYNTHETIC BARRIER MATERIALS..................................................... 10-1
10.2-1 Geomembranes ..................................................................................... 10-2
10.2-2 Thin-Film Geotextile Composites ........................................................ 10-3
10.2-3 Geosynthetic Clay Liners ..................................................................... 10-4
10.2-4 Field-Impregnated Geotextiles ............................................................. 10-5
10.3 APPLICATIONS ................................................................................................. 10-5
10.4 DESIGN CONSIDERATIONS ......................................................................... 10-10
10.4-1 Performance Requirements................................................................. 10-11
10.4-2 In-Service Conditions ......................................................................... 10-11
10.4-3 Durability............................................................................................ 10-12
10.4-4 Installation Conditions........................................................................ 10-12
10.4-5 Peer Review........................................................................................ 10-14
10.4-6 Economic Considerations ................................................................... 10-14
10.5 INSTALLATION .............................................................................................. 10-14
10.6 INSPECTION .................................................................................................... 10-15
10.6-1 Manufacture........................................................................................ 10-16
10.6-2 Field .................................................................................................... 10-16
10.7 SPECIFICATION .............................................................................................. 10-16
10.8 REFERENCES .................................................................................................. 10-17

FHWA NHI-07-092
Geosynthetics Engineering

i-9

August 2008

APPENDICES
Appendix A GEOSYNTHETIC LITERATURE
Appendix B GEOSYNTHETIC TERMS
Appendix C NOTATION AND ACRONYMS
Appendix D AASHTO M288 SPECIFICATION
Appendix E GEOSYNTHETIC TEST STANDARDS
E-1 American Society for Testing and Materials
E-2 Geosynthetic Research Institute
Appendix F REPRESENTATIVE LIST OF GEOSYNTHETIC
MANUFACTURERS AND SUPPLIERS
Appendix G GENERAL PROPERTIES AND COSTS OF GEOTEXTILES
AND GEOGRIDS
Appendix H GEOSYNTHETIC REINFORCEMENT STRUCTURAL
DESIGN PROPERTIES
H.1 BACKGROUND
H.2 TENSILE STRENGTHS
H.3 REDUCTION FACTORS
H.4 IMPLEMENTATION
H.5 ALTERNATIVE LONG-TERM STRENGTH DETERMINATION
H.6 SOIL-REINFORCEMENT INTERACTION
H.7 REFERENCES

FHWA NHI-07-092
Geosynthetics Engineering

i- 10

August 2008

List of Tables

1-1
1-2
1-3
1-4
1-5

Representative Applications and Controlling Functions of Geosynthetics ......................... 1-7


Important Criteria and Principal Properties Required for Evaluation of Geosynthetics ... 1-11
Evaluation of Geosynthetic Property Requirements ......................................................... 1-12
Geosynthetic Properties and Parameters ........................................................................... 1-12
Geosynthetic Field Inspection Checklist ........................................................................... 1-25

2-1 Guidelines for Evaluating the Critical Nature or Severity


of Drainage and Erosion Control Applications.................................................... 2-4
2-2 Geotextile Strength Property Requirements for Drainage Geotextiles ............................. 2-15
3-1 Geotextile Strength Property Requirements for Permanent
Erosion Control Geotextiles................................................................................ 3-7
4-1 Limits of Slope Steepness and Length to Limit Runoff Velocity to 0.3 m/s ...................... 4-4
4-2 Physical Requirements for Temporary Silt Fence Geotextiles ............................................ 4-9
5-1
5-2
5-3
5-4
5-5

Application and Associated Functions of Geosynthetics in Roadway Systems.................. 5-6


Construction Survivability Ratings ................................................................................... 5-16
Geotextile Property Requirements for Stabilization Applications (CBR < 3)................... 5-18
Geotextile Property Requirements for Separation Applications (CBR > 3)...................... 5-19
Geogrid Survivability Property Requirements for
Stabilization and Base Reinforcement Applications.......................................... 5-20
5-6 Bearing Capacity Factors for Different Ruts and Traffic
Conditions Both With and Without Geosynthetics............................................ 5-23
5-7 Recommended mi Values for Modifying Structural Layer Coefficients of
Untreated Base and Subbase Materials in Flexible Pavements ......................... 5-50
5-8 Quality of Pavement Drainage .......................................................................................... 5-50
5-9 Recommended Values of Drainage Coefficient, Cd, for Rigid Pavement Design ............. 5-51
5-10 Qualitative Review of Reinforcement Application
Potential for Paved Permanent Roads................................................................ 5-55
5-11 Recommended Minimum Geotextile Overlap Requirements ......................................... 5-69
6-1 Paving Grade Geotextile Selection.................................................................................... 6-14

FHWA NHI-07-092
Geosynthetics Engineering

i-11

August 2008

7-1 Geosynthetic Properties Required for Reinforcement Applications ................................. 7-25


7-2 Required Degree of Geosynthetic Survivability as a Function of Subgrade Conditions
and Construction Equipment.............................................................................. 7-29
7-3 Required Degree of Geosynthetic Survivability as a Function of Cover Material
and Construction Equipment.......................................................................................... 7-30
7-4 Minimum Geotextile Property Requirements for Geotextile Survivability....................... 7-31
7-5 Geogrid Survivabilty Property Requirements ................................................................... 7-32
8-1
8-2
8-3
8-4
8-5

RSS Slope Facing Options ................................................................................................ 8-21


Default Values For F* and Pullout Factors .................................................................... 8-32
Allowable Geotextile Strength with Various Soil Types .................................................. 8-51
Allowable Geogrid Strength with Various Soil Types...................................................... 8-52
Durability Reduction Factors by Product and Soil Fill pH ............................................... 8-53

9-1
9-2
9-3
9-4

Internal Failure Modes and Required Properties For MSE Walls..................................... 9-12
Default Values For F* And " Pullout Factors ................................................................... 9-30
MSE Soil Fill Requirements.............................................................................................. 9-35
Common LRFD Load Groups for Walls ........................................................................... 9-87

10-1 Common Types of Geomembranes ................................................................................. 10-2


10-2 Recommended Minimum Properties for General
Geomembrane Installation Survivability ......................................................... 10-13
G-1 General Range of Strength and Permeability Properties for
Representative Types of Geotextiles and Geogrids
G-2 General Description of Geotextiles
G-3 Approximate Cost Range of Geotextiles and Geogrids
H-1
H-2
H-3
H-4
H-5
H-6
H-7

Typical Ranges of Creep Reduction Factors


Installation Damage Reduction Factors
Aging Reduction Factors, PET
Anticipated Resistance of Polymers to Specific Environments
Recommended pH Limits for Reinforced Fill Soils
Minimum Requirements for Use of Default Reduction Factor
Basic Aspects of Reinforcement Pullout Performance in
Granular and Low Cohesive Soils
H-8 Default Values for F* and " Pullout Factors

FHWA NHI-07-092
Geosynthetics Engineering

i- 12

August 2008

List of Figures
Figure 1-1
Figure 1-2
Figure 1-3
Figure 2-1
Figure 2-2
Figure 2-3
Figure 2-4
Figure 2-5
Figure 2-6
Figure 2-7
Figure 2-8
Figure 2-9
Figure 2-10
Figure 2-11
Figure 2-12

Figure 3-1
Figure 3-2
Figure 3-3
Figure 3-4
Figure 3-5

Figure 4-1
Figure 4-2
Figure 4-3
Figure 4-4
Figure 4-5

Classification of geosynthetics and other soil inclusions .................................... 1-4


Types of (a) stitches and (b) seams, according to Federal Standard No. 751a ;
and (c) improper seam placement ...................................................................... 1-28
Bodkin connection of HDPE uniaxial geogrid .................................................. 1-29
Grain-size distribution for several soils ............................................................... 2-5
Filter bridge formation......................................................................................... 2-7
Definitions of clogging and blinding ................................................................... 2-7
U.S. Army Corps of Engineers gradient ratio test device.................................. 2-14
Flow chart summary of FHWA filter design procedure .................................... 2-18
Typical gradations and Darcy permeabilities of several aggregate and
graded filter materials ........................................................................................ 2-21
Construction procedure for geotextile-lined underdrains .................................. 2-37
Construction of geotextile drainage systems ..................................................... 2-38
Construction geotextile filters and separators beneath permeable
pavement base.................................................................................................... 2-39
Geocomposite drains.......................................................................................... 2-42
Prefabricated geocomposite edge drain construction using sand fill
upstream of composite ....................................................................................... 2-46
Recommended installation method for prefabricated geocomposite
edge drains ......................................................................................................... 2-47
Flow chart summary of FHWA filter design procedure ...................................... 3-8
Erosion control installations: a) installation in wave protection revetment;
b) river shoreline application; and c) stream application................................... 3-28
Construction of hard armor erosion control systems ......................................... 3-31
Special construction requirements related to specific hard armor erosion
control applications............................................................................................ 3-32
Recommended maximum design velocities and flow durations for erosion
resistance of various surface materials and treatments ...................................... 3-42
Geotextile strength versus post spacing............................................................... 4-7
Post requirements versus post spacing................................................................. 4-8
Typical silt fence installation ............................................................................. 4-14
Installation of a prefabricated silt fence............................................................. 4-15
Recommended maximum design velocities and flow durations for various
classes of erosion control materials .................................................................... 4-19

FHWA NHI-07-092
Geosynthetics Engineering

i-13

August 2008

Figure 5-1
Figure 5-2
Figure 5-3
Figure 5-5
Figure 5-6
Figure 5-7
Figure 5-8
Figure 5-9
Figure 5-10
Figure 5-11
Figure 5-12
Figure 5-13

Potential applications of geosynthetics in a layered pavement system................ 5-2


Geotextile separator beneath permeable base ...................................................... 5-5
Concept of geotextile separation in roadways ..................................................... 5-9
Filtration at the interface of two dissimilar materials (without geosynthetics) . 5-14
U.S. Forest Service thickness design curve for single wheel load..................... 5-25
U.S. Forest Service thickness design curve for tandem wheel load .................. 5-25
Thickness design curves with geosynthetics for a) single and b) dual wheel
loads (modified for highway applications) ........................................................ 5-26
Aggregate loss to weak subgrades ..................................................................... 5-44
Mechanistic-Empirical (M-E) Pavement Design Method showing a) M-E
concept, and b) modified response model for inclusion of reinforcement ........ 5-56
Construction sequence using geosynthetics....................................................... 5-65
Forming curves using geotextiles ...................................................................... 5-68
Repair of rutting with additional material.......................................................... 5-69

Figure 6-1
Figure 6-2
Figure 6-3

Shearing and bending stress in HMA overlay ..................................................... 6-2


Geotextiles (a.k.a. Paving Fabric) in rehabilitated pavement section.................. 6-4
Relationship between the vertical compressive strain at the top of the
subgrade and the number of load applications in geogrid reinforced pavement 6-25

Figure 7-1
Figure 7-2
Figure 7-3
Figure 7-4
Figure 7-5
Figure 7-6
Figure 7-7

Figure 7-9
Figure 7-10
Figure 7-11
Figure 7-12

Reinforced embankment applications.................................................................. 7-3


Reinforced embankments failure modes.............................................................. 7-4
Reinforcement required to provide rotational stability...................................... 7-10
Reinforcement required to limit lateral embankment spreading........................ 7-11
Embankment height versus undrained shear strength of foundation ................. 7-16
Local bearing failure (lateral squeeze)............................................................... 7-17
Construction sequence for geosynthetic reinforced embankments for
extremely weak foundations .............................................................................. 7-48
Placement of fill between toe berms on extremely soft foundations
(CBR < 1) with a mud wave anticipated............................................................ 7-49
Fill placement to tension geotextile on moderate ground conditions ................ 7-50
Reinforced embankment construction ............................................................... 7-51
Reinforced embankment construction for roadway widening ........................... 7-53
Column supported embankment with geosynthetic reinforcement ................... 7-56

Figure 8-1
Figure 8-2
Figure 8-3
Figure 8-4

Use of geosynthetics in engineered slopes........................................................... 8-2


Applications of RSSs: .......................................................................................... 8-3
Requirements for design of a reinforced slope .................................................... 8-8
Critical zone defined by rotational and sliding surface that meet the

Figure 7-8

FHWA NHI-07-092
Geosynthetics Engineering

i- 14

August 2008

Figure 8-5
Figure 8-6
Figure 8-7
Figure 8-8
Figure 8-9
Figure 8-10
Figure 8-11
Figure 8-12

Figure 9-1
Figure 9-2
Figure 9-3
Figure 9-4
Figure 9-5
Figure 9-6
Figure 9-7
Figure 9-8
Figure 9-9
Figure 9-10
Figure 9-11
Figure 9-12
Figure 9-13
Figure 9-14

required safety factor ......................................................................................... 8-10


Rotational shear approach to determine required strength of reinforcement..... 8-12
Sliding wedge approach to determine the coefficient of earth pressure K ........ 8-14
Spacing and embedding requirements for slope reinforcement showing:
primary and intermediate reinforcement layout................................................. 8-16
Developing reinforcement length ...................................................................... 8-18
Cost evaluation of reinforced soil slopes ........................................................... 8-44
Construction of reinforced slopes ...................................................................... 8-57
Reinforced slope construction............................................................................ 8-58
Example of standard design ............................................................................... 8-61

Figure 9-15
Figure 9-16

Component parts of a Reinforced Earth wall....................................................... 9-2


Reinforced retaining wall systems using geosynthetics....................................... 9-3
Examples of geosynthetic MSE walls.................................................................. 9-4
Possible geosynthetic MSE wall facings ............................................................. 9-6
Wall facings ......................................................................................................... 9-8
Actual geosynthetic reinforced soil wall in contrast to the design model ......... 9-11
Geometric and loading characteristics of geosynthetic MSE walls................... 9-14
Example MSE wall drainage blanket detail....................................................... 9-32
Drainage details for MBW faced, MSE wall ..................................................... 9-33
Polyethylene geogrid bodkin connection detail................................................. 9-39
Example MBW mechanical connection............................................................. 9-39
Cost comparison of reinforced systems ............................................................. 9-44
Lift construction sequence for geotextile reinforced soil walls ......................... 9-79
Typical face construction detail for vertical geogrid-reinforced
retaining wall faces ............................................................................................ 9-80
Example of standard MSE wall design.............................................................. 9-83
Typical application of live load surcharge for MSE walls ................................ 9-88

Figure 10-1
Figure 10-2
Figure 10-3
Figure 10-4
Figure 10-5
Figure 10-6
Figure 10-7
Figure 10-8
Figure 10-9

Thin-film geotextile composites ........................................................................ 10-4


Geosynthetic clay liners..................................................................................... 10-4
Control of expansive soils.................................................................................. 10-7
Control of horizontal infiltration of base ........................................................... 10-7
Maintenance of optimum water content ............................................................ 10-8
Waterproofing of tunnels ................................................................................... 10-8
Water conveyance canals................................................................................... 10-9
Secondary containment of underground fuel tanks ........................................... 10-9
Waterproofing of walls ................................................................................... 10-10

FHWA NHI-07-092
Geosynthetics Engineering

i-15

August 2008

FHWA NHI-07-092
Geosynthetics Engineering

i- 16

August 2008

1.0 INTRODUCTION

1.1 BACKGROUND
The objective of this manual is to assist highway design engineers, specification writers,
estimators, construction inspectors, and maintenance personnel with the design, selection,
and installation of geosynthetics. In addition to providing a general overview of these
materials and their applications, step-by-step procedures are given for the cost-effective use
of geosynthetics in drainage and erosion control systems, roadways, reinforced soil
structures, and in containment applications. Although the title refers to the general term
geosynthetic, the appropriate use of the subfamilies of geotextiles, geogrids, geocomposites,
and geomembranes are discussed in specific applications.
The basis for much of this manual is the FHWA Geotextile Engineering Manual (Christopher
and Holtz, 1985). Other sources of technical information include the book by Koerner
(2006) and a number of FHWA reports and publications. If you are not already somewhat
familiar with geosynthetics, you are encouraged to read the books by Ingold and Miller
(1988), Richardson and Koerner (1990), and Fannin (2000). Additional references are in the
geosynthetic bibliographies prepared by Giroud (1993, 1994).
Geosynthetics terminology is defined in Appendix B and ASTM (2006) D 4439 Standard
Terminology for Geosynthetics. Common notation and symbols are used throughout this
manual, and for easy reference a list is provided in Appendix C. The notation and symbols
are generally consistent with the International Geosynthetic Society's (IGS) Recommended
Mathematical and Graphical Symbols (2000).
Sample specifications for each primary application are also included in this manual.
Remember that these specifications are only guidelines and should be modified as
required by project specific design and performance criteria, engineering judgment,
and experience. For the more routine highway applications, specifications are adapted from
the American Association of State Highway and Transportation Officials (AASHTO)
Standard Specification, Designation M 288 (2006). (The AASHTO M 288 specification can
be found in Appendix D.) Other sample specifications were provided by New York and
Washington DOTs, the National Concrete Masonry Association, and the FHWA.
Historically, the AASHTO M 288 specifications were based on a geotextile specification
originally developed by Task Force 25 of the Joint Subcommittee on Materials of AASHTO,
the Association General Contractors (AGC), and the American Road and Transportation
Builders Associations (ARTBA), along with representatives from the geosynthetic industry
FHWA NHI-07-092
Geosynthetics Engineering

1-1

Introduction, Identification, and Evaluation


August 2008

(AASHTO, 1990a). Another important early group was Task Force 27 on soil reinforcing,
sponsored by the same AASHTO-AGC-ARTBA Subcommittee on Materials (AASHTO,
1990b). The FHWA soil reinforcing specifications for walls and slopes are from Elias et al.
(2001).
In this introductory chapter, we define geosynthetics and discuss what they are made of, how
they are made, and how they should be identified. Then we introduce you to the functions
and applications of geosynthetics, and we describe in some detail the methods used to
evaluate their engineering properties. Finally, we provide some general comments about
design, construction, and inspection that apply to all applications.
The remaining chapters of this manual provide specific details about the major application
categories. Each chapter provides a step-by-step systematic approach to design, a design
example, cost considerations, sample specifications, installation procedures, and inspection
suggestions. Proper attention to these details will ensure successful and cost-effective
geosynthetic designs and installations.

1.2 DEFINITIONS, MANUFACTURING PROCESSES, AND IDENTIFICATION


ASTM (2006) D 4439 defines a geosynthetic as a planar product manufactured from a
polymeric material used with soil, rock, earth, or other geotechnical-related material as an
integral part of a civil engineering project, structure, or system. The first to be developed and
most widely used geosynthetic is a geotextile, defined by ASTM as a permeable geosynthetic
comprised entirely of textiles. A number of other geosynthetics are available, including grids,
membranes, nets, meshes, webs, and composites; that are used in combination with or in
place of geotextiles. Geogrids are formed by a regular network of tensile elements with
apertures of sufficient size to interlock with surrounding fill material. Geogrids are primarily
used for reinforcement, geomembranes are low-permeability geosynthetics used as fluid
barriers. Geotextiles and other geosynthetics such as nets and grids can be combined with
geomembranes and other geosynthetics to provide the best attributes of each material. These
products are called geocomposites, and they include geotextile-geonets, geotextile-geogrids,
geotextile-geomembranes, geomembrane-geonets, geotextile-polymeric cores, and even
three-dimensional polymeric cell structures.
A convenient classification scheme for geosynthetics is provided in Figure 1-1. Most
geosynthetics are made from synthetic polymers, and of these, polypropylene, polyester, and
polyethylene are by far the most common. These polymers are normally highly resistant to
biological and chemical degradation. Less-frequently-used polymers include polyamides
FHWA NHI-07-092
Geosynthetics Engineering

1-2

Introduction, Identification, and Evaluation


August 2008

(e.g., nylon, which is not very durable in soil because it softens in the presence of water),
polyvinyl chloride (PVC), and glass fibers. Natural fibers such as cotton, jute, etc., could
also be used to make materials that are similar to geotextiles. Because these products are
biodegradable, they are only for temporary applications. Natural fiber geotextile-related
materials have not been widely utilized in the U.S. For additional information about the
polymeric composition of geosynthetics, see Koerner (2006).
In manufacturing geotextiles, basic elements such as fibers or yarns are combined into planar
textile structures. The fibers can be continuous filaments, which are very long thin strands of
a polymer, or staple fibers, which are short filaments, typically to 6 in. (20 to 150 mm)
long. Sometimes an extruded plastic sheet or film is slit to form thin, flat tapes. With both
continuous filaments and slit tapes, the extrusion or drawing process elongates the polymers
in the direction of the draw and increases the strength of the filament or tape. After the
drawing process, filaments and tapes may also be fibrillated, a process in which the filaments
are split into finer filaments by crimping, twisting, cutting or nipped with a pinned roller.
This process provides pliable, multifilament yarns with a more open structure that are easier
to weave.
Geotextile type is determined by the method used to combine the filaments or tapes into the
planar structure. The vast majority of geotextiles are either woven or nonwoven. Woven
geotextiles are made of monofilament, multifilament, or fibrillated yarns, or of slit film tapes.
The weaving process is as old as Homo Sapiens' have been making clothing and textiles.
Nonwoven textile manufacture is a modern development, a high-tech process industry, in
which synthetic polymer fibers or filaments are continuously extruded and spun, blown or
otherwise laid onto a moving conveyor belt. Then the mass of filaments or fibers are either
needlepunched, in which the filaments are mechanically entangled by a series of small
needles, or heat bonded, in which the individual fibers are welded together by heat and
pressure at their points of contact in the nonwoven mass.
Geogrids with integral junctions are manufactured by extruding and orienting sheets of
polyolefins (polyethylene or polypropylene). These types of geogrids are often called
extruded or integral geogrids. Geogrids may also be manufactured of multifilament
polyester yarns, joined at the crossover points by a knitting or weaving process, and then
encased with a polymer-based, plasticized coating. These types of geogrids are often called
woven or flexible geogrids. A third type, a welded geogrid manufactured, as the name
implies, by welding polymeric strips (e.g., strapping material) together at their cross over
points. All these manufacturing techniques allow geogrids to be oriented such that the
principal strength is in one direction, called uniaxial geogrids, or in both directions (but not
necessarily the same), called biaxial geogrids.
FHWA NHI-07-092
Geosynthetics Engineering

1-3

Introduction, Identification, and Evaluation


August 2008

The manufacture of other geosynthetic products is as varied as the products themselves.


Geonets, geosynthetic erosion mats, geowebs, geomeshes, etc., can be made from large and
often rather stiff filaments formed into a mesh with integral junctions or they can be welded
or glued at the crossover points. Manufacture of geomembranes and other geosynthetic
barriers is discussed in Chapter 10.
Geocomposites result when two or more geosynthetics are combined in the manufacturing
process. Most geocomposites are used in drainage applications and waste containment. A
common example of a geocomposite is a prefabricated drain that consists of a fluted or
dimpled polymeric sheet, which acts as a conduit for water, wrapped with a geotextile that
acts as a filter.
TEXTILES
SYNTHETIC

WEBBINGS
NATURAL

Various polymers
Polypropylene
Polyethylene
Polyester, etc.

IMPERMEABLE

SYNTHETIC

Cotton
Jute
Reeds
Grass

Steel
Polymers

PERMEABLE

CLOSE-MESH

NATURAL
Palm wood
Wood
Bamboo

OPEN MESH

Nets
Mats

Geomembrane polymers:
Polyethylene (HDPE, LLDPE, etc.)
Polyvinyl Chloride (PVC)
Cholosulphonated Polyethylene (CSPE)
Ethylene Interpolymer Alloy (EIA)
Rubber, etc.

Geogrids
Bar mats

Combination
Products

SHEETS

STRIPS

Formed Plastic
with pins, etc.

Reinforced Earth
York System

GEOTEXTILES
NONWOVEN

KNITTED

Continuous Filament
Staple Filament

NEEDLEPUNCHED

WOVEN

Combination Products
(Geocomposites)

CHEMICAL
BONDED
Wet Laid
Resin Bonded

MONOFILAMENT
YARNS

HEATBONDED
Spunbonded

SLIT FILM
YARNS

FIBRILLATED
YARNS

Noncalendered
Calendered

MULTIFILAMENT
YARNS
Noncalendered
Calendered

Figure 1-1. Classification of geosynthetics and other soil inclusions (modified after
Rankilor, 1981).
FHWA NHI-07-092
Geosynthetics Engineering

1-4

Introduction, Identification, and Evaluation


August 2008

Geosynthetics are generically identified by:


1. polymer (descriptive terms, e.g., high density, low density, etc. should be included);
2. type of element (e.g., filament, yarn, tape, strand, rib, coated rib), if appropriate;
3. distinctive manufacturing process (e.g., woven, needlepunched nonwoven,
heatbonded nonwoven, stitchbonded, extruded, knitted, welded, uniaxial, biaxial,
roughened sheet, smooth sheet), if appropriate;
4. primary type of geosynthetic (e.g., geotextile, geogrid, geomembrane, etc.);
5. mass per unit area, if appropriate (e.g., for geotextiles, geogrids, GCLs, erosion
control blankets,) and/or thickness, if appropriate (e.g., for geomembranes); and
6. any additional information or physical properties necessary to describe the material
in relation to specific applications.
Four examples are:
polypropylene staple filament needlepunched nonwoven geotextile, 10 oz/yd2 (340
g/m2);
polyethylene geonet, 200 mil (5 mm) thick;
polypropylene extruded biaxial geogrid, with 1 in. x 1 in. (25 mm x 25 mm)
openings; and
high-density polyethylene roughened sheet geomembrane, 60 mil (1.5 mm) thick.

1.3 FUNCTIONS AND APPLICATIONS


Geosynthetics have six primary functions:
1. filtration
2. drainage
3. separation
4. reinforcement
5. fluid barrier, and
6. protection
Geosynthetic applications are usually defined by their primary function. For example,
geotextiles are used as filters to prevent soils from migrating into drainage aggregate or
pipes, while maintaining water flow through the system. They are similarly used below
riprap and other armor materials in coastal and stream bank protection systems to prevent soil
erosion.
Nonwoven needlepunched geotextiles and geocomposites can also provide drainage, by
allowing water to drain from or through low permeability soils. Geotextile applications
FHWA NHI-07-092
Geosynthetics Engineering

1-5

Introduction, Identification, and Evaluation


August 2008

include dissipation of pore water pressures at the base of roadway embankments. For
situations with higher flow requirements, for example, pavement edge drains, slope
interceptor drains, and retaining wall drains, geocomposite drains are often used. Filtration,
drainage, and erosion control are addressed in Chapters 2, 3, and 4.
Geotextiles are often used as separators to prevent road base materials from penetrating into
the underlying soft subgrade, thus maintaining the design thickness and roadway integrity.
Separators also prevent fine-grained subgrade soils from being pumped into permeable,
granular road bases. Separators are discussed in Chapter 5.
Both geotextiles and geogrids can be used as reinforcement to add tensile strength to a soil
matrix, thereby providing a more competent and stable material. Reinforcement enables
embankments to be constructed over very soft foundations and permits the construction of
steep slopes and retaining walls. Reinforcement applications are presented in Chapters 7, 8,
and 9. Geogrids and geotextiles can also be used as reinforcement in roadway base and
subbase aggregate layers to improve the performance of pavement systems as discussed in
Chapter 5.
Geomembranes, thin-film geotextile composites, geosynthetic clay liners, and field-coated
geotextiles are used as fluid barriers to impede the flow of a liquid or gas from one location
to another. This geosynthetic function has wide application in asphalt pavement overlays,
encapsulation of swelling soils, and waste containment. Pavement overlays are discussed in
Chapter 6. Geomembranes and other geosynthetic barriers are described in Chapter 10.
The sixth function is, protection, in which the geosynthetic acts as a stress relief layer.
Temporary geosynthetic blankets and permanent geosynthetic mats are placed over the soil to
reduce erosion caused by rainfall impact and water flow shear stress. A protective cushion of
nonwoven geotextiles is often used to prevent puncture of geomembranes (by reducing point
stresses) from stones in the adjacent soil or drainage aggregate during installation and while
in service as discussed in Chapter 10. Geotextiles also provide stress relief to retard the
development of reflection cracks in pavement overlays as discussed in Chapter 6
In addition to the primary function, geosynthetics usually perform one or more secondary
functions. The primary and secondary functions make up the total contribution of the
geosynthetic to a particular application. A listing of common applications according to
primary and secondary functions is presented in Table 1-1. Secondary functions can be
equally important as the primary function, and in order to obtain optimum geosynthetic
performance, both much be considered in the design computations and specifications.

FHWA NHI-07-092
Geosynthetics Engineering

1-6

Introduction, Identification, and Evaluation


August 2008

Table 1-1
Representative Applications and
Controlling Functions of Geosynthetics

PRIMARY
FUNCTION

APPLICATION

SECONDARY FUNCTION(S)

Unpaved Roads (temporary &


permanent)
Paved Roads (secondary & primary)
Construction Access Roads
Working Platforms
Railroads (new construction)
Railroads (rehabilitation)
Landfill Covers
Preloading (stabilization)
Marine Causeways
General Fill Areas
Paved & Unpaved Parking Facilities
Cattle Corrals
Coastal & River Protection
Sports Fields

Filter, drains, reinforcement

Filter

Trench Drains
Pipe Wrapping
Base Course Drains
Frost Protection
Structural Drains
Toe Drains in Dams
High Embankments
Filter Below Fabric-Form
Silt Fences
Silt Screens
Culvert Outlets
Reverse Filters for Erosion Control:
Seeding and Mulching
Beneath Gabions
Ditch Armoring
Embankment Protection, Coastal
Embankment Protection, Rivers
& Streams
Embankment Protection, Lakes
Vertical Drains (wicks)

Separation, drains
Separation, drains, protection
Separation, drains
Separation, drainage, reinforcement
Separation, drains
Separation, drains
Drains
Separation, drains
Separation, drains
Separation
Separation
Separation

Drainage-Transmission

Retaining Walls
Vertical Drains
Horizontal Drains
Below Membranes (drainage of gas and
water)
Earth Dams
Below Concrete (decking & slabs)

Separation, filter
Separation, filter
Reinforcement
Reinforcement, protection

Separation

FHWA NHI-07-092
Geosynthetics Engineering

1-7

Filter, drains
Filter, drains, reinforcement
Filter, drains, reinforcement
Filter, drains, reinforcement
Filter, drains, reinforcement
Reinforcement, drains, protection
Reinforcement, drains
Filter, drains, reinforcement
Filter, drains, reinforcement
Filter, drains, reinforcement
Filter, drains, reinforcement
Filter, drains, reinforcement
Filter, drains, protection

Filter
Protection

Introduction, Identification, and Evaluation


August 2008

Table 1-1
Representative Applications and
Controlling Functions of Geosynthetics
(continued)

PRIMARY
FUNCTION
Reinforcement

SECONDARY FUNCTION(S)

APPLICATION

Pavement Overlays
Subbase Reinforcement in Roadways &
Retaining Structures
Membrane Support
Embankment Reinforcement
Fill Reinforcement
Foundation Support
Soil Encapsulation
Net Against Rockfalls
Fabric Retention Systems
Sand Bags
Reinforcement of Membranes
Load Redistribution
Bridging Nonuniformity Soft Soil Areas
Encapsulated Hydraulic Fills
Bridge Piles for Fill Placement

---------R Filter
Drains
Separation, drains, filter, protection
Drains
Drains
Drains
Drains, filter, separation
Drains
Drains
---------Protection
Separation
Separation
Separation
----------

Fluid Barrier

Asphalt Pavement Overlays


Liners for Canals and Reservoirs
Liners for Landfills and Waste
Repositories
Covers for Landfill and Waste
Repositories
Cutoff Walls for Seepage Control
Waterproofing Tunnels
Facing for Dams
Membrane Encapsulated Soil Layers
Expansive Soils
Flexible Formwork

Protection
----------------------------------------------------------------------------------------------------

Protection

Geomembrane Cushion
Asphalt Overlay
Temporary Erosion Control
Permanent Erosion Control

Drains
Fluid barrier
Fluid barrier
Reinforcement, fluid barrier

FHWA NHI-07-092
Geosynthetics Engineering

1-8

Introduction, Identification, and Evaluation


August 2008

1.4 DESIGN APPROACH


Considering the wide variety of geosynthetics available, engineering based on the specific
project conditions and constraints is required in order to obtain the most suitable material for
any application. We recommend the following approach to designing with geosynthetics:
1. Define the purpose and establish the scope of the project.
2. Investigate and establish the geotechnical conditions at the site (geology, subsurface
exploration, laboratory and field testing, etc.).
3. Establish application criticality, severity, and performance criteria. Determine
external factors that may influence the geosynthetic's performance.
4. Formulate trial designs and compare several alternatives.
5. Establish the models to be analyzed, determine the parameters, and carry out the
analysis.
6. Compare results and select the most appropriate design; consider alternatives versus
cost, construction feasibility, etc. Modify the design if necessary.
7. Prepare detailed plans and specifications including: a) specific property requirements
for the geosynthetic; and b) detailed installation procedures.
8. Hold preconstruction meeting with contractor and inspectors.
9. Approve geosynthetic on the basis of specimens' laboratory test results and/or
manufacturer's certification.
10. Monitor construction.
11. Inspect after major events (e.g., 100 year rainfall or an earthquake) that may
compromise system performance.
By following this systematic approach to design and installation of geosynthetics, costeffective designs can be achieved, along with improved performance, increased service life,
and reduced maintenance costs. Good communication and interaction between all concerned
parties is imperative throughout the design and selection process.

1.5 EVALUATION OF PROPERTIES


Today, there are more than 600 different geosynthetic products available in North America.
Because of the wide variety of products, with their different polymers, filaments, weaving (or
nonwoven) patterns, bonding mechanisms, thicknesses, masses, etc., they have a
considerable range of physical and mechanical properties. Thus, the process of comparison
and selection of geosynthetics is not easy. Geosynthetic testing has progressed significantly
since the FHWA Geotextile Engineering Manual (Christopher and Holtz, 1985) was
FHWA NHI-07-092
Geosynthetics Engineering

1-9

Introduction, Identification, and Evaluation


August 2008

published. Specific test procedures for most geosynthetic properties can be found in ASTM
(2006). These procedures have been developed by ASTM Committee D 35 on Geosynthetics
during the past 20 years or so. Because ASTM standards are consensus standards, the
process is often slow, and to help speed up the process, the Geosynthetics Research Institute
(GRI) of Drexel University has issued interim standards for a number of tests. They are only
active until an equivalent ASTM standard is adopted. ASTM (2006) and GRI (2006)
standards are listed in Appendix E. Note that test procedures referred to in the AASHTO
standard Geotextile Specification for Highway Applications, Designation M 288, are
primarily ASTM standard test procedures.
The particular, required design properties of the geosynthetic will depend on the specific
application and the associated function(s) the geosynthetic is to provide. The properties
listed in Table 1-2 cover the range of important criteria and properties required to evaluate a
geosynthetic for most applications in this manual. It should be noted that not all of the listed
requirements will be necessary for all applications. Typically only six to eight properties are
required for a specific application. Also note that in Table 1-2, properties required for
mechanical or hydraulic design are different than those required for constructability
(sometimes called survivability) and longevity or durability.
Table 1-3 lists all the geosynthetics applications included in this manual along with their
associated functions. Use Table 1-3 along with Tables 1-1 and 1-2 to determine the
appropriate properties for each application.
All current geosynthetic properties and parameters are listed in Table 1-4, along with the
ASTM or GRI test procedures for each property and their preferred units of measurement.
All geosynthetic properties can be placed into three basic categories: general, index, and
performance properties. General properties, given in Table 1-4, are usually provided by the
manufacturers or their distributors. Another source of general properties is the Specifier's
Guide published each December in the Geosynthetics magazine (formerly Geotechnical
Fabrics Report), published by the Industrial Fabrics Association International (IFAI). In
addition to general and some index properties for most product types and manufacturers, the
Specifier's Guide also contains a directory of manufacturers, distributors, installers, design
engineers, and testing laboratories. Contact information and web addresses are also
provided.

FHWA NHI-07-092
Geosynthetics Engineering

1-10

Introduction, Identification, and Evaluation


August 2008

Table 1-4 also lists index tests and performance tests. Index tests were originally developed
by manufacturers for quality control purposes, and as the name implies, they give only an
indication or a qualitative assessment of the property of interest. With some exceptions,
index test values are not appropriate for design, although when determined using standard
test procedures, index properties can be used for product comparison, procurement
specifications, and quality control of construction and installation.

Table 1-2
Important Criteria and Principal
Properties Required for Evaluation of Geosynthetics
PROPERTY1

CRITERIA AND
PARAMETER
Design Requirements:

FUNCTION
Filtration

Drainage

Separation

Reinforcement

Barrier

Protection

Wide Width Strength


Wide Width Modulus
Wide Width Strength
Creep Resistance
Creep Resistance
Shear Strength

92

9
9
9
9

9
9
9
9

Permeability
Transmissivity
Apparent Opening Size
Porimetry
Gradient Ratio or LongTerm Flow

9
9
9

9
9

9
9
9
9

Grab Strength
Grab Strength
Burst Strength
Rod or Pyramid
Puncture
Trapezoidal Tear

9
9
9
9

9
9
9
9

9
9
9
9

9
9

9
9
9
9

9
9

Reciprocating Block
Abrasion
UV Resistance
Chemical
Biological
Wet-Dry
Freeze-Thaw

9
9
9
9
9

9
9
9
9

?
?

9
9
9

9
9
9

9
?
?
9

Mechanical Strength
Tensile Strength
Tensile Modulus
Seam Strength
Tension Creep
Compression Creep
Soil-Geosynthetic
Friction
Hydraulic
Flow Capacity
Piping Resistance
Clogging Resistance
Constructability
Requirements:
Tensile Strength
Seam Strength
Bursting Resistance
Puncture Resistance
Tear Resistance
Longevity (Durability):
Abrasion Resistance3
UV Stability4
Soil Environment5

NOTES
1.
See Table 1-4 for specific procedures.
2.
Compression creep is applicable to some geocomposites.
3.
Erosion control applications where armor stone may move.
4.
Exposed geosynthetics only.
5.
Where required.

FHWA NHI-07-092
Geosynthetics Engineering

1-11

Introduction, Identification, and Evaluation


August 2008

Table 1-3
Evaluation of Geosynthetic Property Requirements
Functions
Chapter

Application

Filter

Subsurface Drainage

Prefabricated Drains

Drainage

Separation

Hard Armor Erosion Control

Silt Fence

Subgrade Stabilization

Reinforced Asphalt Overlay


Embankments Over Soft Subgrade

Reinforced Slopes

Reinforced Soil Walls

10

Containment Liners

Asphalt Overlay Stress Relief layer

Protection

Base/Subbase Reinforcement

Barrier

9
;

Subgrade Separation

Reinforcement

;
9

;
;

;
9

; indicates Primary Function


9 indicates Secondary Functions

Table 1-4 Geosynthetic Properties and Parameters


PROPERTY
I. GENERAL PROPERTIES (from manufacturers)
Type and Construction
Polymer
Mass per Unit Area
Thickness (geotextiles and geomembranes)
Roll Length
Roll Width
Roll Weight
Roll Diameter
Specific Gravity & Density
Surface Characteristics

TEST METHOD

UNITS OF MEASUREMENT

N/A
N/A
ASTM D 5261
ASTM D 5199
Measure
Measure
Measure
Measure
ASTM D 792 and D 1505
N/A

--------g/m2
mm
m
m
kg
m
g/m3
-----

ASTM D 4632
ASTM D 6637

N
N

ASTM D 412
ASTM D 638
ASTM D 882

N
N
N

ASTM D 4595
ASTM D 6637
ASTM D 4885
ASTM D 882

N
N
N
N

GRI:GG2

no standard

-----

ASTM D 5262 (Note:


interpretation required)

creep strain: %/hr


creep rupture: kN/m

II. INDEX PROPERTIES


MECHANICAL STRENGTH UNIAXIAL LOADING
a) Tensile Strength (Quality Control)
1) Grab Strength (geotextiles and CSPE reinforced geomembranes)
2) Single Rib Strength (geogrids)
3) Narrow Strip (geomembranes)
- EDPM, CO, IR, CR
- HDPE
- PVC, VLDPE
b) Tensile Strength (Load-Strain Characteristics)
1) Wide-Width Strip (geotextiles)
2) Single or Multi-Rib (geogrids)
3) Wide Strip Strength (geomembranes)
4) 2% Secant Modulus (PE geomembranes)
c) Junction Strength (geogrids)
d) Dynamic and Cyclic Resistance
e) Creep Resistance

FHWA NHI-07-092
Geosynthetics Engineering

1-12

Introduction, Identification, and Evaluation


August 2008

Table 1-4 Geosynthetic Properties and Parameters

(continued)

MECHANICAL STRENGTH UNIAXIAL LOADING (cont.)


PROPERTY

TEST METHOD

UNITS OF MEASUREMENT

GRI:GS7

dimensionless

g) Seam Strength
1) Sewn (geotextiles)
2) Factory Peel and Shear (geomembranes)
3) Field Peel and Shear (geomembranes)

ASTM D 4884
ASTM D 4545
ASTM D 4437

N
N/m
N/m

h) Tear Strength
1) Trapezoid Tearing (geotextiles)
2) Tear Resistance (geomembranes)

ASTM D 4533
ASTM D 1004

N
N

ASTM D 3786
ASTM D 6241

Pa
Pa or N

ASTM D 5514

Pa

ASTM D 4833
ASTM D 5494

N
N

ASTM D 6241

c) Penetration Resistance (Dimensional Stability)

No standard

-----

d) Geosynthetic Cutting Resistance

No standard

-----

ASTM D 1388

Mg/cm2

a) Selecting Test Methods for Experimental Evaluation of


Geosynthetic Durability

ASTM D 5819

--

b) Abrasion Resistance (geotextiles)

ASTM D 4886

c) Ultraviolet (UV) Radiation Stability


1) Xenon-Arc Apparatus (geotextiles)
2) Outdoor Exposure (geotextiles)

ASTM D 4355
ASTM D 5970

%
%

d) Chemical Resistance
1) Geotextiles
2) Geogrids
3) Geomembranes
4) Geonets
5) Chemical ImmersionLaboratory
6) Oxidative Induction Time
7) Environmental Exposure

ASTM D 6389
ASTM D 6213
ASTM D 5747
ASTM D 5288
ASTM D 5322
ASTM D 5885
EPA 9090

% change
% change
% change
% change
temperature & time
minutes
N/A

ASTM D 1987
ASTM G 21 and G 22
ASTM D 3083

m3/s
----% change

No standard

-----

ASTM D 4594
ASTM D 1204

% change
% change

f) Index Friction

MECHANICAL STRENGTH RUPTURE RESISTANCE


a) Burst Strength
1) Mullen Burst (geotextiles)
2) Static Puncture with 50-mm CBR Probe (geotextiles,
geonets, geomembranes)
3) Large Scale Hydrostatic (geomembranes and
geotextiles)
b) Puncture Resistance
1) Index (geotextiles and geomembranes)
2) Pyramid Puncture (geomembranes)
3) Static Puncture with 50-mm CBR Probe
(geotextile, geonets, and geomembranes)

e) Flexibility (Stiffness)
ENDURANCE PROPERTIES

e) Biological Resistance
1) Biological Clogging (geotextile)
2) Biological Degradation
3) Soil Burial
f) Wet and Dry Stability
g) Temperature Stability
1) Temperature Stability (geotextile)
2) Dimensional Stability (geomembrane)

FHWA NHI-07-092
Geosynthetics Engineering

1-13

Introduction, Identification, and Evaluation


August 2008

Table 1-4 Geosynthetic Properties and Parameters

(continued)

II. INDEX PROPERTIES (continued)


PROPERTY
HYDRAULIC
a) Opening Characteristics (geotextiles)
1) Apparent Opening Size (AOS)
2) Porimetry (pore size distribution)Capillary flow or
bubble point test
3) Percent Open Area (POA)
4) Porosity (n)
b) Permeability (k) and Permittivity ( )
c) Soil Retention Ability
d) In-Plane Flow Capacity (Transmissivity, )

TEST METHOD

UNITS OF MEASUREMENT

ASTM D 4751
ASTM D 6767

mm
mm or m

See Christopher & Holtz (1985)


(Vvoids/Vtotal) 100
ASTM D 4491
Empirically Related to
Opening Characteristics
ASTM D 4716

%
%
m/s and s-1
----m2/s

III. PERFORMANCE PROPERTIES


a) Stress-Strain Characteristics:
1) Tension Test in Soil
2) Triaxial Test Method
3) CBR on Soil Fabric System
4) Tension Test in Shear Box
5) Plane Strain In-Soil Device

kN/m and % strain


See Elias et al. (1998)
See Holtz et al.. (1982)
See Christopher & Holtz (1985)
See Christopher & Holtz, (1985)
See Boyle (1995) and
Boyle et al. (1996)

b) Creep Tests:
1) Extension Test in Soil
2) Triaxial Test Method
3) Extension Test in Shear Box
4) Pullout Method

kN/m and % strain


See Elias et al. (1998)
See Holtz et al. (1982)
See Christopher & Holtz (1985)
ASTM D 6706*

c) Friction/Adhesion:
1) Direct Shear (soil-geosynthetic)
2) Direct Shear (geosynthetic-geosynthetic)
3) Pullout Resistance (geogrids)
4) Pullout Resistance (geotextiles)
5) Anchorage Embedment (geomembranes)

ASTM D 5321*
ASTM D 5321*
ASTM D 6706*
ASTM D 6706*
GRI:GM2

degrees ()
degrees ()
dimensionless
dimensionless
kN/m

no standard procedures

N/A

e) Puncture
1) Gravel, truncated cone or pyramid

ASTM D 5494

kPa

f) Chemical Resistance:
1) In-Situ Immersion Testing

ASTM D 5496

N/A

ASTM D 5101

Dimensionless

See Harney and Holtz (2001)


and Harney et al. (2007)
ASTM D 5567
ASTM D 5141

Dimensionless

d) Dynamic and Cyclic Resistance:

g) Soil Retention and Filtration Properties:


1) Gradient Ratio Method for noncohesive sand and
silt type soils
2) Flexible Wall Gradient Ratio Test for fine-grained soils
3) Hydraulic Conductivity Ratio (HCR) - for fine-grained soils
4) Slurry Method for silt fence applications

Dimensionless
%

NOTES:
* -- Interpretation required.
N/A not available or not applicable.

FHWA NHI-07-092
Geosynthetics Engineering

1-14

Introduction, Identification, and Evaluation


August 2008

Performance tests are an attempt to model the soil-geosynthetic interaction. Thus they
require the geosynthetic to be tested together with a sample of the on-site soil in order to
obtain a direct assessment of the property of interest. Since performance tests are
performed under specific design conditions with soils from the construction site,
manufacturers should not be expected to have the capability or the responsibility to
perform such tests. These tests should be performed under the direction of the design
engineer. Performance tests properties are not normally used in specifications; rather, the
geosynthetic are preselected for performance testing based on index values. Sometimes,
performance test results are correlated to index values for use in specifications. How
performance tests can be used in practice is discussed in the section on Specifications later in
this chapter.
Brief descriptions of some of the basic properties of geosynthetics and their tests are
presented below (after Christopher and Dahlstrand, 1989). Note G = general property; I =
index property, and P = performance property.
Mass per Unit Area (G): Area is used as opposed to volume due to variations in thickness
under normal stress. This property is mainly used to identify different materials. See ASTM
D 5261.
Thickness (G): Thickness is not usually required information for geotextiles except in
permeability and hydraulic flow calculations. It may be used for product comparison of
geotextiles or as a primary identifier for geomembranes. When needed, it can be simply
obtained using the procedure in ASTM D 5199. The nominal thickness is measured under a
normal stress of 2 kPa (0.29 psi) for geotextiles and 20 kPa (2.9 psi) for smooth
geomembranes, geonets, etc.
Tensile Strength (I): To understand the load-strain characteristics, it is important to consider
the complete load-strain curve. It is also important to consider the nature of the test and the
testing environment. With most materials, strength and modulus have units of stress.
However, because of the thin, two-dimensional nature of geosynthetics, stress would be
awkward to measure. Therefore, it is conventional with geosynthetics to use force per unit
length measured along the width of the material. Then strength and modulus have units of
FL-1 (i.e., kN/m).
There are several types of uniaxial tensile strength tests. Details about the specific
geosynthetic specimen shapes, clamping devices, and loading rates are given in the ASTM
standards referenced in Table 1-4. The tests all give different results that can only be
compared qualitatively.
FHWA NHI-07-092
Geosynthetics Engineering

1-15

Introduction, Identification, and Evaluation


August 2008

The grab tensile test (ASTM D 4632) is an unusual test, and although it is widely used, its
results are almost universally misused. The grab tensile test normally uses 1-in. (25 mm)
wide jaw clamps to grip a 4-in. (100 mm) wide specimen. The strength is reported as the
total force needed to cause failurenot the force per unit width. It is not obvious how the
tensile force is distributed across the specimen because it is wider than the clamps. The
effect of this difference will depend on the interaction of the geotextile filaments. In
nonwoven geotextiles, these effects are large, but in wovens, they are probably small. The
grab test is useful as an index test, for specifications, and for product comparison. But it is
difficult, if not impossible, to relate grab results to actual tensile strength unless they are
directly correlated with, for example, wide-width tensile tests.
The single rib tensile test (ASTM 6637 Method A) is also an index test because it does not
provide an evaluation of the uniformity of load distribution across a number of elements.
Non-uniformly aligned grid elements could lead to much lower strength values per width of
geogrid than indicated by single rib tests. Therefore single element tests should not be used
to evaluate the strength of geogrids unless the results have been correlated with multi-rib
tensile test results.
Plane-strain represents the loading condition for many practical applications. However,
because there is no standard plane strain test or procedure, in practice, plane strain loading is
approximated by a wide-width strip tensile test (ASTM D 4595 for geotextiles and D 6637
for geogrids). Since narrow strip geosynthetic specimens usually neck when strained, widewidth strip tensile tests are performed on short wide specimens (length to width ratio 2:1).
Although the wide-width test is really an index test, it is commonly used to approximate the
in-soil geosynthetic strength in soil reinforcing applications.
Geosynthetics may have different strengths in different directions. Therefore, tests should be
conducted in both principal directions, and the results clearly stated as to direction of testing
(whether machine direction and/or cross-machine direction). In addition to specimen length
to width ratio, clamping arrangements, and direction of loading, tensile load-strain tests are
influenced by rate of loading, temperature, moisture, lateral restraint, and confinement.
Creep is a time-dependent mechanical property. It is the strain that occurs at constant load.
Creep tests can be run using any type the tensile test, but they are most frequently performed
on a wide-width specimen by applying a constant load for a sustained period. Creep tests are
influenced by the same factors as tensile load-strain tests - specimen length to width ratio,
temperature, moisture, lateral restraint, and confinement.

FHWA NHI-07-092
Geosynthetics Engineering

1-16

Introduction, Identification, and Evaluation


August 2008

Short-term creep strain is strongly influenced by the geosynthetic structure. Geogrids and
woven geotextiles have the least creep, heat-bonded geotextiles are intermediate, and needle
punched geotextiles have the most creep. Longer-term creep rates are controlled by structure
and polymer type. Of the most common polymers, polyester has lower creep rates than
polypropylene.
The creep limit is probably the most important geosynthetic creep characteristic. It is the
sustained load per unit width above which the geosynthetic will creep to rupture. Just as with
creep rates, the creep limit is controlled by the polymer type. It ranges from 20% of the
short-term ultimate strength of low tenacity polypropylene geosynthetics to 60% for
polyesters geosynthetics.
Rupture Resistance: The burst test is performed by applying a normal pressure (by a solid
CBR piston, air or hydraulic fluid) against a geosynthetic specimen clamped in a ring.
Although the burst strength is given in units of pressure, it is not the real stress in the
geosynthetic, but rather it is the normal stress acting on the geosynthetic at failure. Burst
strength is a function of the diameter of the test specimen; therefore, care must be used in
comparing test results on different materials. Because the test pressure is applied to the
specimen in all directions, the ultimate value is controlled by the weakest direction.
Friction: Soil-geosynthetic and geosynthetic-geosynthetic friction are important properties
for many applications. It is common to assume a soil-geosynthetic friction value between 2/3
and one times the soil friction angle. Caution is advised for geomembranes where soilgeosynthetic friction angle may be much lower than the soil friction angle. For important
applications, tests should be run on samples of the proposed geosynthetics and on-site soils.
The direct shear friction test is simple in principle, but numerous details must be considered
for accurate results. The equipment is quite large in order to reduce boundary effects.
According to ASTM D 5321 the minimum shear box size is 12-in. by 12-in. (300 by 300
mm). For many geosynthetics, the measured friction angle depends on the types of soils on
each side of the geosynthetic specimen, as well as on the normal stress; therefore, test
conditions must model the actual field conditions. Since soil is involved, this test is
obviously a performance test.
Durability Properties: In most geosynthetics applications, durability and longevity must be
considered. Exposure to ultraviolet light can weaken and degrade many geosynthetics. The
geosynthetic polymer must be compatible with the chemistry of the environment. The soil
and groundwater should be checked for such items as high and low pH, chlorides, organics,
and oxidation agents. Ferruginous soils (those containing Fe2SO3), calcareous soils, and acid
FHWA NHI-07-092
Geosynthetics Engineering

1-17

Introduction, Identification, and Evaluation


August 2008

sulfate soils can be especially detrimental to geosynthetics over time. Other detrimental
environmental factors include chemical solvents and well as gasoline, diesel, and other fuels.
Each geosynthetic polymer is different in terms of its resistance to aging and attack by
chemical and biological agents. Therefore, each product must be investigated individually to
determine the effects of these durability factors. The best source of this information is
usually the geosynthetic manufacturer. They are usually able to supply the results of product
exposure studies, including, but not limited to, strength reduction due to aging, deterioration
in ultraviolet light, chemical attack, microbiological attack, environmental stress cracking,
hydrolysis, and any possible synergism between individual factors.
Guidelines on soil environments and on geosynthetics properties are presented in the FHWA
Corrosion/Degradation of Soil Reinforcements for Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes (Elias, 1997). This research has been summarized and recommended
aging reduction factorsbasically safety factorsfor soil reinforcement applications is
presented in FHWA (1997)(see Appendix H). With supporting data, a durability reduction
factor as low as 1.1 may be used.
Hydraulic Properties: The hydraulic properties of geosynthetics include their opening
characteristics, hydraulic conductivity (permeability and permittivity), soil retention ability,
clogging potential, and in some cases, their transmissivity. These are all index properties, and
basically they relate to the pore sizes in the geosynthetic. Hydraulic properties may also be
affected by chemical and biological agents. Precipitates as well as slime growth have been
known to clog filter systems (granular filters as well as geotextiles).
The ability of a geotextile to retain soil particles is directly related to its apparent opening
size (AOS) which is the largest hole in the geotextile. The AOS value is equal to the size of
the largest particle that can effectively pass through the geotextile in a dry sieving test
(ASTM D 4751). Another method of obtaining the opening characteristics, including the
AOS value, is the Capillary Flow or Bubble Point Method (ASTM D 6767). This procedure
allows for an evaluation of the complete pore size distribution (PSD), although as indicated
in the ASTM standard, it my not be applicable to more open woven geotextiles with a pore
size of greater than 200 m.
The ability of water to pass through a geotextile is determined from its hydraulic conductivity
(coefficient of permeability, k), as measured in a permeability test. The flow capacity of the
material can then be determined from Darcy's law. Due to the compressibility of geotextiles,
the permittivity (permeability divided by thickness) is usually determined from the test
(ASTM D 4491) and used to directly evaluate flow capacity. Permeability and permittivity
are index properties. The ability of water to pass through a geotextile during the entire life of
FHWA NHI-07-092
Geosynthetics Engineering

1-18

Introduction, Identification, and Evaluation


August 2008

the project is dependent on its filtration potential, that is, its ability not to clog with soil
particles. Essentially, if the finer particles of soil can pass through the geotextile, it should
not clog. Effective filtration can be evaluated through relations between the geotextile's pore
size distribution and the soil's grain size distribution; however, such formulations are still in
the development phase. For a precise evaluation, laboratory performance testing of the
proposed soil and candidate geotextile should be conducted.
For sandy and silty soils (k 10-7 m/s), the only standard filtration test is the gradient ratio
test (ASTM D 5101). In this test, a rigid wall permeameter with strategically located
piezometer ports is used to measure the head loss in the soil alone to the head loss at the soilgeotextile interface under different hydraulic gradients. The ratio of these two gradients is the
gradient ratio. Although ASTM indicates that the test may be terminated after 24 hours, to
obtain meaningful results, the test should be continued until the flow rate has clearly
stabilized. This may occur within 24 hours, but could require several weeks, especially if
significant fines are present in the soil.
A gradient ratio of one or less is preferred. Less than one is an indication that fine soil
particles have passed the filter and that a more open filter bridge has developed in the soil
adjacent to the geotextile. However, a continued decrease in the gradient ratio below one
indicates piping, and an alternate geotextile should be evaluated. On the other hand, a high
gradient ratio indicates that a flow reduction has occurred in the geotextile, most likely due to
geotextile clogging. If the gradient ratio approaches three (the recommended maximum by
the U.S. Army Corps of Engineers, 1977), the flow rate through the system should be
carefully evaluated with respect to the design and system performance requirements. A
continued increase in the gradient ratio indicates clogging, and the geotextile is unacceptable.
For fine-grained soils, the hydraulic conductivity ratio (HCR) test (ASTM D 5567) should be
considered. This test uses a flexible wall permeameter and evaluates the long-term
permeability under increasing gradients with respect to the short-term permeability of the
system at the lowest hydraulic gradient. A decrease in HCR indicates a flow reduction in the
system. Since measurements are not taken near the geotextile-soil interface and soil
permeability is not measured, it is questionable whether an HCR decrease is the result of
flow reduction at the geotextile or blinding within the soil matrix itself. An improvement to
this method would be to include piezometer or transducers within these zones (after the
gradient ratio method) to aid in interpretation of the results. In addition, it is very difficult to
ensure that the specimen is fully saturated without backpressure.
Other filtration tests for clogging potential include the Caltrans slurry filtration test (Hoover,
1982), which was developed by Legge (1990) into the Fine Fraction filtration (F3) test
FHWA NHI-07-092
Geosynthetics Engineering

1-19

Introduction, Identification, and Evaluation


August 2008

(Sansone and Koerner, 1992), and the Long-Term Flow (LTF) test (Koerner and Ko, 1982;
GRI Test Method GT1, 1993). According to Fischer (1994), all of these tests have serious
disadvantages that make them less suitable than the Gradient Ratio (GR) test for determining
the filtration behavior of the soil-geotextile system. The GR test typically must be run longer
than the ASTM-specified 24 hours, and proper attention must be paid to the test details
(Mar, 1994) to get reproducible results.
A recent development is the Flexible Wall GR test (Harney and Holtz, 2001, Bailey et al.,
2005, and Harney et al., 2007). This test combines the best features of the GR test (D 5101)
and the flexible wall permeability test (D 5084). Just as with the GR test, multiple ports are
placed along the soil column to accurately determine head losses. Application of back
pressure ensures that the specimens are 100% saturated. Research indicated that the FWGR
yielded consistent and accurate results, and in significantly less time than the GR, for all
geotextiles tested with fine-grained soils. Preliminary indications of the steady-state
filtration behavior can be obtained in less than 24 hours and all the FWGR tests achieved
constant filtration behavior within five days. The HCR yielded inconclusive and inaccurate
results for most of the soils and geotextiles tested. The GR test can still be used for filtration
testing of coarse-grained soils, but if they contain even a few percent of fines or for finegrained soils, the FWGR is the preferred test.
Additional hydraulic properties that may be required in filtration design are the Percent Open
Area (POA) and the porosity (n). As noted in Table 1-4, there are no standard tests for these
properties, although there is a suggested procedure for POA given by Christopher and Holtz
(1985) and which follows Corps of Engineers procedures. Basically, POA is determined on a
light table or by projection enlargement. Porosity is readily calculated just as it is with soils;
that is, porosity is the volume of the voids divided by the total volume. The total volume is,
for example, 1 m2 times the nominal thickness of the geotextile. The volume of voids is the
total volume minus the volume of the fibers and filaments (solids), or the mass of 1 m2
divided by the specific gravity of the polymer.
National Transportation Product Evaluation Program (NTPEP). Four times a year the
NTPEP evaluates geotextiles and geogrids in accordance with AASHTO M 288. Two state
laboratories conduct the tests. On their website (http://data.ntpep.org/home/index.asp), the
section on Geotextiles And Geosynthetics contains results for geotextile evaluations.
Rolled Erosion Control Products are evaluated using laboratory-scale test methods
established by the Erosion Control Technology Council. Submissions for products are
accepted four times a year, and testing is performed by a private laboratory. The section on
the website Rolled Erosion Control Products contains results these evaluations.
FHWA NHI-07-092
Geosynthetics Engineering

1-20

Introduction, Identification, and Evaluation


August 2008

1.6 SPECIFICATIONS
When highway engineers first started using geosynthetics, their specifications were very
simple: use Brand X or equal. That approach was probably OK when there were only a few
products available, but today, with literally hundreds of different geosynthetics on the market
with a wide variety of properties, specifications should be based on the specific geosynthetic
properties required for design, installation, and durability. The use of standard
geosynthetics may result in uneconomical or unsafe designs. Specifying a particular type of
geosynthetic or its equivalent can also be very misleading. What is equivalent? A contractor
may select a product that has completely different properties than intended by the designer.
Specifications can be classified as generic, performance, approved list, and approved
supplier. For most routine applications, generic specifications are preferred because they are
based on the geosynthetic properties required by the design, installation and construction
conditions, and durability requirements of the project. Performance specifications require
testing of the geosynthetic together with soils from the project. (Recall that the engineer is
responsible for performance tests, not the contractor or manufacturer.) Thus the agency or
owner has to pre-select geosynthetics based on experience or index tests and then obtain
representative samples of soils from the project. In some situations, it may be better to
require the contractor to submit, in advance of construction, samples of the proposed
geosynthetics and soils from the project site or from a proposed borrow area to the engineer
for testing. Realistically, performance testing takes time, often weeks, so the contract must
clearly specify how far in advance of product installation that the samples must be submitted
to the engineer for testing and approval.
Some state agencies use an approved list type of specification. This approach has several
advantages, especially for routine applications, and it minimizes the chances for unwarranted
product substitutions to be made in the field. However, development of an approved list
program will require the agency to do considerable up-front testing to insure that products on
their approved list will actually work for a particular application and for their soils and site
conditions. But once it is established, it provides a simple, convenient method of specifying
geosynthetics with confidence. As new geosynthetics become available, they can be added
to the list after additional testing. The list has to be continually updated too, as the
manufacturing process may have changed since products were first approved. Samples
should be periodically obtained so they can be examined alongside the original tested
specimens to verify consistency in materials and any changes in the manufacturing process.
Note that the purpose of the NTPEP program mentioned above is to reduce the need for each
state to perform its own up-front testing to develop an approved products list.

FHWA NHI-07-092
Geosynthetics Engineering

1-21

Introduction, Identification, and Evaluation


August 2008

Most agencies have used an approved supplier specification for patented reinforced wall
systems because the material suppliers also provide a free design of the wall. However,
this practice has both advantages and disadvantages, as discussed in Chapter 9.
In almost every chapter of this manual, guide specifications are given for the particular
application discussed in the chapter. See Richardson and Koerner (1990) and Koerner and
Wayne (1989) for additional guide specifications.
All geosynthetic specifications should include:
general requirements
specific geosynthetic properties
seams and overlaps
placement procedures
repairs, and
acceptance and rejection criteria
General requirements include the product type(s), acceptable polymeric materials, mass per
unit area, roll dimensions if relevant, etc. Geosynthetic manufacturers and representatives are
good sources of information on these characteristics. Other items that should be specified in
this section are instructions on storage and handling so products can be protected from
ultraviolet exposure, dust, mud, or any other elements that may affect performance.
Guidelines concerning on-site storage and handling of geotextiles are contained in ASTM D
4873, Standard Guide for Identification, Storage, and Handling of Geotextiles. Finally,
certification requirements also should be included in this section.
Specific geosynthetic physical, index, and performance properties as required by the
design must be listed. Properties should be given in terms of minimum (or maximum)
average roll values (MARVs), along with the required test methods. MARVs are simply the
smallest (or largest) anticipated average value that would be obtained for any roll tested
(ASTM D 4439; Koerner, 2006). This average property value must exceed the minimum (or
be less than the maximum) value specified for that property based on a particular standard
test. Ordinarily it is possible to obtain a manufacturer's certification for MARVs.
Seam and overlap requirements should be clearly specified. A minimum overlap of 1 foot
(0.3 m) is recommended for all geotextile applications, but overlaps may be increased due to
specific site and construction requirements. Sewing of seams, discussed in Section 1.8, may
be required for very soft foundation conditions. Also, sometimes geotextiles are supplied
with factory-sewn seams. The seam strengths specified should equal the required strength of
the geosynthetic in the direction perpendicular to the seam length and using the same test
FHWA NHI-07-092
Geosynthetics Engineering

1-22

Introduction, Identification, and Evaluation


August 2008

procedures. For designs where wide width tests are used (e.g., reinforced embankments on
soft foundations), the required seam strength is a calculated design value required for
stability. Therefore, seam strengths should never be specified as a percent of the
geosynthetic strength.
Geogrids and geonets may be overlapped or connected by mechanical fasteners, though the
connection may be either structural or a construction aid (i.e., when strength perpendicular to
the seam length is not required)(see Section 1.8). Geomembranes are normally thermally
bonded (extrusion welded) and, as discussed in Chapter 10, seams are specified in terms of
peel and shear strengths.
For sewn geotextiles, geomembranes, and structurally connected geogrids, the seaming
material (thread, extrudate, or fastener) should consist of polymeric materials that have the
same or greater durability as the geosynthetic being seamed. For example, nylon thread,
unless treated, which is often used for geotextile seams may weaken in time as it absorbs
water. See Section 1.9 for additional information on field seams and anticipated seam
strength values.
Placement procedures should be given in detail in the specifications and on the construction
drawings. These procedures should include grading and ground-clearing requirements,
aggregate specifications, minimum aggregate lift thickness, and equipment requirements.
These requirements are especially important if the geosynthetic was selected on the basis of
survivability. Orientation and direction of geosynthetic placement should also be clearly
specified on the construction drawings. Detailed placement procedures are described in each
application chapter.
Repair procedures for damaged sections of geosynthetics (i.e., rips and tears) should be
detailed. Included are requirements for overlaps, sewn seams, fused seams, or complete
replacement of the damaged product. For overlap repairs, the geosynthetic should extend the
minimum of the overlap length requirement from all edges of the tear or rip (i.e., if a 1 foot
(0.3 m) overlap is required, the patch should extend at least 1 foot (0.3 m) from all edges of
the tear). In reinforcement applications, it is best that the specifications require complete
replacement of a damaged section. Finally, the contract documents should very clearly state
that final approval of the repairs is determined by the engineer, and that payment for repairs
is the responsibility of he contractor.
Acceptance and rejection criteria for the geosynthetic materials should be clearly stated in
the specifications. It is very important that all installations be observed by a designers
representative who is knowledgeable about geotextile placement procedures and who is
FHWA NHI-07-092
Geosynthetics Engineering

1-23

Introduction, Identification, and Evaluation


August 2008

aware of design requirements. Sampling (e.g., ASTM D 4354, Standard Practice for
Sampling of Geosynthetics for Testing) and testing requirements for quality assurance that
are required during construction should also be specified. Guidelines for acceptance and
rejection of geosynthetic shipments are given in ASTM D 4759, Standard Practice for
Determining the Specification Conformance of Geosynthetics.
For small projects, the cost of ASTM acceptance/rejection criterion testing is often a
significant portion of the total project cost and may even exceed the cost of the geosynthetic
itself. In such cases, a certification by the manufacturer should be required. In this case,
collect a few samples from the rolls for future evaluation and confirmation, if required.

1.7 FIELD INSPECTION


Problems with geosynthetics are often due to poor product acceptance and construction
monitoring procedures on the part of the owner, and/or inappropriate installation methods on
the part of the contractor. A checklist for field personnel responsible for observing a
geosynthetic installation is presented in Table 1-5. Recommended installation methods are
presented in the application chapters.

1.8 FIELD SEAMING


Some form of geosynthetic seaming will be necessary for those applications that require
continuity between adjacent rolls, and that includes all the applications discussed in this
manual. Seaming techniques include overlapping, sewing, stapling, tying, heat bonding,
welding, and gluing. Some of these techniques are more suitable for certain types of
geosynthetics than others. For example, the most efficient and widely used methods for
geotextiles are overlapping and sewing, and these techniques are discussed first.
Simple overlap will be suitable for most geotextile and biaxial geogrid projects. The
minimum overlap is 1 foot (0.3 m). Greater overlaps may be required for specific
applications. Note that the only strength provided by an overlap is the friction between
adjacent sheets of geotextiles, or for biaxial geogrids, by friction and fill strike-through of the
apertures. Unless overburden pressures are large and the overlap substantial, very little stress
can actually be transferred through the overlap between adjacent rolls. If overlaps will be
used to transfer stress from one geosynthetic to another, then interface shear strength test
should be performed using the pullout or direct shear methods to evaluate the strength of the
field configuration. In this case, durable ties (e.g., galvanized hog rings or industrial
FHWA NHI-07-092
Geosynthetics Engineering

1-24

Introduction, Identification, and Evaluation


August 2008

polypropylene/polyethylene zip ties) should be used to assure that the geosynthetics maintain
their relative position of the geosynthetics after placement and provide an additional level of
safety.
Table 1-5
Geosynthetic Field Inspection Checklist

1. Read the specifications.


2. Review the construction plans.
3. Determine if the geosynthetic is specified by (a) specific properties or (b) an
approved products list.
(a)For specification by specific properties, check listed material properties of the
supplied geosynthetic from published literature or against the specific property
values specified. Or
(b) Obtain the geosynthetic name(s), type, and style, along with a small sample(s) of
approved material(s) from the design engineer. Check supplied geosynthetic
type and style for conformance to approved material(s). If the geosynthetic is
not listed, reject it. In some cases, you may want to contact the designer with a
description of the material and request an evaluation before permitting it to be
installed.
4. When the geosynthetics are delivered to the site, check the rolls to see that they
are properly stored; check for damage.
5. Check roll and lot numbers to verify whether they match certification documents.
6. Cut two samples 4 to 6 in. (100 mm to 150 mm) square from a roll. Staple one to
your copy of the specifications for comparison with future shipments and send
one to the design engineer for approval or information.
7. Observe materials in each roll to make sure they are the same. Observe rolls for
flaws and nonuniformity.
8. Obtain test samples according to specification requirements from randomly
selected rolls. Mark the machine direction on each sample and note the roll
number. Take at least one archive sample of each geosynthetic, even if testing is
not required.
9. Observe construction to see that the contractor complies with specification
requirements for installation.
10. Check all seams, both factory and field, for any missed stitches in geotextile. If
necessary, either reseam or reject materials.
11. If possible, check geosynthetic after aggregate or riprap placement for possible
damage. This can be done either by constructing a trial installation, or by
removing a small section of aggregate or riprap and observing the geosynthetic
after placement and compaction of the aggregate, at the beginning of the
project. If perforations, tears, or other damage has occurred, contact the design
engineer immediately.
12. Check future shipments against the initial approved shipment and collect
additional test samples. Collect samples of seams, both factory and field, for
testing. For field seams, have the contractor sew several meters of a dummy
seam(s) for testing and evaluation.

FHWA NHI-07-092
Geosynthetics Engineering

1-25

Introduction, Identification, and Evaluation


August 2008

If overlaps become excessive or stress transfer is required between two adjacent rolls, then
sewing offers a practical and economical alternative for geotextiles. For typical projects and
conditions, sewing is generally more economical when overlaps of 3 ft (1 m) or greater are
required. To obtain good-quality, effective seams, the user should be aware of the following
sewing variables (Ko, 1987; Diaz and Myles, 1990; Koerner, 2006):
Thread type:
Kevlar aramid, polyethylene, polyester, or polypropylene (in
approximate order of decreasing strength and cost). Thread durability must be
consistent with project requirements.
Thread tension: Usually adjusted in the field to be sufficiently tight, but not so tight
that the thread cuts the geotextile.
Stitch density: Typically, 200 to 400 stitches per yard (m) of seam are used for
lighter-weight geotextiles, while heavier geotextiles usually allow only 150 to 200
stitches per yard (m).
Stitch type (Figure 1-2(a)): Single- or double-thread chain stitch, Type 101, or a
double-thread lock stitch, Type 401. A lock stitch is preferred because it is less likely
to unravel.
Number of rows: Usually two parallel rows of stitches are preferred for increased
safety.
Seam type (Figure 1-2(b)): Flat or prayer seams, J- or Double J-type seams, or
butterfly seams are the most widely used. Butterfly seams are usually only done in
factories.
When properly made, sewn seams can provide reliable stress transfer between adjacent
sheets of geotextile. However, there are several points with regard to seam strength that
should be understood, as follows.
1. Due to needle damage and stress concentrations at the stitch, sewn seams are weaker
than the geotextile (good, high-quality seams have only about 50% to 80% of the
intact geotextile strength based on wide width tests).
2. Grab strength results are influenced by the stitches, so the test yields artificially high
seam strengths. Grab test should only be used for quality control and not to
determine strength.
3. The maximum seam strengths achievable at this time are on the order of 14,000 lb/ft
(200 kN/m) under factory conditions, using 23,000 lb/ft (330 kN/m) geotextiles.
4. Field seam strengths will most likely be lower than laboratory or factory seam
strengths.
5. All stitches can unravel, although lock-type stitches are less likely to do so.
6. Unraveling can be avoided by utilizing high-quality equipment and proper selection
of needles, thread, seam and stitch type, and by using two or more rows of stitches.
7. Careful inspection of all stitches is essential.
FHWA NHI-07-092
Geosynthetics Engineering

1-26

Introduction, Identification, and Evaluation


August 2008

Field sewing is relatively simple and usually requires two or three laborers, depending on the
geotextile, seam type, and sewing machine. Good seams require careful control of the
operation, cleanliness, and protection from the elements. However, adverse field conditions
can easily complicate sewing operations. Although most portable sewing machines are
electric, pneumatic equipment is available for operating in wet environments.
Since the seam is the weakest link in the geotextile, all seams, including factory seams,
should be carefully inspected. To facilitate inspection and repair, the geotextile should be
placed (or at least inspected prior to placement) with all seams up (Figure. 1-2(c)). Using a
contrasting thread color can facilitate inspection. Procedures for testing sewn seams are
given in ASTM D 4884, Standard Test Method for Seam Strength of Sewn Geotextiles.
Seaming of biaxial geogrids and geocomposites is most commonly achieved by overlaps, and
the remarks above on overlap of geotextiles are generally appropriate to these products.
Again an interface shear test should be performed on adjacent layers if load transfer between
geosynthetic rolls is required. Uniaxial geogrids are normally butted in the along-the-roll
direction. Seams in the roll direction of uniaxial geogrids are made with a bodkin joint for
HDPE geogrids, as illustrated in Figure 1-3, and may be made with overlaps for coated PET
geogrids. Mechanical ties (e.g., plastic ties) can also been used provided the seam strength
has been tested and the ties meet the design life requirements (i.e., same as the geogid).
Seaming of geomembranes and other geosynthetic barriers is much more varied. The method
of seaming is dependent upon the geosynthetic material being used and the project design.
Overlaps of a designated length are typically used for thin-film geotextile composites and
geosynthetic clay liners. Geomembranes are seamed with thermal methods or with solvents.

FHWA NHI-07-092
Geosynthetics Engineering

1-27

Introduction, Identification, and Evaluation


August 2008

(c) Improper seam placement cannot inspect or repair

Figure 1-2.

Types of (a) stitches and (b) seams, according to Federal Standard No. 751a
(1965); and (c) improper seam placement.

FHWA NHI-07-092
Geosynthetics Engineering

1-28

Introduction, Identification, and Evaluation


August 2008

Figure 1-3.

1.9

Bodkin connection of HDPE uniaxial geogrid.

REFERENCES

Reference lists are provided at the end of chapter (and appendix). Note that FHWA
references are generally available at www.fhwa.dot.gov/bridge (under the publications and
geotechnical tabs) and/or at www.nhi.fhwa.dot.gov (under the training and NHI store tabs).
Detailed lists of specific ASTM (2006) and GRI (2006) test procedures are presented in
Appendix E. Koerner (2006) is a recent, comprehensive textbook on geosynthetics and is a
key reference for designers. The bibliographies by Giroud (1993, 1994) contain references
for publications on geosynthetics before January 1, 1993.
AASHTO (2006). Standard Specifications for Geotextiles - M 288, Standard Specifications
for Transportation Materials and Methods of Sampling and Testing, 26th Edition,
American Association of State Transportation and Highway Officials, Washington,
D.C.
AASHTO (1990a). Task Force 25 Report Guide Specifications and Test Procedures for
Geotextiles, Subcommittee on New Highway Materials, American Association of
State Transportation and Highway Officials, Washington, D.C.
AASHTO (1990b). Design Guidelines for Use of Extensible Reinforcements (Geosynthetic)
for Mechanically Stabilized Earth Walls in Permanent Applications, Task Force 27
Report - In Situ Soil Improvement Techniques, American Association of State
Transportation and Highway Officials, Washington, D.C.
FHWA NHI-07-092
Geosynthetics Engineering

1-29

Introduction, Identification, and Evaluation


August 2008

ASTM (2006). Annual Books of ASTM Standards, American International, West


Conshohocken, PA:
Volume 4.08 (I), Soil and Rock
Volume 4.09 (II), Soil and Rock
Volume 4.13 Geosynthetics see Appendix E for full listing
Volume 7.01, 7.02 Textiles
Volume 8.01, 8.02, 8.03, Plastics
Bailey, T. D., Harney, M. D., and Holtz, R. D. (2005). Rapid Assessment of Geotextile
Clogging Potential Using the Flexible Wall Gradient Ratio Test, Proceedings of the
GRI-18 Conference, Austin, Texas (CD-ROM), ASCE.
Boyle, S. R., Holtz, R. D., and Gallagher, M. (1996). Influence of Strain Rate, Specimen
Length, and Confinement on Measured Geotextile Strength Properties,
Geosynthetics International, Vol. 3, No. 2, pp. 205-225.
Boyle, S. R. (1995). Deformation Prediction of Geosynthetic Reinforced Soil Retaining
Walls, PhD Dissertation, University of Washington, Seattle, 392 pp.
Christopher, B.R. and Dahlstrand, T.K. (1989). Geosynthetics in Dams - Design and Use,
Association of State Dam Safety Officials, Lexington, KY, 311 p.
Christopher, B.R. and Holtz, R.D. (1985). Geotextile Engineering Manual, FHWA-TS86/203, 1044 p.
Diaz, V. and Myles, B. (1990). Field Sewing of Geotextiles--A Guide to Seam Engineering,
Industrial Fabrics Association Internationals, St. Paul, MN, 29 p.
Elias, V., Christopher, B.R. and Berg, R.R. (2001). Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes, Design & Construction Guidelines, FHWA-NHI-00-043,
418 p.
Elias, V., Yuan, Z., Swan, R.H. and Bachus, R.C. (1998). Development of Protocols for
Confined Extension/Creep Testing of Geosynthetics for Highway Applications,
FHWA RD-97-143.
Elias, V. (1997). Corrosion/Degradation of Soil Reinforcements for Mechanically Stabilized
Earth Walls and Reinforced Soil Slope, FHWA-SA-96-072, 105 p.
Fannin, R. J. (2000) Basic Geosynthetics: A Guide to Best Practices, BiTech Publishers,
Richmond, BC, 86 p.
Federal Standards (1965). Federal Standard Stitches, Seams, and Stitching, No. 751a, Jan.

FHWA NHI-07-092
Geosynthetics Engineering

1-30

Introduction, Identification, and Evaluation


August 2008

FHWA (1997). Degradation Reduction Factors for Geosynthetics, FHWA Geotechnology


Technical Note, Federal Highway Administration, U.S. Department of
Transportation, May 1997, 5 p. Also reprinted as: Elias, V., DiMaggio, J.A. and
DiMillio, A., FHWA Technical Note on the Degradation-Reduction Factors for
Geosynthetics, Geotechnical Fabrics Report, Vol. 15, No. 6, Industrial Fabrics
Association International, St. Paul, MN, August 1997, pp. 24-26.
Fischer, G.R. (1994). The Influence of Fabric Pore Structure on the Behavior of Geotextile
Filters, Ph.D. Dissertation, University of Washington, 498 p.
Giroud, J.P., with cooperation of Beech, J.F. and Khatami, A. (1994). Geosynthetics
Bibliography, Volume II, IGS, Industrial Fabrics Association International, St. Paul,
MN, 940 p.
Giroud, J.P., with cooperation of Beech, J.F. and Khatami, A. (1993). Geosynthetics
Bibliography, Volume I, IGS, Industrial Fabrics Association International, St. Paul,
MN, 781 p.
GRI (2006). Test Methods & Standards, Geosynthetic Research Institute, Drexel University,
Philadelphia, PA.
GRI (1993). Test Method GT1, Geotextile Filter Performance via Long Term Flow (LTF)
Tests, Standard Test Method, Geosynthetic Research Institute, Drexel University,
Philadelphia, PA.
Harney, M. D. and Holtz, R. D. (2001). Flexible Wall Gradient Ratio Test, Proceedings of
Geosynthetics Conference 2001, Portland, Oregon, Industrial Fabrics Association
International, pp. 409-422.
Harney, M. D., Bailey, T. D., and Holtz, R. D. (2007). Clogging Potential of Geotextile
Filters Using the Flexible Wall Gradient Ratio Test, Geotechnical Testing Journal,
ASTM (accepted for publication).
Holtz, R.D., Tobin, W.R. and Burke, W.W. (1982). Creep Characteristics and Stress-Strain
Behavior of a Geotextile-Reinforced Sand, Proceedings of the Second International
Conference on Geotextiles, Las Vegas, Nevada, Vol. 3, pp. 805-809.
Hoover, T.P. (1982). Laboratory Testing of Geotextile Fabric Filters, Proceedings of the
Second International Conference on Geotextiles, Las Vegas, Nevada, Vol. 3, pp. 839844.
IGS (2000). Recommended Mathematical and Graphical Symbols, International Geosynthetic
Society, 17 p. [http://geosyntheticssociety.org/guideance.htm]
Ingold, T.S. and Miller, K.S. (1988). Geotextiles Handbook, Thomas Telford Ltd., London,
152 p.
FHWA NHI-07-092
Geosynthetics Engineering

1-31

Introduction, Identification, and Evaluation


August 2008

Ko, F.K. (1987). Seaming and Joining Methods, Geotextiles and Geomembranes, Vol. 6,
Nos. 1-3, pp 93-107.
Koerner, R.M. (2006). Designing With Geosynthetics, 5th Edition, Prentice-Hall Inc.,
Englewood Cliffs, NJ, 816 p.
Koerner, R.M. and Wayne M.H. (1989). Geotextile Specifications for Highway Applications,
FHWA-TS-89-026, 90 p.
Koerner, R.M. and Ko, F.K. (1982). Laboratory Studies on Long-Term Drainage Capability
of Geotextiles, Proceedings of the Second International Conference on Geotextiles,
Las Vegas, NV, Vol. I, pp. 91-95.
Legge, K.R. (1990). A New Approach to Geotextile Selection, Proceedings of the Fourth
International Conference on Geotextiles, Geomembranes and Related Products, The
Hague, Netherlands, Vol. 1., pp. 269-272.
Mar, A.D. (1994). The Influence of Gradient Ratio Testing Procedures on the Filtration
Behavior of Geotextiles, MSCE Thesis, University of Washington.
Rankilor, P.R. (1981). Membranes in Ground Engineering, John Wiley & Sons, Inc.,
Chichester, England, 377 p.
Richardson, G.R. and Koerner, R.M., Editors (1990). A Design Primer: Geotextiles and
Related Materials, Industrial Fabrics Association International, St. Paul, MN, 166 p.
Sansone, L.J. and Koerner, R.M. (1992). Fine Fraction Filtration Test to Assess Geotextile
Filter Performance, Geotextiles and Geomembranes, Vol. 11, Nos. 4-6, pp. 371-393.
U.S. Army Corps of Engineers (1977). Civil Works Construction Guide Specification for
Plastic Filter Fabric, Corps of Engineer Specifications No. CW-02215, Office, Chief
of Engineers, U.S. Army Corps of Engineers, Washington, D.C.

FHWA NHI-07-092
Geosynthetics Engineering

1-32

Introduction, Identification, and Evaluation


August 2008

2.0 GEOSYNTHETICS IN SUBSURFACE DRAINAGE SYSTEMS

2.1

BACKGROUND

A major use of geotextiles is as filters in drainage applications such as trench and interceptor
drains, blanket drains, pavement edge drains, structure drains, and beneath permeable
roadway bases. The filter restricts movement of soil particles as water flows into the drain or
drainage layer and is collected and/or transported downstream. Geocomposites consisting of
a drainage core surrounded by a geotextile filter are often used as the drain itself in these
applications. Geotextiles are also used as filters beneath hard armor erosion control systems,
and this application is discussed in Chapter 3.
Because of their comparable performance, improved economy, consistent properties, and
ease of placement, geotextiles have been used successfully to replace graded granular filters
in almost all drainage applications. Thus, they must perform the same functions as graded
granular filters:
allow water to flow through the filter into the drain, and to continue doing this
throughout the life of the project; and
retain the soil particles in place and prevent their migration (piping) through the filter
(if some soil particles do move, they must be able to pass through the filter without
blinding or clogging the downstream media during the life of the project).
Geotextiles, like graded granular filters, require proper engineering design or they may not
perform as desired. Unless flow requirements, piping resistance, clogging resistance and
constructability requirements (defined later) are properly specified, the geotextile/soil
filtration system may not perform properly. In addition, construction must be monitored to
ensure that materials are installed correctly.
In most drainage and filtration applications, geotextile use can be justified over conventional
graded granular filter material use because of cost advantages from:
the use of less-costly drainage aggregate;
the possible use of smaller-sized drains;
the possible elimination of collector pipes;
expedient construction;
lower risk of contamination and segregation of drainage aggregate during
construction;
reduced excavation.
FHWA NHI-07-092
Geosynthetics Engineering

2-1

Subsurface Drainage
August 2008

In addition, geosynthetics often increase drainage system reliability and, considering the
value of drainage in geotechnical engineering, a significant cost-benefit can result when the
designer is assured of a properly performing drain.

2.2

APPLICATIONS

Properly designed geotextiles can be used as a replacement for, or in conjunction with,


conventional graded granular filters in almost any drainage application. Below are a few
examples of drainage applications.
In many of these applications, properly designed
geocomposites can also be employed.
Filters around trench drains
and edge drains to prevent
soil from migrating into the
drainage aggregate or system,
while allowing water to exit
from the soil.

Filters beneath pavement


permeable bases, blanket
drains and base courses.
Prefabricated
geocomposite
drains are also used as blanket
drains and have been used as
horizontal drains in pavement
systems.
Geotextile wraps for slotted or
jointed drain and well pipes -to prevent filter aggregate
from entering the pipe, while
allowing the free flow of water
into the pipe.

FHWA NHI-07-092
Geosynthetics Engineering

2-2

Subsurface Drainage
August 2008

Drains for structures such as


retaining walls and bridge
abutments. They separate the
drainage aggregate or system
from the backfill soil, while
allowing free drainage of
ground and infiltration water.
Geocomposite
drains
are
especially useful in this
application.
Interceptor, toe drains, and
surface drains -- to aid in the
stabilization of slopes by
allowing excess pore pressures
within the slope to dissipate,
and by preventing surface
erosion. Again, geocomposites
have been successfully used in
this application.
Chimney and toe drains for
earth dams and levees -- to
provide seepage control.

Typical geotextile
and geocomposite
pavement
edge
drain applications
(NCHRP 1-37A).

In each of these applications, flow is through the geotextile -- that is, perpendicular to the
plane of the fabric. In other applications, such as vertical drains in soft foundation soils,
lateral drains below slabs and behind retaining walls, and gas transfer media, flow may occur
FHWA NHI-07-092
Geosynthetics Engineering

2-3

Subsurface Drainage
August 2008

both perpendicularly to and transversely in the plane of the geotextile. In many of these
applications, geocomposite drains may be appropriate. Design with geocomposite systems is
covered in Section 2.10.

2.3

GEOTEXTILE FILTER DESIGN PRINCIPLES AND CONCEPTS

All geosynthetic designs should begin with an assessment of the criticality and severity of the
project conditions (see Table 2-1) for a particular application. Although first developed by
Carroll (1983) for drainage and filtration applications, the concept of critical-severe projects
and, thus, the level of engineering responsibility required will be applied to other
geosynthetic applications throughout this manual.
Table 2-1
Guidelines for Evaluating the Critical Nature or Severity
of Drainage And Erosion Control Applications
(after Carroll, 1983)
A. Critical Nature of the Project
Item
Critical
Less Critical
1. Risk of loss of life and/or
structural damage due to
drain failure:

High

None

>>>

= or <

None

Yes

Severe

Less Severe

1. Soil to be drained:

Gap-graded, pipable, or
dispersible

Well-graded or uniform

2. Hydraulic gradient:

High

Low

Dynamic, cyclic, or Pulsating

Steady state

2. Repair costs versus


installation costs of drain:
3. Evidence of drain clogging
before potential
catastrophic failure:
B. Severity of the Conditions
Item

3. Flow conditions:

FHWA NHI-07-092
Geosynthetics Engineering

2-4

Subsurface Drainage
August 2008

A few words about the condition of the soil to be drained (Table 2-1) are in order. First,
grain-size distribution curves for gap-graded, well-graded and uniform soils are illustrated in
Figure 2-1. Certain gap-graded and broadly graded soils may be internally unstable; that is,
they can experience piping or internal erosion. Figure 2-1 also shows two sandy-gravel (GP)
soils that are potentially unstable. Criteria for deciding whether a soil is internally unstable
are discussed in Section 2.4-1.c. On the other hand, a soil is internally stable if it is selffiltering and if its own fine particles do not move through the pores of its coarser fraction
(LaFluer et al., 1993).
Dispersive soils are fine-grained natural soils that deflocculate in the presence of water and,
therefore, are highly susceptible to erosion and piping (Sherard, et al., 1972). See also
Sherard and Decker (1977) for more information on dispersive soils.

Figure 2-1.

Grain-size distributions for several soils.

Designing with geotextiles for filtration is essentially the same as designing graded granular
filters. A geotextile is similar to a soil in that it has voids (pores) and particles (filaments and
fibers). However, because of the shape and arrangement of the filaments and the
compressibility of the structure with geotextiles, the geometric relationships between
filaments and voids is more complex than in soils. In geotextiles, pore size is measured
directly, rather than using particle size as an estimate of pore size, as is done with soils.
Since pore sizes can be directly measured, at least in theory, relatively simple relationships
between the pore sizes and particle sizes of the soil to be retained can be developed. Three
simple filtration concepts are used in the design process:
FHWA NHI-07-092
Geosynthetics Engineering

2-5

Subsurface Drainage
August 2008

1. If the size of the largest pore in the geotextile filter is smaller than the larger particles
of soil, soil particles that tend to move will be retained by the filter. As with graded
granular filters, the larger particles of soil will form a filter bridge over the hole,
which in turn, filters smaller particles of soil, which then retain the soil and prevent
piping (Figure 2-2).
2. If the smaller openings in the geotextile are sufficiently large enough to allow smaller
particles of soil to pass through the filter, then the geotextile will not blind or clog
(see Figure 2-3).
3. A large number of openings should be present in the geotextile so that water flow can
be maintained even if some of the openings later become plugged.
These simple concepts and analogies with soil filter design criteria are used to establish
design criteria for geotextiles. Specifically, these criteria are:
the geotextile must retain the soil particles (retention criterion), while
allowing water to pass (permeability criterion), throughout
the life of the structure (clogging resistance criterion and durability requirements). To
perform effectively, the geotextile must also survive the installation process (survivability or
constructability criterion).

After a detailed study of research carried out both in North America and in Europe on
conventional and geotextile filters, Christopher and Holtz (1985) developed the following
procedure, now called the FHWA filter design procedure, for the design of geotextile filters
for drainage (this chapter) and permanent erosion control applications (Chapter 3). The level
of design and testing required depends on the critical nature of the project and the severity of
the hydraulic and soil conditions (Table 2-1). Especially for critical projects, consideration
of the risks and the consequences of geotextile filter failure require great care in selecting the
appropriate geotextile. For such projects, and for severe hydraulic conditions, very
conservative designs are recommended. Geotextile selection should not be based on cost
alone. The cost of the geotextile is usually minor in comparison to the other components and
the construction costs of a drainage system. Also, do not try to save money by eliminating
laboratory soil-geotextile performance testing when such testing is required by the design
procedure.
A National Cooperative Highway Research Program (NCHRP) study by Koerner et al.
(1994) of the performance of geotextiles in drainage systems indicated that the FHWA
design criteria developed by Christopher and Holtz (1985) were an excellent predictor of
filter performance, particularly for granular soils (<50% passing a No.200 (0.075 mm) sieve).

FHWA NHI-07-092
Geosynthetics Engineering

2-6

Subsurface Drainage
August 2008

Figure 2-2.

Filter bridge formation.

Figure 2-3.

Definitions of clogging and blinding (Bell and Hicks, 1980).

FHWA NHI-07-092
Geosynthetics Engineering

2-7

Subsurface Drainage
August 2008

2.4

FHWA FILTER DESIGN PROCEDURE

Based on the concepts just described, the FHWA filter design procedure has three design
criteria: retention, permeability, and clogging resistance. Survivability and durability also are
part of the design.
2.4-1 Retention Criteria
Because they put different demands on the geotextile filter, two flow conditions, steady state
and dynamic, are considered for retention.
2.4-1.a Steady State Flow Conditions
AOS or O95(geotextile) < B D85 (soil)

[ 2 - 1]

where:
AOS = apparent opening size (see Table 1-3) (mm);
O95
= opening size in the geotextile for which 95% are smaller (mm);
AOS O95;
B
= a coefficient (dimensionless); and
D85
= soil particle size for which 85% are smaller (mm).
The coefficient B ranges from 0.5 to 2 and is a function of the type of soil to be filtered, its
density, the uniformity coefficient Cu if the soil is granular, the type of geotextile (woven or
nonwoven), and the flow conditions.
For sands, gravelly sands, silty sands, and clayey sands (soils with less than 50% passing the
No.200 (0.075 mm) sieve per the Unified Soil Classification System), B is a function of the
uniformity coefficient, Cu. Therefore, for
Cu < 2 or > 8:

B=1

[2 - 2a]

2 < Cu < 4:

B = 0.5 Cu

[2 - 2b]

4 < Cu < 8:

B = 8/Cu

[2 - 2c]

where:
Cu = D60/D10.

FHWA NHI-07-092
Geosynthetics Engineering

2-8

Subsurface Drainage
August 2008

Sandy soils which are not uniform (Figure 2-1) tend to bridge across the openings; thus, the
larger pores may actually be up to twice as large (B 2) as the larger soil particles because,
quite simply, two particles cannot pass through the same hole at the same time. Therefore,
use of the criterion B = 1 would be quite conservative for retention, and this criterion has
been used by, for example, the U.S. Army Corps of Engineers.
If the protected granular soils contain appreciable fines, use only the portion passing the No.4
(4.75 mm) sieve for selecting the geotextile (i.e., scalp off the + No.4 (+4.75 mm) material).
For silts and clays (soils with more than 50% passing the No.200 (0.075 mm) sieve), B is a
function of the type of geotextile:
for woven geotextiles, B = 1; O95 < D85
for nonwoven geotextiles, B = 1.8; O95 < 1.8 D85

[2 - 3]
[2 - 4]

and for both, AOS or O95 < 0.3 mm

[2 - 5]

Due to their random pore characteristics and, in some types, their felt-like nature, nonwoven
geotextiles will generally retain finer particles than a woven geotextile of the same AOS.
Therefore, the use of B = 1 will be even more conservative for nonwoven geotextiles.
2.4-1.b Dynamic Flow Conditions
If the geotextile is not properly weighted down and in intimate contact with the soil to be
protected, or if dynamic, cyclic, or pulsating loading conditions produce high localized
hydraulic gradients, then soil particles can move behind the geotextile. Thus, the use of B =
1 is not conservative, because the bridging network will not develop and the geotextile will
be required to retain even finer particles. When retention is the primary criteria, B should be
reduced to 0.5; or:
O95 < 0.5 D85

[2 -6]

Dynamic flow conditions can occur in pavement drainage applications. For reversing
inflow-outflow or high-gradient situations, it is best to maintain sufficient weight on the filter
to prevent the geotextile from moving. Dynamic flow conditions with erosion control
systems are discussed in Chapter 3.

FHWA NHI-07-092
Geosynthetics Engineering

2-9

Subsurface Drainage
August 2008

2.4-1.c Stable versus Unstable Soils


The above retention criteria assumes that the soil to be filtered is internally stable it will
not pipe internally. If unstable soil conditions are encountered, performance tests should be
conducted to select suitable geotextiles. According to Kenney and Lau (1985, 1986) and
LaFluer, et al. (1989), broadly graded (Cu > 20) soils with concave upward grain size
distributions tend to be internally unstable. The Kenney and Lau (1985, 1986) procedure
utilizes a mass fraction analysis. Research by Skempton and Brogan (1994) verified the
Kenney and Lau (1985, 1986) procedure.
Unstable conditions can also exist in materials such as rubblized base course layers in
roadways, recycled concrete, or dense graded roadway base with erodible fines. When these
materials are used adjacent to a drain (e.g., roadway edgedrains), to avoid clogging the
geotextile filter, the geotextile should not be placed between these unstable materials and the
trench drain aggregate. It should be placed beneath and on the outside of the drain to prevent
infiltration of the subgrade and subbase layers.
2.4-2 Permeability and Permittivity Criteria
We consider two conditions when designing for permeability: (1) less critical/less severe and
(2) critical/severe conditions.
(1) For less critical applications and less severe conditions:
kgeotextile > ksoil

[2 - 7a]

(2) For critical applications and severe conditions:


kgeotextile > 10 ksoil

[2 - 7b]

where:
kgeotextile = Darcy coefficient of permeability (m/sec).
(3) Permittivity requirements (for both critical/severe and less critical/less severe
applications):
The permittivity requirements depend on the percentage of fines in the soil to be filtered.
The more fines in the soil, the greater the permittivity required. The following equations are
recommended based on satisfactory past performance of geotextile filters.
For < 15% passing No.200 (0.075 mm)
0.5 sec-1
[2 - 8a]
-1
For 15 to 50% passing No.200 (0.075 mm) 0.2 sec
[2 - 8b]
-1
For > 50% passing No.200 (0.075 mm)
0.1 sec
[2 - 8c]
FHWA NHI-07-092
Geosynthetics Engineering

2-10

Subsurface Drainage
August 2008

Recall from the discussion on Hydraulic Properties in Sec. 1.5 that geotextile permittivity
= kgeotextile/tgeotextile, where kgeotextile = Darcy coefficient of permeability, and t = geotextile
thickness. Note that permittivity is a function of the hydraulic head as well as the opening
characteristics of the geotextile, and it has units of per seconds (s-1). Basically, permittivity is
flow rate (volume/time) per unit area per unit head, or l/sec = 75 gal/min/ft2/2 in. head or
1200 litres/min/m2/ 50mm head.
Permittivity is good measure of flow capacity, and as such it is a useful qualifier for making
sure the geotextile filter has sufficient flow capacity for a given soil in a particular
application. Because flow capacity of the system depends on the percentage of fines in the
soil to be protected, the minimum permittivity values given above are recommended, in
addition to permeability. Another reason for specifying permittivity is that many
manufacturers give the permittivity value for their products according to ASTM D 4491, and
furthermore, they have products available that meet or exceed these permittivity values.
Therefore, we recommend that, in addition to the minimum permeability of the geotextile,
you always specify permittivity values for your project.
For actual flow capacity, the permeability criteria for noncritical applications is conservative,
since an equal quantity of flow through a relatively thin geotextile takes significantly less
time than through a thick granular filter. Even so, some pores in the geotextile may become
blocked or plugged with time. Therefore, for critical or severe applications, Equation 2-7b is
recommended to provide a large factor of safety and an additional degree of conservatism.
Equation 2-7a may be used where flow reduction is judged not to be a problem, such as in
clean, medium to coarse sands and gravels.
The required flow rate, q, through the system should also be determined, and the geotextile
and drainage aggregate selected to provide adequate capacity. As indicated above, flow
capacities should not be a problem for most applications, provided the geotextile
permeability is greater than the soil permeability. However, in certain situations, such as
where geotextiles are used to span joints in, for example, concrete face panels or where they
are used as pipe wraps, portions of the geotextile may not be available for flow. For these
applications, the following criteria should be used together with the permeability criteria:
qrequired = qgeotextile(Ag/At)

[2 - 9]

where:
Ag
At

= geotextile area available for flow; and


= total geotextile area.

FHWA NHI-07-092
Geosynthetics Engineering

2-11

Subsurface Drainage
August 2008

2.4-3 Clogging Resistance


For clogging resistance, we consider the same two conditions as we did for the permeability
criteria: (1) less critical/less severe and (2) critical/severe conditions.
2.4-3.a Less Critical/Less Severe Conditions
For less critical/less severe conditions:
O95 (geotextile) > 3 D15 (soil)

[2 - 10]

Equation 2-10 applies to soils with Cu > 3. For Cu < 3, select a geotextile with the maximum
AOS value from Section 2.4.1.
In situations where clogging is a possibility (e.g., gap-graded or silty soils), the following
optional qualifiers may be applied:
for nonwoven geotextiles:
porosity of the geotextile, n > 50%

[2 - 11]

for woven monofilament and slit film woven geotextiles:


percent open area, POA > 4%

[2 - 12]

NOTE: See Section 1.5 for comments on porosity and POA.


Most common nonwoven geotextiles have porosities much greater than 70%. Most woven
monofilaments easily meet the criterion of Equation 2-12; tightly woven slit films do not, and
are therefore not recommended for subsurface drainage applications.
For less critical/less severe conditions, a simple way to avoid clogging, especially with silty
soils, is to allow fine particles already in suspension to pass through the geotextile. Then the
bridge network (Figure 2-2) formed by the larger particles retains the smaller particles. The
bridge network should develop rather quickly, and the quantity of fine particles actually
passing through the geotextile is relatively small. This is why the less critical/less severe
clogging resistance criteria requires an AOS (O95) sufficiently larger than the finer soil
particles (D15). Those are the particles that will pass through the geotextile. Unfortunately,
the AOS value only indicates the size and not the number of O95-sized holes available. Thus,
the finer soil particles will be retained by the smaller holes in the geotextile, and if there are
sufficient fines, a significant reduction in flow rate can occur.

FHWA NHI-07-092
Geosynthetics Engineering

2-12

Subsurface Drainage
August 2008

Consequently, to control the number of holes in the geotextile, it may be desirable to increase
other qualifiers such as the porosity and open area requirements. There should always be a
sufficient number of holes in the geotextile to maintain permeability and drainage, even if
some of them clog.
Filtration tests provide another option to consider, especially by inexperienced users.
2.4-3.b Critical/Severe Conditions
For critical/severe conditions, select geotextiles that meet the retention and permeability
criteria in Sections 2.4-1 and 2.4-2. Then perform a filtration test using samples of on-site
soils and hydraulic conditions.
Although several empirical methods have been proposed to evaluate the clogging potential of
geotextiles, the most realistic approach for all filtration applications is to perform a
laboratory test which simulates or models field conditions. We recommend the gradient ratio
test, ASTM D 5101, Measuring the Soil-Geotextile System Clogging Potential by the
Gradient Ratio. This test utilizes a rigid-wall soil permeameter with piezometer taps that
allow for simultaneous measurement of the head losses in the soil and the head loss across
the soil/geotextile interface (Figure 2-4). The ratio of the head loss across this interface
(nominally 1-in. {25 mm}) to the head loss across 2 in. (50 mm) of soil is termed the
gradient ratio. As fine soil particles adjacent to the geotextile become trapped inside or blind
the surface, the gradient ratio will increase. A gradient ratio (GR) less than 3 is
recommended by the U.S. Army Corps of Engineers (1977), based upon limited testing with
severely gap-graded soils (Haliburton and Wood, 1982). Because the test is conducted in a
rigid-wall permeameter, it is most appropriate for sandy soils with k > 10-6 m/sec.
For soils with permeabilities less than about 10-6 m/sec, filtration tests should be conducted
in a flexible wall or triaxial type apparatus to insure that the specimen is 100% saturated and
that flow is through the soil rather than along the sides of the specimen. The soil flexible
wall test is ASTM D 5084, while the Hydraulic Conductivity Ratio (HCR) test (ASTM D
5567) currently is the standard test for geotextiles and soils with appreciable fines. In fact,
ASTM D 5567 states that it is appropriate for soils with permeabilities hydraulic
conductivities) less than 5 x 10-4 m/sec. In Section 1.5 on the hydraulic properties of
geotextiles, we discussed the disadvantages of the HCR test and other filtration tests for fine
grained soils. We described the Flexible Wall GR test that combines the best features of the
GR test (D 5101) and the flexible wall permeability test (D 5084).
Fortunately, very fine-grained, low-permeability soils, especially if they have some plasticity,
rarely present a filtration problem unless they are dispersive (Sherard and Decker, 1977) or
FHWA NHI-07-092
Geosynthetics Engineering

2-13

Subsurface Drainage
August 2008

subject to hydraulic fracturing, such as might occur in dams under high hydraulic gradients
(Sherard, 1986).
Again, we emphasize that these filtration or clogging potential tests are performance tests.
They must be conducted on samples of project site soil by the specifying agency or its
representative. These tests are the responsibility of the engineer because manufacturers
generally do not have soil laboratories or samples of on-site soils. Therefore, realistically,
the manufacturers are unable to certify the clogging resistance of a geotextile.

Figure 2-4.

U.S. Army Corps of Engineers gradient ratio test device.

FHWA NHI-07-092
Geosynthetics Engineering

2-14

Subsurface Drainage
August 2008

2.4-4 Survivability and Durability Criteria


To be sure the geotextile will survive the construction process, certain geotextile strength and
durability properties are required for filtration and drainage applications. The minimum
requirements are given in Table 2-2.
It is important to realize that these minimum survivability values are not based on any
systematic research, but on the properties of existing geotextiles which are known to have
performed satisfactorily in drainage applications. The values are meant to serve as guidelines
for inexperienced users in selecting geotextiles for routine projects. They are not intended to
replace site-specific evaluation, testing, and design.
For critical and/or severe
applications, survivability should be confirmed by test section and exhumation, or
experience with similar installations and construction procedures.
Table 2-2
Geotextile Strength Property Requirements1,2,3,4 for Drainage Geotextiles
(after AASHTO, 2006)
Test
Methods

Units

Grab strength

ASTM D 4632

Sewn seam strengthd

Geotextile Class 2a,b


Elongation< 50%c

Elongation> 50%c

lb (N)

250 (1100)

157 (700)

ASTM D 4632

lb (N)

220 (990)

140 (630)

Tear strength

ASTM D 4533

lb (N)

90 (400) e

56 (250)

Puncture strength

ASTM D 6241

lb (N)

495 (2200)

309 (1375)

b
c
d
e

Required geotextile class is designated in M288 Tables 2, 3, 4, 5, or 6 for the indicated application. The
severity of installation conditions for the application generally dictate the required geotextile class. Class 1 is
specified for more severe or harsh installation conditions where there is a greater potential for geotextile
damage, and Classes 2 and 3 are specified for less severe conditions.
All numeric values represent MARV in the weaker principal direction. (See M 288 Section 8.1.2)
As measured in accordance with ASTM D 4632.
When sewn seams are required. Refer to M288 Appendix for overlap seam requirements.
The required MARV tear strength for woven monofilament geotextiles is 56 lb (250 N).

NOTES:
1. Acceptance of geotextile material shall be based on ASTM D 4759.
2. Acceptance shall be based upon testing of either conformance samples obtained using Procedure
A of ASTM D 4354, or based on manufacturers certifications and testing of quality assurance
samples obtained using Procedure B of ASTM D 4354.
3. Minimum; use value in weaker principal direction. All numerical values represent minimum
average roll value (i.e., test results from any sampled roll in a lot shall meet or exceed the
minimum values in the table). Lot samples according to ASTM D 4354.
4. Woven slit film geotextiles will not be allowed.

FHWA NHI-07-092
Geosynthetics Engineering

2-15

Subsurface Drainage
August 2008

Geotextile durability relates to its longevity. Geotextiles have been shown to be basically
inert materials for most environments and applications. However, certain applications may
expose the geotextile to chemical or biological activity that could drastically influence its
filtration properties or durability. For example, in drains, granular filters and geotextiles can
become chemically clogged by iron or carbonate precipitates, and biologically clogged by
algae, mosses, etc. Biological clogging is a potential problem when filters and drains are
periodically inundated then exposed to air. Excessive chemical and biological clogging can
significantly influence filter and drain performance. These conditions are present, for
example, in landfills.
Biological clogging potential can be examined with ASTM D 1987, Standard Test Method
for Biological Clogging of Geotextile or Soil/Geotextile Filters (1991). If biological
clogging is a concern, a higher-porosity geotextile may be used, and/or the drain design and
operation can include an inspection and maintenance program to flush the drainage system.
2.4-5 Additional Filter Selection Considerations and Summary
Several different geotextiles, ranging from monofilament wovens to an array of light- to
heavy-weight nonwovens, may meet all of the desired design criteria. Depending on the
actual soil and hydraulic conditions, as well as the intended function of the design, it may be
preferable to use one type of geotextile over another to enhance system performance.
Intuitively, the following observations and selection considerations seem appropriate for
these soil conditions:
1. Graded gravels and coarse sands -- Very open monofilament or multifilament woven
geotextiles may be required to permit high rates of flow and low-risk of blinding.
2. Sands and gravels with less than 20% fines -- Open monofilament woven and
needlepunched nonwoven geotextiles with large openings are preferable to reduce the
risk of blinding. For thin, heat-bonded geotextiles and thick, needlepunched
nonwoven geotextiles, filtration tests should be performed.
3. Soils with 20% to 60% fines -- Filtration tests should be performed on all types of
geotextiles especially for critical applications or severe conditions.
4. Soils with greater than 60% fines -- Heavy-weight, needlepunched geotextiles and
heat-bonded geotextiles tend to work best as fines will not pass. If blinding does
occur, the permeability of the blinding cake would equal that of the soil.
5. Gap-graded cohesionless soils -- Consider using a uniform sand filter with a very
open geotextile designed to allow fines to pass.
6. Silts with sand seams -- Consider using a uniform sand filter over the soil with a very
open geotextile, designed to allow the silt to pass but to prevent movement of the
filter sand; alternatively, consider using a heavy-weight (> 10 oz/yd2 {350 g/m2})
FHWA NHI-07-092
Geosynthetics Engineering

2-16

Subsurface Drainage
August 2008

needlepunched nonwoven directly against soil so water can flow laterally through the
geotextile should it become locally clogged.
The above general observations are not meant to serve as recommendations, but are
offered to provide insight for selecting optimum materials. They are not intended to
exclude other possible geotextiles that you may want to consider.
Figure 2-5 is a flow chart summarizing the FHWA filter design process.
For geosynthetics in pavement drainage systems, the requirements can also be evaluated
using the FHWA computer program DRIP along with the effectiveness of the drainage
system and calculate the design requirements for the permeable base design, separator, and
edgedrain design, including retention and permeability requirements. The software can be
downloaded directly from the FHWA Webpage http://www.fhwa.dot.gov/pavement/library
and is included with the NCHRP 1-37A pavement design software.

FHWA NHI-07-092
Geosynthetics Engineering

2-17

Subsurface Drainage
August 2008

O95 < B D85

For
2 < CU < 4
B = 0.5 CU
For
4 < CU < 8
B = 8 / CU

Use
For CU < 3
maximum O95 from
Retention Criteria

Subsurface Drainage
August 2008

SURVIVABILITY and ENDURANCE CRITERIA

2-18

Performance Tests
to Select Suitable
Geotextile

Unstable Soils

Perform filtration test


with on-site soils and
hydraulic conditions

For critical applications and severe conditions:

> 50%
> 0.1 sec-1

Optional Qualifiers for gap-graded or silty soils


For Nonwovens: n > 50%
For Woven monofilament and silt films: POA > 4%

CLOGGING RESISTANCE

Flow chart summary of the FHWA filter design procedure.

FHWA NHI-07-092
Geosynthetics Engineering

Figure 2-5.

For CU > 3
O95 > 3 D15

15% to 50%
> 0.2 sec-1

qrequired = qgeotextile (Ag/At)

< 15%
> 0.5 sec-1

For less critical applications and less severe conditions:

% Passing #200 sieve:


Permittivity Required:

kgeotextile > ksoil

O95 < 0.5 D85

Dynamic Flow

For critical applications and severe


conditions: kgeotextile > 10 ksoil

PERMEABILITY/PERMITTIVITY CRITERIA

B = 1.8 &
O95 < 1.8 D85
and
O95 < 0.3 mm

Nonwovens

B=1&
O95 < D85

Silts and Clays


( > 50% passing No.200 sieve )

Wovens

For less critical applications and less severe conditions:

For
2 > CU > 8
B=1

Sands, Gravelly Sands, Silty Sands & Clayey


Sands (< 50% passing No.200 sieve )

Steady State Flow

RETENTION CRITERIA

2.5

DRAINAGE SYSTEM DESIGN GUIDELINES

In this section, step-by-step drainage design procedures are given. As with a chain, the
integrity of the resulting design will depend on its weakest link; thus, no steps should be
compromised or omitted.
STEP 1.

Evaluate the critical nature and site conditions (see Table 2-1) of the application.

Reasonable judgment should be used in categorizing a project, since there may be a


significant cost difference for geotextiles required for critical/severe conditions. Final
selection should not be based on the lowest material cost alone, nor should costs be
reduced by eliminating laboratory soil-geotextile performance testing, if such testing
is appropriate.
STEP 2.
A.

Obtain soil samples from the site, and:

Perform grain size analyses.

Calculate Cu = D60/D10

Select the worst case soil for retention (i.e., usually the soil with smallest B x D85)

(Eq. 2 - 3)

NOTE: When the soil contains particles 1-in. (25 mm) and larger, use only the gradation
of soil passing the No.4 (4.75 mm) sieve in selecting the geotextile (i.e., scalp off the +
No.4 (+4.75 mm) material).
B.

Perform field or laboratory permeability tests.

Select worst case soil (i.e., soil with highest coefficient of permeability, k).

The permeability of clean sands with 0.1 mm < D10 < 3 mm and Cu < 5 can be
estimated by the Hazen formula, k = (D10)2 (k in cm/sec; D10 in mm). This formula
should not be used for soils with more than about 5% fines.

Laboratory tests for permeability (hydraulic conductivity) are detailed in ASTM D


2434 for granular soils, D 5856 using a compaction-mold permeameter, and in D
5084 using a flexible-wall permeameter for soils with appreciable fines. Field tests
include pumping tests in boreholes and infiltrometer tests. Standard procedures for
several field tests are also in ASTM.

FHWA NHI-07-092
Geosynthetics Engineering

2-19

Subsurface Drainage
August 2008

A good visual classification of the soils a the site will enable an experienced
geotechnical engineer to estimate the permeability to the nearest order of magnitude,
which is often sufficient for geotextile filter design. The following table, adapted
from Casagrande (1938) and Holtz and Kovacs (1981), will give a range of
hydraulic conductivities for different natural soils.
Visual Classification
Clean gravel

C.

Permeability or Hydraulic
Conductivity, k, m/sec
> 0.01

Clean sands and clean sand-gravel mixtures

0.01 < k < 10-5

Very fine sands; silts; mixtures of sand, silt,


and clay; glacial tills; stratified clays
Impervious soils; homogeneous reasonably
intact clays from below zone of weathering
Impervious soils, modified by vegetation,
weathering, fissured, highly OC clays

10-5 < k < 10-9


k > 10-9
5 x 10-5 < k < 5 x 10-8

Select drainage aggregate.

Use free-draining, open-graded material and estimate its permeability (e.g., use
Figure 2-6). If possible, sharp, angular aggregate should be avoided. If it must be
used, then a geotextile meeting the property requirements for high survivability in
Table 2-2 should be specified. For an accurate cost comparison, compare cost of
open-graded aggregate with well-graded, free-draining filter aggregate.

STEP 3.

Calculate anticipated flow into and through drainage system and dimension the
system. Use collector pipe to reduce size of drain.

A. General Case
Use Darcy's Law
q = kiA

[2 - 13]

where:
q =
k =
i =
A=

infiltration rate (m3/sec)


effective permeability of soil (from Step 2B above) (m/sec)
average hydraulic gradient in soil and in drain (m/m)
area of soil and drain material normal to the direction of flow (m2)

FHWA NHI-07-092
Geosynthetics Engineering

2-20

Subsurface Drainage
August 2008

Figure 2-6. Typical gradations and Darcy permeabilities of several aggregate and graded
filter materials (U.S. Navy, 1986).

Use a conventional flow net analysis to calculate the hydraulic gradient (Cedergren, 1989)
and Darcy's Law for estimating infiltration rates into the drain; then use Darcy's Law to
design the drain (i.e., calculate cross-sectional area A for flow through open-graded
aggregate). Note that typical values of hydraulic gradients in the soil adjacent to a geotextile
filter (Giroud, 1988) are:
i < 1 for drainage under roads, embankments, slopes, etc., when the main source of
water is precipitation; and
i = 1.5 in the case of drainage trenches and vertical drains behind walls.
B.

Specific Drainage Systems


Estimates of surface infiltration, runoff infiltration rates, and drainage dimensions can
be determined using accepted principles of hydraulic engineering (Moulton, 1980).
Specific references are:
1. Flow into trenches -- Mansur and Kaufman (1962)
2. Horizontal blanket drains -- Cedergren (1989)
3. Slope drains -- Cedergren (1989)

FHWA NHI-07-092
Geosynthetics Engineering

2-21

Subsurface Drainage
August 2008

C.

Pavement Drainage Systems


The DRIP microcomputer program developed by FHWA can be used to rapidly
evaluate the effectiveness of the drainage system and calculate the design
requirements for the permeable base design, separator, and edgedrain design,
including geotextile filtration requirements. The program can also be used to
determine the drainage path length based on pavement cross and longitudinal slopes,
lane widths, edgedrain trench widths (if applicable), and cross-section geometry
crowned or superelevated. As indicated in Section 2.3, the software can be
downloaded directly from http://www.fhwa.dot.gov/pavement/library.htm and is
included with the NCHRP 1-37A pavement design software.

STEP 4.

Determine geotextile requirements.

A. Retention Criteria Steady State Flow


From Step 2A, obtain D85 and Cu; then determine largest pore size allowed.
AOS < B D85

(Eq. 2 - 1)

where:
For soils with < 50% passing the 0.075 mm sieve:
B=1
for Cu < 2 or > 8
B = 0.5 Cu
for 2 < Cu < 4
B = 8/Cu
for 4 < Cu < 8

(Eq. 2 - 2a)
(Eq. 2 - 2b)
(Eq. 2 - 2c)

and, for soils with > 50% passing the 0.075 mm sieve:
B = 1 for woven geotextiles,
B = 1.8 for nonwoven geotextiles,
and AOS (geotextile) < 0.3 mm

(Eq. 2 - 5)

NOTE: Soils with a Cu of greater than 20 may be unstable (see section 2.41.c): if so, filtration tests should be conducted to select suitable geotextiles.
For dynamic and cyclic flow conditions,
O95 < 0.5 D85

FHWA NHI-07-092
Geosynthetics Engineering

(Eq. 2 -6)

2-22

Subsurface Drainage
August 2008

B. Permeability/Permittivity Criteria
1. Less Critical/Less Severe
kgeotextile > ksoil

(Eq. 2 - 7a)

2. Critical/Severe
kgeotextile > 10 ksoil

(Eq. 2 - 7b)

3. Permittivity Requirements for all criticality and severity conditions


For < 15% passing No.200 (0.075 mm)
0.5 sec-1 (Eq. 2 - 8a)
For 15 to 50% passing No.200 (0.075 mm) 0.2 sec-1 (Eq. 2 - 8b)
For > 50% passing No.200 (0.075 mm)
0.1 sec-1 (Eq. 2 - 8c)
4. Flow Capacity Requirement
qrequired = qgeotextile/(Ag/At), or
(kgeotextile/t) h Ag > qrequired

(Eq. 2 - 9)
[2 14]

where:
qrequired is obtained from STEP 3B (Eq. 2-13) above;
kgeotextile
= permeability of the geotextile;
t
= geotextile thickness;
h
= average head in field;
= geotextile area available for flow (i.e., if 80% of geotextile is
Ag
covered by the wall of a pipe, Ag = 0.2 x total area); and
= total area of geotextile.
At
C. Clogging Criteria
1. Less Critical/Less Severe
a. From Step 2A obtain D15; then determine minimum pore size requirement from
O95 > 3 D15, for Cu > 3
(Eq. 2 - 10)
b. Other qualifiers:
Nonwoven geotextiles:
(Eq. 2 - 11)
Porosity (geotextile) > 50%
Woven geotextiles:
Percent open area > 4%
(Eq. 2 - 12)
Alternative: Run filtration tests

FHWA NHI-07-092
Geosynthetics Engineering

2-23

Subsurface Drainage
August 2008

2. Critical/Severe
Select geotextiles that meet retention, permeability, and survivability criteria,
as well as the criteria in Step 4C.1 above, and perform a filtration test.
Suggested filtration test for sandy soils is the gradient ratio test. The
hydraulic conductivity ratio test is recommended by ASTM for fine-grained
soils, but as noted in Sections 1.5 and 2.4-3, the HCR test has serious
disadvantages.
Alternative: Consider long-term filtration tests, F3 tests, the Flexible Wall GR
test etc.
NOTE: Experience is required to obtain reproducible results from
the gradient ratio test. See Fischer (1994) and Mar (1994).
D. Survivability
Select geotextile properties required for survivability from Table 2-2. Add durability
requirements if appropriate.

STEP 5.

Estimate costs.

Calculate the pipe size (if required), the volume of aggregate, and the area of the
geotextile. Apply appropriate unit cost values.
Pipe (if required) (m)
________________
3
Aggregate (/m )
________________
2
________________
Geotextile (/m )
2
Geotextile placement (/m )
Construction (LS)
________________
Total Cost:
________________
STEP 6.

Prepare specifications.
Include for the geotextile:
A. General requirements
B. Specific geotextile properties
C. Seams and overlaps
D. Placement procedures
E. Repairs
F. Testing and placement observation requirements

FHWA NHI-07-092
Geosynthetics Engineering

2-24

Subsurface Drainage
August 2008

See Sections 1.6 and 2.7 for specification details.


STEP 7.

Collect samples of aggregate and geotextile before acceptance.

STEP 8.

Monitor installation during and after construction.

STEP 9.

Observe drainage system during and after storm events.

2.6

DESIGN EXAMPLE

DEFINITION OF DESIGN EXAMPLE

Project Description:

drains to intercept groundwater are to be placed adjacent to a twolane highway

Type of Structure:

trench drain

Type of Application:

geotextile wrapping of aggregate drain stone

Alternatives:

i) graded soil filter between aggregate and soil being drained; or


ii) geotextile wrapping of aggregate

GIVEN DATA

site has a high groundwater table

drain is to prevent seepage and shallow slope failures, which are currently a maintenance
problem

depth of trench drain is 3 ft (1 m)

soil samples along the proposed drain alignment are nonplastic

gradations of three representative soil samples along the proposed drain alignment

FHWA NHI-07-092
Geosynthetics Engineering

2-25

Subsurface Drainage
August 2008

PERCENT PASSING, BY WEIGHT


SIEVE SIZE
(mm)
25
13
4.76
1.68
0.84
0.42
0.15
0.074

Sample A

Sample B

Sample C

99
97
95
90
78
55
10
1

100
100
100
96
86
74
40
15

100
100
100
100
93
70
11
0

Grain Size Distribution Curve


DEFINE
A. Geotextile function(s)
B. Geotextile properties required
C. Geotextile specification

FHWA NHI-07-092
Geosynthetics Engineering

2-26

Subsurface Drainage
August 2008

SOLUTION
A. Geotextile function(s):
Primary
Secondary
-

filtration
separation

B. Geotextile properties required:


apparent opening size (AOS)
permeability/permittivity
survivability

DESIGN
STEP 1.

EVALUATE CRITICAL NATURE AND SITE CONDITIONS

From given data, assume that this is a noncritical application.


Soils are well-graded, hydraulic gradient is low for this type of application, and flow
conditions are steady state for this type of application.

STEP 2.

OBTAIN SOIL SAMPLES

A.
GRAIN SIZE ANALYSES
Plot gradations of representative soils. The D60, D10, and D85 sizes from the gradation plot are
noted in the table below for Samples A, B, and C. Determine uniformity coefficient, Cu,
coefficient B, and the maximum AOS.
Worst case soil for retention (i.e., smallest B D85) is Soil C, from the following table.

Soil
Sample
A
B
C

D60 D10 = Cu
0.48 0.15 = 3.2
0.25 0.06 = 4.2
0.36 0.14 = 2.6

B=

AOS (mm) < B x D85

0.5 Cu = 0.5 3.2 = 1.6


8 Cu = 8 4.2 = 1.9
0.5 Cu = 0.5 2.6 = 1.3

1.6 1.0 = 1.6


1.9 0.75 = 1.4
1.3 0.55 = 0.72

B.
PERMEABILITY TESTS
Noncritical application, drain will be conservatively designed with an estimated permeability.

FHWA NHI-07-092
Geosynthetics Engineering

2-27

Subsurface Drainage
August 2008

The largest D10 controls permeability; therefore, Soil A with D10 = 0.15 mm controls. For
this example, we will use the Hazen formula, or
k (D10)2 = (0.15)2 = 2 (10)-2 cm/sec = 2 (10)-4 m/sec
Note that this value is a conservative estimate in terms of the visual classification of the soil,
as discussed in Section 2.5, Step 2.
C.
SELECT DRAIN AGGREGATE
Assume drain stone is a rounded aggregate.

STEP 3.
DIMENSION DRAIN SYSTEM
Determine depth and width of drain trench and whether a pipe is required to carry flow - details
of which are not included within this example.

STEP 4.

DETERMINE GEOTEXTILE REQUIREMENTS

A.
RETENTION CRITERIA
Sample C controls (see table above), therefore, AOS < 0.72 mm
B.
PERMEABILITY CRITERIA
From given data, it has been judged that this application is a less critical/less severe
application. Therefore, kgeotextile > ksoil
Soil C controls, therefore, kgeotextile > 2 (10)-4 m/sec
Flow capacity requirements of the system - details of which are not included within this
example.
C.
PERMITTIVITY CRITERIA
All three soils have < 15% passing the 0.075 mm, therefore

> 0.5 sec-1

D.
CLOGGING CRITERIA
From given data, it has been judged that this application is a less critical/less severe
application, and Soils A and B have a Cu greater than 3. Therefore, for soils A and B, O95 >
3 D15. So
O95

3 x 0.15 = 0.45 mm for Sample A


3 x 0.075
= 0.22 mm for Sample B

Soil A controls [Note that sand size particles typically don't create clogging problems,
FHWA NHI-07-092
Geosynthetics Engineering

2-28

Subsurface Drainage
August 2008

therefore, Soil B could have been used as the design control.], therefore,
AOS > 0.45 mm
For Soil C, a geotextile with the maximum AOS value determined from the retention criteria
should be used. Therefore
AOS 0.72 mm
Also,
nonwoven porosity > 50%
and
woven percent open area > 4%
For the primary function of filtration, the geotextile should have 0.45 mm < AOS < 0.72
mm; and kgeotextile > 2 (10)-2 cm/sec and > 0.5 sec-1. Woven slit film geotextiles are not
allowed.

E.
SURVIVABILITY
From Table 2-2, the following minimum values are recommended:
For Survivability, the geotextile shall have the following minimum values (values are
MARV)
Woven Geotextile

Nonwoven Geotextile

250 lb (1100 N)

157 lb (700 N)

Sewn Seam Strength

220 lb (990 N)

140 lb (630 N)

Tear Strength

90 lb (400 N)*

56 lb (250 N)

Puncture Strength

495 lb (2200 N)

309 lb (1375 N)

Grab Strength

*56 lb (250 N) for monofilament geotextiles


NOTE: With lightweight compaction equipment and field inspection, Class 3
geotextile (see Appendix D) could be used.

Complete Steps 5 through 9 to finish design.

FHWA NHI-07-092
Geosynthetics Engineering

2-29

Subsurface Drainage
August 2008

STEP 5.

ESTIMATE COSTS

STEP 6.

PREPARE SPECIFICATIONS

STEP 7.

COLLECT SAMPLES

STEP 8.

MONITOR INSTALLATION

STEP 9.

OBSERVE DRAIN SYSTEM DURING AND AFTER STORM EVENTS

2.7

COST CONSIDERATIONS

Determining the cost effectiveness of geotextiles versus conventional drainage systems is a


straightforward process. Simply compare the cost of the geotextile with the cost of a
conventional granular filter layer, while keeping in mind the following:
Overall material costs including a geotextile versus a conventional system - For
example, the geotextile system will allow the use of poorly graded (less-select)
aggregates, which may reduce the need for a collector pipe, provided the amount
of fines is small (Q decreases considerably if the percent passing the No.200
(0.075 mm) sieve is greater than 5%, even in gravel).
Construction requirements - There is, of course, a cost for placing the geotextile;
but in most cases, it is less than the cost of constructing dual-layered, granular
filters, for example, which are often necessary with conventional filters and finegrained soils.
Possible dimensional design improvements - If an open-graded aggregate is used
(especially with a collector pipe), a considerable reduction in the physical
dimensions of the drain can be made without a decrease in flow capacity. This
size reduction also reduces the volume of the excavation, the volume of filter
material required, and the construction time necessary per unit length of drain.
In general, the cost of the geotextile material in drainage applications will typically range
from $1.00 to $1.50 per square yard, depending upon the type specified and quantity ordered.
Installation costs will depend upon the project difficulty and contractor's experience;
typically, they range from $0.50 to $1.50 per square yard of geotextile. Higher costs should
be anticipated for below-water placement. Labor installation costs for the geotextile are
easily repaid because construction can proceed at a faster pace, less care is needed to prevent
segregation and contamination of granular filter materials, and multilayered granular filters
are typically not necessary.
FHWA NHI-07-092
Geosynthetics Engineering

2-30

Subsurface Drainage
August 2008

2.8

SPECIFICATIONS

The following guide specification is provided as an example. It is a combination of the


AASHTO M288 (2006) geotextile material specification and its accompanying
construction/installation guidelines; developed for routine drainage and filtration
applications. The actual hydraulic and physical properties of the geotextile must be selected
by considering of the nature of the project (critical/less critical), hydraulic conditions
(severe/less severe), soil conditions at the site, and construction and installation procedures
appropriate for the project.

SUBSURFACE DRAINAGE GEOTEXTILES


(after AASHTO M288, 2006)
1. SCOPE
1.1 Description. This specification is applicable to placing a geotextile against the soil to allow longterm passage of water into a subsurface drain system retaining the in-situ soils. The primary
function of the geotextile in subsurface drainage applications is filtration. Geotextile filtration
properties are a function of the in-situ soil gradation, plasticity, and hydraulic conditions.
2. REFERENCED DOCUMENTS
2.1 AASHTO Standards
T88
T90
T99

Particle Size Analysis of Soils


Determining the Plastic Limit and Plasticity Index of Soils
The Moisture-Density Relationships of Soils Using a 5.5 lb (2.5 kg) Rammer and a
12 in. (305 mm) Drop
2.2 ASTM Standards
D 123
Standard Terminology Relating to Textiles
D 276
Test Methods for Identification of Fibers in Textiles
D 4354 Practice for Sampling of Geosynthetics for Testing
D 4355
Test Method for Deterioration of Geotextiles from Exposure to Ultraviolet Light and
Water (Xenon Arc Type Apparatus)
D 4439
Terminology for Geosynthetics
D 4491
Test Methods for Water Permeability of Geotextiles by Permittivity
D 4632
Test Method for Grab Breaking Load and Elongation of Geotextiles
D 4751
Test Method for Determining Apparent Opening Size of a Geotextile
D 4759
Practice for Determining the Specification Conformance of Geosynthetics
D 4873
Guide for Identification, Storage, and Handling of Geotextiles
FHWA NHI-07-092
Geosynthetics Engineering

2-31

Subsurface Drainage
August 2008

D 5141
D 6241

Test Method to Determine Filtering Efficiency and Flow Rate for Silt Fence
Application of a Geotextile Using Site Specific Soil
Test Method for Static Puncture Strength of Geotextiles and Geotextile
Related Products Using a 50-mm Probe

3. PHYSICAL AND CHEMICAL REQUIREMENTS


3.1 Fibers used in the manufacture of geotextiles and the threads used in joining geotextiles by
sewing, shall consist of long chain synthetic polymers, composed of at least 95% by weight
polyolefins or polyesters. They shall be formed into a stable network such that the filaments or
yarns retain their dimensional stability relative to each other, including selvages.
3.2 Geotextile Requirements. The geotextile shall meet the requirements of following Table. Woven
slit film geotextiles (i.e., geotextiles made from yarns of a flat, tape-like character) will not be
allowed. All numeric values in the following table, except AOS, represent minimum average roll
values (MARV) in the weakest principal direction (i.e., average test results of any roll in a lot
sampled for conformance or quality assurance testing shall meet or exceed the minimum values).
Values for AOS represent maximum average roll values.
NOTE: The property values in the following table represent default values which
provide for sufficient geotextile survivability under most conditions. Minimum
property requirements may be reduced when sufficient survivability information is
available [see Note 2 of Table 2-2 and Appendix D]. The Engineer may also specify
properties different from those listed in the following Table based on engineering
design and experience.
Subsurface Drainage Geotextile Requirements
Elongation(1)
ASTM Test
Property
Units
Method
< 50%(1)
> 50%(1)
Grab Strength
D 4632
N
1100
700
(2)
Sewn Seam Strength
D 4632
N
990
630
Tear Strength
D 4533
N
400(3)
250
Puncture Strength
D 6241
N
2200
1375
Percent In-Situ Passing 0.075 mm Sieve(4)
< 15
15 to 50
> 50
-1
Permittivity
D 4491
sec
0.5
0.2
0.1
Apparent Opening Size
D 4751
mm
0.43
0.25
0.22(5)
Ultraviolet Stability
D 4355
%
50% after 500 hours of exposure
NOTES:
(1)
As measured in accordance with ASTM D 4632.
(2)
When sewn seams are required.
(3)
The required MARV tear strength for woven monofilament geotextiles is 250 N.
(4)
Based on grain size analysis of in-situ soil in accordance with AASHTO T88.
(5)
For cohesive soils with a plasticity index greater than 7, geotextile maximum average roll value
for apparent opening size is 0.30 mm.
FHWA NHI-07-092
Geosynthetics Engineering

2-32

Subsurface Drainage
August 2008

4. CERTIFICATION
4.1 The Contractor shall provide to the engineer a certificate stating the name of the manufacturer,
product name, style number, chemical composition of the filaments or yarns and other pertinent
information to fully describe the geotextile.
4.2 The manufacturer is responsible for establishing and maintaining a quality control program to
assure compliance with the requirements of the specification. Documentation describing the
quality control program shall be made available upon request.
4.3 The manufacturers certificate shall state that the furnished geotextile meets MARV requirements
of the specification as evaluated under the manufacturers quality control program. A person
having legal authority to bind the manufacturer shall be attest to the certificate.
4.4 Either mislabeling or misrepresentation of materials shall be reason to reject those geotextile
products.
5. SAMPLING, TESTING, AND ACCEPTANCE
5.1 Geotextiles shall be subject to sampling and testing to verify conformance with this specification.
Sampling shall be in accordance with the most current ASTM D 4354 using the section titled,
Procedure for Sampling for Purchasers Specification Conformation Testing. In the absence of
purchasers testing, verification may be based on manufacturers certifications as a result of a
testing by the manufacturer of quality assurance samples obtained using he procedure for
Sampling or Manufacturers Quality Assurance (MQA) Testing. A lot size shall be considered to
be the shipment quantity of the given product or a truckload of the given product, whichever is
smaller.
5.2 Testing shall be performed in accordance with the methods referenced in this specification for the
indicated application. The number of specimens to test per sample is specified by each test
method. Geotextile product acceptance shall be based on ASTM D 4759. Product acceptance is
determined by comparing the average test results of all specimens within a given sample to the
specification MARV. Refer to ASTM D 4759 for more details regarding geotextile acceptance
procedures.
6. SHIPMENT AND STORAGE
6.1 Geotextile labeling, shipment, and storage shall follow ASTM D 4873. Product labels shall
clearly show the manufacturer or supplier name, style number, and roll number. Each shipping
document shall include a notation certifying that the material is in accordance with the
manufacturers certificate.

FHWA NHI-07-092
Geosynthetics Engineering

2-33

Subsurface Drainage
August 2008

6.2 Each geotextile roll shall be wrapped with a material that will protect the geotextile, including the
ends of the roll, from damage due to shipment, water, sunlight, and contaminants. The protective
wrapping shall be maintained during periods of shipment and storage.
6.3 During storage, geotextile rolls shall be elevated off the ground and adequately covered to protect
them from the following: site construction damage, precipitation, extended ultraviolet radiation
including sunlight, chemicals that are strong acids or strong bases, flames including welding
sparks, temperatures in excess of 71C (160F), and any other environmental condition that may
damage the physical property values of the geotextile.
7.

CONSTRUCTION

7.1 General. Atmospheric exposure of geotextiles to the elements following lay down shall be a
maximum of 14 days to minimize damage potential.
7.2 Seaming.
a. If a sewn seam is to be used for the seaming of the geotextile, the thread used shall consist of
high strength polypropylene, or polyester. Nylon thread shall not be used. For erosion control
applications, the thread shall also be resistant to ultraviolet radiation. The thread shall be of
contrasting color to that of the geotextile itself.
b. For seams which are sewn in the field, the contractor shall provide at least a two m length of
sewn seam for sampling by the engineer before the geotextile is installed. For seams that are
sewn in the factory, the engineer shall obtain samples of the factory seams at random from any
roll of geotextile which is to be used on the project.
b.1 For seams that are field sewn, the seams sewn for sampling shall be sewn using the same
equipment and procedures as will be used for the production of seams. If seams are to be
sewn in both the machine and cross machine directions, samples of seams from both
directions shall be provided.
b.2 The Contractor shall submit the seam assembly description along with the sample of the
seam. The description shall include the seam type, stitch type, sewing thread, and stitch
density.
7.3 Trench. Trench excavation shall be done in accordance with details of the project plans. In all
instances excavation shall be done in such a way so as to prevent large voids from occurring in
the sides and bottom of the trench. The graded surface shall be smooth and free and debris.
7.4 Geotextile Placement.
a. In placement of the geotextile for drainage applications, the geotextile shall be placed loosely
with no wrinkles or folds, and with not void spaces between the geotextile and the ground
FHWA NHI-07-092
Geosynthetics Engineering

2-34

Subsurface Drainage
August 2008

surface. Successive sheets of geotextiles shall be overlapped a minimum of 300 mm, with the
upstream sheet overlapping the downstream sheet.
a.1 In trenches equal to or greater than 300 mm in width, after placing the drainage aggregate the
geotextile shall be folded over the top of the backfill material in a manner to produce a
minimum overlap of 300 mm. In trenches less than 300 mm but greater than 100 mm wide,
the overlap shall be equal to the width of the trench. Where the trench is less than 100 mm
the geotextile overlap shall be sewn or otherwise bonded. All seams shall be subject to the
approval of the engineer.
a.2 Should the geotextile be damaged during installation, or drainage aggregate placement, a
geotextile patch shall be placed over the damaged area extending beyond the damaged area a
distance of 300 mm, or the specified seam overlap, whichever is greater.
7.5 Drainage Aggregate
a. Placement of drainage aggregate should proceed immediately following placement of the
geotextile. The geotextile should be covered with a minimum of 300 mm of loosely placed
aggregate prior to compaction. If a perforated collector pipe is to be installed in the trench, a
bedding layer of drainage aggregate should be placed below the pipe, with the remainder of the
aggregate placed to the minimum required construction depth.
a.1 The aggregate should be compacted with vibratory equipment to a minimum of 95% Standard
AASHTO density unless the trench is required for structural support. If higher compactive
effort is required, a Class 1 geotextile as per Table 1 of the M288 Specification is needed.
8. METHOD OF MEASUREMENT
8.1 The geotextile shall be measured by the number of square meters computed from the
payment lines shown on the plans or from payment lines established in writing by the Engineer.
This excludes seam overlaps, but shall include geotextiles used in crest and toe of slope
treatments.
8.2 Slope preparation, excavation and backfill, bedding, and cover material are separate pay
items.
9. BASIS OF PAYMENT
9.1 The accepted quantities of geotextile shall be paid for per square meter in place.
9.2 Payment will be made under:
Pay Item
Subsurface Drainage Geotextile
FHWA NHI-07-092
Geosynthetics Engineering

Pay Unit
Square Meter
2-35

Subsurface Drainage
August 2008

2.9

INSTALLATION PROCEDURES

For all drainage applications, the following construction steps should be followed:
1. The surface on which the geotextile is to be placed should be excavated to design
grade to provide a smooth, graded surface free of debris and large cavities.
2. Between preparation of the subgrade and construction of the system, the geotextile
should be well-protected to prevent any degradation due to exposure to the elements.
3. After excavating to design grade, the geotextile should be cut (if required) to the
desired width (including allowances for non-tight placement in trenches and overlaps
of the ends of adjacent rolls) or cut at the top of the trench after placement of the
drainage aggregate.
4. Care should be taken during construction to avoid contamination of the geotextile. If
it becomes contaminated, it must be removed and replaced with new material.
5. In drainage systems, the geotextile should be placed with the machine direction
following the direction of water flow; for pavements, the geotextile should be parallel
to the roadway. It should be placed loosely (not taut), but with no wrinkles or folds.
Care should be taken to place the geotextile in intimate contact with the soil so that no
void spaces occur behind it.
6. The ends for subsequent rolls and parallel rolls of geotextile should be overlapped a
minimum of 1 foot (0.3 m) in roadways and 1 to 2 feet (0.3 to 0.6 m) in drains,
depending on the anticipated severity of hydraulic flow and the placement conditions.
For high hydraulic flow conditions and heavy construction, such as with deep
trenches or large stone, the overlaps should be increased. For large open sites using
base drains, overlaps should be pinned or anchored to hold the geotextile in place
until placement of the aggregate. Upstream geotextile should always overlap over
downstream geotextile.
7. To limit exposure of the geotextile to sunlight, dirt, damage, etc., placement of
drainage or roadway base aggregate should proceed immediately following placement
of the geotextile. The geotextile should be covered with a minimum of 1 foot (0.3 m)
of loosely placed aggregate prior to compaction. If thinner lifts are used, higher
survivability fabrics may be required. For drainage trenches, at least 4 in. (0.1 m) of
drainage stone should be placed as a bedding layer below the slotted collector pipe (if
required), with additional aggregate placed to the minimum required construction
depth. Compaction is necessary to seat the drainage system against the natural soil
and to reduce settlement within the drain. The aggregate should be compacted with
vibratory equipment to a minimum of 95% Standard AASHTO T99 density unless the
trench is required for structural support. If higher compactive efforts are required, the
geotextiles meeting the property values listed under the high survivability category in
Table 2-2 should be utilized.
FHWA NHI-07-092
Geosynthetics Engineering

2-36

Subsurface Drainage
August 2008

8. After compaction, for trench drains, the two protruding edges of the geotextile should
be overlapped at the top of the compacted granular drainage material. A minimum
overlap of 1 foot (0.3 m) is recommended to ensure complete coverage of the trench
width. The overlap is important because it protects the drainage aggregate from
surface contamination. After completing the overlap, backfill should be placed and
compacted to the desired final grade.
A schematic of the construction procedures for a geotextile-lined underdrain trench is shown
in Figure 2-7. Construction photographs of an underdrain trench are shown in Figure 2-8,
and diagrams of geosynthetic placement beneath a permeable roadway base are shown in
Figure 2-9.

2.10

FIELD INSPECTION

The field inspector should review the field inspection guidelines in Section 1.7. Special
attention should be given to aggregate placement and potential for geotextile damage. Also,
maintaining the appropriate geotextile overlap at the top of the trench and at roll ends is
especially important.

especially important.

Figure 2-7.

Construction procedure for geotextile-lined underdrains.

FHWA NHI-07-092
Geosynthetics Engineering

2-37

Subsurface Drainage
August 2008

(a)

(b)

(c)

(d)

Figure 2-8.

Construction of geotextile drainage systems: (a) geotextile placement in


drainage ditch; (b) aggregate placement; (c) compaction of aggregate; and
(d) geotextile overlap prior to final cover.

FHWA NHI-07-092
Geosynthetics Engineering

2-38

Subsurface Drainage
August 2008

Figure 2-9.

Construction of geotextile filters and separators beneath permeable pavement


base: (a) geotextile used as a separator; and (b) permeable base and edge
drain combination. (Baumgardner, 1994)

FHWA NHI-07-092
Geosynthetics Engineering

2-39

Subsurface Drainage
August 2008

2.11

IN-PLANE DRAINAGE; PREFABRICATED GEOCOMPOSITE DRAINS

The in-plane drainage ability of geotextiles and prefabricated geocomposite drains is


potentially quite effective in several applications. Virtually all of the examples given in Sec.
2.1 have a lateral transmission component. Specific lateral drainage applications include
interceptor trench drains on slopes, drains behind abutments and retaining structures,
transmission of seepage water below pavement base course layers, pavement edge drains,
vertical drains to accelerate consolidation of soft foundation soils, dissipation of pore water
pressures in embankments and fills, dissipation of seepage forces in earth and rock slopes,
chimney drains in earth dams, leachate collection and gas venting systems for waste
containment systems, etc.
However, it should be realized that the flow quantities transmitted by in-plane flow of typical
geotextiles (on the order of 2 x 10-5 m3/s/linear meter of geotextile under a pressure
equivalent to 0.6 m of soil) are relatively small when compared to the flow capacity of only 6
to 12 in. (0.15 to 0.3 m) of filter sand. Therefore, geotextiles alone should only be used to
replace sand or other drainage layers in situations with small seepage quantities. Remember,
too, that the in-plane seepage quantities of geotextiles are highly affected by compressive
forces, incomplete saturation, and hydraulic gradients. These considerations have led
engineers to use geocomposite drains in many lateral drainage applications.
During the past 20 years or so, a large number of geocomposites drainage products have been
developed, which consist of cores of extruded and fluted plastics sheets, three-dimensional
meshes and mats, plastic waffles, and nets and channels to convey water and geotextiles on
one or both sides to act as a filter. Geocomposite drains may be fabricated on site although
most are manufactured. They generally range in thickness from to 1-in. (5 mm to 25 mm)
or greater and have transmission capabilities of between 0.0002 and 0.01 m3/sec/linear width
of drain. Some geocomposite drains are shown in Figure 2-10.
Prefabricated geocomposite drains are used to replace or support conventional drainage
systems. According to Hunt (1982), prefabricated drains offer a readily available material
with known filtration and hydraulic flow properties; easy installation, and, therefore,
construction economies; and protection of any waterproofing applied to the structure's
exterior. Cost of prefabricated drains typically ranges from $0.75 to $1.00 per square foot.
The high material cost is usually offset by expedient construction and reduction in required
quantities of select granular materials. For example, geocomposites used for pavement edge
drains typically cost $1.00 to $3.00/linear foot installed while a conventional geotextile
wrapped gravel drain with a pipe is on the order of $9.00/linear foot installed.

FHWA NHI-07-092
Geosynthetics Engineering

2-40

Subsurface Drainage
August 2008

Probably the most common uses for geocomposite drains in highways are pavement edge
drains and drains behind retaining walls and abutments. Several states (e.g., Maine,
Wisconsin, and Virginia) have also experimented with the use of horizontal geocomposite
drains selected to be able to handle the estimated flow and support traffic loads. They are
placed either below or above a dense graded base, used as a drainage layer beneath full depth
asphalt, or placed between a crack and seat concrete surface and a new asphalt layer.
Pavement drainage applications are discussed in more detail in Chapter 5, Roadways and
Pavements, and drainage requirements for retaining structures are reviewed in Chapter 9,
Retaining Walls and Abutments
As a soil improvement technique for soft foundations, prefabricated vertical geocomposite
drains, sometimes called PVD or wick drains, have made conventional sand drains obsolete.
PVD drains are reviewed in more detail later in this section.
2.11-1 Design and Selection Criteria
For the design and selection of geotextiles with in-plane drainage capabilities and geotextile
filters for geocomposite drainage systems, there are three basic design considerations:
1.
2.
3.

Adequate filtration without clogging or piping.


Adequate inflow/outflow capacity under design loads to provide maximum
anticipated seepage during design life.
System performance considerations.

2.11-1.a Geotextile Filter


As with conventional drainage systems, geotextile filter design and selection should be based
on the grain size of the material to be protected, permeability requirements, clogging
resistance, and physical property requirements, as described in Section 2.4. For example, in
pavement drainage systems, dynamic loading means severe hydraulic conditions (Table 2-1).
If the geotextile supplied with the geocomposite is not appropriate for your design
conditions, system safety will be compromised and you should specify geotextiles that will
work. This is important especially when prefabricated drains are used in critical situations
and where failure could lead to structure or system failure.
Geotextile filters for prefabricated vertical drains (PVD) or wick drains are a special case.
The objective of projects involving PVDs is to accelerate the consolidation of soft
compressible soils, so the filter should be compatible with the characteristics of the soils to
be drained. All the design procedures described in Sec. 2.4 are appropriate, including long
term filtration (clogging) tests if the project is critical. See also FHWA NHI-06-019 (Elias et
al., 2006), Holtz (1987), and Holtz et al. (1991) for additional information on the design,
FHWA NHI-07-092
Geosynthetics Engineering

2-41

Subsurface Drainage
August 2008

properties, and installation of wick drains. Holtz and Christopher (1987) discuss
specifications for material properties (geotextile filter and core) and installation of wick
drains. Mechanical properties of the drain components are especially important for
successful installation.

Figure 2-10. Geocomposite drains.

FHWA NHI-07-092
Geosynthetics Engineering

2-42

Subsurface Drainage
August 2008

2.11-1.b Flow CapacityShort Term


In order to design the in-plane flow capacity of a geotextile or the flow capacity of the core
or a geocomposite, the maximum seepage flow into the system must be estimated using the
procedure described in Sec. 2.5, Step 4, B.3. Then the geotextile or geocomposite is selected
on the basis of these seepage requirements. The flow capacity of the geocomposite or
geotextile can be determined from the transmissivity of the material. The test for
transmissivity is ASTM D 4716, Constant Head Hydraulic Transmissivity (In-Plane Flow)
of Geotextiles and Geotextile Related Products. The flow capacity per unit width of the
geotextile or geocomposite can then be calculated using Darcy's Law:
q = kp i A = k p i B t

[2 - 15]

q/B = i

[2 - 16]

or,
where:
q
kp
B
t

=
=
=
=
=
=

flow rate (L3/T)


in-plane coefficient of permeability for the geosynthetic (L/T)
width of geosynthetic (L)
thickness of geosynthetic (L)
transmissivity of geosynthetic (= kpt) (L2/T)
hydraulic gradient (L/L)

The flow rate per unit width of the geosynthetic can then be compared with the flow rate per
unit width required of the drainage system. It should be recognized that the in-plane flow
capacity for geosynthetic drains reduces significantly under compression (Giroud, 1980).
Additional decreases in transmissivity may occur with time due to creep of the geotextile into
the core or even the core material itself. Therefore, the composite material should be
evaluated by an appropriate laboratory model (performance) test, under the anticipated
design loading conditions (with a safety factor) for the design life of the project.
2.11-1.c Flow CapacityLong Term
Long-term compressive stress and eccentric loadings on the core of a geocomposite should
be considered during design and selection. Though not yet addressed in standardized test
methods or standards of practice, the following criteria (Berg, 1993) are suggested for
addressing core compression. The design pressure on a geocomposite core should be limited
to either:
i) the maximum pressure sustained on the core in a test of 10,000 hr minimum duration;
or ii) the crushing pressure of a core, as defined with a quick loading test, divided by a
safety factor of five.
FHWA NHI-07-092
Geosynthetics Engineering

2-43

Subsurface Drainage
August 2008

Note that crushing pressure can only be defined for some core types. For cases where a
crushing pressure cannot be defined, suitability should be based on the first criterion, the
maximum load resulting in a residual thickness of the core adequate to provide the required
flow after 10,000 hours.
Intrusion of the geotextiles into the core and long-term outflow capacity should be measured
with a modified transmissivity test similar to ASTM D 4716 (Berg, 1993). The equipment
should be capable of sustained loading and the geotextile should be in contact with a sand
substratum in lieu of closed cell foam rubber. Load should be maintained for at least 300
hours or until equilibrium is reached, whichever is greater.
For PVDs (wick drains), see Holtz et al. (1991) for a discussion of the effects of core
capacity and intrusion on drain performance. They also have a review of testing
proceduresnone are ASTM standards yetthat have been developed to evaluate intrusion,
core kinking, and other detrimental effects.
2.11-1.d System Performance Considerations
Finally, consideration should be given to system performance factors such as distance
between drain outlets, hydraulic gradient of the drains, potential for blockage due to small
animals, freezing, etc. When using geosynthetics to drain earth retaining structures and
abutments, drain location and pressures on the wall or abutment must be properly accounted
for. It is important that the drain be located away from the back of the wall and be
appropriately inclined so it can intercept seepage before it impinges on the back of the wall.
Placement of a thin vertical drain directly against a retaining wall may actually increase
seepage forces on the wall due to rainwater infiltration (Terzaghi et al., 1996; and Cedergren,
1989). For further discussion of this point, see Christopher and Holtz (1985).
2.11-2 Construction Considerations
The following are considerations specific to the installation of geocomposite drains:
1. As with all geosynthetic applications, care should be taken during storage and
placement to avoid damage to the material.
2. Placement of the backfill directly against the geotextile filter must be closely
observed, and compaction of soil with equipment directly against the geocomposite
should be avoided. Otherwise, the filter could be damaged or the drain could even be
crushed. Use of clean granular backfill reduces the compaction energy requirements.
3. At the joints, where the sheets or strips of geocomposite butt together, the geotextile
filter must be carefully overlapped to prevent soil infiltration. Also, the geotextile
should extend beyond the ends of the drain to prevent soil from entering at the edges.

FHWA NHI-07-092
Geosynthetics Engineering

2-44

Subsurface Drainage
August 2008

4. Details must be provided on how the prefabricated drains tie into the collector
drainage systems.
Construction of an edge drain installation is shown in Figures 2-11 and 2-12. Additional
information and recommendations regarding proper edge drain installation can be found in
Koerner, et al. (1994) and in ASTM D 6088 Practice for Installation of Geocomposite Edge
Drains.

2.12

REFERENCES

References quoted within this section are listed below. FHWA references are generally
available at www.fhwa.dot.gov/bridge under the publications and geotechnical tabs and/or at
www.nhi..fhwa.dot.gov under the training and NHI store tabs. A key reference for design is
this manual (FHWA Geosynthetics Manual) and its predecessor Christopher and Holtz
(1985). The NCHRP report (Koerner et al., 1994) specifically addresses pavement edge
drain systems and is based upon analysis of failed systems. It is a key reference for design.
These and other key references are noted in bold type. Detailed lists of specific ASTM and
GRI test procedures are presented in Appendix E.
AASHTO (2006).
Standard Specifications for Geotextiles - M 288, Standard
Specifications for Transportation Materials and Methods of Sampling and Testing,
26th Edition, American Association of State Transportation and Highway Officials,
Washington, D.C.
ASTM (2006). Annual Books of ASTM Standards, American Society for Testing and
Materials, West Conshohocken, PA:
Volume 4.08 (I), Soil and Rock
Volume 4.09 (II), Soil and Rock; Geosynthetics
Baumgardner, R.H. (1994). Geotextile Design Guidelines for Permeable Bases, Federal
Highway Administration, Washington, D.C., June, 33 p.
Bell, J.R. and Hicks, R.G. (1980). Evaluation of Test Methods and Use Criteria for
Geotechnical Fabrics in Highway Applications - Interim Report, FHWA/RD80/021,190 p.
Berg, R.R., (1993). Guidelines for Design, Specification, & Contracting of Geosynthetic
Mechanically Stabilized Earth Slopes on Firm Foundations, FHWA-SA-93-025, 87 p.

FHWA NHI-07-092
Geosynthetics Engineering

2-45

Subsurface Drainage
August 2008

(a) Equipment train used to install PGEDs according to Figure 2-12.

(b) Sand installation and backfilling equipment at end of equipment train (per Figure 2-12).
Figure 2-11. Prefabricated geocomposite edge drain construction using sand fill upstream
of composite (as illustrated in Figure 2-12) (from Koerner, et al., 1994).
FHWA NHI-07-092
Geosynthetics Engineering

2-46

Subsurface Drainage
August 2008

Figure 2-12. Recommended installation method for prefabricated geocomposite edge


drains (from Koerner, et al., 1994).

Carroll, R.G., Jr. (1983). Geotextile Filter Criteria, Engineering Fabrics in Transportation
Construction, Transportation Research Record 916, Transportation Research Board,
Washington, D.C., pp. 46-53.
Casagrande, A. (1938). Notes on Soil Mechanics - First Semester, Harvard University
(unpublished), 129 pp.
Cedergren, H.R. (1989). Seepage, Drainage, and Flow Nets, Third Edition, John Wiley and
Sons, New York, 465 p.
Christopher, B.R. and Holtz, R.D. (1985). Geotextile Engineering Manual, FHWA-TS86/203, 1044 p.
Elias, V., Welsh, J., Warren, J., Lukas, R., Collin, J.G. and Berg, R.R. (2006). Ground
Improvement Methods, FHWA NHI-06-019 (Vol. I) and FHWA NHI-06-020 (Vol. II).
Fischer, G.R. (1994). The Influence of Fabric Pore Structure on the Behavior of
Geotextile Filters, Ph.D. Dissertation, University of Washington, 498 p.
Giroud, J.P. (1988). Review of Geotextile Filter Criteria, Proceedings of First Indian
Geotextiles Conference on Reinforced Soil and Geotextiles, Bombay, India, 6 p.
Giroud, J.P. (1980). Introduction to Geotextiles and Their Applications, Proceedings of the
First Canadian Symposium on Geotextiles, Calgary, Alberta, pp. 3-31.
FHWA NHI-07-092
Geosynthetics Engineering

2-47

Subsurface Drainage
August 2008

Haliburton, T. A. and Wood, P. D. (1982). Evaluation of the U. S. Army Corps of Engineers


Gradient Ratio Test for Geotextile Performance, Proceedings of the Second International
Conference on Geotextiles, Las Vegas, Nevada, Vol. 1, pp.97-101.
Holtz, R. D., Jamiolkowski, M., Lancellotta, R. and Pedroni, S. (1991). Prefabricated
Vertical Drains: Design and Performance, Butterworths/CIRIA co-publication series,
CIRIA, Butterworths-Heinemann, London, England, 131 pp.
Holtz, R. D., (1987). Preloading with Prefabricated Vertical Strip Drains, Geotextiles and
Geomembranes, Vol. 6, Nos. 1-3, pp. 109-131. (Also published in Proceedings of the
First Geosynthetic Research Institute Seminar on Very Soft Soil Stabilization Using High
Strength Geosynthetics, Drexel University, Philadelphia, Pennsylvania, pp. 104-129.)
Holtz, R. D. and Christopher, B. R. (1987). Characteristics of Prefabricated Drains for
Accelerating Consolidation, Proceedings of the Ninth European Conference on Soil
Mechanics and Foundation Engineering, Dublin, Ireland, Vol. 2, pp. 903-906.
Holtz, R. D. and Kovacs, W. D. (1981). An Introduction to Geotechnical Engineering,
Prentice-Hall, p. 210.
Hunt, J.R. (1982). The Development of Fin Drains for Structure Drainage, Proceedings of
the Second International Conference on Geotextiles, Las Vegas, NV, Vol. 1, pp. 25-36.
Kenney, T.C. and Lau, D. (1986). Reply (to discussions), Vol. 23, No. 3, pp. 420-423,
Internal Stability of Granular Filters, Canadian Geotechnical Journal, Vol. 22, No. 2,
1985, pp. 215-225.
Kenney, T.C. and Lau, D. (1985). Internal Stability of Granular Filters, Canadian
Geotechnical Journal, Vol. 22, No. 2, 1985, pp. 215-225.
Koerner, R.M., Koerner, G.R., Fahim, A.K. and Wilson-Fahmy, R.F. (1994). Long
Term Performance of Geosynthetics in Drainage Applications, National Cooperative
Highway Research Program Report No. 367, 54 p.
LeFluer, J., Mlynarek, J. and Rollin, A.L. (1993). Filter Criteria for Well Graded
Cohesionless Soils, Filters in Geotechnical and Hydraulic Engineering, Proceedings of
the First International Conference - Geo-Filters, Karlsruhe, Brauns, Schuler, and
Heibaum Eds., Balkema, pp. 97-106.
LeFluer, J., Mlynarek, J. and Rollin, A.L. (1989). Filtration of Broadly Graded
Cohesionless Soils, Journal of Geotechnical Engineering, American Society of Civil
Engineers, Vol. 115, No. 12, pp. 1747-1768.
Mansur, C.I. and Kaufman, R.I. (1962). Dewatering, Chapter 3 in Foundation Engineering,
G.A. Leonards, Editor, McGraw-Hill, pp. 241-350.
FHWA NHI-07-092
Geosynthetics Engineering

2-48

Subsurface Drainage
August 2008

Mar, A.D. (1994). The Influence of Gradient Ratio Testing Procedures on the Filtration
Behavior of Geotextiles, MSCE Thesis, University of Washington.
Moulton, L.K. (1980). Highway Subdrainage Design, FHWA-TS-80-224.
NCHRP 1-37A (2004). Mechanistic-Empirical Design of New and Rehabilitated Pavement
Structures, Draft Final Report, NCHRP Project 1-37A, National Cooperative Highway
Research Program, National Research Council, Washington, D.C.
Sherard, J.L. (1986). Hydraulic Fracturing in Embankment Dams, Journal of Geotechnical
Engineering, American Society of Civil Engineers, Vol. 112, No. 10, pp. 905-927.
Sherard, J.L. and Decker, R.S., Editors (1977). Dispersive Clays, Related Piping, and
Erosion in Geotechnical Projects, ASTM Special Technical Publication 623, American
Society for Testing and Materials, Philadelphia, PA, 486p.
Sherard, J.L., Decker, R.S. and Ryker, N.L. (1972). Piping in Earth Dams of Dispersive
Clay, Proceedings of the ASCE Specialty Conference on Performance of Earth and Earth
-Supported Structures, American Society of Civil Engineers, New York, Vol. I, Part 1,
pp. 589-626.
Skempton, A.W. and Brogan, J.M. (1994). Experiments on Piping in Sandy Gravels,
Geotechnique, Vol. XLIV, No. 3, pp. 461-478.
Terzaghi, K., Peck, R.B., and Mesri, G. (1996). Soil Mechanics in Engineering Practice,
Third Edition, John Wiley & Sons, New York, pp 330-332.
U.S. Army Corps of Engineers (1977). Civil Works Construction Guide Specification for
Plastic Filter Fabric, Corps of Engineer Specifications No. CW-02215, Office, Chief of
Engineers, U.S. Army Corps of Engineers, Washington, D.C.
U.S. Department of the Navy (1986). Design Manual 7.01 - Soil Mechanics, Department of
the Navy, Naval Facilities Engineering Command, Alexandria, VA. (can be downloaded
from http://www.geotechlinks.com).

FHWA NHI-07-092
Geosynthetics Engineering

2-49

Subsurface Drainage
August 2008

FHWA NHI-07-092
Geosynthetics Engineering

2-50

Subsurface Drainage
August 2008

3.0 GEOTEXTILES IN RIPRAP REVETMENTS AND


OTHER PERMANENT EROSION CONTROL SYSTEMS

3.1 BACKGROUND
As in drainage systems, geotextiles can effectively replace graded granular filters typically
used beneath riprap or other hard armor materials in revetments and other erosion control
systems. This was one of the first applications of geotextiles in the United States; woven
monofilament geotextiles were initially used for this application with rather extensive
installation starting in the early 1960s. Numerous case histories have shown geotextiles to be
very effective compared to riprap-only systems and as effective as conventional graded
granular filters in preventing fines from migrating through the armor system. Furthermore,
geotextiles have proven to be very cost effective in this application.
Since the early developments in coastal and lake shoreline erosion control, the same design
concepts and construction procedures using geotextile filters have subsequently been applied
to stream bank protection (see HEC 11, FHWA, 1989), cut and fill slope protection,
protection of various small drainage structures (see HEC 14, FHWA, 2006) and ditches (see
HEC 15, FHWA, 2005), wave protection for causeway and shoreline roadway embankments,
and scour protection for structures such as bridge piers and abutments (see HEC 18, FHWA,
2001, and HEC 23, FHWA, 2001). Design guidelines and construction procedures with
geotextile filters for these and other similar permanent erosion control applications are
presented in Sections 3.3 through 3.10. Hydraulic design considerations can be found in the
AASHTO Model Drainage Manual (2005) and the above FHWA Hydraulic Engineering
Circulars. Also note that additional information and training are available in another NHI
course and reference manual. The course is entitled Design and Implementation of Erosion
and Sediment Control, and was developed in a joint effort between FHWA and the
Environmental Protection Agency (EPA) course.
Although this chapter focuses on geotextile filters in erosion control systems, there are other
geosynthetics used for permanent erosion protection including geocells and geosynthetic turf
reinforcing mats (TRMs). Geocells are three-dimensional cellular structures made from
formed expandable polyethylene (low and high density) panels. When the expanded panels
are interconnected, they form a 3-D cellular structure that provides confinement and
reinforcement to the free-draining sand and/or gravel infill. Geocells filled with clean gravel
or concrete have been successfully used for all the erosion control applications mentioned
above and as discussed in Section 3.10.

FHWA NHI-07-092
Geosynthetics Engineering

3-1

Erosion Control Systems


August 2008

TRMs are rolled erosion control products (RECPs) composed of nondegradable, threedimensional porous geosynthetic mats that reinforce the roots and help to retain soil and
moisture, thus promoting vegetation growth. These products together with vegetation form
a biocomposite that is very attractive and environmentally friendly. TRMs are a reinforced
grass system capable of withstanding short-term (e.g., 2 hours), high velocity (e.g., 20ft/s {6
m/s}) flows with minimal erosion. TRMs are addressed in Section 3.11.
Erosion control blankets (ECBs), temporary TRMs and other RECPs such as mulch control
nets (MCNs) are covered in Chapter 4.

3.2

APPLICATIONS

Riprap-geotextile systems have


been successful for precipitation
runoff collection and highvelocity diversion ditches.

Geotextiles may be used in


slope protection to prevent or
reduce
erosion
from
precipitation, surface runoff,
and internal seepage or piping.
In this instance, the geotextile
may replace one or more layers
of granular filter materials that
would be placed on the slope in
conventional applications.
Erosion control systems with
geotextiles may be required
along stream banks to prevent
encroachment of roadways or
appurtenant facilities.

FHWA NHI-07-092
Geosynthetics Engineering

3-2

Erosion Control Systems


August 2008

Similarly, they may be used for


scour
protection
around
structures.
A riprap-geotextile system can
also be effective in reducing
erosion caused by wave attack
or tidal variations when
facilities are constructed across
or adjacent to large bodies of
water. Geocells can be used as
an alternate amour system to
replace
riprap
in
slope
protection, diversion ditches
and other runoff applications.
Finally, hydraulic structures
such as culverts, drop inlets, and
artificial stream channels may
require protection from erosion.
In
such
applications,
if
vegetation cannot be established
or the natural soil is highly
erodible, a geotextile can be
used beneath armor materials to
increase erosion resistance.
Geosynthetic erosion control
mats or TRMs are a threedimensional matrix of synthetic
yarns, meshes or webs and that
reinforce the vegetation root
mass and provide tractive
resistance to flowing water on
slopes and in ditches, channels,
and swales.
These threedimensional mats retain soil,
moisture, and seed, and thus
promote vegetative growth.
FHWA NHI-07-092
Geosynthetics Engineering

3-3

Erosion Control Systems


August 2008

3.3

GEOTEXTILE FILTERS BENEATH HARD ARMOR: DESIGN CONCEPTS

Geotextile filter design for hard armor erosion control systems is essentially the same, with a
few exceptions, as the design for geotextile filters in subsurface drainage systems. It would
be a good idea to go back and reread Chapter 2, especially Sections 2.3 and 2.4. This section
highlights those exceptions and discusses the special considerations for geotextile filters
beneath hard armor erosion control systems.
3.3-1 Retention Criteria for Cyclic or Dynamic Flow
Many erosion control situations have cyclic or dynamic flow conditions, so soil particles may
be able to move behind the geotextile if it is not properly weighted down and in intimate
contact with the soil. Thus, unlike conventional filters, using a retention coefficient B = 1
may not be conservative, as the bridging network (Figure 2-2) may not develop and the
geotextile may be required to retain even the finer particles of soil. If there is a risk that
uplift of the armor system can occur, it is recommended that the B value be reduced to 0.5 or
less; that is, the largest hole in the geotextile should be small enough to retain the smaller
particles of soil.
In many erosion control applications it is common to have high hydraulic stresses induced by
wave or tidal action. The geotextile may be loose when it spans between large armor stone
or large joints in block-type armor systems. For these conditions, it is recommended that an
intermediate layer of finer stone or gravel be placed over the geotextile and that riprap of
sufficient weight be placed to prevent wave action from moving either stone or geotextile and
to maintain the intimate contact between the soil and geotextile filter. Geosynthetic
composites (e.g., geonet/geotextile composite) could also be considered beneath block-type
armor systems to prevent movement of the geotextile filter as well as uplift on the block. For
all applications where the geotextile can move, and when it is used as sandbags, it is
recommended that samples of the site soils be washed through the geotextile to determine its
particle-retention capabilities.
3.3-2 Permeability and Effective Flow Capacity Requirements for Erosion Control
In certain erosion control systems, portions of the geotextile may be covered by the armor
stone or concrete block revetment systems, or the geotextile may be used to span joints in
sheet pile bulkheads. For such systems, it is especially important to evaluate the flow rate
required through the open portion of the system and select a geotextile that meets those flow
requirements. Again, since flow is restricted through the geotextile, the required flow
capacity is based on the flow capacity of the area available for flow; or

FHWA NHI-07-092
Geosynthetics Engineering

3-4

Erosion Control Systems


August 2008

qrequired = qgeotextile(Ag/At)
where:

Ag
At

(Eq. 2 - 9)

= geotextile area available for flow, and


= total geotextile area.

The AASHTO M 288 Standard Specification for Geotextiles (2006) presents recommended
minimum permittivity values in relation to percent of in-situ soil passing the No.200 (0.075
mm) sieve. The values are presented in Section 3.4. These permittivity values are based
upon the predominant particle sizes of the in-situ soil and are additional qualifiers to the
permeability criteria.
3.3-3 Clogging Resistance for Cyclic or Dynamic Flow and for Problematic Soils
Since erosion control systems are often used on highly erodible soils with reversing and
cyclic flow conditions, severe hydraulic and soil conditions often exist. Accordingly, designs
should reflect these conditions, and soil-geotextile filtration tests should be conducted. Since
these tests are performance-type tests and require soil samples from the project site, they
must be conducted by the owner or the owners representative and not by geotextile
manufacturers or suppliers. Project specific testing should be performed especially if one or
more of the following problematic soil environments are encountered: unstable or highly
erodible soils such as non-cohesive silts; gap graded soils; alternating sand/silt laminated
soils; dispersive clays; and/or rock flour.
For sandy soils with k > 10-6 m/s the long-term, gradient ratio test (ASTM D 5101) is
recommended, as described in Chapters 1 and 2, and note that the U.S. Army Corps of
Engineers recommends a maximum allowable gradient ratio (GR) of three. For soils with
permeabilities less than about 10-6 m/s, filtration tests should be conducted in a flexible wall
or triaxial type apparatus to ensure that the specimen is 100% saturated and that flow is
through the soil rather than along the sides of the specimen. The soil flexible wall test is
ASTM D 5084, while the Hydraulic Conductivity Ratio (HCR) test (ASTM D 5567)
currently is the standard test for geotextiles and soils with appreciable fines. The HCR test
should be considered only with the modifications and caveats recommended in Chapter 1.
Other filtration tests discussed in Chapters 1 and 2 should also be considered.
3.3-4 Survivability Criteria for Erosion Control
Because the construction procedures for erosion control systems are different than those for
drainage systems, the geotextile property requirements for survivability in Table 3-1 differ
somewhat from those discussed in Section 2.4-4. As placement of armor stone is generally
more severe than placement of drainage aggregate, required property values are higher for
each category of geotextile. Furthermore, the specifications should require the contractor to
FHWA NHI-07-092
Geosynthetics Engineering

3-5

Erosion Control Systems


August 2008

demonstrate in the field that their proposed armoring placement technique will not damage
the geotextile.
Riprap or armor stone should be large enough to withstand wave action and thus not abrade
the geotextile. The specific site conditions should be reviewed, and if such movement cannot
be avoided, then an abrasion requirement based on ASTM D 4886, Standard Test Method for
Abrasion Resistance of Geotextiles should be included in the specifications. Abrasion of
course only affects the physical and mechanics properties of the geotextile. No reduction in
piping resistance, permeability, or clogging resistance should be allowed after exposure to
abrasion.
It is important to realize that the survivability requirements in Table 3-1 are minimum
survivability values and are not based on any systematic research. They are based on the
properties of geotextiles that are known to have performed satisfactorily in various hard
armor erosion control applications. The values in Table 3-1 are meant to serve as guidelines
for inexperienced users in selecting geotextiles for routine projects. They are not intended to
replace site-specific evaluation, testing, and design.
3.3-5 Additional Filter Selection Considerations and Summary
To enhance system performance, special consideration should be given to the type of
geotextile chosen for certain soil and hydraulic conditions. The considerations listed in
Section 2.4-5 also apply to erosion control systems. As mentioned above, special attention
should be given to problematic, unstable, or highly erodible soils. Examples include noncohesive silts, gap graded soils, alternating sands and silts, dispersive clays, and rock flour.
Project specific laboratory testing should be performed especially for critical projects and
severe conditions.
In certain situations, multiple filter layers may be necessary. For example, a sand layer
could be placed on the soil subgrade, with the geotextile designed to filter the sand only but
with sufficient size and number of openings to allow any fines that do reach the geotextile to
pass through it. Another special consideration for erosion control applications relates to a
preference towards felted and rough versus slick surface geotextiles, especially on steeper
slopes where there is a potential for the riprap to slide on the geotextile. Such installations
must be assessed either through field trials or large-scale laboratory tests.
Figure 3-1 is a flow chart summarizing the FHWA filter design process.

FHWA NHI-07-092
Geosynthetics Engineering

3-6

Erosion Control Systems


August 2008

Table 3-1
Geotextile Strength Property Requirements1,2,3,4
for Permanent Erosion Control Geotextiles
(after AASHTO, 2006)
Geotextile Class
Test
Methods
Grab strength

a,b,c

Class 1
Units

Class 2

Elongation

Elongation

Elongation

Elongation

< 50%

> 50%

< 50%

> 50%

ASTM D 4632

lb (N)

315 (1400)

200 (900)

250 (1100)

157 (700)

Sewn seam strength

ASTM D 4632

lb (N)

280 (1260)

180 (810)

220 (990)

140 (630)

Tear strength

ASTM D 4533

lb (N)

110 (500)

80 (350)

90 (400)

Puncture strength

ASTM D 6241

lb (N)

620 (2750)

433 (1925)

495 (2200)

Ultraviolet stability
(retained strength)

ASTM D 4355

a
b

d
e
f

56 (250)
309 (1375)

50% after 500 hours of exposure(5)

Use Class 2 for woven monofilament geotextiles, and Class 1 for all other geotextiles.
As a general guideline, the default geotextile selection is appropriate for conditions of equal or less severity
than either of the following:
a) Armor layer stone weights do not exceed 220 lb (100 kg), stone drop is less than 3.3 ft. (1 m), and no
aggregate bedding layer is required.
b) Armor layer stone weights exceed 220 (100 kg), stone drop height is less than 3.3 ft. (1 m), and the
geotextile is protected by a 6-inch thick aggregate bedding layer designed to be compatible with the
armor layer. More severe applications require an assessment of geotextile survivability based on a
field trial section and may require a geotextile with higher strength properties.
The engineer may specify a Class 2 geotextile based on one or more of the following:
a) The engineer has found Class 2 geotextiles to have sufficient survivability based on field experience.
b) The engineer has found Class 2 geotextiles to have sufficient survivability based on laboratory testing
and visual inspection of a geotextile sample removed from a field test section constructed under
anticipated field conditions.
c) Armor layer stone weighs less than 220 (100 kg), stone drop height is less than 3.3 ft. (1 m), and the
geotextile is protected by a 6-inch thick aggregate bedding layer designed to be compatible with the
armor layer.
d) Armor layer stone weights do not exceed 220 lb (100 kg), stone is placed with a zero drop height.
As measured in accordance with ASTM D 4632.
When sewn seams are required. Refer to Appendix for overlap seam requirements.
The required MARV tear strength for woven monofilament geotextiles is 56 lb (250 N).

NOTES:
1.
Acceptance of geotextile material shall be based on ASTM D 4759.
2.
Acceptance shall be based upon testing of either conformance samples obtained using Procedure A of
ASTM D 4354, or based on manufacturers certifications and testing of quality assurance samples
obtained using Procedure B of ASTM D 4354.
3.
Minimum; use value in weaker principal direction. All numerical values represent minimum average roll
value (i.e., test results from any sampled roll in a lot shall meet or exceed the minimum values in the
table). Lot samples according to ASTM D 4354.
4. Woven slit film geotextiles will not be allowed.
5. The original M288 specifications required 70% strength retention for erosion control applications due to
the potential of UV exposure between riprap.
FHWA NHI-07-092
Geosynthetics Engineering

3-7

Erosion Control Systems


August 2008

O95 < B D85

For
2 < CU < 4
B = 0.5 CU
For
4 < CU < 8
B = 8 / CU
B = 1.8 &
O95 < 1.8 D85

Use
For CU < 3
maximum O95 from
Retention Criteria

SURVIVABILITY and ENDURANCE CRITERIA

3-8

Performance Tests
to Select Suitable
Geotextile

Unstable Soils

Erosion Control Systems


August 2008

Perform filtration test


with on-site soils and
hydraulic conditions

For critical applications and severe conditions:

> 50%
> 0.1 sec-1

Optional Qualifiers for gap-graded or silty soils


For Nonwovens: n > 50%
For Woven monofilament and silt films: POA > 4%

CLOGGING RESISTANCE

Flow chart summary of the FHWA filter design procedure.

FHWA NHI-07-092
Geosynthetics Engineering

Figure 3-1.

For CU > 3
O95 > 3 D15

15% to 50%
> 0.2 sec-1

qrequired = qgeotextile (Ag/At)

< 15%
> 0.5 sec-1

For less critical applications and less severe conditions:

% Passing #200 sieve:


Permittivity Required:

kgeotextile > ksoil

O95 < 0.5 D85

Dynamic Flow

For critical applications and severe


conditions: kgeotextile > 10 ksoil

PERMEABILITY/PERMITTIVITY CRITERIA

and
O95 < 0.3 mm

for Nonwovens

B=1&
O95 < D85

Silts and Clays


( > 50% passing No.200 sieve )

for Wovens

For less critical applications and less severe conditions:

For
2 > CU > 8
B=1

Sands, Gravelly Sands, Silty Sands & Clayey


Sands (< 50% passing No.200 sieve )

Steady State Flow

RETENTION CRITERIA

3.4 GEOTEXTILE DESIGN GUIDELINES


STEP 1. Evaluate critical nature and site conditions.
A. Critical/less critical
1.
If the erosion control system fails, will there be a risk of loss of life?
2.
Does the erosion control system protect a significant structure, or will failure
lead to significant structural damage?
3.
If the geotextile clogs, will failure occur with no warning? Will failure be
catastrophic?
4.
If the erosion control system fails, will the repair costs greatly exceed
installation costs?
B. Severe/less severe
1.
Are soils to be protected gap-graded, pipable, or dispersive?
2.
Do the soils consist primarily of silts and uniform sands with 85% passing the
No.100 sieve?
3.
Will the erosion control system be subjected to reversing or cyclic flow
conditions such as wave action or tidal variations?
4.
Will high hydraulic gradients exist in the soils to be protected? Will rapid
drawdown conditions or seeps or weeps in the soil exist? Will blockage of
seeps and weeps produce high hydraulic pressures?
5.
Will high-velocity conditions exist, such as in stream channels?
NOTE: If the answer is yes to any of the above questions, the design should proceed
under the critical/severe requirements; otherwise use the less critical/less severe
design approach.

STEP 2. Obtain soil samples from the site.


A. Perform grain size analyses
1.
Determine percent passing the No.200 (0.075 mm)sieve.
2.
Determine the plastic index (PI).
3.
Calculate Cu = D60/D10.
NOTE: When the protected soil contains particles passing the No.200 (0.075 mm)
sieve, use only the gradation passing the No.4 (4.75 mm) sieve in selecting the
geotextile (i.e., scalp off the +#4 (+4.75 mm) material).
FHWA NHI-07-092
Geosynthetics Engineering

3-9

Erosion Control Systems


August 2008

4.

Obtain D85 for each soil and select the worst case soil (i.e., soil with smallest B
D85) for retention.

B. Perform field or laboratory permeability tests


1.
Select worst case soil (i.e., soil with highest coefficient of permeability k).
NOTE: The permeability of clean sands (< 5% passing No.200 (0.075 mm) sieve)
with 0.1 mm D10 < 3 mm and Cu < 5 can be estimated by Hazen's formula, k = (D10)2
(k in cm/s; D10 in mm). This formula should not be used if the soil contains more
than 5% fines.
NOTE: Laboratory tests for permeability (hydraulic conductivity) are detailed in
ASTM D 2434 for granular soils, D 5856 using a compaction-mold permeameter, and
in D 5084 using a flexible-wall permeameter for soils with appreciable fines. Field
tests include pumping tests in boreholes and infiltrometer tests. Standard procedures
for several field tests are also in ASTM.
NOTE: A good visual classification of the soils at the site will enable an experienced
geotechnical engineer to estimate the permeability to the nearest order of magnitude,
which is often sufficient for geotextile filter design. The following table, adapted
from Casagrande (1938) and Holtz and Kovacs (1981), gives a range of hydraulic
conductivities for different natural soils.

Visual Classification
Clean gravel
Clean sands and clean sand-gravel mixtures
Very fine sands; silts; mixtures of sand, silt,
and clay; glacial tills; stratified clays
Impervious soils; homogeneous reasonably
intact clays from below zone of weathering
Impervious soils, modified by vegetation,
weathering, fissured, highly OC clays

STEP 3.

Permeability or Hydraulic
Conductivity, k (m/s)
> 0.01
0.01 < k < 10-5
10-5 < k < 10-9
k > 10-9
5 x 10-5 < k < 5 x 10-8

Evaluate armor material and placement.


Design reference: FHWA Hydraulic Engineering Circular No. 15 (FHWA,
2005).

FHWA NHI-07-092
Geosynthetics Engineering

3-10

Erosion Control Systems


August 2008

A. Size armor stone or riprap


Where minimum size of stone exceeds 4 in. (100 mm), or greater than a 4-in. (100
mm) gap exists between blocks, an intermediate gravel layer at least 6 in. (150 mm)
thick should be used between the armor stone and geotextile. Gravel should be sized
such that it will not wash through the armor stone (i.e., D85(gravel) D15(riprap)/5).
B. Determine armor stone placement technique (i.e., maximum height of drop).
C. Consider alternate surface treatments such as with geocells (see section 3.10).
STEP 4. Determine anticipated reversing flow through the erosion control system.
Here we need to estimate the maximum flow from seeps and weeps, maximum flow
from wave action, or maximum flow from rapid drawdown.
A. General case -- use Darcy's law
q = kiA

(Eq. 2 - 12)

where:
q =
k =
i
=

outflow rate (m3/sec)


effective permeability of soil (from Step 2B above) (m/sec)
average hydraulic gradient in soil (e.g., tangent of slope angle for wave
runup)(dimensionless)
A = area of soil and drain material normal to the direction of flow (m2). Can be
evaluated using a unit area.
Use a conventional flow net analysis (e.g., Cedergren, 1989) for seepage through
dikes and dams or from a rapid drawdown analysis.
B. Specific erosion control systems -- Hydraulic characteristics depend on expected
precipitation, runoff volumes and flow rates, stream flow volumes and water level
fluctuations, normal and maximum wave heights anticipated, direction of waves and
tidal variations. Detailed information on determination of these parameters is
available in the FHWA (1989) Hydraulic Engineering Circular No. 11.

STEP 5. Determine geotextile requirements.


A. Retention Criteria
From Step 2A, obtain D85 and Cu; then determine largest opening size allowed.
AOS or O95(geotextile) < B D85(soil)
(Eq. 2 - 1)
where: B = 1 for a conservative design.
FHWA NHI-07-092
Geosynthetics Engineering

3-11

Erosion Control Systems


August 2008

For a less-conservative design and for soils with < 50% passing the No.200 sieve:
B=1
for Cu < 2 or > 8
(Eq. 2 - 2a)
for 2 < Cu < 4
(Eq. 2 - 2b)
B = 0.5 Cu
B = 8/Cu
for 4 < Cu < 8
(Eq. 2 - 2c)
For soils with > 50% passing the No.200 sieve:
B=1
for woven geotextiles
B = 1.8
for nonwoven geotextiles
and AOS or O95 (geotextile) < 0.3 mm

(Eq. 2 - 5)

If geotextile and soil retained by it can move, use:


B = 0.5

(Eq. 2 6)

B. Permeability/Permittivity Criteria
1. Less Critical/Less Severe
kgeotextile > ksoil
2.

(Eq. 2 - 7a)

Critical/Severe
kgeotextile > 10 ksoil

3.

Permittivity Requirement
for < 15% passing No.200 (0.075 mm)
for 15 to 50% passing No.200 (0.075 mm)
for > 50 % passing No.200 (0.075 mm)

4.

(Eq. 2 7b)

0.7 sec-1
0.2 sec-1
0.1 sec-1

Flow Capacity Requirement


qgeotextile > (At/Ag) qrequired
or
(kgeotextile/t) h Ag qrequired
where:

FHWA NHI-07-092
Geosynthetics Engineering

(Eq. 2 - 8a)
(Eq. 2 - 8b)
(Eq. 2 - 8c)

(Eq. 2 - 9)

qrequired is obtained from Step 4 (Eq. 2-13) above (m3/sec)


kgeotextile/t = = permittivity (sec-1)
h =
average head in field (m)
Ag =
area of geotextile available for flow (e.g., if 50% of
3-12

Erosion Control Systems


August 2008

At =

geotextile covered by flat rocks or riprap, Ag = 0.5 total


area) (m2)
total area of geotextile (m2)

C. Clogging Criteria
1. Less critical/less severe
a.

From Step 2A obtain D15; then


1. For soils with Cu > 3, determine minimum pore size requirement, from
O95 > 3 D15
(Eq. 2 - 10)
2. For Cu < 3, specify geotextile with maximum opening size possible
from retention criteria

b.

Other qualifiers
For soils with % passing No.200

> 5%

<5%

Woven monofilament geotextiles: Percent Open Area


Nonwoven geotextiles: Porosity

4%
50%

10%
70%

c. Alternative: Run filtration tests


2.

Critical/severe
Select geotextiles that meet the above retention, permeability, and survivability
criteria; as well as the criteria in Step 5C.1 above; perform a long-term filtration
test.
Suggested filtration test for sandy soils is the gradient ratio (GR) test. The
hydraulic conductivity ratio (HCR) test is recommended by ASTM for finegrained soils, but as noted in Sections 1.5 and 2.4-3, the HCR test has serious
disadvantages.
Alternative: Consider long-term filtration tests, F3 tests and the Flexible Wall
GR test (see Section 1.5).

FHWA NHI-07-092
Geosynthetics Engineering

3-13

Erosion Control Systems


August 2008

D. Survivability
Select geotextile properties required for survivability from Table 3-1. Add
durability requirements if applicable. Don't forget to check for abrasion and check
drop height. Evaluate worst case scenario for drop height.

STEP 6. Estimate costs.


Calculate the volume of armor stone, the volume of aggregate and the area of the
geotextile. Apply appropriate unit cost values.
Grading and site preparation (LS)
__________________
2
2
Geotextile (/yd {/m })
__________________
2
2
__________________
Geotextile placement (/yd {/m })
2
2
In-place aggregate bedding layer (/yd {/m }) __________________
Armor stone (/ton {/kg})
__________________
Armor stone placement (/ton {/kg})
__________________
Total cost
__________________

STEP 7. Prepare specifications.


Include for the geotextile:
A. General requirements
B. Specific geotextile properties
C. Seams and overlaps
D. Placement procedures
E. Repairs
F. Testing and placement observation requirements
See Sections 1.6 and 3.7 for specification details.

STEP 8. Obtain samples of the geotextile before acceptance.

STEP 9. Monitor installation during construction, and control drop height. Observe erosion
control systems during and after significant storm events.

FHWA NHI-07-092
Geosynthetics Engineering

3-14

Erosion Control Systems


August 2008

3.5 GEOTEXTILE DESIGN EXAMPLE

DEFINITION OF DESIGN EXAMPLE

Project Description:

Riprap on slope is required to permit groundwater seepage


out of slope face, without erosion of slope. See figure for project
cross section.

Type of Structure:

small stone riprap slope protection

Type of Application:

geotextile filter beneath riprap

Alternatives:

i.) graded soil filter; or


ii.) geotextile filter between embankment and riprap

GIVEN DATA

see cross section

riprap is to allow unimpeded seepage out of slope

riprap will consist of small stone (2 to 12 in. {50 mm to 0.3 m})

stone will be placed by dropping from a backhoe

seeps have been observed in the existing slope

soil beneath the proposed riprap is a fine silty sand

gradations of two representative soil samples

FHWA NHI-07-092
Geosynthetics Engineering

3-15

Erosion Control Systems


August 2008

Project Cross Section


PERCENT PASSING, BY WEIGHT
SIEVE SIZE
(mm)
4.75
1.68
0.84
0.42
0.15
0.075

FHWA NHI-07-092
Geosynthetics Engineering

Sample A

Sample B

100
96
92
85
43
25

100
100
98
76
32
15

3-16

Erosion Control Systems


August 2008

DEFINE
A. Geotextile function(s)
B. Geotextile properties required
C. Geotextile specification
SOLUTION
A. Geotextile function(s):
Primary
Secondary
-

filtration
separation

B. Geotextile properties required:


apparent opening size (AOS)
permittivity
survivability

DESIGN
STEP 1. EVALUATE CRITICAL NATURE AND SITE CONDITIONS.
From given data, this is a critical application due to potential for loss of life and potential
for significant structural damage.
Soils are reasonably well-graded, hydraulic gradient is low for this type of application, and
flow conditions are steady state.

STEP 2. OBTAIN SOIL SAMPLES.


A. VISUAL CLASSIFICATION AND GRAIN SIZE ANALYSES
Visual classification (ASTM D 2488) indicates the fines are silty; therefore the soils are
classified as silty sands, SM.
Perform grain size analyses of the two soils. Plot gradations of representative soils. The
D60, D10, and D85 sizes from the gradation plot are noted in the table below for Samples A
and B.

FHWA NHI-07-092
Geosynthetics Engineering

3-17

Erosion Control Systems


August 2008

B.

PERMEABILITY TESTS
This is a critical application and soil permeability tests should of course be conducted. In
this example, however, we will use only an estimated permeability, just to show how the
design is done.

STEP 3. EVALUATE ARMOR MATERIAL AND PLACEMENT.


A. Small stone (2 to 12 in. {50 mm to 0.3 m}) riprap will be used (no wave action or
significant surface flow).
B.

A placement drop of less than 3 ft (1 m) will be specified.

STEP 4. CALCULATE ANTICIPATED FLOW THROUGH SYSTEM.


Flow computations are not included within this example. The entire height of the slope face
will be protected. This is conservative and for better appearance.

STEP 5. DETERMINE GEOTEXTILE REQUIREMENTS.


A. RETENTION
AOS < B D85
Determine uniformity coefficient, Cu, coefficient B, and the maximum AOS.
Soil
Sample

D60 D10 = Cu

B=

A
B

0.20 0.045 = 4.4


0.30 0.06 = 5

8 Cu = 8 4.4 = 1.82
8 Cu = 8 5 = 1.6

B D85 > AOS


(mm)
1.82 0.44 = 0.8
1.6 0.54 = 0.86

Worst case for retention is Soil A, so Sample A controls (see table in Step 2.A, above);
therefore,
AOS < 0.8 mm
B.

PERMEABILITY/PERMITTIVITY
This is a critical application, therefore,
kgeotextile > 10 ksoil
For this example, lets estimate the soil permeability (using Hazen's formula, but
recognizing that it is applicable only for clean uniform sands and is much less accurate for
soils with 25 to 15% fines.
k (D10)2

FHWA NHI-07-092
Geosynthetics Engineering

3-18

Erosion Control Systems


August 2008

where:
ksoil

k = approximate soil permeability (cm/sec); and D10 is in mm.

= 2.0 (10)-3 cm/sec for Sample A


= 3.6 (10)-3 cm/sec for Sample B
kgeotextile > 4 (10)-2 cm/sec

Therefore (with rounding the number),

Since 15% to 25% of the soil to be protected is finer than No.200 (0.075 mm), the
permittivity is:
geotextile > 0.2 sec-1

C.

CLOGGING
As the project is critical, a filtration test is recommended to evaluate clogging potential.
Select geotextile(s) meeting the retention and permeability criteria, along with the following
qualifiers:
Minimum Opening Size Qualifier (for Cu > 3):
O95 >

O95 > 3 D15

3 0.057 = 0.17 mm for Sample A


3 0.079 = 0.24 mm for Sample B

Sample A controls, therefore,

O95 > 0.17 mm

Other Qualifiers, since greater than 5% of the soil to be protected is finer than No.200, from
Table 3-1:
for Nonwoven geotextiles - Porosity > 50 %
for Woven geotextiles POA (Percent Open Area) > 4 %
Then run a filtration test to evaluate long-term clogging potential. As the material is quite
silty, the gradient ratio test (ASTM D 5101) may take up to several weeks to stabilize. After
testing, geotextiles that perform satisfactorily can be prequalified. Alternatively, geotextiles
proposed by the contractor must be evaluated prior to installation to confirm compatibility.

D. SURVIVABILITY
A Class 1 geotextile will be specified because this is a critical application. Effect on project
cost is minor. Therefore, from Table 3-1, the following minimum values will be specified
except for the UV resistance. Because this is a critical project and there is a potential for
exposure between riprap, we will increase the UV resistance for this example.
FHWA NHI-07-092
Geosynthetics Engineering

3-19

Erosion Control Systems


August 2008

<50% Elongation
>50% Elongation
Grab Strength
315 lb (1400 N)
200 lb (900 N)
Sewn Seam Strength 280 lb (1260 N)
180 lb (810 N)
Tear Strength
110 lb (500 N)
80 lb (350 N)
Puncture Strength
620 lb (2750 N)
433 lb (1925 N)
Ultraviolet Degradation
70 % strength retained at 500 hours

Complete Steps 6 through 9 to finish design.

STEP 6. ESTIMATE COSTS.


STEP 7. PREPARE SPECIFICATIONS.
STEP 8. COLLECT SAMPLES.
STEP 9. MONITOR INSTALLATION, AND DURING & AFTER STORM EVENTS.

3.6 GEOTEXTILE COST CONSIDERATIONS


The total cost of a riprap-geotextile revetment system will depend on the actual application
and type of revetment selected. The following items should be considered:
1. grading and site preparation;
2. cost of geotextile, including cost of overlapping and pins versus cost of sewn
seams;
3. cost of placing geotextile, including special considerations for below-water
placement;
4. bedding materials, if required, including placement;
5. armor stone, concrete blocks, sand bags, etc.; and
6. placement of armor stone (dropped versus hand- or machine-placed).
For Item No. 2, the cost of overlapping includes the extra material required for the overlap,
cost of pins, and labor considerations versus the cost of field and/or factory seaming, plus the
additional cost of laboratory seam testing. These costs can be obtained from manufacturers,
but typical costs of a sewn seam are equivalent to 1.2 to 1.8 yd2 (1 to 1.5 m2) of geotextile.
Alternatively, the contractor can be required to supply the cost on an area covered or in-place
basis. For example, U.S. Army Corps of Engineers Specifications (CW-02215, 1977) require
FHWA NHI-07-092
Geosynthetics Engineering

3-20

Erosion Control Systems


August 2008

measurement for payment for geotextiles in streambank and slope protection to be on an inplace basis without allowance for any material in laps and seams. Further, the unit price
includes furnishing all plant, labor, material, equipment, securing pins, etc., and performing
all operations in connection with placement of the geotextile, including prior preparation of
banks and slopes. Of course, field performance should also be considered, and sewn seams
are generally preferred to overlaps. For many erosion protection projects, the decision
whether to sew or overlap is left to the contractor, with a bid item for the geotextile based on
the area to be covered.
Another important consideration for Items 2, 4, and 6 is the difference between Moderate
versus High Survivability geotextiles (Table 3-1) and its effect on the cost of bedding
materials and placement of armor stone. Class 1 geotextile materials typically cost 20%
more than Class 2 materials.
To determine cost effectiveness, benefit-cost ratios should be compared for the riprapgeotextile system versus conventional riprap-granular filter systems or other available
alternatives of equal technical feasibility and operational practicality. Average cost of
geotextile protection systems placed above the water level, including slope preparation,
geotextile cost of seaming or securing pins, and placement is approximately $2.50-$5.00 per
square yard, excluding the armor stone. Cost of placement below water level can vary
considerably depending on the site conditions and the contractor's experience. For belowwater placement, it is recommended that prebid meetings be held with potential contractors
to explore ideas for placement and discuss anticipated costs.

3.7 GEOTEXTILE SPECIFICATIONS


In addition to the general recommendations concerning specifications in Chapter 1, erosion
control specifications must include construction details (see Section 3.8), as the appropriate
geotextile will depend on the placement technique. In addition, the specifications should
require the contractor to demonstrate through trial sections that the proposed riprap
placement technique will not damage the geotextile.
Many erosion control projects may be better served by performance-type filtration tests that
provide an indication of long-term performance. Thus, in many cases, approved list-type
specifications, as discussed in Section 1.6, may be appropriate. To develop the list of
approved geotextiles, filtration studies (as suggested in Sections 1.5 and 3.4, Step 5) should
be performed using problem soils and conditions that exist in the localities where geotextiles
will be used. An approved list for each condition should be established. In addition,
FHWA NHI-07-092
Geosynthetics Engineering

3-21

Erosion Control Systems


August 2008

geotextiles should be classified as High or Moderate Survivability geotextiles, in accordance


with the index properties listed in Table 3-1 and construction conditions.
The following example specification is a combination of the AASHTO M288 (2006)
geotextile material specification and its accompanying construction/installation guidelines. It
includes the requirements discussed in Section 1.6 for a good specification. As with the
specification presented in Chapter 2, site-specific hydraulic and physical properties must be
appropriately selected and included.
EROSION CONTROL GEOTEXTILE SPECIFICATION
(after AASHTO M288, 2006)
1. SCOPE
1.1 Description. This specification is applicable to the use of a geotextile between energy absorbing
armor systems and the in-situ soil to prevent soil loss resulting in excessive scour and to prevent
hydraulic uplift pressure causing instability of the permanent erosion control system. This
specification does not apply to other types of geosynthetic soil erosion control materials such as
turf reinforcement mats.
2.

REFERENCED DOCUMENTS

2.1 AASHTO Standards


T88
Particle Size Analysis of Soils
T90
Determining the Plastic Limit and Plasticity Index of Soils
T99
The Moisture-Density Relationships of Soils Using a 2.5 kg Rammer and a 305 mm
Drop
2.2 ASTM Standards
D 123 Standard Terminology Relating to Textiles
D 276 Test Methods for Identification of Fibers in Textiles
D 4354 Practice for Sampling of Geosynthetics for Testing
D 4355 Test Method for Deterioration of Geotextiles from Exposure to Ultraviolet Light and
Water (Xenon Arc Type Apparatus)
D 4439 Terminology for Geosynthetics
D 4491 Test Methods for Water Permeability of Geotextiles by Permittivity
D 4632 Test Method for Grab Breaking Load and Elongation of Geotextiles
D 4751 Test Method for Determining Apparent Opening Size of a Geotextile
D 4759 Practice for Determining the Specification Conformance of Geosynthetics
D 4873 Guide for Identification, Storage, and Handling of Geotextiles
D 5141 Test Method to Determine Filtering Efficiency and Flow Rate for Silt Fence
Applications Using Site Specific Soil
D 5261 Test Method for Measuring Mass per Unit Area of Geotextiles
FHWA NHI-07-092
Geosynthetics Engineering

3-22

Erosion Control Systems


August 2008

D 6241 Test Method for Static Puncture Strength of Geotextiles and Geotextile
Related Products Using a 50-mm Probe
3.

PHYSICAL AND CHEMICAL REQUIREMENTS

3.1 Fibers used in the manufacture of geotextiles and the threads used in joining geotextiles by
sewing, shall consist of long chain synthetic polymers, composed of at least 95% by weight
polyolefins or polyesters. They shall be formed into a stable network such that the filaments or
yarns retain their dimensional stability relative to each other, including selvages.
3.2 Geotextile Requirements. The geotextile shall meet the requirements of following Table.
Woven slit film geotextiles (i.e., geotextiles made from yarns of a flat, tape-like character) will
not be allowed. All numeric values in the following table, except AOS, represent minimum
average roll values (MARV) in the weakest principal direction (i.e., average test results of any
roll in a lot sampled for conformance or quality assurance testing shall meet or exceed the
minimum values). Values for AOS represent maximum average roll values.
NOTE: The property values in the following table represent default values which
provide for sufficient geotextile survivability under most conditions. Minimum
property requirements may be reduced when sufficient survivability information
is available [see Notes a, b of Table 3-1 and Appendix D]. The engineer may also
specify properties different from that listed in the following Table based on
engineering design and experience.
Permanent Erosion Control Geotextile Requirements
Geotextile
All other geotextiles
Property
ASTM Test
Units
Woven
Elongation
Elongation
Method
Monofilament
< 50%(1)
> 50%(1)
Grab Strength
D 4632
N
1100
1400
900
(2)
Sewn Seam Strength
D 4632
N
990
1200
810
Tear Strength
D 4533
N
250
500
350
Puncture Strength
D 6241
N
2200
2750
1925
Percent In-situ Passing No.200 Sieve(3)
< 15
15 to 50
> 50
Permittivity
D 4491
sec-1
0.7
0.2
0.1
Apparent Opening Size
D 4751
mm
0.43
0.25
0.22
Ultraviolet Stability
D 4355
%
70% after 500 hours of exposure
NOTES:
(1) As measured in accordance with ASTM D 4632.
(2) When sewn seams are required.
(3) Based on grain size analysis of in-situ soil in accordance with AASHTO T88.

FHWA NHI-07-092
Geosynthetics Engineering

3-23

Erosion Control Systems


August 2008

4.

CERTIFICATION

4.1 The Contractor shall provide to the engineer a certificate stating the name of the manufacturer,
product name, style number, chemical composition of the filaments or yarns, and other pertinent
information to fully describe the geotextile.
4.2 The manufacturer is responsible for establishing and maintaining a quality control program to
assure compliance with the requirements of the specification. Documentation describing the
quality control program shall be made available upon request.
4.3 The manufacturers certificate shall state that the furnished geotextile meets MARV
requirements of the specification as evaluated under the manufacturers quality control program.
A person having legal authority to bind the manufacturer shall attest to the certificate.
4.4 Either mislabeling or misrepresentation of materials shall be reason to reject those geotextile
products.
5.

SAMPLING, TESTING, AND ACCEPTANCE

5.1 Geotextiles shall be subject to sampling and testing to verify conformance with this
specification. Sampling shall be in accordance with the most current ASTM D 4354 using the
section titled, Procedure for Sampling for Purchasers Specification Conformation Testing. In
the absence of purchasers testing, verification may be based on manufacturers certifications as
a result of a testing by the manufacturer of quality assurance samples obtained using he
procedure for Sampling or Manufacturers Quality Assurance (MQA) Testing. A lot size shall
be considered to be the shipment quantity of the given product or a truckload of the given
product, whichever is smaller.
5.2 Testing shall be performed in accordance with the methods referenced in this specification for
the indicated application. The number of specimens to test per sample is specified by each test
method. Geotextile product acceptance shall be based on ASTM D 4759. Product acceptance is
determined by comparing the average test results of all specimens within a given sample to the
specification MARV. Refer to ASTM D 4759 for more details regarding geotextile acceptance
procedures.
6.

SHIPMENT AND STORAGE

6.1 Geotextile labeling, shipment, and storage shall follow ASTM D 4873. Product labels shall
clearly show the manufacturer or supplier name, style number, and roll number. Each shipping
document shall include a notation certifying that the material is in accordance with the
manufacturers certificate.

FHWA NHI-07-092
Geosynthetics Engineering

3-24

Erosion Control Systems


August 2008

6.2 Each geotextile roll shall be wrapped with a material that will protect the geotextile, and the ends
of the roll, from damage due to shipment, water, sunlight, and contaminants. The protective
wrapping shall be maintained during periods of shipment and storage.
6.3 During storage, geotextile rolls shall be elevated off the ground and adequately covered to protect
them from the following: site construction damage, precipitation, extended ultraviolet radiation
including sunlight, chemicals that are strong acids or strong bases, flames including welding
sparks, temperatures in excess of 71EC (160EF), and any other environmental condition that may
damage the physical property values of the geotextile.
7.

CONSTRUCTION

7.1 General. Atmospheric exposure of geotextiles to the elements following lay down shall be a
maximum of 14 days to minimize damage potential.
7.2 Seaming.
a. If a sewn seam is to be used for the seaming of the geotextile, the thread used shall consist of
high strength polypropylene, or polyester. Nylon thread shall not be used. For erosion control
applications, the thread shall also be resistant to ultraviolet radiation. The thread shall be of
contrasting color to that of the geotextile itself.
b. For seams which are sewn in the field, the contractor shall provide at least a two m length of
sewn seam for sampling by the engineer before the geotextile is installed. For seams which are
sewn in the factory, the engineer shall obtain samples of the factory seams at random from any
roll of geotextile which is to be used on the project.
b.1 For seams that are field sewn, the seams sewn for sampling shall be sewn using the same
equipment and procedures as will be used for the production of seams. If seams are to be
sewn in both the machine and cross machine directions, samples of seams from both
directions shall be provided.
b.2 The contractor shall submit the seam assembly along with the sample of the seam. The
description shall include the seam type, stitch type, sewing thread, and stitch density.
7.3 Geotextile Placement.
a. The geotextile shall be placed in intimate contact with the soils without wrinkles or folds and
anchored on a smooth graded surface approved by the engineer. The geotextile shall be placed
in such a manner that placement of the overlying materials will not excessively stretch so as to
tear the geotextile. Anchoring of the terminal ends of the geotextile shall be accomplished
through the use of key trenches or aprons at the crest and toe of slope. [See Figure 3-2.]
FHWA NHI-07-092
Geosynthetics Engineering

3-25

Erosion Control Systems


August 2008

NOTE 1: In certain applications to expedite construction, 450 mm anchoring pins


placed on 600 to 1800 mm centers, depending on the slope of the covered area,
have been used successfully.
b. The geotextile shall be placed with the machine direction parallel to the direction of water
flow which is normally parallel to the slope for erosion control runoff and wave action (Figure
X1.4), and parallel to the stream or channel in the case of streambank and channel protection
(Figure X1.6). Adjacent geotextile sheets shall be joined by either sewing or overlapping.
Overlapped seams of roll ends shall be a minimum of 300 mm except where placed under water.
In such instances the overlap shall be a minimum of 1 m. Overlaps of adjacent rolls shall be a
minimum of 300 mm in all instances. See Figure 3-3 [this manual].
NOTE 2: When overlapping, successive sheets of the geotextile shall be overlapped
upstream over downstream, and/or upslope over downslope. In cases where wave action or
multidirectional flow is anticipated, all seams perpendicular to the direction of flow shall be
sewn.
c. Armor. The armor system placement shall begin at the toe and proceed up the slope.
Placement shall take place so as to avoid stretching resulting in tearing of the geotextile. Riprap
and heavy stone filling shall not be dropped from a height of more than 300 mm. Stone
weighing more than 450 N shall not be allowed to roll down the slope.
c.1 Slope protection and smaller sizes of stone filling shall not be dropped from a height
exceeding 1 m, or a demonstration provided showing that the placement procedures will not
damage the geotextile. In under water applications, the geotextile and backfill material
shall be placed the same day. All void spaces in the armor stone shall be backfilled with
small stone to ensure full coverage.
c.2 Following placement of the armor stone, grading of the slope shall not be permitted if the
grading results in movement of the stone directly above the geotextile.
d. Damage. Field monitoring shall be performed to verify that the armor system placement does
not damage the geotextile.
d.1 Any geotextile damaged during backfill placement shall be replaced as directed by the
engineer, at the contractors expense.
8.

METHOD OF MEASUREMENT
8.1 The geotextile shall be measured by the number of square yards computed from the payment
lines shown on the plans or from payment lines established in writing by the engineer. This
excludes seam overlaps, but shall include geotextiles used in crest and toe of slope treatments.

FHWA NHI-07-092
Geosynthetics Engineering

3-26

Erosion Control Systems


August 2008

8.2 Slope preparation, excavation and backfill, bedding, and cover material are separate pay
items.

9.

BASIS OF PAYMENT
9.1 The accepted quantities of geotextile shall be paid for per square yard in place.
9.2 Payment will be made under:
Pay Item
Erosion Control Geotextile

Pay Unit
Square Yard

3.8 GEOTEXTILE INSTALLATION PROCEDURES


Construction requirements will depend on specific application and site conditions.
Photographs of several installations are shown in Figure 3-2. The following general
construction considerations apply for most riprap-geotextile erosion protection systems.
Special considerations related to specific applications and alternate riprap designs will
follow.
3.8-1
1.

General Construction Considerations


Grade area and remove debris to provide smooth, fairly even surface.
a. Depressions or holes in the slope should be filled to avoid geotextile bridging
and possible tearing when cover materials are placed.
b. Large stones, limbs, and other debris should be removed prior to placement to
prevent fabric damage from tearing or puncturing during stone placement.

2.

Place geotextile loosely, laid with machine direction in the direction of anticipated
water flow or movement.

3.

Seam or overlap the geotextile as required.


a. For overlaps, adjacent rolls of geotextile should be overlapped a minimum of 1
ft (0.3 m). Overlaps should be in the direction of water flow and stapled or
pinned to hold the overlap in place during placement of stone. Steel pins are
normally 3/16-in. (5 mm) diameter, 18 in. (0.5 m) long, pointed at one end, and
fitted with 1-in. (38 mm) diameter washers at the other end. Pins should be
spaced along all overlap alignments at a distance of approximately 3 ft (1 m)
center to center.

FHWA NHI-07-092
Geosynthetics Engineering

3-27

Erosion Control Systems


August 2008

(a)

(b)

(c)
Figure 3-2. Erosion control installations: a) installation in wave protection revetment; b)
river shoreline application; and c) stream application.

FHWA NHI-07-092
Geosynthetics Engineering

3-28

Erosion Control Systems


August 2008

b.
c.

4.

The geotextile should be pinned loosely so it can easily conform to the ground
surface and give when stone is placed.
If seamed, seam strength should equal or exceed the minimum seam
requirements indicated in Section 1.6 on Specifications.

The maximum allowable slope on which a hard armor-geotextile system can be


placed is equal to the lowest soil-geotextile friction angle for the natural ground or
stone-geotextile friction angle for cover (armor) materials. Additional reductions in
slope may be necessary due to hydraulic considerations and possible long-term
stability conditions. For slopes greater than 2.5 to 1, special construction
procedures will be required, including toe berms to provide a buttress against
slippage, loose placement of geotextile sufficient to allow for downslope movement,
elimination of pins at overlaps, increase in overlap requirements, and possible
benching of the slope. Care should be taken not to put irregular wrinkles in the
geotextile because erosion channels can form beneath the geotextile.
Geocell containment systems, as covered in Section 3.10, and block systems can be
used on steeper slopes with special cables and anchorage systems. Special
construction procedures are also necessary to place the cells and the infill material.
Depending on the site specific runoff conditions, the cells can be filled either with
gravel (hard armor) or soil and vegetated (soft).

5. For streambank and wave action applications, the geotextile must be keyed in at the
bottom of the slope. If the riprap-geotextile system cannot be extended a few yards
above the anticipated maximum high water level, the geotextile should be keyed in at
the crest of the slope. The geotextile should no be keyed in at the crest until after
placement of the riprap. Alternative key details are shown in Figure 3-3.
6.

Place revetment (cushion layer and/or riprap) over the geotextile width, while
avoiding puncturing or tearing it.
a. Revetment should be placed on the geotextile within 14 days.
b. Placement of armor cover will depend on the type of riprap, whether quarry
stone, sandbags (which may be constructed of geotextiles), interlocked or
articulating concrete blocks, soil-cement filled bags, filled geocells, or other
suitable slope protection is used.
c. For sloped surfaces, placement should always start from the base of the slope,
moving up slope and, preferably, from the center outward.
d. In no case should stone weighing more than 90 lbs (40 kg) be allowed to roll
downslope on the geotextile.

FHWA NHI-07-092
Geosynthetics Engineering

3-29

Erosion Control Systems


August 2008

e. Field trials should be performed to determine if placement techniques will


damage the geotextile and to determine the maximum height of safe drop. As a
general guideline, for High Survivability (Class 1) geotextiles (Table 3-1) with
no cushion layer, height of drop for stones less than 220 lbs (100 kg) should be
less than 3 ft (1 m). For High Survivability (Class 1) geotextiles (Table 3-1) or
Moderate Survivability (Class 2) geotextiles with a 6-in. (150 mm) thick small
aggregate cushion layer, height of drop for stones less than 220 lbs (100 kg)
should be less than 3 ft (1 m). Stones greater than 220 lbs (100 kg) should be
placed with no free fall unless field trials demonstrate they can be dropped
without damaging the geotextile.
f. Grading of slopes should be performed during placement of riprap. Grading
should not be allowed after placement if it results in stone movement directly
on the geotextile.
As previously indicated, construction requirements will depend on specific application and
site conditions. In some cases, geotextile selection is affected by construction procedures.
For example, if the system will be placed below water, a geotextile that facilitates such
placement must be chosen. The geotextile may also affect the construction procedures. For
example, the geotextile must be completely covered with riprap for protection from longterm exposure to ultraviolet radiation. Sufficient anchorage must also be provided by the
riprap for weighting the geotextile in below-water applications. Other requirements related
to specific applications are depicted in Figure 3-4 and are reviewed in the following
subsections (from Christopher and Holtz, 1985).
3.8-2
Cut and Fill Slope Protection
Cut and fill slopes are generally protected using an armor stone over a geotextile-type
system. Special consideration must be given to the steepness of the slope. After grading,
clearing, and leveling a slope, the geotextile should be placed directly on the slope. When
possible, geotextile placement should be placed parallel to the slope direction. A minimum
overlap of 1 ft (0.3 m) between adjacent roll ends and a minimum 1 ft (0.3 m) overlap of
adjacent strips is recommended. It is also important to place the up-slope geotextile over the
down-slope geotextile to prevent overlap separation during aggregate placement. When
placing the aggregate, do not push the aggregate up the slope against the overlap. Generally,
cut and fill slopes are protected with armor stone or geocells, and the recommended
placement procedures in Section 3.8-1 should be followed.

FHWA NHI-07-092
Geosynthetics Engineering

3-30

Erosion Control Systems


August 2008

Figure 3-3. Construction of hard armor erosion control systems (a., b. after Keown and
Dardeau, 1980; c. after Dunham and Barrett, 1974). 1 m = 3.3 ft.
FHWA NHI-07-092
Geosynthetics Engineering

3-31

Erosion Control Systems


August 2008

Figure 3-4. Special construction requirements related to specific hard armor erosion control
applications. 1 m = 3.3 ft.
FHWA NHI-07-092
Geosynthetics Engineering

3-32

Erosion Control Systems


August 2008

Figure 3-4. Special construction requirements related to specific hard armor erosion control
applications (cont.).
FHWA NHI-07-092
Geosynthetics Engineering

3-33

Erosion Control Systems


August 2008

3.8-3
Streambank Protection
For streambank protection, selecting a geotextile with appropriate clogging resistance to
protect the natural soil and meet the expected hydraulic conditions is extremely important.
Should clogging occur, excess hydrostatic pressures in the streambank could result in slope
stability problems. Do not solve a surface erosion problem by causing a slope stability
problem!
Detailed data on geotextile installation procedures and relevant case histories for streambank
protection applications are given by Keown and Dardeau (1980). Construction procedures
essentially follow the procedures listed in Section 3.8-1. The geotextile should be placed on
the prepared streambank with the machine direction placed parallel to the bank (and parallel
to the direction of stream flow). Adjacent rolls of geotextile should be seamed, sewed, or
overlapped; if overlapped, secure the overlap with pins or staples. A 1 ft (0.3 m) overlap is
recommended for adjacent roll edges, with the upstream roll edge placed over the
downstream roll edge. Roll ends should be overlapped 3 ft (1 m) and offset as shown in
Figure 3-4a. The upslope roll should overlap the downslope roll.
The geotextile should be placed along the bank to an elevation well below mean low water
level based on the anticipated flows in the stream. Existing agency design criteria for
conventional nongeotextile streambank protection could be utilized to locate the toe of the
erosion protection system. In the absence of other specifications, placement to a vertical
distance of 3 ft (1 m) below mean water level, or to the bottom of the streambed for streams
shallower than 3 ft (1 m), is recommended. Geotextiles should either be placed to the top of
the bank or at a given distance up the slope above expected high water level from the
appropriate design storm event, including whatever requirements are normally used for
conventional (nongeotextile) streambank protection systems. In the absence of other
specifications, the geotextile should extend vertically a minimum of 18 in. (0.5 m) above the
expected maximum water stage, or at least 3 ft (1 m) beyond the top of the embankment if
less than 18 in. (0.5 m) above expected water level.
If strong water movements are expected, the geotextile must be keyed in at the top and toed
in at the bottom of the embankment. The riprap or filled geocells should be extended beyond
the geotextile 18 in. (0.5 m) or more at the toe and the crest of the slope. If scour occurs at
the toe and the surface armor beyond the geotextile is undermined, it will in effect toe into
the geotextile. The whole unit thus drops, until the toed-in section is stabilized. However, if
the geotextile extends beyond the stone and scour occurs, the geotextile will flap in the water
action and tend to accelerate the formation of a scour pit or trench at the toe. Alternative toe
treatments are shown in Figure 3-3. The trench methods in Figures 3-3a and 3-3b require
excavating a trench at the toe of the slope. This may be a good alternative for new
FHWA NHI-07-092
Geosynthetics Engineering

3-34

Erosion Control Systems


August 2008

construction; however, it should be evaluated with respect to slope stability when a trench
will be excavated at the toe of a potentially saturated slope below the water level. Keying in
at the top can consist of burying the top bank edge of the geotextile in a shallow trench after
placement of the armor material. This will provide resistance to undermining from
infiltration of over-the-bank precipitation runoff, and also provide stability should a storm
greater than anticipated occur. However, unless excessive quantities of runoff are expected
and stream flows are relatively small, this step is usually omitted.
The armoring material (e.g., riprap, sandbags, blocks, filled geocells) must be placed to avoid
tearing or puncturing the geotextile, as indicated in Section 3.8-1.
3.8-4
Precipitation Runoff Collection and Diversion Ditches
Runoff drainage from cut slopes along the sides of roads and in the median of divided
highways is normally controlled with one or more gravity flow ditches. Runoff from the
pavement surface and shoulder slopes are collected and conveyed to drop inlets, stream
channels, or other highway drainage structures. If a rock protection-geotextile system is used
to control localized ditch erosion problems, select and specify the geotextile using the
properties indicated in Table 3-1. Geotextile requirements for ditch linings are less critical
than for other types of erosion protection, and minimum requirements for noncritical,
nonsevere applications can generally be followed. If care is taken during construction, the
protected strength requirements appear reasonable. The geotextile should be sized with AOS
to prevent scour and piping erosion of the underlying natural soil and to be strong enough to
survive stone placement.
The ditch alignment should be graded fairly smooth, with depressions and gullies filled and
large stones and other debris removed from the ditch alignment. The geotextile should be
placed with the machine direction parallel to the ditch alignment. Most geotextiles are
available in widths of 6.6 ft (2 m) or more, and, thus, a single roll width of geotextile may
provide satisfactory coverage on the entire ditch. If more than one roll width of geotextile is
required, it is better to sew adjacent rolls together. This can be done by the manufacturer or
on site. Again, for seams, the required strength of the seam should meet the minimum seam
requirements in Table 3-1. The longitudinal seam produced by roll joining will run parallel
with the ditch alignment. Geotextile widths should be ordered to avoid overlaps at the
bottom of the ditch, since this is where maximum water velocity occurs. Roll ends should
also be sewn or overlapped and pinned or stapled. If overlap is used, then an overlap of at
least 3 ft (1 m) is recommended. The upslope roll end should be lapped over the downslope
roll end, to prevent in-service undermining. Pins or staples should be spaced so slippage will
not occur during stone placement or after the ditch is placed in service.

FHWA NHI-07-092
Geosynthetics Engineering

3-35

Erosion Control Systems


August 2008

Cover stone, sandbags, infilled geocells or other material intended to dissipate precipitation
runoff energy should be placed directly on the geotextile, from downslope to upslope. Cover
stone should have sufficient depth and gradation to protect the geotextile from ultraviolet
radiation exposure. Again, the cover material should be placed with care, especially if a
Class 2 geotextiles has been selected. A cross section of the proper placement is shown in
Figure 3-4c. Vegetative cover can be established through the geotextile and stone cover if
openings in the geotextile are sufficient to support growth. If a vegetative cover is desirable,
geotextiles should be selected on the basis of the largest opening possible, or consider the use
of RECPs and TRMs, as discussed in Sec. 3.12.
3.8-5
Wave Protection Revetments
Because of cyclic flow conditions, geotextiles used for wave protection systems in most
cases should be selected on the basis of severe criteria. The geotextile should be placed in
accordance with the procedures listed in Section 3.8-1.
If a geotextile will be placed where existing riprap, rubble, or other armor materials placed
on natural soil have been unsuccessful in retarding wave erosion, site preparation could
consist of covering the existing riprap with a filter sand. The geotextile could then be
designed with less rigorous requirements as a filter for the sand than if the geotextile is
required to filter finer soils.
The geotextile is unrolled and loosely laid on the smooth graded slope. The machine
direction of the geotextile should be placed parallel to the slope direction, rather than
perpendicular to the slope, as was recommended in streambank protection. Thus, the long
axis of the geotextile strips will be parallel to anticipated wave action. Sewing of adjacent
rolls or overlapping rolls and roll ends should follow the steps described in Section 3.8-1,
except that a 3 ft (1 m) overlap distance is recommended by the Corps of Engineers for
underwater placement (Figure 3-4). Again, securing pins (requirements per Section 3.8-1)
should be used to hold the geotextile in place.
If a large part of the geotextile is to be placed below the existing tidal level, special
fabrication and placement techniques may be required. It may be advantageous to pre-sew
the geotextile into relatively large panels and pull the prefabricated panels downslope,
anchoring them below the waterline. Depending upon the placement scheme used, selection
of a floating or nonfloating geotextile may be advantageous. In some cases with very strong
storm waves, composite mats made of geotextiles, fascines, and other bedding materials are
constructed on land, rolled up, and then unrolled off of an offshore barge with divers and
weights facilitating underwater placement.

FHWA NHI-07-092
Geosynthetics Engineering

3-36

Erosion Control Systems


August 2008

Because of potential wave action undermining, the geotextile must be securely toed-in using
one of the schemes shown in Figure 3-3. Also, a key trench should be placed at the top of the
bank, as shown in Figure 3-3a, to prevent revetment stripping should the embankment be
overtopped by wave action during high-level storm events.
Riprap or cover stone should be placed on the geotextile from downslope to upslope, and
stone placement techniques should be designed to prevent puncturing or tearing of the
geotextile. Drop heights should follow the recommendations stated in the general
construction criteria (see 3.8-1). Riprap or cover stone can also be placed underwater by
cranes or bottom dump barges.
3.8-6
Scour Protection
Scour, because of high flows around or adjacent to structures in rivers or coastal areas,
generally requires scour protection for structures. Scour protection systems generally fall
under the critical and/or severe design criteria for geotextile selection.
An extremely wide variety of transportation-associated structures are possible and, thus,
numerous ways exist to protect such structures with riprap geotextile systems. A typical
application is shown in Figure 3-4d. In all instances, the geotextile is placed on a smoothly
graded surface as stated in the general construction requirements (Section 3.8-1). Such site
preparation may be difficult if the geotextile will be placed underwater, but normal stream
action may provide a fairly smooth streambed. In bridge pier protection or culvert approach
and discharge channel protection applications, previous high-velocity stream flow may have
scoured a depression around the structure. Depressions should be filled with granular
cohesionless material. It is usually desirable to place the geotextile and riprap in a shallow
depression around bridge piers to prevent unnecessary constriction of the stream channel.
The geotextile should normally be placed with the machine direction parallel to the
anticipated water flow direction. Seaming and/or overlapping of adjacent rolls should be
performed as recommended in general construction requirements (Section 3.8-1). When roll
ends are overlapped, the upstream ends should be placed over the downstream end. As
necessary and appropriate, the geotextile may be secured in place with steel pins, as
previously described. Securing the geotextile in the proper position may be of extreme
importance in bridge pier scour protection. However, under high-flow velocities or in deep
water, it will be difficult, if not impossible, to secure the geotextile with steel pins alone.
Underwater securing methods must then be developed, and they will be unique for each
project. Alternative methods include floating the geotextile into place, then filling from the
center outward with stones, building a frame to which the geotextile can be sewn; using a
heavy frame to submerge and anchor the geotextile; or constructing a light frame, then
FHWA NHI-07-092
Geosynthetics Engineering

3-37

Erosion Control Systems


August 2008

floating the geotextile and sinking it with riprap. In any case, it may be desirable to specify a
geotextile which will either float or sink, depending upon the construction methods chosen.
In general, geotextiles with a bulk density greater than 1 g/cm3 will sink (i.e., provided the air
contained in the geotextile can be readily removed by submersion) while those less than 1
g/cm3 will float.
Riprap, precast concrete blocks, bedding materials if used, or other elements placed on the
geotextile should be placed without puncturing or tearing the geotextile. Drop heights should
be selected on the basis of geotextile strength criteria, as discussed in the general
construction requirements (Section 8.3-1).

3.9 GEOTEXTILE FIELD INSPECTION


In addition to the general field inspection checklist presented in Table 1-4, the field inspector
should pay close attention to construction procedures. If significant movement (greater than
6 in.) of stone or concrete riprap occurs during or after placement, the blocks should be
removed so that the overlaps can be inspected to ensure they are still intact. As indicated in
Section 3.8, field trials should be performed to demonstrate that placement procedures will
not damage the geotextile. If damage is observed, the engineer should be contacted, and the
contractor should be required to change the placement procedure. This procedure should be
in the construction specifications for every erosion control project.
For below-water placement or placement adjacent to structures requiring special installation
procedures, the inspector should discuss placement details with the engineer, and inspection
requirements and procedures should be worked out in advance of construction.

3.10

GEOCELLS

Geocells infilled with sand, gravel or concrete provide an alternative armor system for
permanent erosion protection. Geocells are three dimensional cellular structures provide
confinement and reinforcement to the infilled soils. Geocells are made from polyethylene
(low and high density) strips 3, 4, 6 or 8 in. (75, 100, 150 or 200 mm) wide that may be solid
or contain small holes. The strips are periodically interconnected to form expandable
rectangular or square panels up to 30 ft (9.1 m) on a side. The panels are flattened for easy
transport and then stretched out and expanded on site to form the cellular or honeycomb
structures. Depending on the climate and type of backfill, vegetation is sometimes
FHWA NHI-07-092
Geosynthetics Engineering

3-38

Erosion Control Systems


August 2008

established on the infilled geocell slopes. The type of infill selected depends on the
hydraulics, soil conditions, and aesthetics.
The various manufacturers can provide materials properties and physical characteristics of
geocells. As far as we know, there are no special design rules for geocell installations.
Conventional geotechnical analyses for slope stability, foundations, compaction, etc., are
appropriate, depending on the specific project requirements. For erosion protection systems,
the same hydraulic considerations mentioned earlier in this chapter apply.
Site preparation is similar to what is normally done for any geosynthetic installation. The
site is cleared and grubbed, tree stumps and boulders are removed, and the surface graded
and smoothed as required. The flattened geocell panels are carried to the site and expanded to
their full panel dimensions. The panels are periodically staked or pinned to the subgrade
slope every two to four cells as well as at the junctions between panels. Some manufactures
have a cable threaded through the each panel that helps to expand and secure them to the
subgrade. Depending on the slope angle and length, an anchor trench at the top of the slope
and along the edges of the protected area may be required. Benches on the slope are
sometimes necessary for stability and to simplify construction.
Placement of the infill material can be by hand, endloader, backhoe, crane-supported
clamshell bucket, or even a conveyor system. Infill materials are usually placed a few cm
above the top of the cells and appropriately compacted. Hydroseeding and mulching may
also be appropriate. Edges must be protected (eg., with toe trenches) to prevent loss of fill
due to erosion and undermining.

3.11 EROSION CONTROL MATS


Unpaved areas are susceptible to erosion by high-velocity flow. Where flow is intermittent, a
grass cover will provide protection against erosion. By reinforcing the grass cover, the
resulting biocomposite layer will enhance the erosion resistance. Geosynthetics used for
erosion control include turf reinforcement mats (TRMs) and extruded plastic mats. Both are
three-dimensional mats made of synthetic filaments, meshes, plastic sheets, or webbings that
serve to reinforce the vegetation root mass, provide tractive resistance to fairly high flows,
and reduce the impact of rainfall while enhancing the establishment of vegetation cover.
TRMs are non-biodegradable rolled erosion control products (RECPs). TRMs are a
reinforced grass system capable of withstanding short-term (e.g., 2 hours), high velocity
(e.g., 20 ft/s {6 m/s}) flows with minimal erosion.

FHWA NHI-07-092
Geosynthetics Engineering

3-39

Erosion Control Systems


August 2008

Erosion control blankets (ECBs) and other geosynthetics such as mulch control nets (MCNs)
are covered in Chapter 4. These are biodegradable RECPs used to enhance the establishment
of vegetation for temporary erosion control applications where vegetation alone can provide
sufficient erosion protection.
The principal applications of TRMs are in highway stormwater runoff ditches, auxiliary
spillways of retention dams, and protection of embankments and reinforced steep slopes
(Chapter 8) against erosion by heavy precipitation or flooding events. TRMs are used for
temporary (e.g., 2 hours), high-velocity flow areas, and is usually not suitable for long-term
high velocity flow applications suited for hard armor systems.
Any waterway lined with TRMs requires inspection and maintenance, and some of the
materials involved may be susceptible to damage, particularly by vandalism. If it is apparent
that these considerations are a problem, then reinforced grass should not be used. However,
the aesthetic advantage of a soft armor lining with a TRM often outweighs potential
disadvantages.
The performance of TRMs is determined by a complex interaction of hydraulic, geotechnical,
and botanical elements. At present, the physical processes and the engineering properties of
geotextiles and grass cannot be quantitatively described. Thus, the design approach is largely
empirical and involves a systematic consideration of each constituent element's behavior
under service conditions. Specific products have been tested in laboratory flume tests to
empirically quantify the tractive shear forces and velocities they can withstand as a function
of flow time.
This section provides only a summary of the design principles and construction procedures
for erosion control mats and TRMs. For detailed information on planning, design,
specifications, construction, on-going management, and support research on TRMs, see
Hewlett, et al. (1987).
Another sources of information on RECPs is the Erosion Control Technology Council
(ECTC). Their website (www.ectc.org) has downloadable documents, design tools, test
methods, case histories, and model specifications for RECPs, for both permanent and
temporary applications. ASTM also has some standard test methods for RECPs. The
Specifier's Guide published each December in the Geosynthetics magazine (formerly
Geotechnical Fabrics Report), published by the Industrial Fabrics Association contains
specific properties of TRMs provided by the manufacturers.

FHWA NHI-07-092
Geosynthetics Engineering

3-40

Erosion Control Systems


August 2008

One of the best sources of performance information is the Texas Department of


Transportation as performed by the Texas Transportation Institute Hydraulics, Sediment &
Erosion Control Laboratory. This agency tests candidate erosion control materials and
categorizes them into classes and types in an approved materials list. Information on the test
program and the results of tests on specific products are available on their web site:
http://www.dot.state.tx.us/services/maintenance/erosion_control.htm.
3.11-1 Summary of Planning, Design, and Installation
The planning stage involves assessing the feasibility of constructing a reinforced grass
system in a particular situation and establishing the basic design parameters. To be
considered, among other things, are design frequency and duration of flow, properties of
subsoil, climate, appearance, usage in no-flow conditions (e.g., agricultural, park, etc.), risk
of vandalism, cost of failure, access to site, method of construction, and capital and
maintenance costs..
The hydraulic design parameters are the velocity, shear stress, and duration of flow, as well
as the erosion resistance of various surface treatments. Figure 3-5 gives approximate
maximum design velocities and flow durations for various types of surfaces, from bare soils
through hard armor systems. For the hydraulics design, see HEC-15 (FHWA, 2005). Note
that the ECTC design procedures also depend on this reference.
The principal geotechnical consideration is the effect that water will have on the subsoil. This
includes seepage from adjacent slopes, rainfall, and flowing water that infiltrates the system.
The stability of slopes during normal or dry conditions as well as during and immediately
following flow should be investigated, and if necessary for increased stability provide
localized drainage to relieve excess pore pressures. Possible settlement of the subsoil should
also be estimated to see whether the armor layer is flexible enough to accommodate that
movement.
Botanical considerations include the type of grass that is appropriate for the soil conditions,
climate, and management requirements. Properties of the TRMs that are important for
stabilization and enhancement of vegetative growth include:
C the tensile strength (required for loading and survivability),
C strength after UV exposure (longevity),
C flexibility (helps maintain intimate contact with the subgrade important for rapid
seedling emergence and minimizing soil loss), and
C construction of the mat (helps to stabilize the vegetation in the matrix).

FHWA NHI-07-092
Geosynthetics Engineering

3-41

Erosion Control Systems


August 2008

Figure 3-5. Recommended maximum design velocities and flow durations for erosion
resistance of various surface materials and treatment (after Hewlet et al., 1987 and
Theisen, 1992).

Other design issues include:


1. Anchorage details for the TRM, including type and length of anchorage pins or
stakes, spacing across and along the edges the mat, roll end anchorage, adjacent rolls
and downslope shingling, and anchorage at the top of the slope or embankment.
2. Avoiding unwanted erosion in the crest, channel, and toe areas as well as the
transitions between two or more plane surfaces.
3. Construction and installation details such as foundation preparation, transition to
adjacent structures, placement requirements, etc.
In areas of sensitive wildlife habitat, a TRM with a very small mesh (<3/16 in2{<5 mm2})
openings is recommended (Barton and Kinkead, 2005).
FHWA NHI-07-092
Geosynthetics Engineering

3-42

Erosion Control Systems


August 2008

Details for each of these requirements are in Hewlett, et al. (1987), on the Texas DOT
website, and on the ECTC website. Specific installation guidelines are usually available
from the manufacturers of TRMs.
3.11-2 Specification
The following example specification for erosion control mats is after the Texas Department
of Transportation (2004) specification for soil retention blankets (SRB). This agency tests
candidate erosion control materials and categorizes them into classes and types in an
approved materials list.
ITEM 169
SOIL RETENTION BLANKETS
169.1. Description. Provide and install soil retention blankets (SRB) as shown on the plans or as
directed.
169.2. Materials. Provide only SR-B that are on the approved list published in "Field Performance
of Erosion Control Products," available from the Maintenance Division. Use material of the
following class and type as shown on the plans and provide a copy of the manufacturer's label for the
selected product.
A. Class I - Slope Protection.
1. Type A. Slopes 3:1 or flatter - clay soils,
2. Type B. Slopes 3:1 or flatter - sandy soils,
3. Type C. Slopes steeper than 3:1 - clay soils, and
4. Type D. Slopes steeper than 3:1 - sandy soils.
B. Class 2 - Flexible Channel Liners.
1. Type E. Biodegradable materials with shear stress less than 1.0 lb. per square foot,
2. Type F. Biodegradable materials with shear stress 1.0 to 2.0 lb. per square foot,
3. Type G. Nonbiodegradable materials with shear stress 2.0 to 5.0 lb. per square foot, and
4. Type H. Nonbiodegradable materials with shear stress equal to or greater than 5.0 lb. per
square foot.
169.3. Construction. Place the SRB within 24 hr. after the seeding or sodding operations, or when
directed, in accordance with the "SRB Product Installation Sheet," available from the Maintenance
Division. Installation includes the repair of ruts, reseeding or resodding, and the removal of rocks,
clods, and other foreign materials which may prevent contact of the blanket with the soil.
169.4. Measurement. This Item will be measured by the square yard of surface area covered.
FHWA NHI-07-092
Geosynthetics Engineering

3-43

Erosion Control Systems


August 2008

169.5. Payment. The work performed and materials furnished in accordance with this Item and
measured as provided under "Measurement" will be paid for at the unit price bid for "Soil Retention
Blankets" of the class and type specified. This price is full compensation for equipment, materials,
labor, tools, and incidentals.

3.12 REFERENCES
AASHTO (2006). Standard Specifications for Geotextiles - M 288, Standard Specifications
for Transportation Materials and Methods of Sampling and Testing, 25th Edition,
American Association of State Transportation and Highway Officials, Washington, D.C.
AASHTO (2005). Model Drainage Manual, 3rd edition, American Association of State
Highway and Transportation Officials, Washington, D.C.
ASTM (2006). Annual Books of ASTM Standards, American Society for Testing and
Materials, West Conshohocken, PA:
Volume 4.08 (I), Soil and Rock
Volume 4.09 (II), Soil and Rock; Geosynthetics
ASTM Test Methods - see Appendix E.
Barton, C. and Kinkead, K. (2005). Do Erosion Control and Snakes Mesh, Journal of Soil
and Water Conservation, Vol. 60, No. 2, 32 pp.
Casagrande, A. (1938). Notes on Soil MechanicsFirst Semester, Harvard University
(unpublished), 129 pp.
Cedergren, H.R. (1989). Seepage, Drainage, and Flow Nets, Third Edition, John Wiley and
Sons, New York, 465p.
Christopher, B.R. and Holtz, R.D. (1985). Geotextile Engineering Manual, FHWA-TS86/203, March, 1044 p.
Dunham, J.W. and Barrett, R.J. (1974). Woven Plastic Cloth Filters for Stone Seawalls,
Journal of the Waterways, Harbors, and Coastal Engineering Division, American Society
of Civil Engineers, New York, February.
FHWA (2006). Hydraulic Design of Energy Dissipators for Culvert and Channels,
Hydraulic Engineering Circular No. 14, Federal Highway Administration, FHWA-NHI06-086.
FHWA (2005). Design of Roadside Channels with Flexible Linings, Hydraulic Engineering
Circular No. 15, Third Edition, Federal Highway Administration, FHWA-IF-05-114.
FHWA NHI-07-092
Geosynthetics Engineering

3-44

Erosion Control Systems


August 2008

FHWA (2001). Bridge Scour and Stream Instability Countermeasures, Experience,


Selection, and Design Guidance, Hydraulic Engineering Circular No. 23, Second Edition,
Federal Highway Administration, Second Edition, FHWA NHI-01-003.
FHWA (2001). Evaluating Scour at Bridges, Hydraulic Engineering Circular No. 18,
Fourth Edition, Federal Highway Administration, FHWA NHI-01-001.
FHWA (1989). Design of Riprap Revetment, Hydraulic Engineering Circular No. 11,
Federal Highway Administration, FHWA-IP-89-016.
Hewlett, H.W.M., Boorman, L.A. and Bramley, M.E. (1987). Design of Reinforced Grass
Waterways Report 116, Construction Industry Research and Information Association,
London, U.K., 116 p.
Holtz, R. D. and Kovacs, W. D. (1981). An Introduction to Geotechnical Engineering,
Prentice-Hall, p. 210.
Keown, M.P. and Dardeau, E.A., Jr. (1980), Utilization of Filter Fabric for Streambank
Protection Applications, TR HL-80-12, Hydraulics Laboratory, U.S. Army Engineer
Waterways Experiment Station, Vicksburg, MS.
Texas Department of Transportation (2004). Standard Specifications for Construction and
Maintenance of Highways, Streets, and Bridges.
Theisen, M.S. (1992). Geosynthetics in Erosion Control and Sediment Control, Geotechnical
Fabrics Report, Industrial Fabrics Association International, St. Paul, MN, May/June pp.
26-35.
U.S. Army Corps of Engineers (1977). Civil Works Construction Guide Specification for
Plastic Filter Fabric, Corps of Engineer Specifications No. CW-02215, Office, Chief of
Engineers, U.S. Army Corps of Engineers, Washington, D.C.

FHWA NHI-07-092
Geosynthetics Engineering

3-45

Erosion Control Systems


August 2008

FHWA NHI-07-092
Geosynthetics Engineering

3-46

Erosion Control Systems


August 2008

4.0 TEMPORARY RUNOFF AND SEDIMENT CONTROL

4.1

INTRODUCTION

Geotextiles, geosynthetic erosion control blankets (ECBs), and other biodegradable rolled
erosion control products (RECPs) such as mulch control nets (MCNs) and turf reinforcing
mats (TRMs) can be used to temporarily control and minimize erosion and sediment
transport during construction. Four specific application areas have been identified:

Geotextile silt fences can be used as a


substitute for hay bales or brush piles to
remove suspended particles from
sediment-laden runoff water.

Geotextiles can be used as a turbidity


curtain placed within a stream, lake, or
other body of water to retain suspended
particles and allow sedimentation to
occur.

Special soil retention blankets, made of


both natural and synthetic grids, meshes,
nets, fibers, and webbings, can be used to
provide tractive resistance and resist
water velocity on slopes. These products
retain seeds and add a mulch effect to
promote the establishment of a vegetative
cover. Light-weight turf reinforcing mats
as described in Section 3.11 may also be
used.

FHWA NHI-07-092
Geosynthetics Engineering

4-1

Temporary Runoff and Sediment Control


August 2008

Geotextiles held in place by pins, riprap,


or sandbags can be used to temporarily
control erosion in temporary diversion
ditches, culvert outfalls, embankment
slopes, etc. Alternatively, ECBs can be
used for temporary erosion control until
vegetation can be established in the
ditch.

The main advantages of using geotextile silt fences over conventional techniques in sediment
control applications include the following:
The geotextile can be designed for the specific application, while conventional
techniques are basically trial-and-error.
Geotextile silt fences in particular often prove to be very cost-effective, especially in
comparison to hay bales, considering ease of installation and material costs.
Control of geosynthetics by material specifications is easier than for hay bales, the
traditional erosion control method.
For runoff control, geosynthetic ECBs are designed to help mitigate immediate erosion
problems and provide long-term stabilization by promoting the establishment of a vegetative
cover. The main advantages of using ECBs include the following:
Vegetative systems have desirable aesthetics.
Products are lightweight and easy to handle and install.
Temporary, degradable products improve establishment of vegetation.
Continuity of protection is generally better over the entire protected area.
Empirically predictable performance; traditional techniques such as seeding, mulch
covers, and brush or hay bale barriers, are often less predictable and reliable.
The following sections review the function, design, selection, specification, and installation
procedure for geosynthetics used as silt fences, turbidity curtains, and erosion control
blankets. Design of geotextiles in temporary runoff control systems to control ditch erosion
is similar to Sec. 3.8. Additional information and training are available in the NHI course
entitled Design and Implementation of Erosion and Sediment Control (No. 142054) (see
www.nhi.fhwa.dot.gov), developed in a joint effort between FHWA and the Environmental
Protection Agency (EPA).

FHWA NHI-07-092
Geosynthetics Engineering

4-2

Temporary Runoff and Sediment Control


August 2008

4.2

FUNCTION OF SILT FENCES

In most applications, a geotextile silt fence is placed downslope from a construction site or
newly graded area to prevent or at least reduce sediment being transported by runoff to the
surrounding environment. Sometimes silt fences are used in permanent or temporary
diversion ditches and swales for the same purpose, although even with special designs, they
have had varying success.
A silt fence primarily functions as a temporary dam (Mallard and Bell, 1981). It retains
water long enough for suspended fine sand and coarse silt particles in the runoff to settle out
before they reach the fence. Generally, a retention time of 25 to 30 minutes is sufficient, so
flow through the geotextile must provide this retention time. Although smaller pore opening
sizes and lower permittivity of the geotextile will allow finer particles to settle out, some
water must be able to pass through the fence to prevent possible overtopping. Appropriate
applications of silt fences are along the site perimeter, below disturbed areas subject to sheet
and rill erosion and sheet flow, and below the toe of exposed and erodible slopes.
With a retention time of 25 to 30 minutes, not all the silt and clay in suspension will settle out
before reaching the fence. Therefore, water flowing through the fence will still contain some
suspended fines. Removal of fines by the geotextile creates a difficult filtration condition. If
the openings in the geotextile (i.e., AOS) are small enough to retain most of the suspended
fines, the geotextile will blind (Fig. 2.3) and its permeability will be reduced so that bursting
or overtopping of the fence could occur. Therefore, it is better to have some geotextile
openings large enough to allow fine silt and clay particles to easily pass through. Even if
some silt passes through the fence, the flow velocity will be small, and some fines may settle
out. If the application is critical, e.g., when the site is immediately adjacent to
environmentally sensitive wetlands or streams, multiple silt fences can be used. A second
fence with a smaller AOS is placed a short distance downslope of the first fence to capture
silt that passed through the first fence.
Wyant (1980) and Allen (1994) have found that the geotextile AOS and permittivity
properties are not directly related to silt fence performance. Their experience indicates that,
in general, most geotextiles have hydraulic characteristics that provide acceptable silt fence
performance for even the most erodible silts. Thus, geotextile selection and specification can
be based on typical properties of silt fence geotextiles known to have performed satisfactorily
in the past.
Because silt fences are only used for temporary erosion control, the fence only needs to work
until the site can be revegetated or otherwise protected from rainfall and erosion. According
FHWA NHI-07-092
Geosynthetics Engineering

4-3

Temporary Runoff and Sediment Control


August 2008

to Richardson and Middlebrooks (1991), silt fences are best applied in situations where sheet
erosion occurs and where flow is not concentrated. Flow velocity should be less than about 1
ft/s (0.3 m/s). Slope grading or check dams may be required to reduce the flow.
Recommendations for allowable slope length versus slope angle to limit runoff velocity are
presented in Table 4-1. Furthermore, the limiting slope angle and velocity requirements
suggest that the drainage areas for overland flow to a fence should be less than about 5 acres
per 100 ft 1 (ha per 30 m) of fence.
Table 4-1
Limits of Slope Steepness and Length
to Limit Runoff Velocity to 1 ft/s (0.3 m/s)
(after Richardson and Middlebrooks, 1991)
Maximum Slope
Length

Slope Steepness
(%)

(ft)
100
80
50
30
15

<2
25
5 10
10 20
> 20

4.3

(m)
30
25
15
10
5

DESIGN OF SILT FENCES

There are two approaches to the design of silt fences. In the first, geotextile selection is
based on the hydraulic properties (i.e., AOS and permittivity) of the geotextile based on
previous experience with successful silt fence installations.
The second design approach is to conduct a performance test that basically is a model test of
a silt fence. The recommended test is ASTM D 5141 (Determining Filtering Efficiency and
Flow Rate of a Geotextile for Silt Fence Applications Using Site-Specific Soil). This test uses
representative samples of soils from the site and candidate geotextiles.
For each approach, designers need to estimate both the runoff volume due to a certain rainfall
intensity and the volume of sediment likely to be generated from the site. The silt fence must
be able to retain both these volumes. Finally, the physical and mechanical properties of the
geotextile must also be appropriately considered.

FHWA NHI-07-092
Geosynthetics Engineering

4-4

Temporary Runoff and Sediment Control


August 2008

4.3-1 Estimates of Runoff and Sediment Volumes


STEP 1.

Estimate runoff volume.

Use any reasonable approach to estimate the runoff volume. The Rational Method for
small watershed areas is one approach:
Q = 2.8 x 10-3 C i A

[4 - 1]

where:
Q
C
i
A

=
=
=
=

runoff (m3/s)
surface runoff coefficient
rainfall intensity (mm/hr)
area (ha)

Use C = 0.2 for rough surfaces, and C = 0.6 for smooth surfaces. A 10-year storm
event is typically used for designing silt fences.
Use the appropriate rainfall intensity factor, i, for the locality. Assume a 10-year
design storm, or use local design regulations. Neglect any concentration times (worst
case). This calculation gives the total storage volume required of the silt fence.

STEP 2.

Estimate sediment volume.

Probably the best way to estimate the volume of sediment generated is to use some
version of the Universal Soil Loss Equation (USLE). This was done for silt fences by
Richardson and Middlebrooks (1991). In the meantime, the USLE procedure has
been considerably revised and updated by the National Resource Conservation
Service (NRCSformerly the Soil Conservation Service) and the Agricultural
Research Service of the USDA. Its current version is called the Revised Universal
Soil Loss Equation, Version 2 (RUSLE2), and put on line. The official NRCS version
of RUSLE2 can be found at
http://fargo.nserl.purdue.edu/rusle2_dataweb/RUSLE2_Index.htm
For details, definitions, documentation and publications, tutorials, etc., go to
http://www.ars.usda.gov/Research/docs.htm?docid=6010
The RUSLE2 predicts an erosion rate in terms of tons of sediment produced per acre
(hectare) for a typical design period. This should provide a reasonable estimate for
FHWA NHI-07-092
Geosynthetics Engineering

4-5

Temporary Runoff and Sediment Control


August 2008

sizing the storage volume behind the silt fence. As the sediments have a very high
water content and a low density, assume a density of about 50 lb/ft3 (800 kg/m3) to
convert the soil loss in metric tons to a volume. Sediment behind a silt fence should
be removed when accumulation reaches approximately one-third to one-half fence
height, and this height should be used for estimating the storage volume required.
The next step is the hydraulic design of the geotextile, and as mentioned above, there are two
approaches to this design. The first is based on the hydraulic properties (i.e., AOS and
permittivity) of geotextiles known to perform satisfactorily. The second approach is to
conduct a performance test to estimate filtration efficiency for specific site conditions. The
next section describes both approaches.
4.3-2 STEP 3 Hydraulic Design of the Geotextile
1.

Nominal AOS and Permittivity Values


Because site specific designs for retention and permittivity are not necessary
for most soils (at least in a practical sense), nominal AOS and permittivity
values for geotextiles known to perform satisfactorily as silt fences may be
used. Suggested values (Richardson and Middlebrooks, 1991) are:
0.15 mm < AOS < 0.60 mm for woven slit films
0.15 mm < AOS < 0.30 mm for all other geotextiles
Permittivity, R > 0.02 sec-1

2.

Using Performance Test ASTM D 5141


This test was developed by Wyant (1980) for the Virginia Highway and
Transportation Research Council (VHTRC) and is based on observed field
performance and laboratory testing. See ASTM D 5141 for details. The
laboratory test is, in effect, a model test that consists of a flume with an
outflow opening similar to the size of a hay bale and fixed at a slope of 8%.
The geotextile is strapped across the end of the flume. A representative soil
sample from the site is then suspended in water to a concentration of about
3000 ppm (equivalent water content is 0.3 percent) and poured into the flume.
Based on their performance, appropriate geotextiles can be selected to provide
filtering efficiencies of 75% or more and flow rates on the order of 0.1
l/min/m2 after three test repetitions representing three storm events.

FHWA NHI-07-092
Geosynthetics Engineering

4-6

Temporary Runoff and Sediment Control


August 2008

The model study approach evaluates the system by utilizing actual soils from
the local area of interest. Thus, it cannot be performed by manufacturers. The
approach lends itself to an approved list-type specification for silt fences. In
this case, the agency or its representatives perform the flume test using their
particular problem soils and prequalifies the geotextiles that meet filtering
efficiency and flow criteria requirements. Qualifying geotextiles can be
placed on an approved list that is then provided to contractors. Geotextiles on
any approved list should be periodically retested because manufacturing
changes often occur.
4.3-3 STEP 4 Physical and Mechanical Properties; Constructability Requirements
The geotextile must be strong enough to support the pooled water and the sediments
collected behind the fence. Minimum strength depends on height of impoundment
and spacing between fence posts. A silt fence geotextile must be strong enough to
enable it to be properly installed.
Use Figure 4-1 to determine required tensile strength for a range of impoundment
heights and post spacings. For geotextiles without wire or plastic mesh backing, limit
impoundment heights to (2 ft) 0.6 m and post spacing to 6.6 ft (2 m); for greater
heights and spacings, use steel or plastic grid/mesh reinforcement to prevent burst
failure of geotextile. Unsupported geotextiles must not collapse or deform, allowing
silt-laden water to overtop the fence. Use Figure 4-2 to design the fence posts.

Figure 4-1.

Geotextile strength versus post spacing (Richardson and Koerner, 1990).


1 m = 3.3 ft.

FHWA NHI-07-092
Geosynthetics Engineering

4-7

Temporary Runoff and Sediment Control


August 2008

Figure 4-2.

Post requirements versus post spacing (Richardson and Koerner, 1990).


1 m = 3.3 ft.

AASHTO M288 property recommendations are indicated in Table 4-2. Realize that these
specifications are not based on research but on properties of geotextiles that have performed
satisfactorily as silt fences. Also given are requirements for resistance to ultraviolet
degradation. Although the applications are temporary (e.g., 6 to 36 months), the geotextile
must have sufficient UV resistance to function throughout its anticipated design life.

FHWA NHI-07-092
Geosynthetics Engineering

4-8

Temporary Runoff and Sediment Control


August 2008

Table 4-2
Physical Requirements1,2,3
for Temporary Silt Fence Geotextiles
(AASHTO, 2006)
ASTM
Test
Method
Maximum Post Spacing
Grab Strength
Machine Direction
X-Machine Direction
Permittivityc
Apparent Opening Size
Ultraviolet Stability
(Retained Strength)
a
b
c

Supported
Silt Fencea

Units

1.2 m

D 4632

D 4491
D 4751

sec-1
mm

D 4355

Requirement
Unsupported Silt Fence
Geotextile
Geotextile
Elongation Elongation
< 50%b
$ 50%b
1.2 m
2m

400
400
0.05
0.60 max.
70% after 500 hours
of exposure

550
550
450
450
0.05
0.05
0.60 max.
0.60 max.
70% after 500 hours
of exposure

Silt fence support shall consist of 14-gage steel wire mesh spacing of 150 mm by 150 mm or prefabricated
polymeric mesh of equivalent strength.
As measured in accordance with ASTM D 4632.
These default filtration property values are based on empirical evidence with a variety of sediments. For
environmentally sensitive areas, a review of previous experience and/or site or regionally specific
geotextile tests should be performed by the agency to confirm suitability of these requirements.

NOTES:
1. Acceptance of geotextile material shall be based on ASTM D 4759.
2. Acceptance shall be based upon testing of either conformance samples obtained using Procedure A of
ASTM D 4354, or based on manufacturers certifications and testing of quality assurance samples obtained
using Procedure B of ASTM D 4354.
3. All numeric values except AOS represent minimum average roll value (i.e., test results from any sampled
roll in a lot shall meet or exceed the minimum values in the table). Lot samples according to ASTM D
4354.
Conversion:

4.4

1 m = 3.3 ft.

1 N = 0.225 lbf

25.4 mm = 1 in.

SPECIFICATIONS

The following Washington State Department of Transportation specification is included


herein for your reference. It is meant to serve as an example for specifying geotextile silt
fence for routine (less critical) projects. It is not intended to replace site-specific evaluation,
testing, and design. Note that today the WSDOT has included specifications for silt fences
along with other geosynthetics materials in their Standard Specifications for Road, Bridge,
and Municipal Construction, N 41-06, 2006. For convenience, we will keep this sample
specification as a separate entity to indicate all the individual items that should be a silt fence
specification.

FHWA NHI-07-092
Geosynthetics Engineering

4-9

Temporary Runoff and Sediment Control


August 2008

CONSTRUCTION GEOTEXTILE FOR SILT FENCE


(after Washington State Department Of Transportation, 2006)
Materials
Geotextiles shall consist only of long chain polymeric fibers or yarns formed into a stable
network such that the fibers or yarns retain their position relative to each other during
handling, placement, and design service life. At least 95 percent by weight of the material
shall be polyolefins or polyesters. The material shall be free from defects or tears. The
geotextile shall also be free of any treatment or coating which might adversely alter its
hydraulic or physical properties after installation.
The geotextile shall conform to the properties as indicated in Table A, and additional tables
as required in the Standard Plans and Special Provisions for each use specified in the Plans.
Thread used for sewing geotextiles shall consist of high strength polypropylene, polyester, or
polyamide. Nylon threads will not be allowed. The thread used to sew permanent erosion
control geotextiles shall also be resistant to ultraviolet radiation. The thread shall be of
contrasting color to that of the geotextile itself.
Table A: Geotextile Property Requirements1
for Temporary Silt Fence
ASTM
Test
Method

Geotextile Property

Geotextile Property Requirements1


Supported Between Posts with
Unsupported Between Posts
Wire or Polymeric Mesh
#30 US sieve (0.60 mm) max.
#30 US sieve (0.60 mm) max.
for slit film wovens
for slit film wovens
#50 US sieve (0.30 mm) max.
#50 US sieve (0.30 mm) max.
for all other geotextile types
for all other geotextile types
#100 US sieve (0.15 mm) min.
#100 US sieve (0.15 mm) min.

AOS

D 4751

Water Permittivity

D 4491

0.02 sec-1 min.

0.02 sec-1 min.

Grab Tensile
Strength, in machine
and x-machine
direction

D 4632

180 lbs. min. in machine direction,


100 lbs. min. in x-machine direction

100 lbs. min.

Grab Failure Strain, in


machine and xmachine direction

D 4632

30% max. at 180 lbs or more

---

Ultraviolet (UV)
Radiation Stability

D 4355

70% Strength Retained min., after


500 hr in weatherometer

70% Strength Retained min., after


500 hr in weatherometer

NOTE:
1. All geotextile properties in Table A are minimum average roll values (i.e., the test result for any sampled roll
in a lot shall meet or exceed the values shown in the table).

FHWA NHI-07-092
Geosynthetics Engineering

4-10

Temporary Runoff and Sediment Control


August 2008

Geotextile Approval and Acceptance


Source Approval
The Contractor shall submit to the Engineer the following information regarding each geotextile
proposed for use:
Manufacturer's name and current address,
Full product name,
Geotextile structure, including fiber/yarn type,
Geosynthetic polymer type(s), and
Proposed geotextile use(s).
If the geotextile source has not been previously evaluated, or is not listed in the current WSDOT
Qualified Products List (QPL), a sample of each proposed geotextile shall be submitted to the State
Materials Laboratory in Tumwater for evaluation. After the sample and required information for each
geotextile type have arrived at the State Materials Laboratory in Tumwater, a maximum of 14
calendar days will be required for this testing. Source approval will be based on conformance to the
applicable values from Table A and additional tables as specified in the Special Provisions. Source
approval shall not be the basis of acceptance of specific lots of material unless the lot sampled can be
clearly identified and the number of samples tested and approved meet the requirements of WSDOT
Test Method 914.
Acceptance Samples
Samples will be randomly taken by the Engineer at the job site to confirm that the geotextile meets
the property values specified.
Approval will be based on testing of samples from each lot. A "lot" shall be defined for the purposes
of this specification as all geotextile rolls within the consignment (i.e., all rolls sent to the project site)
that were produced by the same manufacturer during a continuous period of production at the same
manufacturing plant and have the same product name. After the samples have arrived at the State
Materials Laboratory in Tumwater, a maximum of 14 calendar days will be required for this testing.
If the results of the testing show that a geotextile lot, as defined, does not meet the properties required
for the specified use as indicated in Table A and additional tables as specified in the Special
Provisions, the roll or rolls which were sampled will be rejected. Two additional rolls for each roll
tested which failed from the lot previously tested will then be selected at random by the Engineer for
sampling and retesting. If the retesting shows that any of the additional rolls tested do not meet the
required properties, the entire lot will be rejected. If the test results from all the rolls retested meet the
required properties, the entire lot minus the roll(s) that failed will be accepted. All geotextile that has
defects, deterioration, or damage, as determined by the Engineer, will also be rejected. All rejected
geotextile shall be replaced at no cost to the Contracting Agency.
Acceptance by Certificate of Compliance
When the quantities of geotextile proposed for use in each geotextile application are less than or equal
to the following amounts, acceptance shall be by Manufacturers Certificate of Compliance:
Application: Temporary Silt Fence
Geotextile Quantities: All quantities
FHWA NHI-07-092
Geosynthetics Engineering

4-11

Temporary Runoff and Sediment Control


August 2008

The manufacturer's certificate of compliance shall include the following information about each
geotextile roll to be used:
Manufacturer's name and current address,
Full product name,
Geotextile structure, including fiber/yarn type,
Polymer type,
Geotextile roll number,
Proposed geotextile use(s), and
Certified test results.
Approval of Seams
If the geotextile seams are to be sewn in the field, the Contractor shall provide a section of sewn seam
that can be sampled by the Engineer before the geotextile is installed.
The seam sewn for sampling shall be sewn using the same equipment and procedures as will be used
to sew the production seams. If production seams will be sewn in both the machine and crossmachine directions, the Contractor must provide sewn seams for sampling which are oriented in both
the machine and cross-machine directions. The seams sewn for sampling must be at least 2 yards in
length in each geotextile direction. If the seams are sewn in the factory, the Engineer will obtain
samples of the factory seam at random from any of the rolls to be used. The seam assembly
description shall be submitted by the Contractor to the Engineer and will be included with the seam
sample obtained for testing. This description shall include the seam type, stitch type, sewing thread
type(s), and stitch density.
Construction Geotextile (Installation Requirements)
Silt fence shall be installed in accordance with the Plans.
When backup support is used, steel wire shall have a maximum mesh spacing of 2-inches by 4inches. and the plastic mesh shall be as resistant to ultraviolet radiation as the geotextile it supports.
The geotextile shall be attached to the posts and support system using staples, wire, or in accordance
with the manufacturer's recommendations.
The geotextile shall be sewn together at the point of manufacture, or at a location approved by the
Engineer, to form geotextile lengths as required. All sewn seams and overlaps shall be located at a
support post.
Posts shall be either wood or steel. Wood posts shall have minimum dimensions of 1-inches by 1inches by the minimum length shown in the Plans. Steel posts shall have a minimum weight of 0.90
lbs/ft.
When sediment deposits reach approximately one-third the height of the silt fence, the deposits shall
be removed and stabilized.

FHWA NHI-07-092
Geosynthetics Engineering

4-12

Temporary Runoff and Sediment Control


August 2008

4.5

INSTALLATION PROCEDURES

Silt fences are quite simple to construct; the normal construction sequence is shown in Figure
4-3. Installation of a prefabricated silt fence is shown is Figure 4-4.
1.

2.
3.
4.

5.

Install wooden or steel fence posts or large wooden stakes in a row, with
normal spacing between 1.5 to 10 ft (0.5 to 3 m), center to center, and to a
depth of 15 to 18 in. (0.4 to 0.6 m). Most pre-fabricated fences have posts
spaced approximately 6 to 10 ft (2 to 3 m) apart, which is usually adequate
(Step 1).
Construct a small (minimum 6 in. (150 mm) deep and 4 in. (100 mm) wide)
trench on the upstream side of the silt fence (Step 2).
Attach reinforcing wire, if required, to the posts (Step 3).
If a prefabricated silt fence is not being used, the geotextile must be attached
to the posts using staples, reinforcing wire, or other attachments provided by
the manufacturer. Geotextile should be extended at least 6 in. (150 mm)
below the ground surface (Steps 4 and 5).
Bury the lower end of the geotextile in the upstream trench and backfill with
natural material, tamping the backfill to provide good anchorage (Step 6).

Silt fence ends should be turned uphill to ensure they capture runoff water and prevent flow
around the ends. The groundline at the fence ends should be at or above the elevation of the
lowest portion of the fence top. Measures should be taken to prevent erosion along the fence
backs that run downhill for a significant distance. Gravel check dams at approximately 6 to
10 ft. (2 to 3 m) intervals along the back of the fence can be used.
The field inspector should review the field inspection guidelines in Section 1.7.

4.6

INSPECTION AND MAINTENANCE

Silt fences should be checked periodically, especially after a rainfall or storm event.
Excessive buildup of sediment must be removed so the silt fence can function properly.
Generally, sediment buildup behind the fence should be removed when it reaches to of
the fence height. Repair or replace any split, torn, slumping or weathered geotextile. The toe
trench should also be checked to ensure that runoff is not piping under the fence.

FHWA NHI-07-092
Geosynthetics Engineering

4-13

Temporary Runoff and Sediment Control


August 2008

Figure 4-3.

Typical silt fence installation.

FHWA NHI-07-092
Geosynthetics Engineering

4-14

Temporary Runoff and Sediment Control


August 2008

Figure 4-4.

4.7

Installation of a prefabricated silt fence . 1 m = 3.3 ft.

SILT AND TURBIDITY CURTAINS

At construction sites adjacent to open water, silt and turbidity curtains may be necessary to
prevent sediment from entering the lake, stream, or river. The curtain is suspended from
floats at the waters surface and weighted down at the bottom, and then stretched across the
stream or in a semicircle around the end of the site. A tension cable is often built into the
curtain immediately above or below the flotation segments to absorb stress imposed by
currents, wave action, and wind. Often the curtains are fabricated on-shore, then deployed
from boats or barges, and in deep water, divers may be necessary to attach the bottom of the
curtain to concrete blocks or other weights. In shallow water, curtains can be constructed
similar to on-shore silt fences. Geotextiles have also been attached to soldier piles and
draped across riprap barriers as semi-permanent curtains, and in side sheet-pile cofferdams to
retain and fines in the backfill.

FHWA NHI-07-092
Geosynthetics Engineering

4-15

Temporary Runoff and Sediment Control


August 2008

The U.S. Army Corps of Engineers (1977) indicates that silt and turbidity curtains are not
appropriate for:

operations in open ocean;

operations in currents exceeding 0.5 m/s (1.6 ft/s);

in areas frequently exposed to high winds and large breaking waves;

use near hopper or cutter head dredges where frequent curtain movement
would be necessary.
Silt and turbidity curtains perform essentially the same function in water as silt fences do on
land; that is, they intercept sediment-laden water while allowing clear water to pass. Thus,
for maximum efficiency, a silt or turbidity curtain should pass a maximum amount of water
while retaining a maximum amount of sediment. Unfortunately, such optimum performance
is normally not possible because sediments will eventually blind or clog (Figure 2-3) the
geotextile.
To maximize the geotextile's efficiency, the following soil, site, and
environmental conditions should be established, and the geotextile selected should provide a
specific filtering efficiency while maintaining the required flow rate (Bell and Hicks, 1980).
1.
Determine the grain size distribution of soil to be filtered.
2.
Estimate the soil volume to be filtered during construction.
3.
Estimate the flow conditions, anticipated runoff, and water level fluctuations.
4.
Consider the environmental conditions, temperature, and duration of sunlight
exposure.
5.
Determine the velocity, direction, and quantity of discharge water.
6.
Determine water depth, levels of turbidity, bottom sediments, and vegetation
at the site.
7.
Determine wind conditions.
On the basis of these considerations, the geotextile can then be selected either according to
the properties required to maximize particle retention and flow capacity while resisting
clogging, or by performing filtration model studies such as ASTM D 5141. The first
approach follows the criteria developed in Chapter 2 for drainage systems. Silt and turbidity
curtains are generally concerned with fine-grained soils; therefore, the following criteria
could be considered when selecting the geotextile.
A.

Retention Criteria
AOS
AOS

=
=

FHWA NHI-07-092
Geosynthetics Engineering

D85 for woven geotextiles.


1.8 D85 for nonwovens.

4-16

Temporary Runoff and Sediment Control


August 2008

NOTE: The D85 is a characteristic large-grain size appropriate to the suspended


sediment grain size distribution. It will be strongly influenced by items Nos.
1, 3, 5, and 6 above.
B.

Flow Capacity Criterion


= (10 q ) A (per 1 foot unit head)
where:

q
A
10

C.

=
=
=
=

permittivity of geotextile (sec-1)


flow rate (ft3/sec)
cross-sectional area silt curtain (ft2)
factor of safety

Clogging Resistance
Maximize AOS requirements using largest opening possible from criterion A
above.

Both woven and nonwoven geotextiles have been used for silt curtains, but experience has
shown that slit-film and monofilament wovens, because of their slick surface texture, tend to
be easier to maintain. They are also less susceptible to possible biological clogging.
For silt and turbidity curtain construction, the geotextile forming the curtain is held vertical
by flotation segments at the top and ballast along the bottom (Bell and Hicks, 1980). A
tension cable is often built into the curtain immediately above or below the flotation
segments to absorb stress imposed by currents, wave action, and wind. Barrier sections are
usually about 100 ft (30 m) long, with the width determined by the water depth. The sections
are seamed appropriately for the full length of the curtain. They can be made to any
reasonable width, depending on the water depth. Curtains can also be constructed within
shallow bodies of water using silt fence-type construction methods. Geotextiles have also
been attached to soldier piles and draped across riprap barriers as semi-permanent curtains.
Fabrication and installation of silt and turbidity curtains in deeper water or at complicated
sites are often done by specialized contractors.

FHWA NHI-07-092
Geosynthetics Engineering

4-17

Temporary Runoff and Sediment Control


August 2008

4.8

EROSION CONTROL BLANKETS

In freshly graded areas that are not to be immediately paved or covered, the soils are
susceptible to erosion by rainfall and runoff. Hydraulic seeding and mulching techniques as
well as biodegradable erosion control blankets and other similar products can be used to
enhance the establishment of vegetation after construction is completed. Erosion protection
must be provided for three distinct phases, namely:
1. prior to vegetation growth
2. during vegetation growth
3. after vegetation is fully established
Erosion control blankets provide protection during the first two phases. After vegetation is
established, and if further protection is necessary, TRMs that reinforce the vegetation root
mass are necessary. Permanent erosion control with TRMs was discussed in Sec. 3.11.
Erosion control blankets (ECBs) belong to a class of geosynthetics called rolled erosion
control products (RECPs), which are three-dimensional porous geosynthetic mats that
reinforce the roots and promote vegetation growth. These products together with vegetation
form an attractive and environmentally friendly biocomposite. Biodegradable RECPs are
used to enhance the establishment of vegetation for temporary erosion control applications
and where vegetation alone provides sufficient erosion protection after the temporary
products degrade. Also included in these biodegradable geosynthetic are mulch control nets
(MCNs) and open weave textiles (OWTs).
MCNs are manufactured of light-weight polymer net(s), typically polypropylene, and a
bedding or matrix of degradable organic materials such as aspen excelsior, straw, or coir
(coconut fibers). The matrix material protects the soil against erosion and helps retain
moisture, seeds, and soil to promote vegetation growth. These polymer materials are
typically not stabilized against ultraviolet light, and are designed to degrade over time.
Erosion control blankets have design lives that vary between approximately 6 months to 5
years. Some blankets are provided with grass seeds encased in paper, and some
biodegradable products may provide additional nutrients to the soil.
Erosion control blankets provide protection against moderate-flow velocities for short
periods of time. They are typically used on moderate slopes and low velocity intermittent
flow channels. Flows up to 5 ft/s (1.5 m/s) and durations of 0.5 to approximately 5 hr can be
withstood, as illustrated in Figure 4-5.

FHWA NHI-07-092
Geosynthetics Engineering

4-18

Temporary Runoff and Sediment Control


August 2008

Figure 4-5. Recommended maximum design velocities and flow durations for various
classes of erosion control materials (after Hewlett et al., 1987 and Theisen, 1992).

ECBs are usually evaluated by field trial sections, although there are very few published
records of comparative use. Therefore selection is usually based on local experience. Users
should be aware that a variety of products and systems exist, and as an aid to selecting the
best system, consult manufacturers and other agencies about their experiences. For example,
some ECB manufacturers have actual flume test data and design recommendations for their
specific products. One of the best sources of performance information is the Texas
Department of Transportation as performed by the Texas Transportation Institute Hydraulics,
Sediment & Erosion Control Laboratory.
This agency tests candidate erosion control
materials and categorizes them into classes and types in an approved materials list.
Information on the test program and the results of tests on specific products are available on
their web site: http://www.dot.state.tx.us/services/maintenance/erosion_control.htm. The
Erosion Control Technology Council (ECTC) is also a good source of information on ECBs.

FHWA NHI-07-092
Geosynthetics Engineering

4-19

Temporary Runoff and Sediment Control


August 2008

Since the design of erosion control blankets is empirical, specification by index properties is
not easily accomplished. Also, most current ASTM standards for RECPs are for TRMs, and
relatively few test methods have been standardized for erosion control blankets, although the
ECTC has developed some suggested test procedures that can be used for ECBs. Therefore,
it is recommended that specifications use an approved products list.
Construction plans and specifications should detail and note installation requirements.
Details such as anchoring in trenches, use of pins, pin length, pin spacing, roll overlap
requirements, and roll termination should be addressed.
Specifications for RECPs and TRMs can be found on the Texas DOT website or the Erosion
Control Technology Councils website: www.ectc.org/specifications.asp

4.9

REFERENCES

AASHTO (2006). Standard Specifications for Geotextiles - M 288, Standard Specifications


for Transportation Materials and Methods of Sampling and Testing, 25th Edition,
American Association of State Transportation and Highway Officials, Washington,
D.C.
Allen, T. (1994). Personal Communication.
ASTM (2006). Annual Books of ASTM Standards, American Society for Testing and
Materials, West Conshohocken, PA:
Volume 4.08 (I), Soil and Rock
Volume 4.09 (II), Soil and Rock; Geosynthetics
ASTM Test Methods - see Appendix E.
Bell, J.R. and Hicks, R.G. (1980). Evaluation of Test Methods and Use Criteria for
Geotechnical Fabrics in Highway Applications - Interim Report, Report No.
FHWA/RD-80/021, June, 190 p.
Hewlett, H.W.M., Boorman, L.A. and Bramley, M.E. (1987). Design of Reinforced Grass
Waterways Report 116, Construction Industry Research and Information
Association, London, U.K., 116 p.
Mallard, P. and Bell, J.R. (1981). Use of Fabrics in Erosion Control, Transportation
Research Report 81-4, FHWA, January.
Richardson, G.N. and Middlebrooks, P. (1991). A Simplified Design Method for Silt Fences,
Proceedings of the Geosynthetics '91 Conference, Industrial Fabrics Association
International, St. Paul, MN, pp. 879-888.
FHWA NHI-07-092
Geosynthetics Engineering

4-20

Temporary Runoff and Sediment Control


August 2008

Richardson, G.N. and Koerner, R.M., Editors (1990). A Design Primer: Geotextiles and
Related Materials, Industrial Fabrics Association International, St. Paul, MN, 166 p.
Theisen, M.S. (1992). Geosynthetics in Erosion Control and Sediment Control, Geotechnical
Fabrics Report, Industrial Fabrics Association International, St. Paul, MN, May/June
pp. 26-35.
U.S. Army Corps of Engineers (1977). Civil Works Construction Guide Specification for
Plastic Filter Fabric, Corps of Engineers Specifications No. CW-02215, Office,
Chief of Engineers, U.S. Army Corps of Engineers, Washington, D.C.
Washington State Department Of Transportation (2006). Construction Geotextile For Silt
Fence, Standard Specifications for Road, Bridge, and Municipal Construction, N 4106.
Wyant, D.C. (1980). Evaluation of Filter Fabrics for Use as Silt Fences, Report No.
VHTRC 80-R49, Virginia Highway and Transportation Research Council,
Charlottesville, VA.

FHWA NHI-07-092
Geosynthetics Engineering

4-21

Temporary Runoff and Sediment Control


August 2008

[ BLANK ]

FHWA NHI-07-092
Geosynthetics Engineering

4-22

Temporary Runoff and Sediment Control


August 2008

5.0 GEOSYNTHETICS IN ROADWAYS AND PAVEMENTS

5.1

INTRODUCTION

Geosynthetics have been found to provide significant improvement in pavement construction


and performance. Figure 1 illustrates a number of potential geosynthetic applications in a
layered pavement system to improve its performance. The most common of all uses of
geosynthetics is in road and pavement construction. Geotextiles placed at the subgrade
increase stability and improve performance of pavement constructed on high fines subgrade
soils (i.e., soils containing high quantities of silt and/or clay factions) during construction,
primarily by separating the aggregate from the subgrade. In addition, geogrids and some
geotextiles can provide strength through friction or interlock developed between the
aggregate and the geosynthetic. Geotextiles can also provide filtration and drainage by
allowing excess pore water pressures in wet subgrade soils to dissipate into the aggregate
base course and, in cases of poor-quality aggregate, through the geotextile plane itself. The
separation, filtration and reinforcement functions combine to provide a method of mechanical
stabilization for weak subgrade soils.
Geosynthetics can also provide long-term benefits and improve the performance of the road
over its design life. Geotextiles used as separators over high fines subgrades prevent the
migration of fines into base/subbase materials, especially into open graded bases,
maintaining the support and drainage characteristics of the base over the life of a pavement
system. As discussed in Chapter 2, geotextile filters and geocomposite drains are key
components in pavement drainage systems. Adequate drainage can directly provide
improvement in long-term flexible and rigid pavement performance (AASHTO, 1986;
NCHRP 137-A, 2002; NHI 131026 Participants Notebook {ERES, 1999}; and, FHWA
NHI-05-037 {Christopher et al., 2006}). Geogrids and geotextiles can also be used to
reinforce the base course of flexible pavement to improve its serviceability.
In this chapter, each of the geosynthetic functions will be discussed and related to design
concepts and performance properties. The advantages, limitations and potential cost benefits
of using geosynthetics in each of these application areas will be reviewed. Selection,
specification, and construction procedures will also be presented. As shown in Figure 1,
there is also a potential to improve the pavement performance by incorporating geosynthetics
into the design of surficial asphalt layers. These applications will be covered in Chapter 6.

FHWA NHI-07-092
Geosynthetics Engineering

5-1

Roadways and Pavements


August 2008

Figure 5-1.

Potential applications of geosynthetics in a layered pavement system

5.2 APPLICABILTY AND BENEFITS OF GEOSYNTHETICS IN ROADWAYS


Roads and highways are broadly classified into two categories: permanent and temporary,
depending on their service life, traffic applications, or desired performance. Permanent roads
include both paved and unpaved systems which usually remain in service 10 years or more.
Permanent roads may be subjected to more than a million load applications during their
design lives. On the other hand, temporary roads are, in most cases, unpaved. They remain
in service for only short periods of time (often less than 1 year), and are usually subjected to
fewer than 10,000 load applications during their services lives. Temporary roads include
detours, haul and access roads, construction platforms, and stabilized working tables required
for the construction of permanent roads, as well as embankments over soft foundations.
5.2-1 Temporary Roads and Working Platforms
Geosynthetics are used in temporary roads to reduce rutting of the gravel surface and/or to
decrease the amount of gravel required to support the anticipated traffic. Furthermore, the
geosynthetic helps to maintain the aggregate thickness over the life of the temporary road.
Where the soils are normally too weak to support the initial construction work, geosynthetics
in combination with gravel provide a working platform to allow construction equipment
access to sites. This is one of the more important uses of geosynthetics. Even if the finished
roadway can be supported by the subgrade, it may be virtually impossible to begin
construction of the embankment or roadway. Such sites require stabilization by dewatering,
demucking, excavation and replacement with select granular materials, utilization of
stabilization aggregate, chemical stabilization, etc. Geosynthetics can often be a costeffective alternate to these expensive foundation treatment procedures.

FHWA NHI-07-092
Geosynthetics Engineering

5-2

Roadways and Pavements


August 2008

5.2-2 Permanent Paved and Unpaved Roads


For permanent road construction, the temporary working platform described in the previous
section also may be used to provide an improved roadbed. This mechanically stabilized layer
enables contractors to meet minimum compaction specifications for the first two or three
aggregate lifts. This is especially true on very soft, wet subgrades, where the use of ordinary
compaction equipment is very difficult or even impossible. Long term, a geosynthetic acts to
maintain the roadway design section and the base course material integrity. Thus, the
geosynthetic will ultimately increase the life of the roadway.
As previously indicated, geotextiles can be placed beneath free draining (i.e., open graded)
base as a separator/filter to maintain drainage and improve pavement performance over the
life of both paved and unpaved roadways. The use of free draining base is encouraged for
optimum performance of pavement systems. However, where free draining base is not used,
not readily available and/or relatively expensive, geosynthetics with lateral drainage
capability (transmission function) may be placed at the subgrade-base interface to: (i)
provide drainage of the subgrade when a poorly draining, dense graded base course is used;
and/or (ii) to promote quicker drainage of the base course. For the first case of poorly
draining bases, where water in a wet subgrade cannot readily drain upward into the base,
geosynthetics that allow some drainage in its plane (i.e., thick nonwoven geotextiles or
geocomposites) would also allow for pore water pressure dissipation. For the second case,
drainage geocomposites with sufficient compressive strength to support traffic without
excessive deformation and with adequate flow capacity can be used to enhance drainage of
the roadway system. Improved drainage applications are reviewed in the permanent roadway
application section of this chapter (Section 5.6-4).
Another geosynthetic application in roadways is to place geosynthetics (primarily geogrids
and geocomposites) at the bottom of or within the base course to provide reinforcement
through lateral confinement of the aggregate layer. Lateral confinement arises from the
development of interface shear stresses between the aggregate and the reinforcement and
occurs during placement, compaction, and traffic loading. A small residual restraint remains
after each load application, thus increasing the lateral confinement of the aggregate with
increasing load applications. Base reinforcement thus improves the long-term structural
support for the base materials and reduces permanent deformation in the roadway section and
has been found under certain conditions to provide significant improvement in pavement
performance. Increases in traffic volume up to a factor of 10 to reach the same distress level
(1 in. {25-mm} rutting) have been observed for reinforced sections versus unreinforced
sections of the same design asphalt and base thickness (Berg et al., 2000). This application is
reviewed later in the permanent roadway application section of this chapter (Section 5.7).

FHWA NHI-07-092
Geosynthetics Engineering

5-3

Roadways and Pavements


August 2008

5.2-3 Subgrade Conditions in which Geosynthetics are Useful


Geosynthetics have a 30+ year history of successful use for the stabilization of very soft wet
subgrades. Based on experience and several case histories summarized by Haliburton,
Lawmaster, and McGuffey (1981) and Christopher and Holtz (1985), the following subgrade
conditions are considered optimum for using geosynthetics in roadway construction:
Poor soils
(USCS: SC, CL, CH, ML, MH, OL, OH, and PT)
(AASHTO: A-5, A-6, A-7-5, and A-7-6)
Low undrained shear strength

f = cu < 2000 psf (90 kPa)

CBR < 3
(Note: Soaked Saturated CBR as determined with ASTM D 4429)
R-value (California) < 20
MR < 4500 psi (30 MPa)
High water table
High sensitivity

Under these conditions, multiple functions are possible. Geosynthetics function as


separators to prevent intermixing of roadway aggregate and the subgrade. Filtration is
required because soils below a CBR of 3 are typically wet and saturated. This water must be
allowed to pass up through the geosynthetic into the aggregate, such that destabilizing pore
pressure in the subgrade generated from wheel loads can rapidly dissipate. Pore pressure
dissipation will also allow for strength gains in the subgrade over time. Some level of
reinforcement may also be provided through lateral restraint of the roadway aggregate placed
directly above the geosynthetic, which in turn reduces the stresses on the subgrade and
improves bearing capacity. If large ruts develop during placement of the first aggregate lift,
then some membrane reinforcing effect is also present.
As the geosynthetic provides multiple functions, which both benefit construction and allow
for subgrade improvement with time, AASHTO M288 has identified applications where the
undrained shear strength is less than about 2000 psf (90 kPa) (CBR about 3) as a form of
mechanical stabilization. From a foundation engineering point of view, clay soils with
undrained shear strengths of 2000 psf (90 kPa), or higher, are considered to be stiff clays
(Terzaghi and Peck, 1967) and are generally quite good foundation materials. Allowable
footing pressures on such soils can be around 3000 psf (150 kPa) or greater. Simple stress
distribution calculations show that for static loads, such soils will readily support reasonable
truckloads and tire pressures, even under relatively thin granular bases.

FHWA NHI-07-092
Geosynthetics Engineering

5-4

Roadways and Pavements


August 2008

Dynamic loads and high tire pressures are another matter. Some rutting will probably occur
in such soils, especially after a few hundred passes (Webster, 1993). If traffic is limited, as it
is in many temporary roads, or if shallow (< 3 in. {75 mm}) ruts are acceptable, as in most
construction operations, a maximum undrained shear strength of approximately 2000 psf (90
kPa) (CBR = 3) for geosynthetic use in highway construction seems reasonable. However,
for soils that are seasonally weak (e.g., from frost heave) or for high fines content soils which
are susceptible to pumping, a geotextile separator may be of benefit in preventing migration
of fines at a much higher subgrade undrained shear strength. This is especially the case for
permeable base applications. Significant fines migration has been observed with a subgrade
CBR as high as 8 (e.g., Al-Qadi et al., 1998). On firm subgrades, a geotextile placed beneath
the base functions as a separator and filter, as illustrated in Figure 5-2. A greater range of
geotextile applicability is recognized in the M288 specification (AASHTO, 2006) with a
CBR 3 the geotextile application is identified as separation. The complete M288
specification is presented in Appendix D.
The subgrade conditions for separation applications apply equally to filtration and drainage
applications. Soils with high fines are poorly draining and the strength of such soils is highly
influenced by moisture. Base reinforcement has also been found under certain conditions to
provide significant improvement in performance of pavement sections over a wide range of
subgrade condition (up to a CBR of 8 or greater) (Berg et al, 2000).
As a summary, the application areas and functions in Table 5-1 have been identified as
appropriate for the corresponding subgrade conditions.

Figure 5-2.

Geotextile separator beneath permeable base (Baumgartner, 1994).

FHWA NHI-07-092
Geosynthetics Engineering

5-5

Roadways and Pavements


August 2008

Table 5-1
Application and Associated Functions of Geosynthetics in Roadway Systems
Application
Separator

Function(s)

Subgrade Strength

Separation

2000 psf cu 5000 psf

Secondary: filtration*

(90 kPa cu 240 kPa)


3 CBR 8

Qualifier
Soils containing high
fines (SC, CL, CH, ML,
MH, SM, SC, GM,GC)

4500 psi MR 11,600 psi


(30 MPa MR 80 MPa)
Stabilization

Separation, filtration and


some reinforcement
(especially CBR <1)
Secondary: Transmission

Base
Reinforcement

cu < 2000 psf (90 kPa)


CBR < 3
MR < 4500 psi (30 MPa)

Reinforcement

600 psf cu 5000 psf

Secondary: separation

(30 kPa cu 240 kPa)


3 CBR 8

Wet, saturated fine


grained soils (i.e., silt,
clay and organic soils)

All subgrade conditions.


Reinforcement located
within 6 to 12 in. (150 to
300 mm) of pavement

1500 psi MR 11,600 psi


(10 MPa MR 80 MPa)
Drainage

Transmission and filtration

not applicable

Poorly draining subgrade

Secondary: separation
*always evaluate filtration requirements

5.2-4 Benefits
Geosynthetics used in both temporary and permanent roadways on soft subgrades, may
provide several cost and performance benefits, including the following.
1. Reducing the intensity of stress on the subgrade (function: reinforcement).
2. Preventing the base aggregate from penetrating into soft subgrades (function:
separation).
3. Preventing subgrade fines from pumping or otherwise migrating up into the base
(function: separation and filtration).
4. Preventing contamination of the base materials which may allow more open-graded,
free draining aggregates to be considered in the design (function: filtration).
5. Reducing the depth of excavation required for the removal of soft, unsuitable
subgrade materials (function: separation and reinforcement).
6. Reducing the thickness of aggregate required to stabilize soft subgrades (function:
separation and reinforcement).
FHWA NHI-07-092
Geosynthetics Engineering

5-6

Roadways and Pavements


August 2008

7. Reducing disturbance of soft or otherwise sensitive subgrade during construction


(function: separation and reinforcement).
8. Allowing an increase in soft subgrade strength over time (function: filtration).
9. Providing more uniform support by reducing the differential settlement of roadways
constructed over variable subgrade (i.e., soft to firm) conditions, which helps
maintain pavement integrity and uniformity (function: reinforcement). Geosynthetics
will also aid in reducing differential settlement in transition areas from cut to fill.
{NOTE: Consolidation settlements are not reduced by the use of geosynthetic
reinforcement.}
10. Reducing maintenance, extending the life of the pavement system, and maintaining
the long-term integrity of base/subbase layer(s) for pavement surface rehabilitation
projects (functions: all).
Geosynthetics may also provide cost and performance benefits when used in roadways with
firm, fairly competent subgrades (CBR ranging from 3 to 8), but containing a high quantity
of fines, which are often sensitive to seasonal environmental conditions, including the
following.
1. Maintaining the structural and drainage characteristics of free draining, more open
graded base and subbase layers by reducing fine grain soil intrusion from the
subgrade over the life of the road, (functions: separation and filtration).
2. Improving deformation response of base/subbase by providing lateral restraint
(function: reinforcement).
3. Extending the design life or increasing structural support through improved drainage
when used for (i.e., geocomposite drains) or as part of the roadway drainage systems
(functions: filtration and drainage).
4. Using a free draining stone layer sandwiched between geotextile filters or a
geocomposite to provide a capillary break to reduce frost action in frost-susceptible
soils (function: drainage).
5. To provide fully or partially membrane-encapsulated soil layers (MESL) to reduce
the effects of seasonal water content changes on roadways on swelling clays
(function: barrier).

FHWA NHI-07-092
Geosynthetics Engineering

5-7

Roadways and Pavements


August 2008

5.3 ROADWAY DESIGN USING GEOSYNTHETICS


Certain design principles are common to all types of roadways, regardless of the design
method or the type of geosynthetic (i.e., geotextile or geogrid). Basically, the design of any
roadway involves a study of each of the components of the system, (surface, aggregate base
courses and subgrade) detailing their behavior under traffic load and their ability to carry that
load under various climatic and environmental conditions. All roadway systems, whether
permanent or temporary, derive their support from the underlying subgrade soils. Thus,
when placed at the subgrade interface, the geosynthetic functions are similar for either
temporary or permanent roadway applications. However, due to different performance
requirements, design methodologies for temporary roads should not be used to design
permanent roads. Temporary roadway design usually allows some rutting to occur over the
design life, as ruts will not necessarily impair service. Obviously, ruts are not acceptable in
permanent roadways.
For temporary roads, our design basically uses geosynthetics for the construction and traffic
support of the roadway section allowing for a specific tolerable amount of rutting.
Recommended design procedures for temporary roads are presented in Sections 5.4 for
geotextiles and 5.5 for geogrids. For permanent road and pavement design, approaches for
using geotextiles for separation, stabilization, and improved drainage applications are
presented in Sections 5.6. Approaches for using geogrids in permanent roads for
stabilization and base reinforcement are covered in Section 5.7. Design for each application
is based on the function(s) of the geosynthetic and the properties required to perform the
intended functions as covered in the following sections.
5.3-1 Functions of Geosynthetics in Roadways and Pavements
A geosynthetic placed at the interface between the aggregate base course and the subgrade
functions as a separator to prevent two dissimilar materials (subgrade soils and aggregates)
from intermixing. Geotextiles perform this function by preventing penetration of the
aggregate into the subgrade (localized bearing failures) and prevent intrusion of subgrade
soils up into the base course aggregate (Figure 5-3). Geogrids can also prevent aggregate
penetration into the subgrade, depending on the ability of the geogrid to confine and prevent
lateral displacement of the base/sub-base. However, the geogrid does not prevent intrusion
of subgrade soils up into the base/sub-base course, which must have a gradation that is
compatible with the subgrade based on standard geotechnical graded granular filer criteria
when using geogrids alone. Localized bearing failures and subgrade intrusion occur in very
soft, wet, weak subgrades.

FHWA NHI-07-092
Geosynthetics Engineering

5-8

Roadways and Pavements


August 2008

Figure 5-3.

Concept of geotextile separation in roadways (after Rankilor, 1981).

Subgrade intrusion can also occur under long term dynamic loading due to pumping and
migration of fines, especially when open-graded base courses are used. It only takes a small
amount of fines to significantly affect the structural characteristics of select granular
aggregate (e.g., see Jornby and Hicks, 1986). Therefore, separation is important to maintain
the design thickness and the stability and load-carrying capacity of the base course. Soft
subgrade soils are most susceptible to disturbance during construction activities such as
clearing, grubbing, and initial aggregate placement. Geosynthetics can help minimize
subgrade disturbance and prevent loss of aggregate during construction. Thus, the primary
geotextile function in this application is separation, and can in some cases be considered a
secondary geogrid function.
The system performance may also be influenced by functions of filtration and drainage (see
Table 1-1). The geotextile acts as a filter to prevent fines from migrating up into the
aggregate due to high pore water pressures induced by dynamic wheel loads. It also acts as a
drain, allowing the excess pore pressures to dissipate through the geotextile and the subgrade
soils to gain strength through consolidation and improve with time.
System performance may also be improved through reinforcement. Geogrids and geotextiles
may provide reinforcement through three possible mechanisms.
1. Lateral restraint of the base and subgrade through friction (geotextiles) and interlock
(geogrids) between the aggregate, soil and the geosynthetic (Figure 5-4a).
2. Increase in the system bearing capacity by forcing the potential bearing capacity failure
surface to develop along alternate, higher shear strength surfaces (Figure 5-4b).
3. Membrane support of the wheel loads (Figure 5-4c).
FHWA NHI-07-092
Geosynthetics Engineering

5-9

Roadways and Pavements


August 2008

Figure 5-4.

Possible reinforcement functions provided by geosynthetics in roadways:


(a) lateral restraint, (b) bearing capacity increase, and (c) membrane
tension support (after Haliburton, et al., 1981).

When an aggregate layer is loaded by a vehicle wheel or dozer track, the aggregate tends to
move or shove laterally, as shown in Figure 5-4a, unless it is restrained by the subgrade or
geosynthetic reinforcement. Soft, weak subgrade soils provide very little lateral restraint, so
when the aggregate moves laterally, ruts develop on the aggregate surface and also in the
subgrade. A geogrid with good interlocking capabilities or a geotextile with good frictional
capabilities can provide tensile resistance to lateral aggregate movement. Another possible
geosynthetic reinforcement mechanism is illustrated in Figure 5-4b. Using the analogy of a
wheel load to a footing, the geosynthetic reinforcement forces the potential bearing capacity
failure surface to follow an alternate higher strength path. This tends to increase the bearing
capacity of the subgrade soil.
FHWA NHI-07-092
Geosynthetics Engineering

5-10

Roadways and Pavements


August 2008

A third possible geosynthetic reinforcement function is membrane-type support of wheel


loads, as shown conceptually in Figure 5-4c. In this case, the wheel load stresses must be
great enough to cause plastic deformation and ruts in the subgrade. If the geosynthetic has a
sufficiently high tensile modulus, tensile stresses will develop in the reinforcement, and the
vertical component of this membrane stress will help support the applied wheel loads. As
tensile stress within the geosynthetic cannot be developed without some elongation, wheel
path rutting (in excess of 4 in. {100 mm}) is required to develop membrane-type support.
Therefore, this mechanism is generally limited to temporary roads or the first aggregate lift in
permanent roadways.
5.3-2 Possible Failure Modes of Permanent Roads
Yoder and Witczak (1975) define two types of pavement distress or failure. The first is a
structural failure, in which a collapse of the entire structure or a breakdown of one or more of
the pavement components renders the pavement incapable of sustaining the loads imposed on
its surface. The second type failure is a functional failure; it occurs when the pavement, due
to its roughness, is unable to carry out its intended function without causing discomfort to
drivers or passengers or imposing high stresses on vehicles. The cause of these failure
conditions may be due to excessive loads, climatic and environmental conditions, poor
drainage leading to poor subgrade conditions, and disintegration of the component materials.
Excessive loads, excessive repetition of loads, and high tire pressures can cause either
structural or functional failures.
Pavement failures may occur due to the intrusion of subgrade soils into the granular base,
which results in inadequate drainage and reduced stability. A small increase in fines can
substantially reduce the ability of the base to drain, e.g., an increase from 5% to 10% fines
(particles < No. 200 {0.075mm} sieve) can reduce the permeability of gravel by several
orders of magnitude. Either increasing fines or the moisture in the base can reduce its
resilient modulus, but a combination of both can be disastrous. Distress may also occur due
to excessive loads that cause a shear failure in the subgrade, base course, or the surface.
Other causes of failures are surface fatigue and excessive settlement, especially differential
settlement of the subgrade. Volume change of subgrade soils due to wetting and drying,
freezing and thawing, or improper drainage may also cause pavement distress.
Inadequate drainage of water from the base and subgrade is a major cause of pavement
problems (Cedergren, 1987). If the subgrade is saturated, excess pore pressures will develop
under traffic loads, resulting in subsequent softening of the subgrade. Under dynamic
loading, fines can be literally pumped up into the subbase or base.

FHWA NHI-07-092
Geosynthetics Engineering

5-11

Roadways and Pavements


August 2008

Improper construction practices may also cause pavement distress. Wetting of the subgrade
during construction may permit water accumulation and subsequent softening of the
subgrade in the rutted areas after construction is completed. Use of dirty aggregates (e.g.,
aggregates containing more than 5% fines) or contamination of the base aggregates during
construction may produce inadequate drainage, instability, and frost susceptibility.
Reduction in design thickness during construction due to insufficient subgrade preparation
may result in undulating subgrade surfaces, failure to place proper layer thicknesses, and
unanticipated loss of base materials due to subgrade intrusion. Yoder and Witczak (1975)
state that a major cause of pavement deterioration is inadequate observation and field control
by qualified engineers and technicians during construction.
After construction is complete, improper or inadequate maintenance may also result in
pavement distress. Sealing cracks and joints at proper intervals must be performed to prevent
surface water infiltration. Maintenance of shoulders will also affect pavement performance.
As indicated in the list of possible benefits resulting from geosynthetic use in permanent
roadway systems (section 5.2-4), properly designed geosynthetics can enhance pavement
performance and reduce the likelihood of failures. Considering that a modern pavement
section costs on the order of $25/yd2 ($30/m2) for secondary roads and up to $100/yd2
($120/m2) for a primary road (i.e., based on five hundred thousand to several million dollars
per mile), only a year or two of extended life will easily cover the cost of the geosynthetic,
typically on the order of 1 to 3 $/yd2 (1.20 to 3.60 $/m2) installed. A recent study in Virginia
on monitored pavement sections found an anticipated improvement of over 100 % in traffic
loading for sections with geotextile separation layers as compared to control sections (Bhutta,
1998; and, Al-Qadi and Appea, 2003).
Perhaps of greater value, the geosynthetics at the subgrade-base interface and/or within the
base increase the reliability of the base and subgrade support, allowing rehabilitations of the
riding layer (i.e., asphalt) and extending pavement life before complete reconstruction is
required. A competent subgrade/base support is critical to realizing life-cycle cost benefits of
surface rehabilitations over the life of a pavement structure.
5.3-3 Design for Separation
For separation design, the base course thickness required to adequately carry the design
traffic loads for the design life of the pavement is not reduced due to the use of a
geosynthetic. Recall that geotextile separators help prevent pavement failures due to the
intrusion of finer subgrade soil fractions into the granular base layer(s). Geotextiles
separation layers may also be used between dense and open graded base layers.

FHWA NHI-07-092
Geosynthetics Engineering

5-12

Roadways and Pavements


August 2008

Most any geotextile will work in this application as long as it is strong enough to survive
construction. As indicated in Table 5-1, filtration is a secondary function in this application.
Therefore the geotextile should have small enough openings to prevent contamination of the
base and subbase pavement layers from the subgrade materials and be sufficiently permeable
(i.e., more permeable than the subgrade) to prevent the development of pore water pressure
in the subgrade.
5.3-4 Design for Stabilization
In stabilization design, the geosynthetic (geotextile or geogrid) and aggregate thickness
required to stabilize the subgrade and provide an adequate roadbed are evaluated. Recall that
this application is primarily for construction expedience. For design of permanent roads, this
stabilization lift also provides an improved roadbed (i.e., less subgrade disturbance, a gravel
layer that will not be contaminated due to intermixing with the subgrade, and a potential for
subgrade improvement of time). The base course thickness required to adequately carry the
design traffic loads for the design life of the pavement may be reduced due to the improved
roadbed condition, provided an assessment is made of the improvement.
As indicated in Table 5-1, geosynthetics used in this application perform multiple functions
of separation, filtration and reinforcement. Separation design requirements were discussed in
previous section. Because the subgrade soils are generally wet and saturated in this
application, the filtration design principles developed in Chapter 2 are applicable.
With respect to reinforcement requirements, there are two main approaches to stabilization
design. The first approach inherently includes the reinforcement function through improved
bearing capacity and there is no direct reinforcing contribution (or input) for the strength
characteristics of the geosynthetic. In this method, geotextiles act as a separator and filter
only. When this approach is used for geogrids, a geotextile or graded granular soil separation
layer is also required to address these functional requirements. The second approach
considers a possible reinforcing effect due to the geosynthetic. It appears that the separation
function is more important for roadway sections with relatively small live loads where ruts,
approximating 2 in to 4 in. (50 to 100 mm) are anticipated. In these cases, a design which
assumes no reinforcing effect is generally conservative. On the other hand, for large live
loads on thin roadways where deep ruts (> 4 in. {100 mm}) may occur, and for thicker
roadways on softer subgrades, the reinforcing function becomes increasingly more important
if stability is to be maintained. It is for these latter cases that reinforcing analyses have been
developed and are appropriate.

FHWA NHI-07-092
Geosynthetics Engineering

5-13

Roadways and Pavements


August 2008

The reinforcing mechanisms mobilized in subgrade stabilization are different between


geogrids and geotextiles. Due to the open structure and large apertures, the geogrids
interlock with base course aggregate and change the stress and strain conditions in the
vicinity of the geogrid. The efficiency of the geogrid-aggregate interlock depends on the
relationship between aperture size and aggregate particle size, and the in-plane stiffness of
the geogrid ribs and junctions. A design method that recognizes these distinctions is
presented in Section 5.5 along with the empirical method in Section 5.4 for geotextiles,
modified for geogrids.
The separation function of geogrids, which is considered secondary in geogrid reinforced
stabilization applications, is less obvious mainly because of the geogrids open structure. It
is recommended that a geotextile be used as a separator beneath a geogrid to prevent
migration of fines into the aggregate layers over time. However, it is possible to eliminate
the geotextile by designing the gradation of the base or a subbase layer to provide separation
based on well known graded granular filter design principals (e.g., see Cedegren, 1989).
Graded granular subbase layers are conventionally used in roadway design (e.g., see the
FHWA Geotechnical Aspects of Pavements Manual, Christopher et al., 2006). Geogrids
provide a stable platform for the base aggregate, which may be sized to adequately filter the
subgrade fines to prevent pumping (see Figure 5.5 and Anderson, 2006). The movement of
fine grained soils into coarse aggregates can be prevented if the pore spaces of the aggregates
are small enough to hold the particles in place. When a geogrid is present at the subgradebase course interface, the relative movement of the soil particles is further constrained due to
confinement provided by the geogrid-aggregate interlock. As a result, the possibility of soil
migration is further reduced. The application of the graded granular filter criteria is shown in
the example for geogrid-reinforced unpaved roads in Section 5.5-3.

Figure 5-5. Filtration at the interface of two dissimilar materials (without geosynthetics)
(after Cedegren, 1989).
FHWA NHI-07-092
Geosynthetics Engineering

5-14

Roadways and Pavements


August 2008

5.3-5 Reinforced Base/Subbase Design


Geogrids have been used for reinforcement of aggregate layers within the pavement system
since their introduction in the early 1980s. The predominant reinforcing mechanism
associated with this application is base course lateral restraint (Figure 5-4a). The base course
lateral restraint develops through interlock between the aggregate, soil and the geogrid
because four reinforcement effects: (1) prevention of lateral spreading of aggregate; (2)
confinement of aggregate resulting in increased strength/stiffness of aggregate in the vicinity
of the geogrid; (3) reduction of vertical stresses on top of the subgrade; and (4) reduction of
shear stress on the subgrade.
Despite many successful projects and research it is recognized that the use of geogrid
reinforcement in paved roadways is relatively limited compared to other geosynthetics
applications. Berg et al. (2000), Perkins et al. (2005c), and Gabr et al. (2006) indicated the
following major reasons for the relatively limited used of geogrids in base reinforcement
applications:
1. Lack of an accepted design method. Currently the use of base reinforcement applications
is based on prior experience and empirically based design approaches which are limited
to the conditions of the related experiments.
2. Existing numerical models for pavement design without geosynthetics are complicated
(NCHRP, 2002) and the perception is that the inclusion of geosynthetics will complicate
them further. A recent movement toward adoption of a mechanistic-empirical design
approach recognized that this will allow quantification of the geogrid benefits in a
rational and consistent way.
3. Few studies provide comparison for the full range of available geogrids (i.e., woven vs.
extruded, different aperture stiffness/stability, etc.), and some types of geogrids have
been studied more often than others.
4. Geogrids (and geosynthetics in general) are perceived as special materials that are
considered only if problem areas need to be fixed (Gabr et al., 2006).
5. Lack of a uniform method for cost-benefit analysis.
It is recognized that the development of a design method within the framework of the
mechanistic-empirical design method will address the limitations of the current design
approaches and lead to a broader use of geosynthetics in base reinforcement applications.
Two approaches are presented for base reinforcement in Section 5.7. The first uses an
empirical procedure based on current AASHTO and the improved traffic benefit derived
from using the geosynthetic. The other method is based on the AASHTO mechanisticempirical design approach (AASHTO, 2008).

FHWA NHI-07-092
Geosynthetics Engineering

5-15

Roadways and Pavements


August 2008

5.3-6 Material Properties used in Design


As with any geosynthetic applications, the material properties required for design are based
on the properties required to perform the primary and secondary function(s) for the specific
application over the life of the system and the properties required to survive installation. The
separation and filtration functions are related to opening characteristics and are determined
based on the gradation of the adjacent layers (i.e., subgrade, base and/or subbase layers).
Some strength is, of course, required for the reinforcing function, which is based on the
requirements in the specific design approach. If the roadway system is designed correctly,
then the stress at the top of the subgrade due to the weight of the aggregate and the traffic
load should be less than the bearing capacity of the soil plus a safety factor, which is
generally a relatively low value compared to the strength of most geosynthetics. However,
the stresses applied to the subgrade and the geotextile during construction may be much
greater than those applied in-service. Therefore, the strength of the geotextile or geogrid in
roadway applications is usually governed by the anticipated construction stresses and the
required level of performance. This is the concept of geosynthetic survivability -- the
geosynthetic must survive the construction operations if it is to perform its intended function.
Table 5-2 relates the elements of construction (i.e., equipment, aggregate characteristics,
subgrade preparation, and subgrade shear strength) to the severity of the loading imposed on
the geosynthetic. If one or more of these items falls within a particular severity category
(i.e., moderate or high), then geosynthetics meeting those survivability requirements should
be selected.
Table 5-2
Construction Survivability Ratings (after Task Force 25, AASHTO, 1990)
Site Soil CBR at Installation1
Equipment Ground Contact
Pressure

<1

1 to 2

>3

> 50 psi
(350 kPa)

< 50 psi
(350 kPa)

> 50 psi
(350 kPa)

< 50 psi
(350 kPa)

> 50 psi
(350 kPa)

< 50 psi
(350 kPa)

4 in. (100 mm)3,4

NR5

NR

15

25

6 in. (150 mm)

NR

NR

12 in. (300 mm)

NR

18 in. (450 mm)

Cover Thickness2 (compacted)

NOTES:
1. Assume saturated CBR unless construction scheduling can be controlled.
2. Maximum aggregate size not to exceed one-half the compacted cover thickness.
3. For low-volume, unpaved roads (ADT < 200 vehicles).
4. The 4 in. (100 mm) min. cover is limited to existing road bases & not intended for use in new construction.
5. NR = NOT RECOMMENDED; 1 = high survivability Class 1 geotextiles per AASHTO M288 (2006).;
and, 2 = moderate survivability Class 2 geotextiles per AASHTO M288 (2006).

FHWA NHI-07-092
Geosynthetics Engineering

5-16

Roadways and Pavements


August 2008

For the high category in Table 5-2, geosynthetics that can survive the most severe conditions
anticipated during construction should be used and are designated as Class 1 geosynthetics in
the following geosynthetic property requirements tables. Geosynthetics that can survive
normal construction conditions are Class 2 geosynthetics and may be considered for the
moderate category. Variable combinations indicating a NOT RECOMMENDED rating
suggests that one or more variables should be modified to assure a successful installation.
Some judgment is required in using these criteria.
Table 5-3 and 5-4 list the geotextile property requirements from AASHTO M288 (2006) for
stabilization and separation applications. Geotextiles that meet or exceed these survivability
requirements can be considered acceptable for most projects. The selected geotextile must
also retain the underlying subgrade soils, allowing the subgrade to drain freely, consolidate,
and gain strength. Thus, the geotextile must be checked, using the drainage and filtration
requirements in Chapter 2. Default geotextile requirements are presented in each table.
The survivability requirements in these tables were based on both research and on the
properties of geotextiles which have performed satisfactorily as separators in roads and in
similar applications. In the absence of any other information, they should be used as
minimum property values. Judgment and experience may be used to reduce the geotextile
requirements as indicated by AASHTO M288 or possibly increase the geotextile
requirements for very harsh construction conditions (e.g., NR is indicated by Table 5-2).
Table 5-5 lists the survivability requirements for geogrids in stabilization and base
reinforcement applications. A national guide of practice has not been established for
geogrids. Therefore the recommended requirements were developed specifically for this
manual and were based on a review of research on construction survivability (e.g., GMA,
1999), a review of state and federal agency specifications on geogrids (e.g., Christopher et
al., 2001 and USCOE, 2003), and on the properties of geogrids which have performed
satisfactorily in these applications (e.g., Berg et al., 2000). The specific property
requirements were conservatively selected with consideration for high reliability required on
public sector projects. Field trials or construction survivability tests following the
recommendations in note 5 of Table 5-5 for both the material and junction strength could be
used to reduce this conservatism.

FHWA NHI-07-092
Geosynthetics Engineering

5-17

Roadways and Pavements


August 2008

Table 5-3
Geotextile Property Requirements1,2,3
for Stabilization Applications (CBR < 3)
(after AASHTO, 2006)
ASTM
Test
Method

Property

Units

Requirement
Geotextile Class 14

SURVIVABILITY
< 50%5

Elongation

> 50%5

Grab Strength

D 4632

lb (N)

315 (1400)

200 (900)

Sewn Seam Strength6

D 4632

lb (N)

280 (1260)

180 (810)

Tear Strength

D 4533

lb (N)

110 (500)

80 (350)

Puncture Strength

D 6241

lb (N)

620 (2750)

433 (1925)

Ultraviolet Stability
(Retained Strength)

D 4355

50% after 500 hours of exposure

DRAINAGE AND FILTRATION7


Apparent Opening Size

D 4751

mm

0.43 for < 50% passing No. 200 (0.075 mm) sieve
< 0.3 for > 50% passing No. 200 (0.075 mm) sieve

Permittivity

D 4491

sec-1

0.5 for < 15% passing No. 200 (0.075 mm) sieve
0.2 for 15 to 50%passing No. 200 (0.075 mm) sieve
0.1 for > 50% passing No.200 (0.075 mm) sieve

NOTES:
1. Acceptance of geotextile material shall be based on ASTM D 4759.
2. Acceptance shall be based upon testing of either conformance samples obtained using Procedure A of
ASTM D 4354, or based on manufacturers certifications and testing of quality assurance samples obtained
using Procedure B of ASTM D 4354.
3. Minimum; use value in weaker principal direction. All numerical values represent minimum average roll
value (i.e., test results from any sampled roll in a lot shall meet or exceed the minimum values in the table).
Lot samples according to ASTM D 4354.
4. Default geotextile selection. The engineer may specify a Class 2 geotextile (see Appendix D) for moderate
survivability conditions, see Table 5-2.
5. As measured in accordance with ASTM D 4632.
6. When seams are required. Values apply to both field and manufactured seams.
7. The geotextile permeability should be greater than the soil permeability.
8. Due to filtration and drainage requirements, woven slit film geotextiles should not be allowed.

FHWA NHI-07-092
Geosynthetics Engineering

5-18

Roadways and Pavements


August 2008

Table 5-4
Geotextile Property Requirements1,2,3
for Separation Applications (CBR 3)
(after AASHTO, 2006)
ASTM
Test
Method

Property

Units

Requirement
Geotextile Class 24

SURVIVABILITY
< 50%5

Elongation

> 50%5

Grab Strength

D 4632

lb (N)

250 (1100)

157 (700)

Sewn Seam Strength6

D 4632

lb (N)

220 (990)

140 (630)

Tear Strength

D 4533

lb (N)

90 (400)

56 (250)

Puncture Strength

D 6241

lb (N)

495 (2200)

309 (1375)

Ultraviolet Stability
(Retained Strength)

D 4355

50% after 500 hours of exposure

DRAINAGE AND FILTRATION7


Apparent Opening Size

D 4751

mm

< 0.6 for < 50% passing No. 200 (0.075 mm) sieve
< 0.3 for > 50%passing No. 200 (0.075 mm) sieve

Permittivity

D 4491

sec-1

> 0.02 and > Permittivity of soil

NOTES:
1. Acceptance of geotextile material shall be based on ASTM D 4759.
2. Acceptance shall be based upon testing of either conformance samples obtained using Procedure A of
ASTM D 4354, or based on manufacturers certifications and testing of quality assurance samples obtained
using Procedure B of ASTM D 4354.
3. Minimum; use value in weaker principal direction. All numerical values represent minimum average roll
value (i.e., test results from any sampled roll in a lot shall meet or exceed the minimum values in the table).
Lot samples according to ASTM D 4354.
4. Default geotextile selection. The engineer may specify a Class 3 geotextile (see Appendix D) for moderate
survivability conditions, see Table 5-2.
5. As measured in accordance with ASTM D 4632.
6. When seams are required. Values apply to both field and manufactured seams.
7. Also, the geotextile permeability should be greater than the soil permeability.

FHWA NHI-07-092
Geosynthetics Engineering

5-19

Roadways and Pavements


August 2008

Table 5-5
Geogrid Survivability Property Requirements1,2,3
For Stabilization and Base Reinforcement Applications
Property

Test
Method

Units

Requirement
Geogrid Class

SURVIVABILITY
Ultimate Multi-Rib Tensile
Strength
Junction Strength5
Ultraviolet Stability
(Retained Strength)

CLASS 14

CLASS 2

CLASS 3

ASTM
D 6637

lb/ft
(kN/m)

1230 (18)

820 (12)

820 (12)

GSI GRI
GG2

lb (N)

255 (1105)

25 (110)

8 (35)

ASTM
D 4355

50% after 500 hours of exposure

OPENING CHARACTERISTICS
Aperture Size

Direct
measure

in. (mm)

0.5 to 3 in. (12.5 to 75 mm) and


Aperture Size D50 of aggregate above geogrid
Aperture Size 2D85 of aggregate above geogrid

Separation

ASTM
D 422

mm

D85 of aggregate above geogrid < 5D85 subgrade


Other wise use separation geotextile with geogrid

NOTES:
1. Acceptance of geogrid material shall be based on ASTM D 4759.
2. Acceptance shall be based upon testing of either conformance samples obtained using Procedure A of
ASTM D 4354, or based on manufacturers certifications and testing of quality assurance samples obtained
using Procedure B of ASTM D 4354.
3. Minimum; use value in weaker principal direction. All numerical values represent minimum average roll
value (i.e., test results from any sampled roll in a lot shall meet or exceed the minimum values in the table).
Lot samples according to ASTM D 4354.
4. Default geogrid selection. The engineer may specify a Class 2 or 3 geogrid for moderate survivability
conditions, see Table 5-2, based on one or more of the following:
a) The Engineer has found the class of geogrid to have sufficient survivability based on field experience.
b) The Engineer has found the class of geogrid to have sufficient survivability based on laboratory testing
and visual inspection of a geotextile sample removed from a field test section constructed under
anticipated field conditions (see note 5).
5. Junction strength requirements have not been fully supported by data, and until such data is established,
manufacturers shall submit data from full scale installation damage tests in accordance with ASTM D 5818
documenting integrity of junctions. For soft soil applications, a minimum of 6 in. (150 mm) of cover
aggregate shall be placed over the geogrid and a loaded dump truck used to traverse the section a minimum
number of passes to achieve 4 in. (100 mm) of rutting. A photographic record of the geogrid after
exhumation shall be provided, which clearly shows that junctions have not been displaced or otherwise
damaged during the installation process.

FHWA NHI-07-092
Geosynthetics Engineering

5-20

Roadways and Pavements


August 2008

Survivability of geogrids and geotextiles for major projects should be verified by conducting
field tests under site-specific conditions. These field tests should involve trial sections using
several geosynthetics on typical subgrades at the project site and implementing various types
of construction equipment. After placement of the geosynthetics and aggregate, the
geosynthetics are exhumed to see how well or how poorly they tolerated the imposed
construction stresses. These tests could be performed during design or after the contract was
let, similar to the recommendations for riprap placement (Section 3.8-1). In the latter case,
the contractor is required to demonstrate that the proposed subgrade condition, equipment,
and aggregate placement will not significantly damage the geotextile or geogrid. If
necessary, additional subgrade preparation, increased lift thickness, and/or different
construction equipment could be utilized. In rare cases, the contractor may even have to
supply a different geosynthetic.

5.4 DESIGN GUIDELINES FOR USE OF GEOTEXTILES IN TEMPORARY AND


UNPAVED ROADS
Geotextiles have been extensively used for over four decades to facilitate the construction
and improve the performance of unpaved low-volume roads on weak subgrades. The
benefits of geotextiles used in unpaved low-volume roads have been well documented in
numerous case histories, full-scale laboratory experiments, and instrumented field studies
(e.g., Stewart et al, 1977; Bender and Barenberg, 1978; Haliburton and Barron, 1983; AlQadi, et al., 1994; Austin and Coleman, 1993; Wen-Sen Tsai, 1995; Fannin and Sigurdsson,
1996; Christopher and Lacina, 2008; just to name just a few). This historical data base
provides the support for the recommended design methods covered in this section as well as
the extension of these procedures for designing geotextiles in permanent roads in Section 5.6.
For unpaved road design, the design method presented in this section considers mainly the
separation and filtration functions of the geotextile with some reinforcing benefit. It was
selected because it has a long history of successful use, it is based on principles of soil
mechanics, and it has been calibrated by full-scale field tests. It can also be adapted to a
wide variety of conditions. Although the reinforcement function is inherently included in
this method through improved bearing capacity, there is no direct reinforcing contribution (or
input) for the strength characteristics of the geotextile. Most of the above referenced research
has found that significant rutting (i.e., > 4 in. {100 mm}) is required to obtain additional
reinforcement benefit. Other methods considering reinforcement functions are described by
Koerner (2005), Giroud and Han (2004), Christopher and Holtz (1985), and Giroud and
Noiray (1981). For roadways where stability of the embankment foundation is questionable
(i.e., ((H/c) > 3), refer to Chapter 7 for information on reinforced embankments.
FHWA NHI-07-092
Geosynthetics Engineering

5-21

Roadways and Pavements


August 2008

The following design method was developed by Steward, Williamson, and Mohney (1977)
for the U.S. Forest Service (USFS). It allows the designer to consider:

vehicle passes;
equivalent axle loads;
axle configurations;
tire pressures;
subgrade strengths; and
rut depths.

The following limitations apply:

the aggregate layer must be


a) high quality fill (e.g., laboratory CBR based on ASTM D 1883 80),
b) cohesionless (nonplastic);
vehicle passes less than 10,000;
geotextile survivability criteria must be considered; and
subgrade undrained shear strength less than about 2000 psf (90 kPa) (CBR < 3).

As discussed in Section 5.2-3, for subgrades stronger than about 2000 psf (90 kPa) (CBR >
3), geotextiles are rarely required for stabilization, although geotextiles may provide
separation benefits and should be used to enhance drainage and filtration (i.e., allowing the
use of open graded aggregate). Where drainage and filtration are likely to be important, the
principles developed in Chapter 2 are applicable, just as they are for weaker subgrades where
drainage and filtration are likely to be very important.
Based on both theoretical analysis and empirical (laboratory and full-scale field) tests on
geotextiles, Steward, Williamson and Mohney (1977) determined that a certain amount of
rutting would occur under various traffic conditions, both with and without a geotextile
separator and for a given stress level acting on the subgrade. They present this stress level in
terms of bearing capacity factors, similar to those commonly used for the design of shallow
foundations on cohesive soils. These factors and conditions are given in Table 5-6.
The following design procedure is recommended:

FHWA NHI-07-092
Geosynthetics Engineering

5-22

Roadways and Pavements


August 2008

Table 5-6
Bearing Capacity Factors for Different Ruts and Traffic
Conditions both With and Without Geotextiles
(after Steward, Williamson, and Mohney, 1977)

Ruts
in. (mm)

Traffic
(Passes of 18 kip {80 kN}
axle equivalents)

Bearing Capacity
Factor, Nc

Without Geotextiles

< 2 in. (50 mm)


> 4 in. (100 mm)

>1000
<100

2.8
3.3

With Geotextiles

< 2 in. (50 mm)


> 4 in. (100 mm)

>1000
<100

5.0
6.0

Condition

STEP 1. Determine soil subgrade strength.


Determine the subgrade soil strength in the field using the field CBR, cone penetrometer,
vane shear, resilent modulus, or any other appropriate test. The undrained shear strength
of the soil, c, can be obtained from the following relationships:
for field CBR, c in psi = 4.3 x CBR (c in kPa = 30 x CBR);
for the WES cone penetrometer, c = cone index divided by 10 or 11,
depending on the soil type; and
for the vane shear test, c is directly measured.
Other in-situ tests, such as the static cone penetrometer test (CPT) or dilatometer (DMT),
may be used, provided local correlations with undrained shear strength exist. Use of the
Standard Penetration Test (SPT) is not recommended for soft clays.
Determine subgrade strength at several locations and at different times of the year. Make
strength determinations at several locations where the subgrade appears to be the
weakest. Strengths should be evaluated at depth of 0 in. to 8 in. (0 to 200 mm) and from
8 in. to 20 in. (200 - 500 mm); six to ten strength measurements are recommended at each
location to obtain a good average value. Tests should also be performed when the soils
are in their weakest condition, when the water table is the highest, etc. Alternatively, a
saturated soaked laboratory CBR test (ASTM D1883) could be performed to model wet
conditions in the field (e.g., for compacted soils that will be exposed to wet conditions).
STEP 2. Determine wheel loading.
Determine the maximum single wheel load, maximum dual wheel load, and the
maximum dual tandem wheel load anticipated for the roadway during the design period.
For example, a 10 yd3 (7.6 m3) dump truck with tandem axles will have a dual wheel load
of approximately 8,000 lbf (35 kN). A motor grader has a wheel load of 5,000 to 10,000
lbf (22 to 44 kN).
FHWA NHI-07-092
Geosynthetics Engineering

5-23

Roadways and Pavements


August 2008

STEP 3. Estimate amount of traffic.


Estimate the maximum amount of traffic anticipated for each design vehicle class.
STEP 4. Establish tolerable rutting.
Establish the amount of tolerable rutting during the design life of the roadway. For
example, a rut of 2 in. to 3 in. (50 to 75 mm) is generally acceptable during construction.
STEP 5. Obtain bearing capacity factor(s).
Obtain appropriate subgrade stress level in terms of the bearing capacity factors in Table
5-6. Values may be obtained for both the conditions with geotextiles and without
geotextiles for estimating the cost effectiveness of using a geotextile.
STEP 6. Determine required aggregate thickness(es).
Determine the required aggregate thickness(es) from the USFS design chart (Figures 5-6,
5-7 or 5-8) for each maximum loading. Enter the curve with appropriate bearing capacity
factors (Nc) multiplied by the design subgrade undrained shear strength (c) to evaluate
each required stress level (cNc).
STEP 7. Select design thickness(es).
Select the design thickness based on the design requirements (e.g., for the maximum
loading condition, with and without geotextiles, as required). The design thickness(es)
should be given to the next higher 1 in. (25 mm).
STEP 8. For geotextile, check geotextile drainage and filtration characteristics.
Check the geotextile drainage and filtration requirements. Use the gradation and
permeability of the subgrade, the water table conditions, and the retention and
permeability criteria given in Chapter 2. From Chapter 2, that criteria is:
(Wovens)
(Eq. 2-3)
AOS < D85
(Eq. 2-4)
AOS < 1.8 D85 (Nonwovens)
kgeotextile > ksoil
(Eq. 2-7a)
-1
< 0.1 sec
(Eq. 2-8c)
STEP 9. Determine geotextile survivability requirements.
Check the geotextile survivability strength requirements as discussed in Section 5.3-6.
STEP 10. Specify geotextile property requirements.
Specify geotextiles that meet or exceed the survivability criteria from Step 9.
STEP 11. Specify construction requirements. (Follow the procedures in Section 5.8.)
FHWA NHI-07-092
Geosynthetics Engineering

5-24

Roadways and Pavements


August 2008

Figure 5-6.

U.S. Forest Service thickness design curve for single wheel load (Steward et
al., 1977).

Figure 5-7.

U.S. Forest Service thickness design curve for tandem wheel load (Steward et
al., 1977).

FHWA NHI-07-092
Geosynthetics Engineering

5-25

Roadways and Pavements


August 2008

(a)

(b)
Figure 5-8. Thickness design curves with geotextiles for a) single and b) dual wheel loads
(after Steward et al., 1977 & FHWA NHI-95-038, 1998; modified for highway applications).
FHWA NHI-07-092
Geosynthetics Engineering

5-26

Roadways and Pavements


August 2008

5.4-1

Temporary Road Design Example

DEFINITION OF DESIGN EXAMPLE


Project Description:
A haul road over wet, soft soils is required for a highway
construction project.

Type of Structure:

Type of Application:
geotextile for stabilization of subgrade (functions separation,
filtration, and some reinforcement)

Alternatives

GIVEN DATA
subgrade

traffic

ruts

temporary unpaved road

:i) excavate unsuitable material and increased aggregate thickness


ii.) geotextile between aggregate and subgrade
iii.) use an estimated depth of aggregate and maintain as required

cohesive subgrade soils


high water table
average undrained shear strength about 600 psf (30 kPa) or CBR = 1
approximately 5000 passes
20,000 lbf (90 kN) single axle truck
80 psi (550 kPa) tire pressure
maximum of 2 in to 4 in. (50 to 100 mm)

REQUIRED
Design the roadway section.
Consider: 1) design without a geotextile; and, 2) alternate with geotextile.
DEFINE
A. Geotextile function(s):
B. Geotextile properties required:
C. Geotextile specification:
SOLUTION
A. Geotextile function(s):
Primary
- separation and filtration
Secondary - drainage, reinforcement
B. Geotextile properties required:
survivability
apparent opening size (AOS)
FHWA NHI-07-092
Geosynthetics Engineering

5-27

Roadways and Pavements


August 2008

DESIGN

Design roadway with and without geotextile inclusion. Compare options.

STEP 1.

DETERMINE SOIL SUBGRADE STRENGTH

given - CBR 1
Assume that CBR 1 is taken from area(s) where the subgrade appears to be the weakest.
STEP 2.
given

STEP 3.
given
STEP 4.
given

DETERMINE WHEEL LOADING


-

20,000 lbf (90 kN) single-axle truck, with 80 psi (550 kPa) tire pressure
therefore, 10,000 lbf (45 kN) single wheel load

ESTIMATE AMOUNT OF TRAFFIC


-

5,000 passes

ESTABLISH TOLERABLE RUTTING


-

2 in. to 4 in. (50 to 100 mm)

STEP 5.
OBTAIN BEARING CAPACITY FACTOR
Without a geotextile: - 2.8 < Nc < 3.3
- assume Nc . 3.0 for 5,000 passes and 2 in to 4 in. (50 to 100 mm) ruts
with a geotextile:

STEP 6.

- 5.0 < Nc < 6.0


- assume Nc . 5.5 for 5,000 passes and 2 in to 4 in. (50 to 100 mm) ruts

DETERMINE REQUIRED AGGREGATE THICKNESSES

without a geotextile (FROM FIGURE 5-6)


- cNc = 600 psf x 3.0 = 1800 psf
(cNc = 30 kPa x 3.0 = 90 kPa)
depth of aggregate 19 in. (475 mm)
with a geotextile
- cNc = 600 psf x 5.5 = 3500 psf
(cNc = 30 kPa x 3.0 = 165 kPa)
Depth of aggregate 13 in. (325 mm)

FHWA NHI-07-092
Geosynthetics Engineering

5-28

Roadways and Pavements


August 2008

STEP 7.

SELECT DESIGN THICKNESS

Use 13 in. (325 mm) and a geotextile

STEP 8.

CHECK GEOTEXTILE DRAINAGE AND FILTRATION CHARACTERISTICS

Use AOS < 0.012 in. (0.3 mm) and permittivity 0.1 sec-1, per requirement of Table 5-3 since
soil has > 50% passing the No. 200 (0.075 mm) sieve. Permeability of geotextile must be greater
than soil permeability.

STEP 9.

DETERMINE GEOTEXTILE SURVIVABILITY REQUIREMENTS

Use Table 5-2: with CBR = 1, dump truck contact pressure > 80 psi (550 kPa), and 13 in. (325
mm) cover thickness, and find a MODERATE survivability to NOT RECOMMENDED rating.
Use a HIGH, or Class 1, survivability geotextile, or greater.

STEP 10.

SPECIFY GEOTEXTILE PROPERTY REQUIREMENTS

From Table 5-3; geotextile shall meet or exceed the minimum average roll values, with
elongation at failure determined with the ASTM D 4632 test method, of:

Property
Grab Strength
Sewn Seam Strength
Tear Resistance
Puncture Strength
Ultraviolet Stability

ASTM
Test Method
D 4632
D 4632
D 4533
D 6241
D 4355

Elongation
Elongation
< 50%
> 50%
315 lb (1400 N)
200 lb (900 N)
270 lb (1200 N)
180 lb (810 N)
110 lb ( 500 N)
80 lb (350 N)
620 lb (2750 N)
433 lb (1925 N)
50% strength retained after 500 hours

The geotextile shall have an AOS < 0.3 mm, > 0.1 sec-1, and the permeability shall be ______.

STEP 11. SPECIFY CONSTRUCTION REQUIREMENTS


See Section 5.8.

FHWA NHI-07-092
Geosynthetics Engineering

5-29

Roadways and Pavements


August 2008

5.5

DESIGN GUIDELINES FOR USE OF GEOGRIDS IN TEMPORARY AND


UNPAVED ROADS

Geogrids are commonly used to facilitate the construction and improve the performance of
unpaved low-volume roads on weak subgrades. As previously indicated in Section 5-3, the
primary function of the geogrid in this application is reinforcement leading to reduced
amount of aggregate needed, less maintenance, extended service life or a combination of
these. A secondary function is fill/subgrade separation.
The benefits of geogrids in unpaved low-volume roads have been shown in numerous
laboratory and full-scale experiments (e.g., Haas et al., 1988; Webster, 1993; Collin et al.,
1996; Fannin and Sigurdsson, 1996; Knapton and Austin, 1996; Gabr et al., 2001; and, Leng
and Gabr, 2002). Some experimental programs investigated the performance of different
geogrids (extruded, woven or welded) and the results showed that the stiffer geogrids
performed better (Webster, 1993; Collin et al., 1996). These experiments served as a basis
for the development of the empirical design methods for geogrid-reinforced unpaved lowvolume roads.
Historically the geogrids were introduced to the market in the early 1980s and by that time
geotextiles were used at the base-subgrade interface for separation, filtration and some
reinforcement. As a result, the first empirical design procedures of Barenberg et al. (1975)
and Steward et al. (1977) (as described in Section 5.4) were developed for geotextilesreinforced unpaved roads using solutions based on the limit equilibrium bearing capacity
theory. The solution of Steward et al. (1977) was modified by Tingle and Webster (2003) for
geogrid reinforcement and the proposed modification was adopted in the COE method for
design of geotextile- and geogrid-reinforced unpaved roads (USCOE, 2003). This approach
is described in Section 5.5-1.
Utilizing previous research, Giroud and Han (2004) developed theoretically based and
experimentally calibrated design method for geogrid-reinforced unpaved roads that reflects
the improvements due to the geogrid-aggregate interlock. The method can also be utilized
for analysis of unreinforced and geotextile-reinforced unpaved roads, or temporary platforms.
This approach will be presented in Section 5.5-2.
5.5-1 Empirical Design Method: Modified Steward et al. (1977)
Tingle and Webster (2003) used full-scale experiments to evaluate the applicability of the
design procedure for geogrid-reinforced unpaved roads and further validated the bearing
capacity factors of 5.0 and 6.0 for geotextile-reinforced unpaved roads (as discussed in
Section 5.4). Their analysis concluded that the bearing capacity factor of 2.8 for
FHWA NHI-07-092
Geosynthetics Engineering

5-30

Roadways and Pavements


August 2008

unreinforced roads was acceptable. For the geogrid-reinforced case they suggested a bearing
capacity factor of 5.8 and recommended the use of geotextile as a separator. Application of
the modified Steward et al. (1977) design method to geogrid-reinforced unpaved roads is the
same as the method outlined in Section 5.4 and using a bearing capacity factor of 5.8. The
geogrid is based on the properties listed in Table 5-5. The area of applicability and
limitations of this design method are the same as those presented in Section 5.4 and are not
repeated here. It is recommended that a geotextile be used as a separator beneath a geogrid
unless the gradation of the aggregate can act as a separator for the subgrade (Section 5.3-4).
5.5-2 Empirical Design Method of Giroud and Han (2004)
Giroud and Han (2004) developed a theoretically based and empirically calibrated design
method specifically designed for geogrid-reinforced unpaved roads and areas. They built
upon earlier design methods developed by Giroud and Noiray (1981) and Giroud et al.
(1985) using recent field and laboratory test data. Giroud and Noiray (1981) developed an
empirical solution for unreinforced unpaved roads using field test data and quantified the
benefits resulting from geotextile reinforcement. The solution was based on the limit
equilibrium bearing capacity theory with a modification to consider the benefit of the tension
membrane effect. The Giroud-Han theoretical formulation takes into account the distribution
of stresses, strength of base course material, geogrid-aggregate interlock, and geogrid inplane stiffness in addition to conditions considered in earlier methods (traffic volume, wheel
loads, tire pressure, subgrade strength, rut depth and influence of reinforcing geosynthetics of
the failure mode of unpaved roads). The influence of different factors on the theoretical
formulations, the assumptions and the limitations of the Giroud-Han design method are
briefly presented below.
The properties of the base course material are considered in the solution which is an
advancement compared to previous methods. The base course material is characterized by its
CBR using the AASHTO chart for correlation with the resilient modulus for subbase
(AASHTO, 1993).
The subgrade soil is assumed to be saturated and exhibit undrained behavior under traffic
loading. The subgrade soil modulus is used based on correlation between the field CBR and
the field resilient modulus for fine grained soils (Heukelom and Klomp, 1962). Other
relationships can also be used to derive the resilient modulus of the subgrade soil. In the
formulation of the design equation, the ratio of the resilient modulus of base course to
subgrade soil is limited to 5. Additional data are necessary to justify the use of higher values
for stiff geogrids which appear to improve the compaction of base course material even on
very soft subgrades.

FHWA NHI-07-092
Geosynthetics Engineering

5-31

Roadways and Pavements


August 2008

Serviceability Criterion Based on Rut Depth. Failure of the unpaved roads is assumed to
be controlled by the shear failure or the excessive deformation of the subgrade. The
formulation of the design method is based on a typical surface rut depth of 3 in. (75 mm)
which a serviceability criterion. It allows for rut depths between 2 and 4 in. (50 and 100 mm)
to be analyzed. Additional field data are needed to support the use of the method beyond
these limits.
Characterization of Geogrid Reinforcement. The properties of geogrids relate to their
ability to interlock with the base course material and provide confinement. Based on
research by Kinney (1995) and Collin et al. (1996), the aperture stability modulus was the
stiffness property selected, based on correlation with measured performance in roads. The
aperture stability modulus is obtained by measuring the in-plane torsional behavior directly
across the junction of a biaxial geogrid. It is a direct measure of the in-plane stiffness and
stability of the ribs and junctions of the geogrid. The method was calibrated using data for
stiff biaxial geogrids with aperture stability modulus of 0.32 and 0.65 N-m/deg (Kinney,
2000). In the design method the aperture stability modulus can vary from zero to a
maximum value based on the data used in the calibration (Giroud and Han, 2004b). A draft
test method for determining the aperture stability modulus of a geogrid has been developed
by Kinney (2000) and a standard method is currently under development by ASTM.
Bearing Capacity Factors. The bearing capacity factors for unreinforced unpaved roads as
presented in Section 5.4 ranged from 2.8 to 3.3. Giroud and Han (2004a) adopted a bearing
capacity factor of 3.14 (i.e., ) which is the value of the elastic limit for saturated undrained
subgrade soil for plain-strain and axisymmetric conditions and zero interface shear stress. As
discussed earlier the strike through and the interlock at the geogrid-reinforced interface
resists the lateral movement at the top of the subgrade, and creates inward shear stresses on
the subgrade. The theoretical value of the ultimate bearing capacity factor for axisymmetric
conditions and maximum inward shear stress of 5.71 (i.e., 3/2) is adopted for the geogridreinforced unpaved roads. For the case when the base course is separated by a geotextile and
there is no interlock, Giroud and Han adopted the value of 5.14 (i.e., +2) initially proposed
by Giroud and Noiray (1981), which is the ultimate bearing capacity factor for plain-strain
conditions and zero shear stress at the base-subgrade interface.
Equation for Required Thickness of Base Course. The thickness of the base course
material was determined on the basis of the bearing capacity theory to prevent the
development of rut depths exceeding the predetermined serviceability criterion. The
deformation of the subgrade depends on the stresses applied at the base-subgrade interface
and the development of the rut depth as a function of the stresses at the base-subgrade
interface and the bearing capacity of the subgrade. The influence of traffic, properties of
FHWA NHI-07-092
Geosynthetics Engineering

5-32

Roadways and Pavements


August 2008

base course material, and geogrid properties are expressed through two important parameters
the Bearing Capacity Mobilization Coefficient (m), and the Stress Distribution Angle ().
The Bearing Capacity Mobilization Coefficient defines the level of mobilized bearing
capacity, which depends on the deflection at the top of subgrade when the surface rutting
reaches the allowable rut depth. The Stress Distribution Angle defines the capability of the
base course material to transfer traffic loads to the subgrade. The effect of traffic and
geogrid on the rate of change of stress distribution angle as the unpaved roads deteriorate
under repeated loading is considered in the formulation.
The following design equation for base course thickness was developed through calibration
and verification with laboratory and field data (Giroud and Han, 2004b):

1.5

r

P

0.868 + (0.661 1.006 J 2 ) log N


2
h

r
h=
1 r
2

[1 + 0.204(RE 1)]

1
0
.
9
e
N c f c CBR sg
f

where:
(0.661-1.006 J2) > 0
h = required base course thickness (in. or m)
J = aperture stability modulus in metric units (N-m/degree)
P = wheel load (lbs or kN)
r = radius of tire print (in. or m)
N = number of axle passes
RE = modulus ratio = Ebc/Esg = 3.28 CBRbc0.3 / CBRsg 5
Ebc = base course resilient modulus (psi or MPa})
Esg = subgrade soil resilient modulus (psi or MPa)
CBRbc = aggregate CBR
CBRsg = subgrade CBR
fs = rut depth factor
s = maximum rut depth (in. or m)
Nc = bearing capacity factor
= 3.14 for unreinforced roads
= 5.14 for geotextile reinforced roads
= 5.71 for geogrid reinforced roads
fc = factor relating subgrade CBR to undrained cohesion, cu = 4.3 psi (30 kPa)

(1)

Limitations of the Design Method. The validity of the Giroud and Han method is limited
by the following conditions:
FHWA NHI-07-092
Geosynthetics Engineering

5-33

Roadways and Pavements


August 2008

Rut depth from 2 to 4 in. (50 to 100 mm);


Field subgrade CBR less than 5;
Maximum ratio of base course modulus Ebc to subgrade soil modulus Esg of 5;
Maximum number of passes Based on the current state of practice, the trafficking for
unpaved roads is limited to 10,000 ESALs.
The tension membrane effect was not taken into account since it is negligible for rut
depths less than 4 in. (100 mm);
The influence of geogrid reinforcement is considered through a bearing capacity factor of
Nc = 5.71, and the aperture stability module (J) of geogrid;
The influence of geotextile reinforcement is considered through a bearing capacity factor
of Nc = 5.14, and aperture stability module equal to zero;
For the unreinforced unpaved roads, the solution is valid for bearing capacity factor of Nc
= 3.14, and aperture stability module equal to zero;
Minimum thickness of 4 in. (100 mm) of base course aggregate.

Giroud and Han (2004b) suggest that these limitations may change as additional empirical
data become available.
Design Procedure. The design steps from the previous Section 5-4 should be followed.
Steps 4 6 are replaced for a geogrid-reinforced alternative using the Giroud and Han (2004)
procedure as follows:
STEP 4: Preliminary calculations
Select allowable rut depth depending on the road use
Calculate the radius of the equivalent rut depth
r =

where: P = wheel load (lb or kN)


r = radius of tire contact (in. or m)
p = tire pressure (psi or kN/m2)
If necessary determine the undrained shear strength of the subgrade soil from available
data or correlations.
STEP 5: Check capacity of subgrade soil to support wheel load without reinforcement
P h = 0 , unreinf

FHWA NHI-07-092
Geosynthetics Engineering

s
r 2 N c c u
=

fs

5-34

Roadways and Pavements


August 2008

where:
Ph = support capacity of subgrade (lb or kN)
s = the allowable rut depth (in. or mm)
fs = 3 in. (75 mm)
r = radius of tire contact (in. or m)
Nc = 3.14 bearing capacity factor for unreinforced case
cu = subgrade undrained shear strength (psi or kN/m2)
If P < Ph=0, unreinf the subgrade soil can support the wheel load and a minimum thickness of 4
in. (100 mm) base course is recommended to prevent disturbance of the subgrade. If P >
Ph=0, unreinf the use of reinforcement is required and the solution continues to the next step.
STEP 6: Determine the required base course thickness for reinforced or unreinforced roads
using Equation (1). The calculation of the base course thickness requires iteration.
The minimum thickness of the base course is 4 in. (100 mm).
The Giroud and Han method will be illustrated in the example presented in the next section.
5.5-3 Design Examples for Geogrid Reinforced Unpaved Road
The design of geogrid-reinforced unpaved road will be illustrated with two examples. The
first example is based on the Giroud and Han method (2004a,b), where the geogrid
reinforcement benefits are considered through the bearing capacity factor (Nc) and the
aperture stability of the geogrid (J). An important feature of the Giroud and Han is that it can
differentiate the benefits of different types of geogrids.
The second example is based on the Modified Steward et al., 1977 method (USCOE, 2003),
where the geogrid reinforcement benefits are considered only through the bearing capacity
factor, Nc = 5.8, derived from empirical studies for extruded biaxial geogrids under laid with
a geotextile separators.

DESIGN EXAMPLE 1: GIROUD AND HAN METHOD (2004 a, b)


PART I: GEOGRID REINFORCEMENT
Determine an appropriate aggregate thickness for a haul road over weak subgrade that is required for
a highway construction project. Investigate a conventional unreinforced solution and a geogridreinforced alternative, using the Giroud and Han method (2004 a, b) for the given set of design
parameters.

FHWA NHI-07-092
Geosynthetics Engineering

5-35

Roadways and Pavements


August 2008

DESIGN INPUT
Traffic Load:
Axle load = 18 kip (80 kN)
Tire pressure = 80 psi (550 kPa)
Number of axle passes = 5000
Failure Criteria:
Maximum rut depth = 3 in. (75 mm)
Aggregate and Subgrade Soil Strength:
Aggregate fill CBR = 15
Field subgrade CBR = 1
Geosynthetic Reinforcement:
Extruded Biaxial Geogrid with Aperture Stability Modulus, J = 0.32 N-m/degree
Bearing capacity factors:
Nc = 3.14 for unreinforced road section
Nc = 5.71 for geogrid-reinforced road section

DESIGN CALCULATIONS
STEP 4: PRELIMINARY CALCULATIONS
Wheel load, P = 9,000 lbs (40 kN)
Allowable rut depth, s = 3 in. (75 mm)
Radius of tire contact:
Ratio

of

base

E bc 3.48CBRbc
=
E sg
CBR sg

40
= 0.152 m = 6 in.
3.14 x 550

r=

course
0.3

(3.48)(15)
(1.0)

modulus
0.3

to

subgrade

modulus:

= 7. 8 > 5

The ratio of base course modulus to subgrade modulus of 5 is used in the calculations.

STEP 5: CHECK CAPACITY OF SUBGRADE SOIL TO SUPPORT WHEEL LOAD WITHOUT


REINFORCEMENT
P h = 0 , unreinf

75
=
(0 . 152
75

)2 (3 . 14 )(30

* 1 . 0 ) = 6 . 83

kN

P = 40 kN > 6.83 kN = Ph, unreinf


The subgrade soil cannot support the wheel load and use of reinforcement is required.

FHWA NHI-07-092
Geosynthetics Engineering

5-36

Roadways and Pavements


August 2008

STEP 6: CALCULATION OF THE REQUIRED BASE COURSE THICKNESS.


Giroud and Han design equation (1) is used to determine the required aggregate thickness (h)
for each of the unreinforced and reinforced cases. In order to calculate a required thickness
using the iterative Giroud-Han equation, it is necessary to substitute for the thickness, h, until
both sides of the equation are numerically the same.
Case 1: Unreinforced Unpaved Road
Using Equation (1) for J = 0, and Nc=3.14, and after two or three iteration cycles the right
side of the equation is approximately the same as the left side for h = 20 in. (505 mm).

0.152
0.868 + 0.661
log 5000

550
0.5045
0.152 = 0.5045 m

h=
1

[1 + 0.204(5 1)]

0.152

75
0.5045

1 0.9e
3.14 x 30 x 1

75

1.5

Therefore the calculated thickness for the unreinforced case is 20 in. (510 mm).
Case 2: Unpaved Road Reinforced with Stiff Biaxial Geogrid
Using Equation (1) for J = 0.32 N-m/degree, and Nc = 5.71, and after two or three iteration
cycles the right side of the equation is approximately the same as the left side for h = 12 in.
(300 mm).

1 .5

0 .152
0 .868 + 0 .661 1 .006 x 0 .32 2

0 .3054
h=
[1 + 0.204 (5 1)]

log 5000

550
1 0.152 = 0 .3054 m
2

0 .152

75
0 .3054 5 .71 x 30 x 1

1
0
.
9
e

75

The calculated thickness for the geogrid reinforced unpaved road is 12 in. (300 mm).
For J = 0, and Nc = 5.14, Equation (1) can be used to calculate the required base course
thickness for the case of geotextile-reinforced unpaved road. In this case the required
thickness will be 14 in. (360 mm).
STEP 7: SELECT BASE COURSE THICKNESS.
The geogrid-reinforced option for the unpaved road has been selected for:
Aggregate thickness = 12 in. (300 mm)
FHWA NHI-07-092
Geosynthetics Engineering

5-37

Roadways and Pavements


August 2008

STEP 8: SUBGRADE SEPARATION


Use of a geotextile separator is recommended unless the aggregate meets the natural filter
criteria for the subgrade. For the geotextile requirements, see Step 8 in Section 5-4.
The application of filter criteria for the case of geogrid-reinforced unpaved roads is illustrated
in the following calculation for subgrade separation (Cedergren, 1989, Berg et al., 2000). It
was discussed in Section 5.3-4 that the geogrid-aggregate interlock prevents the relative
movement of the soil particles and therefore further reduces the possibility of migration of
fine particles into the coarser material. In addition to the effect of proper gradations, there is
an effect of reduced pressures and deflections in the subgrade that results of the mechanical
interlock and lateral confinement of the aggregate provided by the geogrid. However, the
separation function of the geogrid has not been quantified, and if the natural filter gradation
requirements are not met, a geotextile separator should be specified.
DESIGN INPUT FOR SUBGRADE SEPARATION
The proposed unpaved road will be built on fine-grained subgrade, and aggregate material
from two different sources has been considered. In order to prevent contamination of the
base aggregate, for each of the aggregate sources, check the potential for migration of the
subgrade soil particles under the mechanical action of construction and operating traffic.
The following information has been provided by the geotechnical engineer for the existing
subgrade soil and the two aggregate materials that are being considered:
Soil

Aggregate

Subgrade

Classification per USCS

Option 1

Option 2

ML

SP-SC

SP

Low plasticity silt

Poorly graded sand


with clay and gravel

Poorly graded
sand with gravel

with sand
3 in. (75 mm)

100

97

97

No. 4 (4.75 mm)

88

71

77

No. 40 (0.425 mm)

28

39

No. 200 (0.075 mm)

78

11

No. 400 (0.038 mm)

41

0.01 mm

LL

33

32

PI

16

Non Plastic

Coefficient of Uniformity, Cu

4.8

5.4

Coefficient of Curvature, Cc

2.9

3.6

Gradation
(%
Passing)

Plasticity

FHWA NHI-07-092
Geosynthetics Engineering

5-38

Roadways and Pavements


August 2008

SUBGRADE SEPARATION CHECK USING NATURAL FILTER CRITERIA


The subgrade soil has been identified by the geotechnical engineer as Low Plasticity Silt with
Sand (ML) and the following filter criteria apply (Cedergren, 1989):
D15Fill
D85 Subgrade

and

D50Fill
D50 Subgrade

25

The calculations for both options are presented in the following table:
Subgrade
Soil Type

Characteristic
Size (mm)

ML

Particle

D15
D50
D85

Piping Ratio = (D15 Fill)/(D85 Subgrade)


(D50 Fill)/(D50 Subgrade)
Aggregate Option 1:
D15Fill
D85 Subgrade

Low plasticity
silt with sand
0.045
1.37
-

Aggregate
Option 1

Option 2

SP-SC
Sand with clay and
gravel (poorly graded)
0.11
1.46
0.08
32

SP
Sand with gravel
(poorly graded)
0.13
0.86
0.1
19

Sand with Clay and Gravel (SP-SC)

= 0.1 5 (OK )

D50Fill
= 32.5 > 25 ( not satisfied)
D50 Subgrade

The calculation for Aggregate Option 1 indicates that the filter gradation requirements are not
satisfied and there is a potential for migration of fine particles from the silty soil subgrade
into the aggregate layer. If this material is selected, a layer of filter fabric meeting the
requirements of the AASHTO M288 specification is recommended for separation.
Aggregate Option 2:
D15Fill
D85 Subgrade
D50Fill
D50 Subgrade

Sand with Gravel (SP)

= 0.1 5 (OK )

= 19.0 < 25 ( OK )

The calculation for Aggregate Option 2 indicates that the filter gradation requirements are
satisfied and migration of fine particles into the aggregate layer will not occur.
SELECTED AGGREGATE
Based on the above analysis, Aggregate Option 2, the sand with gravel (SP) is selected.
Additional measures for subgrade separation are not required.
STEPS 9 and 10: SPECIFY GEOGRID PROPERTIES.
See Table 5-5 and Section 5.9-2.
FHWA NHI-07-092
Geosynthetics Engineering

5-39

Roadways and Pavements


August 2008

STEP 11: SPECIFY CONSTRUCTION REQUIREMENTS.


(See Section 5.8.)

DESIGN EXAMPLE 2: MODIFIED STEWART ET AL. (1977) BASED ON USCOE (2003)


Step 5 9 of the example in Section 5.4-1 will be reworked for a geogrid-reinforced alternative using
the modified Stewart et al. (1977) based on USCOE (2003) for the given set of design parameters.
STEPS 1 4: DESIGN INPUT
Subgrade Soil :
Field CBR = 1; c = 600 psf (30 kPa)
Traffic Load:
Axle load = 18 kips (80 kN)
Tire pressure = 80 psi (550 kPa)
Number of axle passes = 5000
Failure Criteria :
Maximum rut depth = 3 in. (75 mm)
Geosynthetic Reinforcement:
Stiff Biaxial Geogrid
DESIGN CALCULATIONS
STEP 5: OBTAIN BEARING CAPACITY FACTOR
Bearing capacity factors (From USCOE,2003):
Nc = 2.8 for unreinforced road section
Nc = 5.8 for geogrid-reinforced road section
STEP 6A: REQUIRED AGGREGATE THICKNESS FOR UNREINFORCED ROAD SECTION
Using Figure 5-6 (see below), the required aggregate thickness is as follows:
cNc = 600 psf x 2.8 = 1,680 psf = 11.7 psi
tunreinf = 19 in. (475 mm)
STEP 6B: REQUIRED AGGREGATE THICKNESS FOR GEOGRID-REINFORCED ROAD
SECTION
Using Figure 5-6 (see below), the required aggregate thickness is as follows:
cNc = 600 psf x 5.8 = 3,480 psf = 24.1 psi
tgeogrid-reinf = 12 inches (300 mm)

FHWA NHI-07-092
Geosynthetics Engineering

5-40

Roadways and Pavements


August 2008

Figure 5-6 (redrawn) Aggregate-surfaced pavement design curves for single-wheel roads
(after Figure 4, USCOE 2003)

STEP 7: SELECT DESIGN THICKNESS


Use 12 in. (300 mm) and a layer of geogrid placed on top of the subgrade.
STEP 8: CHECK SUBGRADE SEPARATION.
The initial design assumed that a geotextile be used beneath the geogrid as a separator. However,
upon detailed examination it was found that for the selected aggregate and subgrade soils, the
filter criteria are satisfied and migration of fine particles is not anticipated. Therefore, a
geotextile separator is not required. (To check subgrade separation refer to the analysis in Part II
of Example 1, and to Section 5.3-4)
STEP 9 & 10: SPECIFY GEOGRID PROPERTIES.
See Table 5-5 and Section 5.9-2.
STEP 11: SPECIFY CONSTRUCTION REQUIREMENTS.
See Section 5.8.

FHWA NHI-07-092
Geosynthetics Engineering

5-41

Roadways and Pavements


August 2008

5.6 DESIGN GUIDELINES FOR USE OF GEOTEXTILES IN PERMANENT


PAVED ROADWAYS
As indicated in Section 5-1 and 5-2, geotextiles can be used to improve the performance of
permanent roads:
as separation layers,

in conjunction with gravel as a form of mechanical stabilization of the subgrade


and/or the base course layers, and

to improve drainage.

In this section, design guidance is provided for each of the these application areas.
5.6-1 Separation
Separation design is straight forward.
STEP 1. Assess the need for a geotextile separator.
Determine if subgrade conditions warrant the use of a geotextile separator. Referring to
Sections 5-1 and 5-2.
STEP 2. Determine geotextile survivability and opening requirements.
Check the geotextile strength requirements for survivability as discussed in Section 5.3
and listed in Table 5-4.
The maximum opening and minimum permittivity requirements are also listed in Table 54 and should be evaluated with respect to filtration requirements when geotextile is used
as separation for open graded or otherwise free draining base layers as follows:
AOS < D85 subgrade
AOS < 1.8 D85 subgrade
kgeotextile > ksoil

(Wovens)
(Nonwovens)

(Eq. 2-3)
(Eq. 2-4)
(Eq. 2-7a)

STEP 3. Specify geotextiles that meet or exceed those survivability criteria.


STEP 4. Follow the construction recommendations in Section 5.8.
5.6-2 Stabilization
The recommended design method for using geotextiles for permanent pavements is that
developed by Christopher and Holtz (1985; 1991). It is based on the following concepts:
1. Standard methods are used to design the pavement system (i.e., AASHTO, CBR, Rvalue, resilient modulus, etc.).
2. The geotextile is assumed to provide no structural support, therefore, no reduction is
allowed in aggregate thickness required for structural support.

FHWA NHI-07-092
Geosynthetics Engineering

5-42

Roadways and Pavements


August 2008

3. Aggregate savings is achieved through a reduction in the stabilization aggregate


required for construction and possibly through improved subgrade conditions.
4. The recommended method is used to design the first construction lift, which is called
the stabilization lift since it sufficiently stabilizes the subgrade to allow access by
normal construction equipment.
5. Once the stabilizer lift is completed, construction proceeds using the stabilized lift as
the roadbed surface layer.
The design of the geotextile for stabilization is completed using the design-by-function
approach in conjunction with AASHTO M288 in the steps outlined below. The conventional
design method assumes that the stabilizer lift is an unpaved road which will be exposed to
relatively few vehicle passes (i.e., construction equipment only) and which can tolerate 2 in
to 3 in. (50 to 75 mm) of rutting under the equipment loads. A key feature of this method is
the assumption that the structural pavement design is not modified at all in the procedure.
The pavement design proceeds exactly according to standard procedures as if the geotextile
was not present. The geotextile instead replaces additional unbound material that might be
placed to support construction operations, and replaces no part of the pavement section itself.
However, this unbound layer will provide some additional support. If the soil has a CBR of
less than 3 and the aggregate thickness is determined based on a low rutting criteria in the
following steps (i.e., rutting < 2 in. {50 mm} using an Nc = 5.0), the support for the
composite system is theoretically equivalent to a CBR = 3 (resilient modulus 4500 psi {30
MPa}). The equivalent CBR = 3 is based on a conservative soil strength value in Figures 56, 5-7 and 5-8 where low rutting would be anticipated with a geotextile and no gravel. As
with thick aggregate fill used for stabilization, the support value should be confirmed though
field testing using for example, a plate load test or FWD to verify that a minimum composite
subgrade modulus has been achieved. Note that the FHWA procedure is controlled by soil
CBR as measured using ASTM D 4429 Bearing Ratio of Soils In-Place, which should be
performed on the wettest, weakest soil condition anticipated during construction, or a
saturated soaked laboratory CBR test (ASTM D1883) for predicting the performance of soils
that will likely be wet during construction.
The design consists of the following steps:
STEP 1. Assess the subgrade conditions.
Identify properties of the subgrade, including CBR, depth of groundwater table,
AASHTO and/or USCS classification, and sensitivity.
STEP 2. Assess need for geotextile.
Estimate the need for a geotextile based on the subgrade strength (CBR 3) and by
past performance in similar types of soils.
FHWA NHI-07-092
Geosynthetics Engineering

5-43

Roadways and Pavements


August 2008

STEP 3. Design pavement without a geotextile.


Design the roadway for structural support using normal pavement design methods;
provide no allowance for the geotextile.
STEP 4. Determine need for additional aggregate.
Determine the need for additional imported aggregate to ameliorate mixing at the
base/subgrade interface due to susceptibility of soils to pumping and base course
intrusion. Use local practice or see Figure 5-9 to determine if additional aggregate above
that required for structural support is needed. If so, determine its thickness, t1, and reduce
the thickness by 50% and include a geotextile at the base/subgrade interface.

Figure 5-9.

Aggregate loss to weak subgrades (Christopher and Holtz, 1989).

STEP 5. Determine aggregate depth required to support construction equipment.


Determine the additional aggregate thickness t2 required for stabilization of the subgrade
during construction activities from Figures 5-7 or 5-8. Select Nc based on allowable
subgrade ruts for the anticipated construction equipment. Refer to the procedures
outlined in Section 5.4, where:
Nc = 5 for small rutting (< 2 in. {< 50 mm}),
Nc = 5.5 for moderate rutting (2 to 4 in. {50 to 100 mm}), and
Nc = 6 for large rutting (> 4 in. {> 150 mm}).
(For comparison, without a geotextile: Nc = 2.8, 3.0, or 3.3, respectively, for small to
large ruts.)
Alternatively, local policies or charts may be used.
FHWA NHI-07-092
Geosynthetics Engineering

5-44

Roadways and Pavements


August 2008

Note: As previously indicated, if a lower rutting criteria of < 2 in. (50 mm) is used, there
is a potential to improve subgrade support conditions to an equivalent CBR of 3
((resilient modulus of 4500 psi {30 MPa}) for pavement design, provided field tests (e.g.,
Falling Weight Deflectometer, FWD, tests on a trial section) are performed to confirm
the improved support conditions.
STEP 6. Compare thicknesses.
Compare the aggregate-geotextile system thicknesses determined in Steps 4 and 5. Select
the greater of t2 (from Step 5) or 50% t1 (from Step 4).
STEP 7. Check geotextile filtration requirements.
Check the geotextile filtration characteristics using the gradation and permeability of the
subgrade, the water table conditions, and the retention and permeability criteria. From
Chapter 2, that criteria is:
AOS < D85
AOS < 1.8 D85
kgeotextile > ksoil
> 0.1 sec-1

(Wovens)
(Nonwovens)

(Eq. 2-3)
(Eq. 2-4)
(Eq. 2-7a)
(Eq. 2-8c)

STEP 8. Determine geotextile survivability requirements.


Check the geotextile strength requirements for survivability as discussed in Section 5.3-6.
STEP 9. Specify geotextiles that meet or exceed those survivability criteria.
STEP 10. Follow the construction recommendations in Section 5.8.
5.6-3

Permanent Road Subgrade Stabilization Design Example (Christopher and


Holtz, 1991)

DEFINITION OF DESIGN EXAMPLE

Project Description: New public street and service drive for a suburban Washington, D.C.,
townhouse development. State of Virginia DOT regulations apply.
Type of Structure:
Category IV street (permanent road)
Type of Application: stabilization with geotextiles
Alternatives:
i) excavate unsuitable material and increase subgrade aggregate
thickness; or
ii.) geotextile between aggregate and subgrade

FHWA NHI-07-092
Geosynthetics Engineering

5-45

Roadways and Pavements


August 2008

GIVEN DATA
subgrade

traffic

geotextile -

surficial soils: micaceous silts (CBR = 2)


local areas of very poor soils)
low-lying topography
poor drainage
other nearby streets and roads require frequent maintenance
maximum 300 vehicles per day
96% passenger, 5% single-axle, 1 multiaxle
equivalent daily 18,000 lbf (80 kN) single-axle load (EAL) applications = 10
a geotextile was used for this project

REQUIRED
Design the pavement section.
Consider: 1) standard AASHTO design; and

2) alternate with geotextile

DEFINE
A. Geotextile function(s):
B. Geotextile properties required:
C. Geotextile specification:
SOLUTION
A. Geotextile function(s):
Primary
- separation and filtration
Secondary - reinforcement
B. Geotextile properties required:
survivability
apparent opening size (AOS)
permeability
DESIGN

Design pavement with and without geotextile inclusion. Compare options.

STEP 1.

ESTIMATE NEED FOR GEOTEXTILE

Ideal conditions for considering a geotextile; e.g., low CBR, saturated subgrade, and poor
performance history with conventional design.

FHWA NHI-07-092
Geosynthetics Engineering

5-46

Roadways and Pavements


August 2008

STEP 2.

DESIGN WITHOUT GEOTEXTILE

The structural design for the pavement section was based on the AASHTO Guide for Design of
Pavement Structures (1972) using an equivalent design structural number for the anticipated
loading and soil support conditions. (NOTE: this case history used the AASHTO guide, 1972,
which was current at time of design. Today we would use AASHTO guide, 1993.)
For the traffic - as given, determine structural number, SN:
from AASHTO design charts for 20 years, CBR = 2, EAL = 10, and Regional Factor = 2
SN is equal to 2.9
Compute pavement thickness for structural support on a stable subgrade (i.e., no fines pumped
into aggregate subbase and no aggregate loss down into the subgrade):
Assume 2.5 inches asphaltic concrete surface and 8 inches aggregate base course

SN =
2.9 =

surface
a1 D1
0.4 x 2.5 in.

+
+
+

base
+
a2 D2
+
0.14 x 8 in. +

subbase
a3 D3
0.13 x D3

Therefore, D3 = 6 inches (150 mm) required for subbase.


Structural design:

STEP 3.

2.5 in. (65 mm) asphaltic concrete


8 in. (200 mm) aggregate base
6 in. (150 mm) aggregate subbase

ADDITIONAL AGGREGATE FOR PUMPING AND INTRUSION

By local experience in this area, and for subgrades of CBR 2, an additional 8 in. (200 mm) of
aggregate subbase is required (stabilization aggregate).
For the geotextile stabilization alternate, this entire stabilization layer could be eliminated.
However, some very poor soils are anticipated, and some conservatism can be applied.
Therefore, reduce subbase aggregate thickness to 4 inches (100 mm) with use of a geotextile
separator.
Total design base/subbase thickness without geotextile = 8 in. + 6 in. + 8 in. = 22 in. (560 mm)
Total trial base/subbase thickness with geotextile
= 8 in. + 6 in. + 4 in. = 18 in. (460 mm)

FHWA NHI-07-092
Geosynthetics Engineering

5-47

Roadways and Pavements


August 2008

STEP 4.

DESIGN FOR CONSTRUCTABILITY USING A GEOTEXTILE

Use temporary road design procedures.


Assume:
CBR = 2
loaded dump trucks
< 100 passes
2 in. (50 mm) rut depth acceptable
Use Figure 5-7 (use 40 kN load) and
c
= 600 psf (30 kPa) x CBR
= 1200 psf (60 kPa)
Nc = 6
cNc = 7500 psf (360 kPa)
aggregate depth
= 4 in. (100 mm)
STEP 5.

COMPARE THE THICKNESSES DETERMINED IN STEPS 3 AND 4

The two thicknesses are equal; therefore, use 4 in. (100 mm) of stabilization aggregate with a
geotextile separator.
Note that the minimum recommended thickness of aggregate for a construction over a geotextile
is 6 in. (150 mm), therefore the contractor should be required to place an initial subbase lift
thickness of 6 in. (150 mm) or greater (i.e., up to the required 10 in. {250 mm} subbase thickness
depending on the minimum compaction lift thickness requirements).
STEP 6.

CHECK GEOTEXTILE FILTRATION CHARACTERISTICS

Silt type soils have > 50% passing the No. 200 (0.075 mm) sieve.
-1

Therefore, per Table 5-3 use AOS < 0.3 mm and permittivity 0.1 sec .
The permeability of geotextile must be > the soil permeability per Table 5-3. The estimate silt
type soil permeability is ~10-5 cm/sec, therefore, a geotextile permeability > 10-5 cm/sec is
required; however, all geotextiles will meet this low permeability value and permittivity controls.
STEP 7.

DETERMINE GEOTEXTILE SURVIVABILITY REQUIREMENTS

Use Table 5-2, with CBR = 2, dump truck contact pressure > 7500 psf (350 kPa), and 6 in. (150
mm) cover thickness (note that 4 in. (100 mm) cover is limited to existing road bases, therefore 6
in. (150 mm) minimum compacted lift thickness is recommended), and determine that a
geotextile with a HIGH, or Class 1, survivability rating is required.
FHWA NHI-07-092
Geosynthetics Engineering

5-48

Roadways and Pavements


August 2008

STEP 8.

SPECIFY GEOTEXTILE PROPERTY REQUIREMENTS

From Table 5-3; geotextile shall meet or exceed the minimum average roll values, with
elongation at failure determined with the ASTM D 4632 test method, of:

Property
Grab Strength
Sewn Seam Strength
Tear Resistance
Puncture Strength
Ultraviolet Stability

ASTM
Test Method
D 4632
D 4632
D 4533
D 6241
D 4355

Elongation
< 50%

Elongation
> 50%

315 lb (1400 N)
200 lb (900 N)
270 lb (1200 N)
180 lb (810 N)
110 lb ( 500N)
80 lb (350 N)
620 lb (2750N)
433 lb (1925 N)
50% strength retained after 500 hours

The geotextile shall have an AOS < 0.3 mm and > 0.1 sec-1.
STEP 9.
SPECIFY CONSTRUCTION REQUIREMENTS
See Section 5.8

5.6-4 Improved Drainage


As reviewed in Chapter 2, geotextiles and geocomposites are often incorporated into the
roadway drainage system. There are three distinct applications of geosynthetics in drainage
systems:
1. Geotextiles as filters for the edge drains, underdrains, and free draining base
2. Geocomposites (placed vertical) as edge drains
3. Geocomposites (placed horizontal beneath the base/subbase or pavement) as a
horizontal base drain layer.
The design requirements for the geotextile in each of these applications were covered in
Chapter 2, however, the pavement design with respect to improved drainage and the design
requirements for the drainage system to meet specific levels of improvement were not
reviewed. The benefits of improved drainage in roadway systems is covered in AASHTO
(1986 and 1993) and NCHRP 1-37A pavement design guides.
From the AASHTO Guide (1993), modified layer coefficients determine the treatment for the
expected drainage level for flexible pavements. The factor for modifying the layer
coefficient is referred to as an mi value, thus the structural number (SN) becomes:

FHWA NHI-07-092
Geosynthetics Engineering

5-49

Roadways and Pavements


August 2008

SN = a1 D1 + a 2 D 2 m 2 + a 3 D 3 m 3
where:
D1,D2,D3

a1,a2,a3
m2,m3

=
=

thicknesses of existing pavement surface, base, and


subbase layers;
corresponding structural layer coefficients; and
drainage coefficients for granular base and subbase.

The recommended mi values are presented in Table 5-7 as a function of the drainage quality
and the percent of time during the year the pavement structure is near saturation. Definitions
of quality of drainage are presented in Table 5-8.
From the AASHTO guide (1993), the drainage coefficient, Cd, in the performance equation
determines the treatment for the expected drainage level for rigid pavements. The
performance equation is used to calculate a design slab thickness for a rigid pavement.
Recommended values for Cd are presented in Table 5-9.
Table 5-7
Recommended mi Values for Modifying Structural Layer Coefficients
of Untreated Base and Subbase Materials in Flexible Pavements
(from AASHTO, 1993)

Percent of Time Pavement Structure is Exposed


to Moisture Levels Approaching Saturation

Quality of
Drainage
Excellent
Good
Fair
Poor
Very Poor

<1%

1 to 5 %

5 to 25 %

> 25 %

1.40 - 1.35
1.35 - 1.25
1.25 - 1.15
1.15 - 1.05
1.05 - 0.95

1.35 - 1.30
1.25 - 1.15
1.15 - 1.05
1.05 - 0.80
0.95 0.75

1.30 - 1.20
1.15 - 1.00
1.00 - 0.80
0.80 - 0.60
0.75 - 0.40

1.20
1.00
0.80
0.60
0.40

Table 5-8
Quality of Pavement Drainage
(from AASHTO, 1993)

Quality of Drainage

Water Removed Within

Excellent
Good
Fair
Poor
Very Poor

2 hours
1 day
1 week
1 month
(water will not drain)

FHWA NHI-07-092
Geosynthetics Engineering

5-50

Roadways and Pavements


August 2008

Table 5-9
Recommended Values of Drainage Coefficient, Cd,
for Rigid Pavement Design
(from AASHTO, 1993)

Quality of
Drainage

Percent of Time Pavement Structure is Exposed


to Moisture Levels Approaching Saturation
<1%

1 to 5 %

5 to 25 %

> 25 %

Excellent

1.25 - 1.20

1.20 - 1.15

1.15 - 1.10

1.10

Good

1.20 - 1.15

1.15 - 1.10

1.10 - 1.00

1.00

Fair

1.15 - 1.10

1.10 - 1.00

1.00 - 0.90

0.90

Poor

1.10 - 1.00

1.00 - 0.90

0.90 - 0.80

0.80

Very Poor

1.00 - 0.90

0.90 - 0.80

0.80 - 0.70

0.70

The NCHRP 1-37A method requires a computer model that is still under development as of
the publication of this manual. However, an initial evaluation of the influence drainage can
be found in NHI Course Number 13204 Geotechnical Aspects of Pavement. Using either the
AASHTO, 1993 or NCHRP 1-37A method, the influence on design can be significant. For
example, in high rainfall areas, the base section of a flexible pavement system (with a
relatively thick base layer) can be reduced in thickness by as much as a factor of two, or the
design life extended by an equivalent amount, if excellent drainage is provided versus poor
drainage. Likewise, an improvement in drainage leads to a reduction in Portland cement
concrete (PCC) slab thickness.
The design of the system components to meet specific level of improvement are well covered
in NHI Course Number 131026 on Pavement Subsurface Drainage Design, NHI Course
Number 13204 Geotechnical Aspects of Pavements, and the associated reference manuals
(ERES, 1999 and Christopher et al., 2006).

FHWA NHI-07-092
Geosynthetics Engineering

5-51

Roadways and Pavements


August 2008

5.7

DESIGN GUIDELINES FOR USE OF GEOGRIDS IN PERMANENT PAVED


ROADWAYS

Geogrid reinforcement is used in permanent paved roadways in two major application areas
base reinforcement and subgrade stabilization. In base reinforcement applications, the
geogrids are placed within or at the bottom of unbound layers of a flexible pavement system
and improve the load-carrying capacity of the pavement under repeated traffic. In subgrade
stabilization applications, the geogrids are used to build a construction platform over weak
subgrades (CBR 3) to carry equipment and facilitate the construction of the pavement
system without excessive deformations of the subgrade. The design guidelines for use of
geogrids in subgrade stabilization applications are discussed in Section 5.5. The current
design practice and the recent developments for the use of geogrids in base reinforcement
applications are discussed in this section (Holtz et al., 1998; Berg et al., 2000; AASHTO,
2001; Perkins et al., 2005a; Gabr et al., 2006).
The state of practice for design of geogrid-reinforced base courses in flexible pavements is in
accordance with the AASHTO Guide for Design of Pavement Structures (1993 or earlier
editions) and the AASHTO Provisional Standard PP 46-01 Recommended Practice for
Geosynthetic Reinforcement of the Aggregate Base Course of Flexible Pavement Structures
(2001). The AASHTO (1993) method, and its regional adaptations, is widely used by
pavement designers in various public agencies and private companies. It is empiricallybased and models the flexible pavement as a series of layers which have a combined
structural capacity to carry a certain number of traffic loads (ESALs) with pre-determined
minimum levels of serviceability and statistical confidence. Based on the AASHTO, 1993
design guide, the overall structural contribution of geosynthetic reinforcement is considered
in the design through either of the following factors that are derived from empirical productspecific data:
Traffic Benefit Ratio (TBR) the ratio of the number of load applications necessary to
reach a specific failure state in a geosynthetic-reinforced pavement to the number of load
applications required to reach the same failure state in an unreinforced section (i.e., the
same pavement section but without reinforcement).
Base Course Reduction Factor (BCR) the percent reduction in the thickness of base or
subbase material in a reinforced pavement compared to an unreinforced one, given that
the trafficking capacity for a defined failure state remains the same.
To assess and characterize the appropriate TBR or BCR values, the user is advised to refer to
product-specific studies and test sections that demonstrate the value added by the
geosynthetic reinforcement in pavement structures. The base course reinforcement
FHWA NHI-07-092
Geosynthetics Engineering

5-52

Roadways and Pavements


August 2008

specifications established by the Federal Aviation Administration (FAA, 1994) recommend


evaluation of alternative materials based on full-scale trials, independent certified test results
for performance based geogrid properties, and successful performance demonstrated in
projects of comparable size and application, and verified after a minimum of one year of
service life. As discussed earlier, the lack of such data often limits the use of geogrid
reinforcement in flexible pavements.
A major initiative in the area of pavement design was the AASHTO decision to develop and
subsequently adopt a mechanistic-empirical (M-E) method for design of pavement systems.
The initiative was conceived in the early 1990s and led by AASHTO committees with the
help of research institutions. While the M-E Pavement Design Guide (MEPDG) has recently
been officially adopted by AASHTO (2008), some work is still ongoing (NCHRP, 2008). In
the original scope of the M-E Design Guide, geosynthetics materials have not been included
but the opportunity to utilize the M-E concept to rationalize the design has been recognized
by the geosynthetics community. A series of research projects designed to define and
quantify the benefits of geosynthetics within the M-E design framework had been initiated
and the results to date are promising (Perkins et al., 2004; Perkins et al., 2005b; Kwon et al.,
2005; Al-Qadi et al., 2007; Al-Qadi et al., 2008).
The AASHTO PP 46-01 guidelines for use of geosynthetic reinforcement in aggregate base
courses of flexible pavements, and the recent developments based on the M-E concept are
presented in the subsequent section.
5.7-1 Empirical Design Method from AASHTO PP 46-01 (2001)
AASHTO PP 46-01 (2001) provides guidelines for design of geogrid-reinforced base courses
in flexible pavements. The guidelines are empirical in nature and the design steps follow a
procedure that was initially reported by Berg et al. (2000).
In the AASHTO, 1993 Guide, pavement design is based on a reference to the serviceability
of the pavement system expressed through measurements of roughness and different types of
distress (cracking, rutting, etc.). The load carrying capacity of a pavement is expressed with
the number of equivalent standard axial loads (ESALs) at which the permanent deformation
at the surface reaches specific value (allowable rut depth). The number of equivalent
standard axial loads (ESALs) is calculated using the AASHTO, 1993 equation shown below.
PSI
log10

4.2 1.5 + 2.32 log M 8.07


log10 (W18 ) = Z R S o + 9.36 log10 ( SN + 1) 0.20 +
10
R
1094
0.40 +
( SN + 1) 5.19

FHWA NHI-07-092
Geosynthetics Engineering

5-53

(5.7-1)

Roadways and Pavements


August 2008

Where:
W18
ZR

= Allowable trafficking (ESALs)


= Standard Normal Deviate; based on Table 4.1 in Part I of the AASHTO
(1993) guidelines, for a Reliability of 95%, the Standard Normal Deviate,
ZR = -1.282.
So
= Standard Deviation = 0.49
SN
= Structural Number
PSI = Change in present Serviceability Index
MR
= Resilient Modulus of subgrade or base being considered (psi)

Using Equation 5.7-1, the Traffic Benefit Ratio (TBR) can be applied directly to the
calculated number of ESALs or can be used to adjust the structural number. The Base
Course Reduction Factor (BCR) is used to directly reduce the required thickness of the
unreinforced base course. It is recommended that agencies with limited experience with
geosynthetic reinforcement primarily use the reinforcement to improve the service life of
pavement structures, and limit reduction of the structural section until more experience is
gained (Berg et al., 2000).
Design Procedure
The design steps for use of geosynthetic base reinforcement for flexible pavements are
outlined below. For details refer to Berg et al. (2000).
Step 1: Initial assessment of geosynthetic applicability
Requires assessment of: subgrade strength, aggregate thickness required for
unreinforced section, characteristic of base/subbase materials, seasonal variation in
moisture levels, reinforcing mechanisms and value added by geosynthetics.
Step 2: Design of unreinforced pavement section
Using an established method for design of unreinforced pavements, the structural
layers, the type of material, and the thicknesses are determined for a pavement section
without geosynthetics.
Step 3: Investigate potential benefits of using geosynthetics reinforcement
Requires review of available data to define potential and target benefits for the
specific project. The conditions for which various geosynthetic products should be
considered for this application are summarized in Table 5-10.
Step 4: Define reinforcement benefits in terms of Traffic Benefit Ratio (TBR) or Base Course
Reduction Factor (BCR)
Requires review of successful applications, field studies, and lab test results.
FHWA NHI-07-092
Geosynthetics Engineering

5-54

Roadways and Pavements


August 2008

Table 5-10
Qualitative Review of Reinforcement Application Potential for Paved Permanent Roads
(after Berg et al., 2000).
Roadway Design
Conditions
Subgrade

Base /
Subbase
Thickness1
(mm)

Geosynthetic Type
Geotextile
Nonwoven

Woven

Geogrid 2
Knitted
Extruded
or
Woven

GG-GT Composite
Open
Well Graded
Graded
Base
3
Base

Soft
(CBR < 3)
(MR <30 MPa)

150 300

> 300

Firm - Very Stiff


(3 CBR 8)

150 300

> 300

(30 MR 80)
KEY: usually applicable
applicable for some conditions
usually not applicable
insufficient information at this time see note
NOTES: 1. Total base or subbase thickness with geosynthetic reinforcement. Reinforcement may be placed at
bottom of base or subbase, or within base for thicker (usually > 12 in {300 mm})) thicknesses.
Thicknesses less than 6 in. (150 mm) not recommended for construction over soft subgrades.
Placement of less than 6 in. (150 mm) over a geosynthetic not recommended.
2. For open graded base or thin bases over wet, fine grained subgrades, a separation geotextile should be
considered with geogrid reinforcement.
3. Potential assumes base placed directly on subgrade. A subbase also may provide filtration.
Reinforcement usually applicable, but typically addressed as subgrade stabilization.
Geotextile component of composite likely is not required for filtration with a well graded base course,
therefore, composite reinforcement usually not applicable.

Step 5: Design of reinforced pavement section


The use of the Traffic Benefit Ratio (TBR) for design of reinforced pavement section
based on AASHTO, 1993 Guide is illustrated in the design example, Section 5.7-4.
Step 6: Cost-benefit analysis
Evaluation of benefits in terms of reduction in initial construction cost or reduction in
life cycle costs can be used to justify the use of geosynthetics depending on the
priorities of the user.
Step 7 - 8: Develop specifications, bid documents and construction drawings
Step 9: Monitor construction and document performance
The AASHTO PP 46-01 document and Berg et al., 2000 provide guidelines and represent the
state of practice for design of geogrid-reinforced base courses in flexible pavements.
FHWA NHI-07-092
Geosynthetics Engineering

5-55

Roadways and Pavements


August 2008

5.7-2 Mechanistic-Empirical Approach for Pavement Design


As mentioned above, the initiative to adopt a mechanistic-empirical (M-E) method for design
of pavement systems has been gaining momentum over the last decades, and the concept was
included in the AASHTO, 1993, Pavement Design Guide. Since then the NCHRP Project 137A Mechanistic-Empirical Pavement Design Guide (MEPDG) (AASHTO, 2008) has been
completed along with the supporting software (MP-PDG, Version 1.00,
www.trb.org/mepdg/software.htm). The MEPDG approach is shown in Figure 5-10.

(a)

New Features

Shear Interaction
Reinforcement Layer

Existing Features

Asphalt Concrete
Unbound Aggregate

Material
Models

Subgrade

Accompanying
material models

82
(b)
Figure 5-10: Mechanistic-Empirical (M-E) Pavement Design Method showing: a) M-E
concept (FHWA/DGIT), and b) modified response model for inclusion of reinforcement.

FHWA NHI-07-092
Geosynthetics Engineering

5-56

Roadways and Pavements


August 2008

The major steps of the MEPDG are (AASHTO, 2008):


Selection of pavement structure (layers, type of materials, thicknesses);
Characterization of climate, traffic, and materials for the specific project location;
Analysis of the mechanistic model of the pavement structure
Calculation of critical responses (stresses, strains)
Evaluation of the accumulated damage and associated distress with reference to preset
criteria.
The design may require several iterations considering different pavement structures.
Design is completed when for a specific section the distress levels do not exceed the
acceptable levels for the design life of the structure.
Important components of the MEPDG method are (1) a mechanistic model to calculate the
critical responses of the system, and (2) empirical performance or damage models that relate
the critical responses to the accumulated damage and distress levels. The development of a
mechanistic response model based on the finite element method that account for the
influence of reinforcement on the surrounding base aggregate has been the focus of several
studies.
Recently FHWA sponsored a study at Montana State University in cooperation with the
Finnish, Norwegian and Swedish National Road Administrations to develop an interface for
including geosynthetic base reinforcement in mechanistic empirical design, consistent with
the MEPDG model (Perkins et al., 2004). The mechanistic model developed from this study
consists of a compaction module and three subsequent trafficking response modules to
describe the effect of geogrid reinforcement during the initial construction operations and inservice traffic loading. An empirical damage model was first developed for the permanent
deformation accumulation of the unbound aggregate layer within a zone influenced by the
geogrid reinforcement. The compaction module treated the developed lateral stress due to
geogrid reinforcement as the stress developed in the presence of the permanent interface
shear stress (Perkins and Svano, 2004). Then, the finite element response models were
utilized at several damage steps to account for the effect of increasing lateral confinement
with increasing traffic load repetitions. The combined use of these response models with the
empirical damage model resulted in a rational method for showing the benefit of geogrid
reinforcement and evaluating the performance of reinforced pavements. Important feature of
the developed mechanistic model is that it has been partially calibrated using available full
scale test data and is based on laboratory test methods that are under development by ASTM
(Perkins et al., 2005).
FHWA NHI-07-092
Geosynthetics Engineering

5-57

Roadways and Pavements


August 2008

Another mechanistic response model that accounts for the confinement effects in geogridreinforced flexible pavements was developed by a research team at the University of Illinois
(Kwon et al., 2005 and 2007, Al-Qadi et al., 2008). The effects of interlock between the
aggregate and geogrid were simulated by considering locked-in horizontal residual stresses in
the granular base as initial stresses in the pavement response analysis. Increased confinement
resulted in this stiffening effect above and below the geogrid reinforcement and the predicted
modulus values were shown to increase significantly especially around the geogrid. The
developed mechanistic model can be used as main structural analysis model for mechanistic
based design of base reinforcement as well as for FWD back calculation. The developed
mechanistic model has been partially calibrated using available test data (Al-Qadi et al., 2007
and 2008).
The distress models are the other major component of the M-E based solution and they relate
the critical responses to the accumulated damage and distress levels. The development of
distress models requires instrumented full-scale sections that are trafficked until failure.
Perkins et al., (2005) reported that several research groups have conducted studies and
significant progress has been made in some areas; however, the calibration of performance
models requires significant investments in order to advance the implementation of the M-E
design procedure in base course reinforcement applications.
5.7-3 Design Example for Geogrid Reinforced Paved Roadway
DEFINITION OF DESIGN EXAMPLE
Project Description:
A 30-mile (48 km) section of a two-lane state highway is scheduled for reconstruction. The
initial traffic estimates indicated 1,000,000 equivalent single axle loads (ESALs) over the 20 year
performance period based on a 2% growth factor. The geotechnical evaluation indicates that the
subgrade soil is low plastic clay (CL) with CBR= 4, Mr = 6000 psi (40 MPa).
Due to budget constraints, high cost of local aggregate and environmental concerns, the owner is
considering the use of geogrid reinforcement to either reduce the thickness of the base course or
extend the performance period of the pavement structure. The type of application proposed for
consideration is geogrid base reinforcement. A 1-mile (1.6 km) section is scheduled for
reconstruction to help establish the design criteria for the future reconstruction of the entire
roadway section.

Design Alternatives:
(I) Unreinforced conventional roadway section with a 20 year design life
(II) Geogrid reinforcement in the base layer to extend the performance period.
(III) Geogrid reinforcement in the base layer to reduce the initial cost of construction by
reducing the required base course thickness (20 year design life).

FHWA NHI-07-092
Geosynthetics Engineering

5-58

Roadways and Pavements


August 2008

REQUIRED
Design of flexible pavement section for the three alternatives based on AASHTO, 1993 Design Guide
and AASHTO PP 46-01 Standard of Practice, Geosynthetic Reinforcement of the Aggregate Base
Course of Flexible Pavement Structures.
DESIGN
STEP 1. INITIAL ASSESSMENT OF BASE REINFORCEMENT APPLICABILITY.
A. In-situ strength of subgrade soil: Low plastic clay (CL)
CBR= 4, Mr = 6000 psi (40 MPa), moderate strength
B. Typical thickness of unreinforced base course for local conditions
C. Estimated thickness of geogrid reinforced base course for comparable conditions.
D. Evaluation of separation and drainage conditions
E. Upon detailed examination it was found that for aggregates from local sources and subgrade
soils, the filter criteria are satisfied and migration of fine particles is unlikely. Therefore, a
geotextile separator is not required. (For subgrade separation design refer to the analysis in
Part II of Example 1 in Section 5.5 and to Section 5.3-4). For the local conditions and free
draining aggregated base, the drainage coefficient is m = 1.2.
F. From Table 5-1 and Table 5-10, for a CBR of 4, extruded geogrid reinforcement is selected
for this application.
STEP 2. DESIGN OF THE UNREINFORCED PAVEMENT SECTION ALTERNATIVE I.
The unreinforced pavement design cross section will be evaluated using the equation from the
AASHTO, 1993 design guide.
PSI
log 10

4 .2 1 .5 + 2 .32 log M 8 .07


log 10 (W18 ) = Z R S o + 9 .36 log 10 ( SN + 1) 0 .20 +
10
R
1094
0 .40 +
( SN + 1) 5 .19

Where:
W18 = Allowable trafficking (ESALs)
ZR = Standard Normal Deviate; based on Table 4.1 in Part I of the AASHTO
(1993) guidelines, for a Reliability of 95%, the Standard Normal Deviate, ZR = -1.282.
So
= Standard Deviation = 0.49
SN = Structural Number
PSI = Change in Present Serviceability Index = 4.2 2.0 = 2.2
MR = Resilient Modulus of subgrade or base being considered (psi) = 6000 psi
FHWA NHI-07-092
Geosynthetics Engineering

5-59

Roadways and Pavements


August 2008

The pertinent data and the corresponding structural numbers for the unreinforced pavement
section are as follows:
Layer
Drainage
Structural
Thickness,
Layer
Material Type
coefficient, coefficient,
number
t (in.)
ai
mi
(t x ai x mi)
Asphalt wearing
1.0
0.42
N/A
0.42
1
course
Asphalt binder
2.5
0.40
N/A
1.0
2
Aggregate base
11.0
0.14
1.2
1.85
3
course
Subbase course
6.0
0.10
0.7
0.42
4
3.69
Overall structural number:
The calculated traffic for the unreinforced section based on the AASHTO, 1993 equation is
1,100,000 ESALs, which meets the design traffic of 1,000,000 ESALs and the unreinforced
section meets the design requirements.
STEP 3. Qualitative benefits of geogrid reinforcement.
Two main benefits of the geogrid reinforcement will be considered:
extended life of the pavement (i.e., additional vehicle passes), and
reduced base aggregate thickness (i.e., reduced undercut, aggregate quantities and initial
construction cost)
STEP 4. Geogrid reinforcement benefits in terms of Traffic Benefit Ratio (TBR)
Webster (1993) reported TBR data from instrumented sections of full-scale field test for 30-kips
single wheel load, 68 psi tire pressure, and traffic wandering. For Type 2 extruded stiff biaxial
geogrid the reported data are:
TBR = 4.7 for CBR = 3, Base thickness = 14 in. (360 mm), Asphalt thickness = 2.4 in. (60 mm)
TBR = 6.7 for CBR = 8, Base = 10 in. (250 mm), Asphalt thickness = 2.4 in. (60 mm)
The research data were evaluated with regard to the project conditions and the performance of
existing geogrid-reinforced projects, and the following geogrid reinforcement was selected for the
project:
Extruded biaxial geogrid (punched & drawn sheet) with aperture stability modulus of 0.65 mN/deg (Kinney, 2000)
Traffic Benefit Ratio (TBR) based on 1 in. rut depth equal to 4.0.
STEP 5. Design of the geogrid reinforced pavement.
Alternative II: Extend the performance period by using geogrid reinforcement

FHWA NHI-07-092
Geosynthetics Engineering

5-60

Roadways and Pavements


August 2008

From Step 2, the calculated traffic for the unreinforced section is:
T I, unreinf = 1,100,000 ESALs.
For TBR = 4.0, the calculated traffic for the geogrid reinforced section will be:
T II, reinf = 1,100,000 x 4.0 = 4,400,000 ESALs.
The trafficking calculated for the geogrid-reinforced section is based on the limitation of the
surface rutting to 1.0 in. On the basis of the TBR, a road with the Alternative 1 design with the
geogrid could easily have a 40 years or greater design life. The deterioration of the asphalt
surface due to environmental factors has to be addressed in the maintenance program, which will
be evaluated in the cost analysis in Step 6.
Alternative III: Reduce the required base course thickness by using geogrid reinforcement as
follows.
The equivalent geogrid structural number is calculated based on a T II, reinf = 4,400,000 ESALs,
which provides a structural number of 4.45 Based on the AASHTO, 1993 equation. Thus, an
equivalent SN 0.7 is estimated for the geogrid. This value must be confirmed through a field
evaluation program and the 1 mile initial test section affords the opportunity to do so.

Layer

Material Type

1
2
3
4
5

Asphalt wearing course


Asphalt binder
Aggregate base course
Geogrid*
Subbase course

Layer
Drainage
coefficient, coefficient,
ai
mi
1.0
0.42
N/A
2.5
0.40
N/A
7.0
0.14
1.2
Equivalent SN for TBR = 4
6.0
0.10
0.7
Overall structural number:

Thickness,
t (in)

Structural
number
(t x ai x mi)
0.42
1.0
1.18
0.7
0.42
3.72

The calculated traffic for the geogrid-reinforced section for Alternative III, based on the
AASHTO, 1993 equation again is 1,100,000 ESALs.
The calculated traffic of 1,100,000 ESALs exceeds the design traffic of 1,000,000 ESALs and the
calculated traffic of the unreinforced section and meets the design requirements. The geogrid
reinforcement reduced the thickness of the base course by 4.0 in. and increased the allowable
traffic capacity with approximately 10%.
STEP 6. Cost-Benefit Analysis.
INITIAL CONSTRUCTION COSTS
The comparison of the initial construction costs for Alternative I (unreinforced road) and
Alternative II and III (geogrid-reinforced road) is done for the following cost of materials based
on local sources:
FHWA NHI-07-092
Geosynthetics Engineering

5-61

Roadways and Pavements


August 2008

Layer

Cost*
($/ton)
55.00
55.00
22.00
12.00

Material Type

Asphalt wearing course


Asphalt binder
Aggregate base course
Subbase course
Biaxial Geogrid
5
(incl. installation cost)
* Average cost in 2008 from a Southeast DOT
1
2
3
4

Cost
($/cy)
107.70
107.70
35.60
21.80
$4.25 /SY

Summary of initial construction costs for 1-mile of road for Alternatives I, II and III.
Alternative I:
Unreinforced
$260,400
$291,700
$69,500
$0.0
$0.0
$621,600
$42.40/SY

Expenses
Asphalt
Aggregate base
Subbase
Geogrid
Undercut/Fill
Total Costs
Unit Costs

Alternative II:
Geogrid-reinforced
$260,400
$291,700
$69,500
$67,100
$0.0
$688,700
$47.00/SY

Alternative III:
Geogrid-reinforced
$260,400
$185,600
$69,500
$67,100
$0.0
$582,600
$39.70/SY

The analysis of the initial construction costs indicate that the geogrid-reinforced alternative (III)
leads to overall savings of 6.4 % relative to the unreinforced one (I).
LONG-TERM COSTS
The benefits of extended design life of Alternative II vs. Alternative I and III can be evaluated by
life-cycle cost analyses for the unreinforced and reinforced pavement using the FHWA program
RealCost available at http://www.fhwa.dot.gov/infrastructure/asstmgmt/lccasoft.cfm or other
software.
Parameters Used in Life-Cycle Cost Examples
Parameter

Value

Initial Serviceability

4.2

Terminal Serviceability

Reliability Level

95

Overall Standard Deviation

0.49

Subgrade Resilient Modulus

6000 psi

Structural Design Number


Maintenance Annual cost
initiates 5 yrs after construction or rehabilitation

3.72
$160/lane km*

Discount Rate

3.50

Evaluation Method

NPV

Salvage Value
FHWA NHI-07-092
Geosynthetics Engineering

5-62

Roadways and Pavements


August 2008

Summary of Pavement Design for Life-Cycle Cost Analyses


ESAL/Analysis Period

2,200,000 / 40 years

Design Option

Alternative I
Unreinforced

Alternative II
Performance Period
Extension
W/ geosynthetic

Alternative III
Reinforced w/ reduced
section

1,000,000

4,000,000

1,000,000

20 w/ 10 yr repair

40 w/ 20 yr repair

20 w/ 10 yr repair

Pavement Option
ACC Surface
ACC Binder
Base Course
Subbase Course

1 in.
2.5 in.
11 in.
6 in.

1 in.
2.5 in.
11 in.
6 in.

1 in.
2.5 in.
7 in.
6 in

Geosyn. Reinforcement
In-Place Cost
TBR Value

none
n/a
n/a

yes
4.25
4

yes
4.25

1,000,000

4,000,000

1,000,000

Performance Period (yrs)

20 w/ 10 yr repair

40 w/ 20 yr repair

20 w/ 10 yr repair

Initial Construction Cost


($/mile)

$621,600

$688,700

$582,600

868,800

777,800

830,000

10.5 %

4.5 %

Design ESAL
Performance Period (yrs)

Design ESAL

Total Life-cycle Cost


($/mile)
Percent Savings Compared to
Unreinforced Design

Note: a. In todays dollars.

Alternative III, geogrid-reinforced pavement section for 1,000,000 ESALs, 20 year design life, is
selected for the construction of the first 1-mile section. Its performance will be documented for
the purposes of establishing the design criteria for the future reconstruction of the entire roadway
section. In addition, the data will be used to develop a database of projects and respective
materials properties (aggregate, subgrade, etc.), for future use in assessing long-term
reinforcement benefits in regard to soil and environmental conditions, materials, traffic loads, and
key properties of geosynthetics reinforcement.
STEP 7. Development of a project specification.
See Section 5.9-4 for performance property requirements and Table 5-5 for survivability
requirements.
STEP 8. Development of construction drawings and bid documents.
STEP 9. Construction of the roadway.
FHWA NHI-07-092
Geosynthetics Engineering

5-63

Roadways and Pavements


August 2008

5.8 INSTALLATION PROCEDURES


5.8-1 Roll Placement
Successful use of geosynthetics in pavements requires proper installation, and Figure 5-11
shows the proper sequence of construction. Even though the installation techniques appear
fairly simple, most geosynthetic problems in roadways occur as the result of improper
construction techniques.
If the geosynthetic is ripped or punctured during construction activities, it will not likely
perform as desired. If either a geotextile or geogrid is placed with a lot of wrinkles or folds,
it will not be in tension, and, therefore, cannot provide a reinforcing effect. Other problems
occur due to insufficient cover over the geotextiles or geogrids, rutting of the subgrade prior
to placing the geosynthetic, and thick lifts that exceed the bearing capacity of the soil. The
following step-by-step procedures should be followed, along with careful observations of all
construction activities.
1.
The site should be cleared, grubbed, and excavated to design grade, stripping all
topsoil, soft soils, or any other unsuitable materials (Figure 5-11a). If moderate site
conditions exist, i.e., CBR greater than 1, lightweight proofrolling operations should
be considered to help locate unsuitable materials. Isolated pockets where additional
excavation is required should be backfilled to promote positive drainage. Optionally,
geotextile wrapped trench drains could be used to drain isolated areas.
2.
During stripping operations, care should be taken not to excessively disturb the
subgrade. This may require the use of lightweight dozers or grade-alls for lowstrength, saturated, noncohesive and low-cohesive soils. For extremely soft ground,
such as peat bog areas, do not excavate surface materials so you may take advantage
of the root mat strength, if it exists. In this case, all vegetation should be cut at the
ground surface. Sawdust or sand can be placed over stumps or roots that extend
above the ground surface to cushion the geotextile or geogrid. Remember, the
subgrade preparation must correspond to the survivability properties of either the
geotextile or geogrid.
3.
Once the subgrade along a particular segment of the road alignment has been
prepared, the geotextile or geogrid should be rolled in line with the placement of the
new roadway aggregate (Figure 5-11b). Field operations can be expedited if the
geotextile or geogrid is manufactured to design widths in the factory so it can be
unrolled in one continuous sheet. Geogrids should be placed directly on top of
geotextiles when used together. The geosynthetic should not be dragged across the
subgrade. The entire roll should be placed and rolled out as smoothly as possible.
Wrinkles and folds in the geotextile or geogid should be removed by stretching and
staking as required.
FHWA NHI-07-092
Geosynthetics Engineering

5-64

Roadways and Pavements


August 2008

a. Prepare the ground by removing stumps,


boulders, etc.; fill in low spots.

PREPARE THE GROUND

c. Back dump aggregate onto previously place


aggregate. Do not drive on the geosynthetic.
Maintain 6 to 12 inches (150 300 mm) cover
between truck tires and geosynthetic.

BACK DUMP AGGREGATE

b. Unroll the geosynthetic directly over the


ground to be stabilized. If more than one roll is
required, overlap rolls. Inspect geosynthetic.

UNROLL THE GEOSYNTHETIC

d. Spread the aggregate over the


geosynthetic to the design thickness.

SPREAD THE AGGREGATE

e. Compact the aggregate using dozer tracks


or smooth drum vibratory roller.

COMPACT THE AGGREGATE

Figure 5-11. Construction sequence using geosynthetics.

FHWA NHI-07-092
Geosynthetics Engineering

5-65

Roadways and Pavements


August 2008

4.
5.

6.

7.

8.

9.

Parallel rolls of geotextiles or geogrids should be overlapped, sewn, or joined as


required. (Specific requirements are given in Sections 5.8-2 and 5.8-3.)
For curves, the geotextile should be folded or cut and overlapped in the direction of
the turn (previous fabric on top) (Figure 5-12). Folds in the geotextile should be
stapled or pinned approximately 2 ft (0.6 m) on centers. Geogrids should be cut and
overlapped in the direction of the turn.
When the geotextile or geogrid intersects an existing pavement area, the geosynthetic
should extend to the edge of the old system. For widening or intersecting existing
roads where geotextiles or geogrids have been used, consider anchoring the geotextile
or geogrid at the roadway edge. Ideally, the edge of the roadway should be excavated
down to the existing geosynthetic and the existing geosynthetic mechanically
connected to the new geosynthetic (i.e., sewn to the new geotextile or mechanically
connected with plastic ties to the geogrid). Overlaps, staples, and pins could also be
utilized.
Before covering, the condition of the geotextile or geogrid should be checked for
excessive damage (i.e., holes, rips, tears, etc.) by an inspector experienced in the use
of these materials. If excessive defects are observed, the section of the geosynthetic
containing the defect should be repaired by placing a new layer of geosynthetic over
the damaged area. The minimum required overlap required for parallel rolls should
extend beyond the defect in all directions. Alternatively, the defective section can be
replaced.
The base aggregate should be end-dumped on the previously placed aggregate (Figure
5-11c). For very soft subgrades, pile heights should be limited to prevent possible
subgrade failure. The maximum placement lift thickness for such soils should not
exceed the design thickness of the road.
The first lift of aggregate should be spread and graded to 12 in. (300 mm), or to the
design thickness if less than 12 in. (300 mm), prior to compaction (Figure 5-11d). At
no time should traffic be allowed on a soft roadway with less than 8 in. (200 mm), (or
6 in. {150 mm} for CBR 3) of aggregate over the geotextile. Equipment can
operate on the roadway without aggregate for geotextile or geocomposite installation
under permeable bases, if the subgrade is of sufficient strength. For extremely soft
soils, lightweight construction vehicles will likely be required for access on the first
lift. Construction vehicles should be limited in size and weight so rutting in the initial
lift is limited to 3 in. (75 mm). If rut depths exceed 3 in. (75 mm), it will be
necessary to decrease the construction vehicle size and/or weight or to increase the
lift thickness. For example, it may be necessary to reduce the size of the dozer
required to blade out the fill or to deliver the fill in half-loaded rather than fully
loaded trucks.

FHWA NHI-07-092
Geosynthetics Engineering

5-66

Roadways and Pavements


August 2008

10.

11.

12.

13.

14.

The first lift of base aggregate should be compacted by tracking with the dozer, then
compacted with a smooth-drum vibratory roller to obtain a minimum compacted
density (Figure 5-11e). For construction of permeable bases, compaction shall meet
specification requirements. For very soft soils, design density should not be
anticipated for the first lift and, in this case, compaction requirements should be
reduced. One recommendation is to allow compaction of 5% less than the required
minimum specification density for the first lift.
Construction should be performed parallel to the road alignment. Turning should not
be permitted on the first lift of base aggregate. Turn-outs may be constructed at the
roadway edge to facilitate construction.
On very soft subgrades, if the geotextile or geogrid is to provide some reinforcing,
pretensioning of the geosynthetic should be considered. For pretensioning, the area
should be proofrolled by a heavily loaded, rubber-tired vehicle such as a loaded dump
truck. The wheel load should be equivalent to the maximum expected for the site.
The vehicle should make at least four passes over the first lift in each area of the site.
Alternatively, once the design aggregate has been placed, the roadway could be used
for a time prior to paving to prestress the geotextile-aggregate or geogrid-aggregate
system in key areas.
Any ruts that form during construction should be filled in, as shown in Figure 5-13 to
maintain adequate cover over the geotextile or geogrid. Ruts should never be bladed
down, as this would decrease the amount of aggregate cover between the ruts.
All remaining base aggregate should be placed in lifts not exceeding 10 in. (250 mm)
in loose thickness and compacted to the appropriate specification density.

5.8-2a Geotextile Overlaps


Overlaps can be used to provide continuity between adjacent geotextile rolls through
frictional resistance between the overlaps. Also, a sufficient overlap is required to prevent
soil from squeezing into the aggregate at the joint. The amount of overlap depends primarily
on the soil conditions and the potential for equipment to rut the soil. If the subgrade does not
rut under construction activities, only a minimum overlap is required to provide some pullout
resistance. As the potential for rutting and squeezing of soil increases, the required overlap
increases. Since rutting potential can be related to CBR, it can be used as a guideline for the
minimum overlap required, as shown in Table 5-11.

FHWA NHI-07-092
Geosynthetics Engineering

5-67

Roadways and Pavements


August 2008

2 ft (0.6 m)

2 ft (0.6 m)

Figure 5-12. Forming curves using geotextiles.

FHWA NHI-07-092
Geosynthetics Engineering

5-68

Roadways and Pavements


August 2008

Figure 5-13. Repair of rutting with additional material.


Table 5-11
Recommended Minimum Geotextile Overlap Requirements
CBR

Minimum Overlap

> 2
1-2
0.5 - 1
< 0.5
All roll ends

12 18 in. (300 - 450 mm)


24 36 in. (600 - 900 mm)
36 in. (900 mm) or sewn1
sewn1
36 in. (900 mm) or sewn1
NOTE: 1. See Section 5.8-3.

The geotextile can be stapled or pinned at the overlaps to maintain their position during
construction activities. Nails 10 to 12 in. (250 to 300 mm) long should be placed at a
minimum of 50 ft (15 m) on centers for parallel rolls and 5 ft (1.5 m) on centers for roll ends.
Geotextile roll widths should be selected so overlaps of parallel rolls occur at the roadway
centerline and at the shoulders. Overlaps should not be placed along anticipated primary
wheel path locations. Overlaps at the end of rolls should be in the direction of the aggregate
placement (previous roll on top).
5.8-2b Geogrid Overlaps
Overlap recommendations are the same as for Geotextile Overlaps, except sewing, which in
most cases is not an option with geogrids. For geogrids that can be sewn (e.g., woven
geogrids), the recommendations for geotextile seams in the following section should be
followed.

FHWA NHI-07-092
Geosynthetics Engineering

5-69

Roadways and Pavements


August 2008

5.8-3 Seams
When sewn seams are required, they should meet the same tensile strength requirements for
survivability (Table 5-2) as those of the geosynthetics perpendicular to the seam (as
determined by the same testing methods). Seaming is discussed in detail in Section 1.8. All
factory or field seams should be sewn with thread as strong and durable as the material in the
geosynthetic. J-seams with interlocking stitches are recommended. Alternatively, if bagtype stitches, which can easily unravel, or butt-type seams are used, seams should be doublesewn with parallel stitching spaced no more than to in (6 to 13 mm) apart. Double
sewing is required to safeguard against undetected missed stitches. The geosynthetic
strength may actually have to exceed the specifications in order to provide seam strengths
equal to the specified tensile strength.
For certain geogrids, overlap joints, tying or interlocking with wire cables, plastic pipe, hog
rings, or bodkin joints may be required. Geotextile seam strength requirements should also
be applied to overlapped or mechanically fastened geogrids. Consult the manufacturer for
specific recommendations and strength test data.
5.8-4

Field Inspection

The field inspector should review the field inspection guidelines in Section 1.7. Particular
attention should be paid to factors that affect geosynthetic survivability: subgrade condition,
aggregate placement, lift thickness, and equipment operations.

5.9 SPECIFICATIONS
Specifications should generally follow the guidelines in Section 1.6.
The main
considerations include the minimum geosynthetic requirements for design and those obtained
from the survivability, retention, and filtration requirements in (Sections 5.3 - 5.7), as well as
the construction requirements covered in Section 5.8. As with other applications, it is very
important that an engineer's representative be on site during placement to observe that the
correct geosynthetic has been delivered, that the specified construction sequence is being
followed in detail, and that no damage to the geotextile is occurring. Specifications are
provided in this section for geotextiles in separation and stabilization applications, geogrids
in stabilization applications and geosynthetics in base reinforcement applications.
5.9-1 Geotextile for Separation and Stabilization Applications
The following example specification is a combination of the AASHTO M288 (2006)
geotextile material specification and its accompanying construction/installation guidelines.
FHWA NHI-07-092
Geosynthetics Engineering

5-70

Roadways and Pavements


August 2008

SPECIFICATION FOR GEOTEXTILES USED IN


SEPARATION AND STABILIZATION APPLICATIONS
(after AASHTO M288, 2006)

1. SCOPE
1.1 Description. This specification is applicable to the use of a geotextile to prevent mixing of a
subgrade soil and an aggregate cover material (i.e., separation application); and to the use of a
geotextile in wet, saturated conditions to provide the coincident functions of separation and
filtration (i.e., stabilization application). In some stabilization applications, the geotextile can
also provide the function of reinforcement.
1.2 Separation. The separation application is appropriate for pavement structures constructed over
soils with a California Bearing Ratio greater than or equal to three (CBR > 3) (shear strength
greater than approximately 2000 psf {90 kPa}). It is appropriate for unsaturated subgrade soils.
The primary function of a geotextile in this application is separation.
1.3 Stabilization. The stabilization application is appropriate for subgrade soils which are saturated
due to a high groundwater table or due to prolonged periods of wet weather. Stabilization is
applicable to pavement structures constructed over soils with a CBR between one and three (1 <
CBR <3) (shear strength between approximately 600 psf {30 kPa} and 2000 psf {90 kPa}). This
specification is not appropriate for embankment reinforcement where stress conditions may cause
global subgrade foundation or stability failure. Reinforcement of the pavement section is a sitespecific design issue.
2. REFERENCED DOCUMENTS
2.1 AASHTO Standards
T88
Particle Size Analysis of Soils
T90
Determining the Plastic Limit and Plasticity Index of Soils
T99
The Moisture-Density Relationships of Soils Using a 5.5 lb (2.5 kg) Rammer and a 12 in.
(305 mm) Drop
2.2 ASTM Standards
D 123 Standard Terminology Relating to Textiles
D 276 Test Methods for Identification of Fibers in Textiles
D 4354 Practice for Sampling of Geosynthetics for Testing
D 4355 Test Method for Deterioration of Geotextiles from Exposure to Ultraviolet Light and
Water (Xenon Arc Type Apparatus)
D 4439 Terminology for Geosynthetics
D 4491 Test Methods for Water Permeability of Geotextiles by Permittivity
D 4632 Test Method for Grab Breaking Load and Elongation of Geotextiles
D 4751 Test Method for Determining Apparent Opening Size of a Geotextile
D 4759 Practice for Determining the Specification Conformance of Geosynthetics
D 4833 Test Method for Index Puncture Resistance of Geotextiles, Geomembranes and Related
Products
FHWA NHI-07-092
Geosynthetics Engineering

5-71

Roadways and Pavements


August 2008

D 4873 Guide for Identification, Storage, and Handling of Geotextiles


D 5261 Test Method for Measuring Mass per Unit Area of Geotextiles
D 6241 Test Method for Static Puncture Strength of Geotextiles and Geotextile
Related Products Using a 50-mm Probe
3. PHYSICAL AND CHEMICAL REQUIREMENTS
3.1 Fibers used in the manufacture of geotextiles and the threads used in joining geotextiles by
sewing, shall consist of long chain synthetic polymers, composed of at least 95% by weight
polyolefins or polyesters. They shall be formed into a stable network such that the filaments or
yarns retain their dimensional stability relative to each other, including selvages.
3.2 Geotextile Requirements. The geotextile shall meet the requirements of following Table. Woven
slit film geotextiles (i.e., geotextiles made from yarns of a flat, tape-like character) will not be
allowed. All numeric values in the following table, except AOS, represent minimum average roll
values (MARV) in the weakest principal direction (i.e., average test results of any roll in a lot
sampled for conformance or quality assurance testing shall meet or exceed the minimum values).
Values for AOS represent maximum average roll values.
4. CERTIFICATION
4.1 The Contractor shall provide to the engineer a certificate stating the name of the manufacturer,
product name, style number, chemical composition of the filaments or yarns and other pertinent
information to fully describe the geotextile.
4.2 The manufacturer is responsible for establishing and maintaining a quality control program to
assure compliance with the requirements of the specification. Documentation describing the
quality control program shall be made available upon request.
4.3 The manufacturers certificate shall state that the furnished geotextile meets MARV requirements
of the specification as evaluated under the manufacturers quality control program. A person
having legal authority to bind the manufacturer shall attest to the certificate.
4.4 Either mislabeling or misrepresentation of materials shall be reason to reject those geotextile
products.
5. SAMPLING, TESTING, AND ACCEPTANCE
5.1 Geotextiles shall be subject to sampling and testing to verify conformance with this specification.
Sampling shall be in accordance with the most current ASTM D 4354 using the section titled,
Procedure for Sampling for Purchasers Specification Conformation Testing. In the absence of
purchasers testing, verification may be based on manufacturers certifications as a result of a
testing by the manufacturer of quality assurance samples obtained using he procedure for
Sampling or Manufacturers Quality Assurance (MQA) Testing. A lot size shall be considered to
be the shipment quantity of the given product or a truckload of the given product, whichever is
smaller.

FHWA NHI-07-092
Geosynthetics Engineering

5-72

Roadways and Pavements


August 2008

Table 1. Geotextile Property Requirements


Separation Application
Class 2(1)
Property

ASTM Test
Method

Units

Grab Strength

D 4632

Sewn Seam Strength(4)

Stabilization Application
Class 1(2)

Geotextile
Elongation
< 50%(3)

Geotextile
Elongation
> 50%(3)

Geotextile
Elongation
< 50%(3)

Geotextile
Elongation
> 50%(3)

1100

700

1400

900

D 4632

990

630

1260

810

Tear Strength

D 4533

400(6)

250

500

350

Puncture Strength

D 6241

2200

1375

2750

1925

Permittivity

D 4491

sec-1

0.02(5)

0.05(5)

Apparent Opening Size

D 4751

mm

0.60 max.

0.43 max.

Ultraviolet Stability
(Retained Strength)

D 4355

50% after 500 hours of exposure

NOTES:
(1) Default geotextile selection. Class 1 should be specified for more severe or harsh conditions where there is a
greater potential for geotextile damage. The engineer may specify a Class 3 geotextile [Appendix D] based
on one or more of the following:
a) The Engineer has found Class 3 geotextiles to have sufficient survivability based on field experience.
b) The Engineer has found Class 3 geotextiles to have sufficient survivability based on laboratory testing
and visual inspection of a geotextile sample removed from a field test section constructed under
anticipated field conditions.
c) Aggregate cover thickness of the first lift over the geotextile exceeds 12 in. (300 mm) and aggregate
diameter is less than 2 in. (50 mm).
d) Aggregate cover thickness of the first lift over the geotextile exceeds 6 in. (150 mm), aggregate
diameter is less than 1.2 in. (30 mm), and construction equipment contact pressure is less than 80 psi
(550 kPa).
(2) Default geotextile selection. The Engineer may specify a Class 2 or 3 geotextile [Appendix D] based on one
or more of the following:
a) The engineer has found the class of geotextile to have sufficient survivability based on field experience.
b) The engineer has found the class of geotextile to have sufficient survivability based on laboratory testing
and visual inspection of a geotextile sample removed form a field test section constructed under
anticipated field conditions.
(3) As measured in accordance with ASTM D 4632.
(4) When sewn seams are required.
(5) Default value. Permittivity of the geotextile should be greater than that of the soil (g > s). The Engineer
may also require the permeability of the geotextile to be greater than that of the soil (kg > ks).
(6) The required MARV tear strength for woven monofilament geotextiles is 250 N.

FHWA NHI-07-092
Geosynthetics Engineering

5-73

Roadways and Pavements


August 2008

5.2 Testing shall be performed in accordance with the methods referenced in this specification for the
indicated application. The number of specimens to test per sample is specified by each test
method. Geotextile product acceptance shall be based on ASTM D 4759. Product acceptance is
determined by comparing the average test results of all specimens within a given sample to the
specification MARV. Refer to ASTM D 4759 for more details regarding geotextile acceptance
procedures.
6. SHIPMENT AND STORAGE
6.1 Geotextile labeling, shipment, and storage shall follow ASTM D 4873. Product labels shall
clearly show the manufacturer or supplier name, style number, and roll number. Each shipping
document shall include a notation certifying that the material is in accordance with the
manufacturers certificate.
6.2 Each geotextile roll shall be wrapped with a material that will protect the geotextile, including the
ends of the rolls, from damage due to shipment, water, sunlight, and contaminants. The
protective wrapping shall be maintained during periods of shipment and storage.
6.3 During storage, geotextile rolls shall be elevated off the ground and adequately covered to protect
them from the following: site construction damage, precipitation, extended ultraviolet radiation
including sunlight, chemicals that are strong acids or strong bases, flames including welding
sparks, temperatures in excess of 71C (160F), and any other environmental condition that may
damage the physical property values of the geotextile.
7. CONSTRUCTION
7.1 General. Atmospheric exposure of geotextiles to the elements following lay down shall be a
maximum of 14 days to minimize damage potential.
7.2 Seaming.
a. If a sewn seam is to be used for the seaming of the geotextile, the thread used shall consist of
high strength polypropylene, or polyester. Nylon thread shall not be used. For erosion control
applications, the thread shall also be resistant to ultraviolet radiation. The thread shall be of
contrasting color to that of the geotextile itself.
b. For seams which are sewn in the field, the Contractor shall provide at least a two m length of
sewn seam for sampling by the engineer before the geotextile is installed. For seams that are
sewn in the factory, the engineer shall obtain samples of the factory seams at random from any
roll of geotextile which is to be used on the project.
b.1 For seams that are field sewn, the seams sewn for sampling shall be sewn using the same
equipment and procedures as will be used for the production of seams. If seams are to be
sewn in both the machine and cross machine directions, samples of seams from both
directions shall be provided.
b.2 The Contractor shall submit the seam assembly along with the sample of the seam. The
description shall include the seam type, stitch type, sewing thread, and stitch density.

FHWA NHI-07-092
Geosynthetics Engineering

5-74

Roadways and Pavements


August 2008

7.3 Site Preparation. The installation site shall be prepared by clearing, grubbing, and excavation or
filling the area to the design grade. This includes removal of top soil and vegetation.
NOTE: Soft spots and unsuitable areas will be identified during site preparation or
subsequent proof rolling. These areas shall be excavated and backfilled with select
material and compacted using normal procedures.
7.4 Geotextile Placement.
a. The geotextile shall be laid smooth without wrinkles or folds on the prepared subgrade in the
direction of construction traffic. Adjacent geotextile rolls shall be overlapped, sewn or joined
as required in the plans. Overlaps shall be in the direction as shown on the plans. See
following Table for overlap requirements.
Overlap Requirements
SOIL CBR

MINIMUM OVERLAP

Greater than 3

12 18 in. (300 - 450 mm)

13

2 3 ft (0.6 - 1 m)

0.5 1

3 ft (1 m) or sewn

Less than 0.5

Sewn

All roll ends

3 ft (1 m) or sewn

a.1 On curves the geotextile may be folded or cut to conform to the curves. The fold or overlap
shall be in the direction of construction and held in place by pins, staples, or piles of fill or
rock.
a.2 Prior to covering, the geotextile shall be inspected to ensure that the geotextile has not been
damaged (i.e., holes, tears, rips) during installation. The inspection shall be done by the

engineer or the engineers designated representative. It is recommended that the


designated representative be a certified inspector. Damaged geotextiles, as identified by
the Engineer, shall be repaired immediately. Cover the damaged area with a geotextile patch
which extends an amount equal to the required overlap beyond the damaged area.
b. The subbase shall be placed by end dumping onto the geotextile from the edge of the
geotextile, or over previously placed subbase aggregate. Construction vehicles shall not be
allowed directly on the geotextile. The subbase shall be placed such that at least the minimum
specified lift thickness shall be between the geotextile and equipment tires or tracks at all times.
Turning of vehicles shall not be permitted on the first lift above the geotextile.
NOTE: On subgrades having a CBR values of less than one, the subbase aggregate
should be spread in its full thickness as soon as possible after dumping to minimize
the potential of localized subgrade failure due to overloading of the subgrade.

FHWA NHI-07-092
Geosynthetics Engineering

5-75

Roadways and Pavements


August 2008

b.1 Any ruts occurring during construction shall be filled with additional subbase material, and
compacted to the specified density.
b.2 If placement of the backfill material causes damage to the geotextile, the damaged area shall
be repaired as previously described. The placement procedures shall then be modified to
eliminate further damage from taking place. (i.e., increased initial lift thickness, decrease
equipment loads, etc.)
NOTE: In stabilization applications, the use of vibratory compaction equipment is
not recommended with the initial lift of subbase material, as it may cause damage to
the geotextile.
8. METHOD OF MEASUREMENT
8.1 The geotextile shall be measured by the number of square meters computed from the
payment lines shown on the plans or from payment lines established in writing by the Engineer.
This excludes seam overlaps, but shall include geotextiles used in crest and toe of slope
treatments.
8.2 Slope preparation, excavation and backfill, bedding, and cover material are separate pay
items.
9. BASIS OF PAYMENT
9.1 The accepted quantities of geotextile shall be paid for per square yard (sq. meter) in place.
9.2 Payment will be made under:
Pay Item

Pay Unit

Separation Geotextile

Square Yard (Square Meter)

Stabilization Geotextile

Square Yard (Square Meter)

5.9-2 Geogrids for Subgrade Stabilization


AASHTO (i.e., AASHTO Provisional Standard PP 46-01, 2001) and the US Army Corp of
Engineers (UFGS-02375) have established geogrid specification for base stabilization
applications. The following specification was developed based on the attributes of these two
specifications with modifications to conform to the definitions and design methods presented
in this manual.

FHWA NHI-07-092
Geosynthetics Engineering

5-76

Roadways and Pavements


August 2008

Standard Specification
for
Geogrid Subgrade Stabilization
of Pavement Structures for Highway Applications
1. SCOPE
1.1 This is a material specification covering the use of a geogrid between aggregate cover soil
(e.g., subbase or construction platform) and soft subgrade soils (typically wet and saturated) to
provide the coincident functions of separation by preventing aggregate intrusion into the subgrade and
reinforcement of the aggregate layer to restrain subgrade movement (i.e., mechanical stabilization
application). This is a material purchasing specification and design review of use is recommended.
1.2 The stabilization application is appropriate for subgrade soils which are saturated due to a
high groundwater table or due to prolonged periods of wet weather. Stabilization is applicable to
pavement structures constructed over soils with a CBR between one and three (1 < CBR <3) (shear
strength between approximately 600 psf {30 kPa} and 2000 psf {90 kPa}). This specification is not
appropriate for embankment reinforcement where stress conditions may cause global subgrade
foundation or stability failure. Reinforcement of the pavement section is a site-specific design issue.
1.2 This is not a construction specification. This specification is based on required geogrid
properties defined by subgrade stabilization design and by geogrid survivability from installation
stresses.
2. REFERENCED DOCUMENTS
2.1 ASTM Standards:1
ASTM D 4355 (2005) Deterioration of Geotextiles from
Exposure to Light, Moisture and Heat in a Xenon-Arc Type Apparatus
ASTM D 4873 Identification, Storage, and Handling of Geosynthetic Rolls and Samples
ASTM D 5818 Obtaining Samples of Geosynthetics from a Test Section for Assessment of
Installation Damage
ASTM D 6637 Tensile Properties of Geogrids by the Single or Multi-Rib Tensile Method
2.2 GRI Standards:2
GSI GRI GG1 Geogrid Rib Tensile Strength
GSI GRI GG2 Individual Geogrid Junction Strength
GSI GRI GG4a Determination of the Long-Term Design Strength of Stiff Geogrids
GSI GRI
Determination of the Long-Term Design Strength of Flexible Geogrids
GG4b
GSI GRI GG6 Grip Types for Use in Wide Width Testing of Geotextiles and Geogrids
1
2

Available from ASTM, 100 Barr Harbor Drive, West Conshohocken, PA 19428.
Available from Geosynthetic Research Institute, Drexel University, Philadelphia, PA.

FHWA NHI-07-092
Geosynthetics Engineering

5-77

Roadways and Pavements


August 2008

2.3 U.S. Army Corps of Engineers CW-00215 Test Method for Determining Percent Open Area
(Modified for Geogrids)
3. PHYSICAL REQUIREMENTS
3.1 Polymers used in the manufacture of geogrids shall consist of long-chain synthetic polymers,
composed of at least 95 percent by weight of polyolefins, polyesters, or polyamides. They shall be
formed into a stable network such that the ribs, filaments or yarns retain their dimensional stability
relative to each other, including selvages.
3.2 Geogrids used for subgrade stabilization shall conform to the physical requirements of Section 7.
3.3 All property values, with the exception of the coefficient of interaction, coefficient of direct
shear, and ultraviolet stability, in these specifications represent minimum average roll values
(MARV) in the weakest principle direction (i.e., average test results of any roll in a lot sampled for
conformance or quality assurance testing shall meet or exceed the minimum values provided herein).
4. CERTIFICATION AND SUBMITTAL
4.1 The contractor shall provide to the Engineer, a certificate stating the name of the
manufacturer, product name, style number, chemical composition of the geogrid product and physical
properties applicable to this specification.
4.2 The Manufacturer is responsible for establishing and maintaining a quality control program to
assure compliance with the requirements of the specification. Documentation describing the quality
control program shall be made available upon request.
4.3 The Manufacturers certificate shall state that the furnished geogrid meets MARV
requirements of the specification as evaluated under the Manufacturers quality control program. The
certificate shall be attested to by a person having legal authority to bind the Manufacturer.
4.4 Either mislabeling or misrepresentation of materials shall be reason to reject those geogrids.
4.5 The contractor shall provide approximately ___ of the aggregate cover layer (e.g., subbase)
for the agency to test.
Note: The approximate amounts of subbase required by test are: gradation 18 lb (8.0 kg) for 1.5
in. (38 mm) max size, 130 lb (60.0 kg) for 3 in, (75 mm) max size; proctor (6 in. {152 mm}
diameter) 65 lb (29 kg); lab CBR 65 lb (29 kg); and moisture 11 lb (5 kg).
5. SAMPLING, TESTING, AND ACCEPTANCE
5.1 Geogrids shall be subject to sampling and testing to verify conformance with this
specification. Sampling for testing shall be in accordance with ASTM D 4354. Acceptance shall be
based on testing of either conformance samples obtained using Procedure A of ASTM D 4354, or
based on manufacturers certifications and testing of quality assurance samples obtained using
Procedure B of ASTM D 4354. A lot size for conformance or quality assurance sampling shall be
considered to be the shipment quantity of the given product or a truckload of the given product,
whichever is smaller.

FHWA NHI-07-092
Geosynthetics Engineering

5-78

Roadways and Pavements


August 2008

5.2 Testing shall be performed in accordance with the methods referenced in this specification for
the indicated application. The number of specimens to test per sample is specified by each test
method. Geogrid product acceptance shall be based on ASTM D 4759. Product acceptance is
determined by comparing the average test results of all specimens within a given sample to the
specification MARV. Refer to ASTM D 4759 for more detail regarding geogrid acceptance
procedures.
6. SHIPMENT AND STORAGE
6.1 Geogrids labeling, shipment, and storage shall follow ASTM D 4873. Product labels shall
clearly show the manufacturer or supplier name, style name, and roll number. Each shipping
document shall include a notation certifying that the material is in accordance with the manufacturers
certificate.
6.2 During storage, geogrid rolls shall be elevated off the ground and adequately covered to
protect them from the following: site construction damage, precipitation, extended ultraviolet
radiation including sunlight, chemicals that are strong acids or strong bases, flames including welding
sparks, temperatures in excess of 71EC (160EF), and any other environmental condition that may
damage the physical property values of the geogrid.
7. GEOGRID PROPERTY REQUIREMENTS FOR SUBGRADE RESTRAINT
7.1 The reinforcement shall meet the requirements of Table 1.
7.1.1 All numeric values in Table 1 represent MARVs with the exception of the ultraviolet light
stability. All numeric strength values are for the weaker principal direction, unless noted otherwise.
7.1.2 Index, survivability property values in Table 1 represent default values which provide
sufficient geogrid survivability under most construction conditions. The geogrid properties required
for survivability are dependent upon geogrid elongation.
7.1.3 The geogrid is assumed to be placed with the machine direction (MD - roll length) parallel
with the centerline of the roadway alignment. If the geogrid is placed with the machine direction
transverse to the centerline of the roadway alignment, the machine (MD) and cross machine direction
(XD) tensile strength requirements listed in Table 1 shall be reversed.

FHWA NHI-07-092
Geosynthetics Engineering

5-79

Roadways and Pavements


August 2008

Table 1. Geogrid Property Requirements


Property

Test Method

Units

Required
Value1,2

ASTM D 6637

lb/ft (kN/m)

820 (12)

CW-02215

50

Minimum Aperture Size

Direct measure

in. (mm)

____ (____)

Maximum Opening Size

Direct measure

in. (mm)

____ (____)

Reinforcement Properties
Ultimate Tensile Strength
Geogrid Percent Open Area

Class 23

Survivability Index Values


Ultimate Tensile Strength
Junction Strength
Ultraviolet Stability

ASTM D 66373

lb/ft (kN/m)

820 (12)

GRI GG2

lb (N)

25 (110)

ASTM D4355

% at 500 hrs

50

Notes
1 Values, except Ultraviolet Stability, are MARVs (average value minus two
standard deviations).
2 Minimum strength direction
3 Default geotextile selection. The Engineer may specify a Class 3 geogrids [See
Section 5.3-6] based on one or more of the following:
a) The Engineer has found Class 3 geogrids to have sufficient survivability
based on field experience.
b) The Engineer has found Class 3 geogrids to have sufficient survivability
based on laboratory testing and visual inspection of a geotextile sample
removed from a field test section constructed under anticipated field
conditions.
4 Minimum opening size must be D50 of aggregate above geogrid to provide
interlock, but not less than in. (25 mm).
5 Maximum opening size must be 2D85 to prevent aggregate from penetrating
into the subgrade, but no greater than 3 in. (75 mm).

5.9-3 Geosynthetics for Base Reinforcement of Pavement Structures


Specifications for base reinforcement in pavement structures have also been established by
AASHTO. A typical, generic type material specification for geosynthetic reinforcement for
pavements will be difficult to develop because of: the proprietary nature (i.e., current product
patents) of biaxial geogrids and some geocomposites; a lack of a thorough understanding of
the mechanistic benefits of geosynthetic reinforcement; lack of performance documentation;
and inability to measure contribution of geosynthetic reinforcement to pavement structure
with non-destructive testing methods.

FHWA NHI-07-092
Geosynthetics Engineering

5-80

Roadways and Pavements


August 2008

A geosynthetic reinforcement specification should allow use of other geosynthetic products


that either: (i) meet the physical properties defined by the key property requirements; or (ii)
by demonstrating performance equivalency through full-scale laboratory testing, in-ground
testing of pavements, and prior projects. The following Specification is from AASHTO 4E
and can be used for geogrids, geotextiles or geogrid-geotextile composites used in base
reinforcement applications.
Standard Specification
for
Geosynthetic Base Reinforcement of Pavement Structures for Highway Applications
1. SCOPE
1.1 This is a material specification covering [geogrid] [geotextile] [geogrid-geotextile composite]
geosynthetics for use as reinforcement of base OR subbase layers of pavement structures. This is a
material purchasing specification and design review of use is recommended.
__________________________
NOTE: Eliminate non applicable geosynthetic type(s) in [ ] throughout specification.
1.2 This is not a construction specification. This specification is based on required geosynthetic
properties defined by pavement design and by geosynthetic survivability from installation stresses.
2. REFERENCED DOCUMENTS
2.1 ASTM Standards:1
2.2 GRI Standards:2
2.3 U.S. Army Corps of Engineers
3. PHYSICAL REQUIREMENTS
3.1 Polymers used in the manufacture of [geogrids][geotextiles][geogrid-geotextile composites]
shall consist of long-chain synthetic polymers, composed of at least 95 percent by weight of
polyolefins, polyesters, or polyamides. They shall be formed into a stable network such that the ribs,
filaments or yarns retain their dimensional stability relative to each other, including selvages.
3.2 [Geogrids][Geotextiles][Geogrid-geotextile composites] used for reinforcement of base or
subbase layers shall conform to the physical requirements of Section 7.
3.3 All property values, with the exception of the coefficient of interaction, coefficient of direct
shear, and ultraviolet stability, in these specifications represent minimum average roll values
(MARV) in the weakest principle direction (i.e., average test results of any roll in a lot sampled for
conformance or quality assurance testing shall meet or exceed the minimum values provided herein).

__________________________
1. Available from ASTM, 100 Barr Harbor Drive, West Conshohocken, PA 19428.
2. Available from Geosynthetic Research Institute, Drexel University, Philadelphia, PA.

FHWA NHI-07-092
Geosynthetics Engineering

5-81

Roadways and Pavements


August 2008

4. CERTIFICATION AND SUBMITTAL


4.1 The contractor shall provide to the Engineer, a certificate stating the name of the
manufacturer, product name, style number, chemical composition of the
[geogrid][geotextile][geogrid-geotextile composite] product and physical properties applicable to this
specification.
4.2 The Manufacturer is responsible for establishing and maintaining a quality control program to
assure compliance with the requirements of the specification. Documentation describing the quality
control program shall be made available upon request.
4.3 The Manufacturers certificate shall state that the furnished [geogrid][geotextile][geogridgeotextile composite] meets MARV requirements of the specification as evaluated under the
Manufacturers quality control program. The certificate shall be attested to by a person having legal
authority to bind the Manufacturer.
4.4 Either mislabeling or misrepresentation of materials shall be reason to reject those
[geogrids][geotextiles][geogrid-geotextile composites].
4.5 The contractor shall provide to the Engineer, a submittal of the material values for the
properties listed in Table 1, for information purposes.
__________________________
NOTE: Eliminate non applicable tables.
4.6 The contractor shall provide approximately ___ kg of base (and subbase if applicable) for the
agency to test.
__________________________
NOTE: The approximate amounts of base (and subbase) required by test are: gradation 8.0 kg for
38 mm max size, 60.0 kg for 75 mm max size; proctor (152 mm diameter) 45 kg; lab
CBR 29 kg; moisture 5 kg; direct shear 50 kg*; pullout 1250 kg.*
*Dependent upon box size of test apparatus.
4.7 The contractor shall provide access to the agency for retrieval of asphalt concrete cores.
Contractor shall patch core holes.
__________________________
NOTE: Cores are required for Marshall stability or elastic modulus, thickness, and in-place bulk
density testing. Approximately ____ cores are required.
5. SAMPLING, TESTING, AND ACCEPTANCE
5.1 [Geogrids][Geotextiles][Geogrid-geotextile composites] shall be subject to sampling and
testing to verify conformance with this specification. Sampling for testing shall be in accordance
with ASTM D 4354. Acceptance shall be based on testing of either conformance samples obtained
using Procedure A of ASTM D 4354, or based on manufacturers certifications and testing of quality
assurance samples obtained using Procedure B of ASTM D 4354. A lot size for conformance or
quality assurance sampling shall be considered to be the shipment quantity of the given product or a
truckload of the given product, whichever is smaller.
5.2 Testing shall be performed in accordance with the methods referenced in this specification for
the indicated application. The number of specimens to test per sample is specified by each test
method. [Geogrid][Geotextile][Geogrid-geotextile composite] product acceptance shall be based on
ASTM D 4759. Product acceptance is determined by comparing the average test results of all
specimens within a given sample to the specification MARV. Refer to ASTM D 4759 for more detail
regarding acceptance procedures.

FHWA NHI-07-092
Geosynthetics Engineering

5-82

Roadways and Pavements


August 2008

Table 1a. Geogrid Property Submittal Requirements for


Base Reinforcement of Pavements
Test
Method

Property

Units

Reinforcement Properties
2% & 5 % secant moduli
Coef of Pullout Interact.
Coef of Direct Shear
Aperture Size
Percent Open Area

Value
MD

ASTM D6637
GRI GG5
ASTM D5321
Direct measure
COE CW-02215

lb/ft (kN/m)
4

degrees
in. (mm)
%

ASTM D6637
GRI GG2
ASTM D4355

lb/ft (kN/m)
lb (kN)
%

XD

Survivability Index Values


Ult Tensile Strength
Junction Strength
Ultraviolet Stability

Notes:
1 Values, except Ultraviolet Stability, Aperture Size, and Coefs are MARVs.
2 MD - machine, or roll, direction; XD - cross machine, or roll, direction
3 The stiffness properties of flexural rigidity and aperture stability are currently
being evaluated by the geosynthetic industry, in regards to this application.
4 Dimensionless.

Table 1b. Geogrid-Geotextile Composite Property Submittal


Requirements for Base Reinforcement of Flexible Pavements
Test
Method

Property

Units

Reinforcement Properties

2% & 5 % Secant Moduli


Flexural Rigidity
Coef of Pullout Interact.
Coef of Direct Shear
Permittivity
Apparent Opening Size

MD
4

ASTM D4595
4
ASTM D5732
GRI GG5
ASTM D5321
ASTM D4491
ASTM D4751

Elong
< 50%
4

ASTM D4595
GRI GG2
ASTM D 4355

XD

lb/ft (kN/m)
2
2
oz/in. (mg/cm )
5

degrees
in. (mm)
-1
sec

Survivability Index Values


Ult Tensile Strength
Junction Strength
Ultraviolet Stability

Required Value

lb (N)
lb (kN)
%

Elong
> 50%

>__%, ___ hrs

Notes
1 Values, except Ultraviolet Stability, Apparent Opening Size, ____ and ____, are
MARVs (average value minus two standard deviations).
2 MD - machine, or roll, direction; XD - cross machine, or roll, direction
3 The stiffness properties of flexural rigidity and aperture stability are currently
being evaluated by the geosynthetic industry, in regards to this application.
4 Modified test method.
5 Dimensionless.

FHWA NHI-07-092
Geosynthetics Engineering

5-83

Roadways and Pavements


August 2008

6. SHIPMENT AND STORAGE


6.1 [Geogrids][Geotextiles][Geogrid-geotextile composites] labeling, shipment, and storage shall
follow ASTM D 4873. Product labels shall clearly show the manufacturer or supplier name, style
name, and roll number. Each shipping document shall include a notation certifying that the material
is in accordance with the manufacturers certificate.[For geotextiles and geocomposites, each roll
shall be wrapped with a material that will protect the geotextile from damage due to shipment, water,
sunlight, and contaminants. The protective wrapping shall be maintained during periods of shipment
an storage.]
6.2 During storage, [geogrid][geotextile][geogrid-geotextile composite] rolls shall be elevated off
the ground and adequately covered to protect them from the following: site construction damage,
precipitation, extended ultraviolet radiation including sunlight, chemicals that are strong acids or
strong bases, flames including welding sparks, temperatures in excess of 71EC (160EF), and any other
environmental condition that may damage the physical property values of the
[geogrid][geotextile][geogrid-geotextile composite].
7. [GEOGRID][GEOTEXTILE][GEOGRID-GEOTEXTILE COMPOSITE] PROPERTY
REQUIREMENTS FOR BASE REINFORCEMENT
******************************************************
Two approaches to specification may be used. An approved products list should be used for designs
based upon product-specific data. Generic material specification should be used for designs based
upon generic properties. Approved products list approach is presented.
******************************************************
7.1 The [geogrid][geotextile][geogrid-geotextile composite] reinforcements approved for use on
this project are listed in Table 1.
7.2 Equivalent Products
7.2.1 Products submitted as equivalent for approval to use shall have documented equivalent, or
better, performance in base reinforcement or subgrade restraint in laboratory tests, full-scale field
tests, and completed project experience for the project conditions (base course material and
thickness, failure criterion, subgrade strength, etc).
7.2.2 Products submitted as equivalent shall have a documented TBR value equal or greater than
___, BCR value equal or greater than ___, or LCR value equal or greater than ___, for the
project conditions: base course thickness = ___, subbase thickness = ___, asphalt thickness =
___, failure criterion = ___ mm rut depth, and subgrade strength = ___ CBR.
Table 2. Approved [geogrid][geotextile][geogrid-geotextile composite] Reinforcement Products
Manufacturer or Distributor

Specific Product Name

******************************************************
Equivalent material description (Table 2) may not be desired, or required. Particularly if more than
one [geogrid][geotextile][geogrid-geotextile composite] is listed on the approved product list, or if a
single [geogrid][geotextile][geogrid-geotextile composite] is bid against a thicker unreinforced
pavement structure option.
******************************************************

FHWA NHI-07-092
Geosynthetics Engineering

5-84

Roadways and Pavements


August 2008

5.9-4 Selection Considerations


For a geosynthetic to perform its intended function in a roadway, it must be able to tolerate
the stresses imposed on it during construction; i.e., the geosynthetic must have sufficient
survivability to tolerate the anticipated construction operations. Geotextile or geogrid
selection for roadways is usually controlled by survivability, and the guidelines given in
Section 5.3-6 are important in this regard. The specific geotextile property values given in
Table 5-3, 5-4 and geogrid property values given in Table 5-5 are minimums. For important
projects, you are strongly encouraged to conduct your own field trials, as described in
Section 5.3-6.

5.10

COST CONSIDERATIONS

Estimation of construction costs and benefit-cost ratios for geosynthetic-stabilized road


construction is straight-forward and basically the same as that required for alternative
pavement designs. Primary factors include the following:
1. cost of the geosynthetic;
2. cost of constructing the conventional design versus a geosynthetic design (i.e.,
stabilization requirements for conventional design versus geosynthetic design),
including
a) stabilization aggregate requirements,
b) excavation and replacement requirements,
c) operational and technical feasibility, and
d) construction equipment and time requirements;
3. cost of conventional maintenance during pavement service life versus improved
service anticipated by using geosynthetic (estimated through pavement management
programs); and
4. regional experience.
Annual cost formulas, such as the Baldock method (Illinois DOT, 1982), can be applied with
an appropriate present worth factor to obtain the present worth of future expenditures.
Cost tradeoffs should also be evaluated for different construction and geosynthetic
combinations. This should include subgrade preparation and equipment control versus
geosynthetic survivability. In general, higher-cost geosynthetics with a higher survivability
on the existing subgrade will be less expensive than the additional subgrade preparation
necessary to use lower-survivability geosynthetics.

FHWA NHI-07-092
Geosynthetics Engineering

5-85

Roadways and Pavements


August 2008

With the significant history of use and advancement of geosynthetics, numerous research
efforts are ongoing to quantify the cost-benefit life cycle ratio of using geosynthetics in
permanent roadway systems (e.g., Yang, 2006). In any case, the in-place cost of a separation
geosynthetic is generally on the order of $1/yd2 ($1.20/m2), stabilization geosynthetics on the
order of 1 to 3 $/yd2 (1.20 to 3.60 $/m2), and reinforcement geosynthetics on the order of 2 to
5 $/yd2 (2.40 to 6.00 $/m2). As previously indicated the cost of the pavement section is
generally $25/yd2 ($30/m2) to $100/yd2 ($120/m2), which implies that the geosynthetic cost
ranges from less than 1% to up to 5% of the initial construction cost. For any of these
applications, the geosynthetic easily extends the life of a pavement by more than 5% and will
more than make up for the cost of the geosynthetic. The ability of a geosynthetic to prevent
premature failure of the subgrade, prevent contamination of the base and/or provided
improved base support provides a low-cost insurance that planned surface rehabilitations can
be performed and design pavement life reached. Again, it is noted that competent
subgrade/base support is critical to realizing life-cycle cost benefits of surface rehabilitations
over the life of a pavement structure.

5.11

REFERENCES

AASHTO (1972). Interim Guide for the Design of Pavement Structures, American
Association of State Transportation and Highway Officials, Washington, D.C.
AASHTO (1986). AASHTO Guide for Design of Pavement Structures, American
Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO (1990). Task Force 25 Report Guide Specifications and Test Procedures for
Geotextiles, Subcommittee on New Highway Materials, American Association of State
Transportation and Highway Officials, Washington, D.C.
AASHTO (1993). AASHTO Guide for Design of Pavement Structures, American
Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO (2001). Geosynthetic Reinforcement of the Aggregate Base Course of Flexible
Pavement Structures PP 46-01, Standard Specifications for Transportation Materials
and Methods of Sampling and Testing, 26th Edition, and Provisional Standards,
American Association of State Transportation and Highway Officials, Washington, D.C.
AASHTO (2006). Standard Specifications for Geotextiles - M 288, Standard Specifications
for Transportation Materials and Methods of Sampling and Testing, 26th Edition,
American Association of State Transportation and Highway Officials, Washington, D.C.

FHWA NHI-07-092
Geosynthetics Engineering

5-86

Roadways and Pavements


August 2008

AASHTO (2008). Mechanistic-Empirical Pavement Design Guide, Interim Edition: A


Manual of Practice, the AASHTO Mechanistic-Empirical Pavement Design Guide,
Interim Edition. American Association of State Transportation and Highway Officials,
Washington, D.C. https://bookstore.transportation.org/item_details.aspx?ID=1249
Al-Qadi, I. L. and Appea, A. K. (2003). Eight-Year Field Performance of A Secondary Road
Incorporating Geosynthetics at the Subgrade-Base Interface, Transportation Research
Board-82nd Annual Meeting, January 12-16, Washington, D.C.
Al-Qadi, I.L., Brandon, TlL., Bhutta, S.A., and Appea, A.K. (1998). Geosynthetics
Effectiveness in Flexible Secondary Road Pavements, The Charles E. Via Department of
Civil Engineering, Virginia Polytechnic Institute and State University, Blacksburg, VA.
Al-Qadi, I. L., Brandon, T. L., Valentine, R. J., Lacina, B. A., and Smith, T. E. (1994).
Laboratory Evaluation of Geosynthetic-Reinforced Pavement Sections, Transportation
Research Record 1439, TRB, National Research Council, Washington, D. C., pp. 25-31.
Al-Qadi, I.L., Tutumluer, E., Dessouky, S. and Kwon, J. (2007). Responses of geogridreinforced flexible pavement to accelerated loading, Proceedings of the International
Conference on Advanced Characterization of Pavement and Soil Engineering Materials,
Athens, Greece.
Al-Qadi, I.L., Tutumluer, E., Dessouky, S., and Kwon, J. (2007). Mechanistic response
measurements of geogrid reinforced flexible pavements to vehicular loading. In CDROM Proceedings of the Geosynthetics 2007 Conference, Washington, D.C.
Al-Qadi, I., Tutumluer, E., Kwon, J., and Dessouky, S., (2008). Geogrid in Flexible
Pavements: Validated Mechanism, Accepted for publication of Transportation Research
Board, Washington, D.C.
Anderson, R.P. (2006). Geogrid Separation, Proceedings of the International Conference on
New Developments in Geoenvironmental and Geotechnical Engineering, Incheon, South
Korea, pp.472-480.
Austin, D. N. and Coleman, D. M. (1993). A Field Evaluation of Geosynthetic-Reinforced
Haul Roads Over Soft Foundation Soils, Geosynthetics 93, Vancouver, Canada, pp. 6580.
ASTM (2006). Annual Books of ASTM Standards,
Conshohocken, Pennsylvania:
Volume 4.08 (I), Soil and Rock
Volume 4.09 (II), Soil and Rock; Geosynthetics
Volume 4.13 Geosynthetics

ASTM International, West

Barenberg, E.J., Hales, J., and Dowland, J. (1975). Evaluation of Soil-Aggregate Systems
with MIRAFI Fabrics, UILU-ENG-75-2020 Report for Celanese Fibers Marketing
Company, University of Illinois, Urbana.
FHWA NHI-07-092
Geosynthetics Engineering

5-87

Roadways and Pavements


August 2008

Barksdale, R.D., Brown, S.F. and Chan, F. (1989). Potential Benefits of Geosynthetics in
Flexible Pavement Systems, National Cooperative Highway Research Program Report
315, Transportation Research Board, Washington, D.C., 56 p.
Baumgartner, R.H. (1994). Geotextile Design Guidelines for Permeable Bases, Federal
Highway Administration, June, 33 p.
Berg, R.R., Christopher, B.R. and Perkins, S.W. (2000). Geosynthetic Reinforcement of the
Aggregate Base/Subbase Courses of Pavement Structures GMA White Paper II,
Geosynthetic Materials Association, Roseville, MN, 176 p.
(www.gmanow.com/pdf/WPIIFINALGMA.pdf)
Bhutta, S.A. (1998).
Mechanistic-Empirical Pavement Design Procedure for
Geosynthetically Stabilized Flexible Pavements, Ph.D. Thesis, Virginia Polytechnic
Institute and State University, Blacksburg, VA. (http://www.vtti.vt.edu)
Cedergren, H.R. (1987). Drainage of Highway and Airfield Pavements, Krieger, 289 p.
Cedergen, H.R. (1989). Seepage, Drainage, and Flow Nets, John Wiley & Sons, New York,
pp. 153-156.
Chan, F., Barksdale, R.D. and Brown, S.F. (1989). Aggregate Base Reinforcement of
Surface Pavements, Geotextiles and Geomembranes, Vol. 8, No. 3, pp. 165-189.
Christopher, B.R., Berg, R.R., Perkins, S.W. (2001). Geosynthetic Reinforcements in
Roadway Sections National Cooperative Highway Research Program, NCHRP Project
20-7, Task 112, FY2000, Transportation Research Board, National Academy Press,
Washington, D.C.
Christopher, B.R. and Holtz, R.D. (1985). Geotextile Engineering Manual, Federal
Highway Administration, FHWA-TS-86/203, March, 1044 p.
Christopher, B.R. and Holtz, R.D. (1989). Geotextile Design and Construction Guidelines,
U.S. Department of Transportation, Federal Highway Administration, Washington DC,
Report No. HI-89-050, 265 p.
Christopher, B.R. and Holtz, R.D. (1991). Geotextiles for Subgrade Stabilization in
Permanent Roads and Highways, Proceedings of Geosynthetics '91, Atlanta, GA, Vol. 2,
pp. 701-713.
Christopher, B.R. and Lacina, B. (2008). Roadway Subgrade Stabilization Study,
Proceedings of GeoAmericas 2008, Cancun, Mexico, pp. 1013 -1021.
Christopher, B.R., Schwartz, C., Boudreau, R. (2006). Geotechnical Aspects of Pavements,
U.S. Department of Transportation, National Highway Institute, Federal Highway
Administration, Washington DC, FHWA-NHI-05-037, 874 p.
FHWA NHI-07-092
Geosynthetics Engineering

5-88

Roadways and Pavements


August 2008

Collin, J.G., Kinney, T.C., and Fu, X. (1996). Full Scale Highway Load Test of Flexible
Pavement Systems with Geogrid Reinforced Base Courses, Geosynthetics International,
3(4), pp. 537-549.
ERES Consultants, Inc. (1999). Pavement Subsurface Drainage Design, Participants
Reference Manual for NHI Course Number 131026, National Highway Institute, Federal
Highway Administration, Washington, DC.
FAA (1994). Engineering Brief No. 49, Geogrid Reinforced Base Course, Federal Aviation
Administration, U.S. Department of Transportation.
Fannin, R.J. and Sigurdsson, O. (1996). Field Observations on Stabilization of Unpaved
Roads with Geosynthetics, Journal of Geotechnical Engineering, 122(7), pp. 544-553.
FHWA (1999). Pavement Subsurface Drainage Design, Participants Reference Manual for
NHI Course Number 131026, National Highway Institute, FHWA, ERES.
Gabr, M., Robinson, B., Collin, J.G. and Berg, R.R. (2006). Promoting Geosynthetics Use
on Federal Lands Highway Projects, Federal Highway Administration, Central Federal
Lands Highway Division, Lakewood, Colorado, FHWA-CFL/TD-06-009, p. 116.
Gabr, M. Leng, J. and Ju, T.J. (2001). Response and Characteristics of Geogrid-Reinforced
Aggregate Under Cyclic Plate Load, Research Report Submitted to Tensar Earth
Technologies, NC State University, 40 pp.
GeoServices, Inc. (1989). Geotextile Design Examples, Federal Highway Adminstration,
FHWA Contract No. DTFH-86-R-00102.
Giroud, J.P. and Bonaparte, R. (1985). Design of Unpaved Roads and Trafficed Areas with
Geogrids, Foundations for Roads and Loaded Areas, Polymer Grid Reinforcement,
Thomas Telford Ltd., London, pp. 116-127.
Giroud, J.P. and Noiray, L. (1981). Geotextile-Reinforced Unpaved Roads, Journal of the
Geotechnical Engineering Division, American Society of Civil Engineers, Vol. 107, No
GT 9, pp. 1233-1254.
Giroud, J.P. and Han, J. (2004a). Design Method for Geogrid-Reinforced Unpaved Roads:
I Development of Design Method, Journal of Geotechnical and GeoEnvironmental
Engineering, Vol. 130, No. 8, pp. 775- 786.
Giroud, J.P. and Han, J. (2004b). Design Method for Geogrid-Reinforced Unpaved Roads:
II Calibration and Applications, Journal of Geotechnical and GeoEnvironmental
Engineering, Vol. 130, No. 8, pp. 787- 797.
GMA (1999). Geosynthetics in Pavement Systems Applications, Section One: Geogrids,
Section Two: Geotextiles, prepared for AASHTO, Geosynthetics Materials Association,
Roseville, MN, 46 p.
FHWA NHI-07-092
Geosynthetics Engineering

5-89

Roadways and Pavements


August 2008

Haas, R., Walls, J. and Carroll, R.G. (1988). Geogrid Reinforcement of Granular Base in
Flexible Pavements, presented at the 67th Annual Meeting, Transportation Research
Board, Washington, D.C.
Haliburton, T.A., Lawmaster, J.D. and McGuffey, V.C. (1981). Use of Engineering Fabrics
in Transportation Related Applications, Federal Highway Administration, FHWA
DTFH61-80-C-00094.
Hamilton, J.S. and Pearce, R.A. (1981). Guidelines for Design of Flexible Pavements using
Mirafi Woven Stabilization Fabrics, Law Engineering Testing Co. Report to Celanese
Corp., 47 p.
Heukelom, W. and Klomp, A.J.G. (1962). Dynamic Testing as a Means of Controlling
Pavements during and after Construction, Proceedings, 1st Int. Conf. on Structural
Design of Asphalt Pavement, Univ. of Michigan, pp. 667-679.
Helwany, S., Dyer, J., and Leidy, J. (1998). Finite Element Analyses of Flexible Pavements,
Journal of Transportation Engineering, Vol. 124, No. 5, pp. 491-499.
Holtz, R.D., Christopher, B.R. and Berg, R.R. (1998). Geosynthetic Design and
Construction Guidelines, FHWA-HI-98-038, U.S. Department of Transportation,
Federal Highway Administration, Washington, D.C., pp. 460.
Illinois Department of Transportation (1982). Design Manual - Section 7: Pavement
Design, I-82-2, Bureau of Design, Springfield.
Jorenby, B.N. and Hicks, R.G. (1986). Base Coarse Contamination Limits, Transportation
Research Record, No. 1095, Washington, D.C.
Kinney, T.C. and Xiaolin, Y. (1995). Geogrid Aperture Rigidity by In-Plane Rotation,
Proceedings of Geosynthetics 95, Nashville, TN, pp. 525 537.
Kinney, T.C. (2000). Standard Test Method for Determining the Aperture Stability Modulus
of a Geogrid, Shannon & Wilson, Seattle, WA.
Knapton, J. and Austin, R.A. (1996). Laboratory testing of reinforced unpaved roads, Earth
Reinforcement, H. Ochiai, N. Yasufuku, and K. Omine eds., Balkema, Rotterdam, The
Netherlands, pp. 615-618.
Koerner, R.M. (1990). Editor, The Seaming of Geosynthetics, Special Issue, Geotextiles and
Geomembranes, Vol. 9, Nos. 4-6, pp. 281-564.
Koerner, R.M. (1994). Designing With Geosynthetics, 3rd Edition, Prentice-Hall Inc.,
Englewood Cliffs, NJ, 783p.
Koerner, R.M. (2005). Designing With Geosynthetics, 5th Edition, Prentice-Hall Inc.,
Englewood Cliffs, NJ, 796p.
FHWA NHI-07-092
Geosynthetics Engineering

5-90

Roadways and Pavements


August 2008

Kwon, J., Tutumluer, E., and Kim, M. (2005). Development of a Mechanistic Model for
Geogrid Reinforced Flexible Pavements. Geosynthetics International, 12:6, pp. 310-320.
Kwon, J., Tutumluer, E., and Konietzky, H. (2007). Aggregate Base Residual Stresses
Affecting Geogrid Reinforced Flexible Pavement Response, International Journal of
Pavement Engineering.
Leng, J. and Gabr, M. (2002). Characteristics of Geogrid-Reinforced Aggregate under
Cyclic Load, Journal of Transportation Research Board, No. 1786, National Research
Council, Washington, D.C., pp. 29-35.
Mery, J. (1995) Field Studies on the Mechanical Behavior of Geosynthetic-Reinforced
Unpaved Roads, Sixth International Conference on Low Volume Roads, Vol. 2, pp. 234239.
Miuara, N., Sakai, A., Taesiri, Y., Yamanouchi, T. and Yasuhara, K. (1990). Polymer Grid
Reinforced Pavement on Soft Clay Grounds, Geotextiles and Geomembranes, Vol. 9,
No. 1, pp. 99-123.
NCHRP (2002). Design Guide Design of New and Rehabilitated Pavement Structures,
NCHRP 1-37A Project, Draft Final Report, Part 1 Introduction and Part 2 Design
Inputs, Prepared for the National Cooperative Highway Research Program by ERES
Division of ARA.
NCRHP (2008). NCHRP Research Field 1Pavements, the National Cooperative Highway
Research
Program,
Transportation
Research
Board,
Washington,
D.C.
http://www.trb.org/CRP/NCHRP/NCHRPProjects.asp
Perkins, S.W. (1999). Geosynthetics Stabilization of Flexible Pavements: Laboratory Based
Pavement Test Sections, FHWA Report Reference Number MT-99-001/8138, 140 p.
Perkins, S.W., Bowders, J.J., Christopher, B.R., and Berg, R.R. (2005a). Advances in
Geosynthetic Reinforcement in Pavement Systems, Proceedings of the GeoFrontiers
Conference, Austin, Texas.
Perkins, S.W., Bowders, J.J., Christopher, B.R., and Berg, R.R. (2005c). Geosynthetic
Reinforcement for Pavement Systems: US Perspectives, Geotechnical Special
Publication 141, Proceedings of the GeoFrontiers Conference, Austin, Texas.
Perkins, S.W., Christopher, B.R., Cuelho, E.L., Eiksund, G.R., Hoff, I., Schwartz, C.W.,
Svan, G., and Want, A. (2004). Development of Design Methods for Geosynthetic
Reinforced Flexible Pavements, FHWA Report Reference Number DTFH61-01-X00068, 63p. (http://www.coe.montana.edu/wti/wti/pdf/426202_Final_Report.pdf)
Perkins, S.W., Christopher, B.R., Eiksund, G.R., Schwartz, C.W., and Svano, G. (2005b).
Modeling Effects of Reinforcement on Lateral Confinement of Roadway Aggregate,
Geotechnical Special Publication 130, GeoFrontiers 2005, pp. 283-296.
FHWA NHI-07-092
Geosynthetics Engineering

5-91

Roadways and Pavements


August 2008

Perkins, S.W. and Salvano, G. (2004). Assessment of Interface Shear Growth from
Measured Geosynthetic Strains in a Reinforced Pavement Subject to Repeated Loads,
Proceedings of the 8th International Conference on Geosynthetics, Yokahama, Japan.
Rankilor, P.R (1981). Membranes in Ground Engineering, John Wiley & Sons, Inc.,
Chichester, England, 377 p.
Steward, J., Williamson, R. and Mohney, J. (1977). Guidelines for Use of Fabrics in
Construction and Maintenance of Low-Volume Roads, USDA, Forest Service, Portland,
OR. Also reprinted as Report No. FHWA-TS-78-205.
Terzaghi, K. and Peck, R.B. (1967). Soil Mechanics in Engineering Practice, John Wiley &
Sons, New York, 729p.
Tingle, J.S. and Webster, S.L. (2003). Review of Corps of Engineers Design of Geosynthetc
Reinforced Unpaved Roads, Annual meeting CD-ROM, TRB, Washington, D.C.
Tsai, W. (1995). Evaluation of Geotextiles as Separators in Roadways, Ph.D. Thesis,
University of Washington, Seattle, Washington, 172 p.
United States Army Corps of Engineers (2003). Use of Geogrids in Pavement Construction,
ETL 1110-1-189, Department of the Army, U.S. Army Corps of Engineers, Washington,
D.C., 37 p.
Webster, S.L. (1993). Geogrid Reinforced Base Courses for Flexible Pavements for Light
Aircraft: Test Section Construction, Behavior Under Traffic, Laboratory Tests, and
Design Criteria, USAE Waterways Experiment Station, Vicksburg, MS, Technical
Report DOT/FAA/RD-92/25, 100 p.
Yang, S.H. (2006). Effectiveness of Using Geotextiles in Flexible Pavements: Life-Cycle
Cost Analysis, MSCE Thesis, Virginia Polytechnic Institute and State University,
Blacksburg, VA.
Yoder, E.J. and Witczak, M.W. (1975). Principles of Pavement Design, Second Edition,
Wiley, 711 p.

FHWA NHI-07-092
Geosynthetics Engineering

5-92

Roadways and Pavements


August 2008

6.0 PAVEMENT OVERLAYS

6.1

BACKGROUND

The second largest application of geosynthetics in North America is in asphalt overlays of


asphalt concrete (AC) and Portland cement concrete (PCC) pavement structures. (The
largest single application is separation/stabilization, which utilizes an estimated 120 million
square yards {100 million square meters} of geotextile and geogrids.) It is estimated that
more than 100 million square yards (91 million square meters) of geotextiles are used in
overlays per year or about the equivalent of 140,000 lane miles (225,000 lane kilometers)
(GMA, 2007). This is indeed an impressive statistic. Geogrids and geocomposites have also
made recent inroads into this application.
The history of geosynthetics in this application accounts for much of the confusion and
skepticism. Promotion of geotextiles in overlays in the 1970s and early 1980s claimed that
the geotextile reinforced the pavement and that the reinforcement prevented cracks in the old
pavement from reflecting up through the new overlay. These claims are rarely, if ever,
presented today. The tensile moduli of the light-weight nonwoven geotextiles typically used
in this application are too low to mobilize significant tension under acceptable pavement
deflections for the geotextile to act as reinforcement. It is also commonly accepted that
geotextiles do not prevent, but rather retard reflection cracking from occurring. Geogrids are
currently being promoted on the same level as geotextiles of the 1970s. Are either of these
geosynthetics beneficial in overlay pavement construction? The answer to this question is
yes and no: they provide benefit in some applications and in others, they should not be used.
In this chapter, the functions of geotextiles, geogrids and geocomposites in pavement
overlays will be identified. Application of these mechanisms will then be examined with
respect to pavement type and condition along with the cost-benefit of the appropriate use of
these materials in reducing the rate or severity of reflection cracks. Emphasis will be placed
on proper construction to achieve optimum benefit.

6.2

PAVEMENT OVERLAYS AND REFLECTIVE CRACKING

When pavements have reached their design performance period, the surface asphalt or
concrete pavement often contains predictive fatigue cracks. The surface pavement may also
contain cracks for reasons such as excess traffic loads, thermal cracks, cracks due to frost
heave, or water problems. Cracks are problematic due to reduction in structural capacity as
FHWA NHI-07-092
Geosynthetics Engineering

6-1

Pavement Overlays
August 2008

well as allowing water to enter into other (often moisture sensitive) layers of the pavement.
While a number of methods exist for rehabilitation of pavements (covered in detail in
FHWA-NHI 131008, 1998), due to the ease of installation and initial cost, the most common
treatment to extend the life of the old pavement is to overlay the surface with a new hot mix
asphalt (HMA) layer. The HMA overlay provides improved structural strength and ride
quality. Important factors in the decision to use and design the overlay are covered in
FHWA-NHI 131008 (FHWA, 1998).
A major problem encountered with HMA overlays is the potential for cracks in the old
asphalt or concrete to reflect through the new pavement. Reflective cracks are caused by
shear, tensile and bending stresses created in the new HMA overlay due to lateral, vertical
and/or differential crack or joint movements in the old pavement resulting from traffic
loading and changes in temperature. Low temperatures cause the underlying pavement to
contract, increasing joint and crack openings, thus, creating tensile stress in an overlay.
Traffic loadings produce a completely different type of movement in which the differential
vertical deflection created by traffic passing over a joint or working crack creates shearing
and bending stress in the overlay as shown in Figure 6-1. Movement can also be created by
foundation problems, as well as heave from expansive subgrade soils or frost; however, these
special problematic conditions cannot usually be handled with just crack reflection
treatments or even an overlay for that matter.

Figure 6-1. Shearing and bending stress in HMA overlay (Jayawickrama and Lytton, 1987).

FHWA NHI-07-092
Geosynthetics Engineering

6-2

Pavement Overlays
August 2008

The key to preventing reflection cracking is to eliminate the deformation and the stresses;
however, it is highly unlikely that this can be accomplished. Therefore the best that can be
achieved is usually a reduction in the rate of appearance and the severity of the reflection
cracking. A number of different treatments are listed in FHWA-NHI 131008, including the
use of geosynthetics. As reviewed in FHWA-NHI 131008, there has been varying success
with these treatments and thus far there is no conclusive evidence to which one works the
best. It should be noted that in extreme cases (such as large temperature variations), no
treatment has been found to offer relief and the best that can be expected is temporary
mitigation. On the other hand, in mild climates where the underlying pavement does not
experience large vertical movement, a number of treatments have the potential to succeed.
Since no treatment stops reflective cracking, effectiveness of the treatment should be based
on its ability to:
Retard the rate of reflection cracking.
Reduce the severity of the cracks once they occur.
Provide other benefits, such as reducing the overlay thickness or enhancing the
waterproofing capabilities of the pavement.
Two basic approaches will be reviewed in this chapter including:
1. Stress-relieving interlayer in which the stresses are dissipated at the joint or crack
before they create stress in the overlay: Nonwoven geotextiles and geosynthetic
composite strips are used for this approach.
2. Stress-resistance layer in which a high tensile modulus reinforcement is used to
resist tensile stress in the new HMA overlay, a relatively new approach not covered
in the NHI 131008 manual: Geogrids are used for this approach.
Multilayer geocomposites may provide elements of both approaches.
Reflection cracking also leads to increased infiltration of surface water into the pavement,
which in turn weakens the supporting layers. Both nonwoven geotextiles and the
geocomposites used in this application provide the added benefit of enhancing the
waterproofing capabilities of the pavement, even after the reflection cracks return. As
discussed in the following sections, this added benefit may offer equal or even greater
benefit to the longevity of the pavement than retarding the reflection cracks.

FHWA NHI-07-092
Geosynthetics Engineering

6-3

Pavement Overlays
August 2008

6.3

GEOTEXTILES (a.k.a. Paving Fabric Interlayers)

6.3-1 Functions
Geotextiles, typically needlepunched nonwovens, can be installed over an asphalt tack coat
beneath a layer of HMA, and in the process, become uniformly saturated with asphalt
cement, as shown in Figure 6-2. If this asphalt saturated geotextile layer is sufficiently thick
(as defined later in this section), it will provide a cushion layer to absorb some movement
from the old pavement without transferring it to the new overlay and act as a stress-relieving
interlayer. Properly installed, asphalt-saturated geotextiles also function as a moisture barrier
that protects the underlying pavement structure from further degradation due to infiltration of
surface water. When properly installed, both primary functions,
Protection as a stress relief interlayer; and
Fluid barrier as a water proofing membrane
combine to extend the life of the overlay and the pavement section.
Thus, geotextiles can be used as alternatives to stress relieving rubberized asphalts, stressarresting granular layers and seal coats for retarding reflection cracks and controlling surface
moisture infiltration in pavement overlays. Geotextiles can also be used as interlayers on
new pavements to reduce water infiltration. The conditions of using this treatment with
HMA overlays over old asphalt concrete pavements and rigid concrete pavements as well as
use with chip seals over unpaved and paved roads along with design, installation, and
specification are reviewed in the following sections.

Figure 6-2. Geotextiles (a.k.a. Paving Fabric) in rehabilitated pavement section.


FHWA NHI-07-092
Geosynthetics Engineering

6-4

Pavement Overlays
August 2008

6.3-2 Asphalt Concrete (AC) Pavement Applications


When geotextiles are used with HMA overlays of AC pavements, they can be effective in
controlling (retarding) reflection cracking of low- and medium-severity alligator-cracked
pavements. They also may be useful for controlling reflection of thermal cracks, although
they are not as effective in retarding reflection of cracks due to significant horizontal or
vertical movements. (AASHTO, 1993)
The variable performance of geotextile overlays, and the divergent opinions regarding
benefits, is strongly influenced by the following factors (Barksdale, 1991):

Type and extent of existing pavement distress, including crack widths.


Extent of remedial work performed on the old pavement, such as crack sealing and/or
filling, pothole repair, and replacement of failed base and subgrade areas.
Overlay thickness.
Variability of pavement structural strength from one section to another.
Climate.

Obviously, an additional factor is the geotextile. These factors are summarized below.
Distress Type (Barksdale, 1991): Geotextiles generally have performed best when used for
load-related fatigue distress (e.g., closely spaced alligator cracking). Fatigue cracks should
be less than 0.12 in (3 mm) wide for best results. Cracks greater than 0.4 in (10 mm) wide
require stiff fillers. Geotextiles used to retard thermal cracking have, in general, been found
to be ineffective.
Remediation of Old Pavement (AASHTO, 1993): Much of the deterioration that occurs in
overlays is the result of unrepaired distress in the existing pavement prior to the overlay.
Distressed areas of the existing pavement should be repaired if the distress is likely to affect
performance of the overlay within a few years. The amount of pre-overlay repair is related to
the overlay design. The engineer should consider the cost implications of pre-overlay repair
versus overlay design. Guidelines on pre-overlay repair techniques are included in FHWANHI 131008.
Crack Movement (FHWA-NHI 131008): Several studies have indicated that the effectiveness of geosynthetics is related to the magnitude of differential vertical movement at the joint
or crack, with a range of movement on the order of 3 mil to 8 mil (0.08 mm to 0.2 mm)
providing effective performance. The lower end is consistent with the Asphalt Institutes
recommendation of limiting the differential vertical deflection at the joints to 2 mil (0.05
mm) to achieve good performance from an overlay if no treatment is used. Where existing
FHWA NHI-07-092
Geosynthetics Engineering

6-5

Pavement Overlays
August 2008

joints or cracks experience large horizontal or differential vertical movement, cracks tend to
reflect through the geosynthetic, eliminating its effectiveness. However, if the geotextile
does not rupture, some degree of waterproofing will still be provided even if the overlay is
cracked.
Effect of Overlay Thickness: Pavement performance is quite sensitive to the overlay
thickness, either with or without a geotextile interlayer. Correspondingly, the benefits of a
geotextile in retarding reflective cracking will increase with increasing thickness of the
overlay. Geotextiles are most effective in retarding reflective cracking in overlays that are
1.5 in (38 mm) thick or greater. These observations, as reported by Barksdale (1991), are
based upon extensive research conducted by Caltrans (Predoehl, 1990). The California
results imply that a geotextile interlayer is equivalent to adding 1.2 in (30 mm) of asphalt
concrete to the overlay.
Variability of Pavement Structural Strength (Barksdale, 1991): The structural strength
of existing pavement, and, therefore, required overlay thickness, often varies greatly along a
roadway. The significant effect of such variation on overlay performance has not often been
considered in the past. This oversight likely contributes to some of the diverse opinions
regarding geotextile benefits in asphalt overlays. (Variation of pavement strength along a
roadway should be addressed for future demonstration or test sections of overlays.)
Climate: It has been observed (Aldrich, 1986) that geotextile interlayers have generally
performed better in warm and mild climates than in cold climates. However, the beneficial
effects of reducing water infiltration - a principal function of the geotextile - were not
considered in Aldrich's (1986) study. Successful installation and beneficial performance of
geotextile interlayers in cold regions, such as Alaska, challenge the generality regarding
climate.
Geotextile: Lightweight (e.g., 4 oz/yd2 {140 g/m2}) nonwoven geotextiles are typically used
for asphalt overlays. These asphalt-impregnated geotextiles primarily function as a moisture
barrier. Lighter weight geotextiles provide little, if any, stress relief. Use of heavier,
nonwoven geotextiles can provide greater cushioning or stress-relieving membrane
interlayer-like benefits, in addition to moisture-barrier functions. However, the weight
should be limited to no greater than about 6 oz/yd2 (200 g/m2) to avoid the potential for
delaminating and shearing in the plane of the geotextile.
6.3-3 Portland Cement Concrete Pavement Applications
Geotextiles are used with HMA overlays of crack/seat-fractured plain Portland cement
concrete (PCC) pavement to help control reflection cracking (AASHTO, 1993). A review of
FHWA NHI-07-092
Geosynthetics Engineering

6-6

Pavement Overlays
August 2008

various studies by Maxim Technologies (1997) confirms the successful use in this
application. Caltrans has found a successful practice for this type of overlay to consist of
placing approximately 1 in. (25 mm) of an asphalt concrete leveling course, a non-woven
fabric saturated with an AR-4000 asphalt cement which serves as an interlayer, 4 in. (100
mm) of dense graded asphalt concrete, and 1 in. (25 mm) of an open-graded asphalt concrete
surface course. The expected satisfactory performance period for this 6 in. (150 mm) thick
overlay is about 10 years. An evaluation of this overlay design with an increased thickness
for heavily trafficked freeways is provided by Monismith and Long (1999).
Geotextiles also may be used with HMA overlays of jointed plain concrete pavement (JPCP)
and of jointed reinforced concrete pavement (JRCP) to control reflective cracking. The
effectiveness with these pavements is listed as questionable in the AASHTO Guide for
Design of Pavement Structures (1993); however, Maxim Technologies (1997) identified
several studies that demonstrated successful performance including use to correct a specific
distress condition (i.e., pop-outs). Again, geotextiles were not effective in retarding
transverse cracks or joints when there are existing excessive vertical movements.
Important factors in assessing applicability and potential benefits of using a geotextile
interlayer with PCC pavements include (Barksdale, 1991):
Existing structural strength of the pavement.
Slab preparation.
Crack movement.
Geotextile installation.
Required overlay thickness.
Climate.
Economics of geotextile overlay versus other design alternatives.
6.3-4 HMA-Overlaid PCC Pavements
Geotextiles are also used with new HMA overlays of AC-overlaid Portland cement concrete
pavements (AC/PCC), where the original pavements may be JPCP, JRCP or continuously
reinforced concrete pavement (CRCP). Some pavements are constructed as AC/PCC,
although most are PCC pavements that have already been overlaid with AC. In addition to
controlling reflecting cracking, an impregnated geotextile can help control surface water
infiltration into the pavement. Water infiltration can result in loss of bond between AC and
PCC, stripping in the AC layers, progression of D cracking or reactive aggregate distress (in
pavements with these problems), and weakening of the base and subgrade materials.

FHWA NHI-07-092
Geosynthetics Engineering

6-7

Pavement Overlays
August 2008

6.3-5 Chip Seals for Unpaved Roads and AC Pavements


Use of geotextile treatments with chip seals is a simple, innovative application that is not
broadly used in the U.S., but is increasing (e.g., in several counties in California and Texas)
and is used in several other countries (e.g., Australia, Canada, France, New Zealand, South
Africa, and the United Kingdom). Geotextiles have been used with chip seals for both
surface dressing of unpaved roads and overlaying of secondary paved roads. Maxim
Technologies (1997) reports on a number of studies demonstrating superior performance as
compared to control sections with normal chip seal construction procedures. However, some
unsuccessful trials have been reported in Montana and Nevada (Gransberg and James, 2005).
The primary benefit in both paved and unpaved roads is waterproofing (i.e., maintaining
uniformity and controlling moisture contents in the pavement and roadbed layers during
seasonal wet and dry periods). The main challenge is to maintain long term bonding of the
stone to the geosynthetic (a problem with chip seals in general), which has been
accomplished through the use of appropriate AC sealants based on testing and field trials
with the specific geosynthetic.
In paved roads, Brown (2003) studied field trials and experimentation of double chip seal
with geotextiles over a period of 19 years and found that this treatment substantially adds to
the pavement life through the retardation of reflection cracks and water/air proofing at a
lower cost than other standard overlays. He noted that the treatment reduced further
deterioration of the old pavement due to oxidation and stripping, and reduced crack reflection
more than other conventional methods including asphalt overlays with paving fabric.
Additional information on the current practice is available in the Transportation Research
Board National Cooperative Highway Research Program (NCHRP) Synthesis 342 on Chip
Seal Best Practices 342 (Gransberg and James, 2005).
6.3-6 Advantages and Potential Problems
Advantages
An asphalt-impregnated geotextile functioning as a moisture barrier and a stress-relieving
interlayer provides several possible advantages to their use. Retardation of reflection cracks
will:
Increase the overlay and the roadway life.
Decrease roadway maintenance costs.
Increase pavement serviceability.

FHWA NHI-07-092
Geosynthetics Engineering

6-8

Pavement Overlays
August 2008

Reflection cracking can have a considerable, often controlling, influence on the life of an
HMA overlay (AASHTO, 1993). After a pavement cracks, its longevity is quickly reduced.
Deteriorated reflection cracks require more frequent maintenance, such as sealing and
patching. Reflection cracks also permit water to enter the pavement structure, which can
weaken the base layers and subgrade, and decrease the structural capacity of the pavement.
Base and subgrade will be weakened by ingress of water if the base does not have excellent
(i.e., water removed within 2 hours) or good (i.e., water removed within 1 day) drainage.
Water infiltration causes a reduction in shear strength and the subgrade, which in turn leads
to a rapid deterioration of the pavement system. The sealing function of the asphaltimpregnated geotextile is intended to reduce surface water infiltration through reflection
cracks (when they eventually reappear at the surface of the overlay) and through thermalinduced cracks. Both laboratory and field pavement cores indicate that the presence of a
properly installed geotextile interlayer system reduces the permeability of a pavement by one
to three orders of magnitude, thus becoming an efficient moisture barrier to enhance
pavement performance (Marienfeld and Baker, 1999).
Reduction in surface water entering PCC pavements potentially provides additional benefits
such as:
Reduction or elimination of pumping (i.e., no water, no pumping).
Decreased slab movements through reduced erosion of fines from beneath the slab
(lower moisture gradients might also reduce slab warping).
Increased subgrade strength through a decrease in moisture (Barksdale, 1991).
Potential Problems
Correct construction of the asphalt-impregnated geotextile and HMA overlay is paramount to
it functioning as designed. Too little asphalt in the tack coat can result in a partially saturated
geotextile, which in turn can absorb moisture and lead to spalling or popping off of the
surface treatment due to freeze-thaw action within the geotextile. Bleeding occurs with too
much asphalt which can result in overlay slippage, as well as potential pavement slippage
planes. Bleeding also can cause difficulty with installation, as it can result in the geotextile
sticking to and being pulled up by the tires and tracks of the asphalt trucks and paving
vehicle. The HMA overlay must be placed below the specified temperature, which requires
inspection and control. HMA placement significantly above the specified temperature can
result in the asphalt tack coat being drawn out of the geotextile which can result in shrinkage
or even melting of the geotextile. Shrinkage and melting is a concern for a polypropylene
geotextile which has a typical melt temperature of 330F (165C). It is not a concern for a
polyester geotextile which has a typical melt temperature of 435F (225C). Improper
pavement preparation and crack filling can also decrease the effectiveness of the geotextile
moisture barrier.
FHWA NHI-07-092
Geosynthetics Engineering

6-9

Pavement Overlays
August 2008

Geotextiles cannot be expected to perform well when the roadway being overlaid is
structurally inadequate. Nor will such surface treatments do anything to solve groundwater
problems, subgrade softening, base course contamination, or freeze-thaw problems. These
problems must be corrected before resurfacing, independently of the geotextile used.
Moisture barriers can also trap water in pavements, which could potentially create problems
such as stripping and pumping. Overlays (with or without geotextiles) should not be placed
directly over wet pavement sections (e.g., as indicated by water in cracks or core holes). In
all cases water should be allowed to drain from the pavement sections. Drainage systems
(e.g., edge drains) should be installed to facilitate the removal of water prior and subsequent
to the HMA overlay. Drainage is also important where there is a potential for water to move
into the pavement system due to capillary rise.
In general, geotextiles have not been found to be effective in reducing thermal cracking
within the overlay itself (Maxim Technologies, 1997). Pavement overlay systems have also
had limited success in areas of heavy rainfall and regions with significant freeze-thaw. Part
of the problem in cold regions is that when reflective cracks return, there is a potential for
water to be trapped in surficial cracks and to freeze (i.e., expanding and contracting to the
detriment of the overlay). Agencies in cold regions that have proactive crack sealing
programs (i.e., annual programs where any new or recurring cracks are sealed before the
winter) have reported success with geotextiles in overlays.
6.3-7 Design
General
Design of HMA overlays is thoroughly presented in the AASHTO Guide for Design of
Pavement Structures (1993). To have a high probability of success, a carefully planned and
executed study is required to develop an engineered overlay design using a geotextile
(Barksdale, 1991). A carefully planned and executed study also is required for successful,
alternative (i.e., non-geotextile) overlay designs.
The steps required to develop an overlay design for flexible pavements with a geotextile, as
summarized from Barksdale (1991), are as follows.

STEP 1.
Pavement condition evaluation.
The results of a general pavement condition survey are valuable in establishing the
type, severity, and extent of pavement distress. Candidate pavements should be
divided into segments, and a thorough visual evaluation made of each segment to
determine the type, extent, and severity of cracking, and to classify the present
distress as: alligator cracking, block cracking, transverse cracks, joint cracking,
FHWA NHI-07-092
Geosynthetics Engineering

6-10

Pavement Overlays
August 2008

patching, potholes, widening drop-offs, etc. Crack widths should be measured. (See
AASHTO Guide for Design of Pavement Structures (1993) for guidance.)
STEP 2.
Structural strength.
The overall structural strength of the pavement should be evaluated along its length,
using suitable nondestructive techniques, such as the Benkelman beam, falling-weight
deflectometer, Dynaflect, or Road Rater.
STEP 3.
Base/subgrade failure.
Areas with base or subgrade failures should be identified. Benkelman beam
pavement deflections greater than approximately 25 mil (0.6 mm) are indicative of
failure, as is excessive rutting.
STEP 4.
Remedial pavement treatment.
The results of the pavement condition survey and deflection measurements should be
used to develop a pavement repair strategy for each segment.
STEP 5.
Overlay design.
A realistic overlay thickness must be selected to ensure a reasonable overlay life.
Design methodologies are presented in the AASHTO Guide for Design of Pavement
Structures (1993). The overlay thickness with a geotextile should be determined as if
the interlayer is not present.
STEP 6.
Geotextile selection.
Geotextile selection requirements are reviewed in Section 6.3-8. A critical
performance element is the determination of the appropriate amount of tack coat,
which will depend on the amount required to saturate the old pavement and the
amount required to saturate the geotextile. The geotextile must be saturated to
provide adequate bond and optimum waterproofing and the amount will depend on
the geotextile type and mass per unit area. The heat from the new asphalt overlay
draws the tack coat up through the geotextile, saturating it and bonding it to the old
asphalt and overlay,
STEP 7.
Performance monitoring.
Performance monitoring during the life of the overlay is highly desirable for
developing a local data bank of performance histories using geotextiles in overlays.
Using a control section without a geotextile interlayer, with all other items equal, will
yield valuable comparative data.

FHWA NHI-07-092
Geosynthetics Engineering

6-11

Pavement Overlays
August 2008

The steps in developing an overlay design for PCC pavements where a geotextile may be
used are generally similar to those for flexible pavements (Barksdale, 1991). Vertical joint
deflection surveys should be performed. Full-width geotextiles should not be used when
vertical joint deflections are greater than about 8 mil (0.2 mm), unless corrective measures
such as undersealing, are taken to reduce joint movement. Horizontal thermal joint
movement should be less than about 50 mil (1.3 mm). As before, the thickness of the overlay
is not reduced with the use of a geotextile interlayer.
For chips seal applications, a tack coat is applied to the surface of the paved or unpaved road
as was done for the flexible pavement. The installed geotextile should be prepared to accept
the chip seal procedure by rolling it with a rubber tire roller. Rolling is required to help presaturate the geotextile. The chip seal is then installed on the geotextile in the same manner as
if the surface was a normal asphalt pavement.
Drainage Considerations
A primary function of the geotextile in an overlay is to minimize infiltration of surface water
into the pavement structure. The benefits of this are normally not quantified and, if
incorporated into the design, are only subjectively treated. One method to objectively
quantify the benefits of a moisture barrier is to estimate the effects of the asphaltimpregnated geotextile barrier on the drainage characteristics of the pavement structure. As
reviewed in Chapter 5, the 1993 AASHTO Guide for Design of Pavement Structures presents
provisions for modifying pavement design equations to take advantage of performance
improvements due to good drainage. Although not discussed in the design guide, a geotextile
overlay could be considered a method to improve drainage via reduced infiltration. As
previously indicated in Section 6.3-6, a properly installed geotextile interlayer system
reduces the permeability of a pavement by one to three orders of magnitude, thus becoming
an efficient moisture barrier to enhance pavement performance. The modified layer
coefficients for drainage m2 and m3 in the AASHTO (1993) equation for calculating the
structural number of a flexible pavement section and drainage coefficient Cd in the AASTHO
performance equation for calculating the thickness of a rigid pavement would be increased.
Again, these modification factors are not discussed for use with geotextile overlays in the
AASHTO design guide. However, if the inflow into a pavement system is reduced by an
order of magnitude through the use of a geotextile in the overlay, it follows that the percent
of time that the pavement structure is exposed to moisture levels approaching saturation is
reduced by at least one level (e.g., 5 - 25% to 1 - 5%). The drainage coefficients can be
accordingly increased to the values in the new level shown in Table 5-8 or 5-9 in Chapter 5.
These hypothesized values could aid in objectively estimating the structural benefit of a
geotextile moisture barrier.
FHWA NHI-07-092
Geosynthetics Engineering

6-12

Pavement Overlays
August 2008

6.3-8. Geotextile Selection


The selected paving grade geotextile must have the ability to absorb and retain the asphalt
tack coat to effectively form a waterproofing and stress-relief interlayer. The most common
paving grade geotextiles are needlepunched, nonwoven materials, with a mass per unit area
of 4 to 6 oz/yd2 (135 to 200 g/m2). These types of geotextiles are very porous and have a
high asphalt retention property that benefits the waterproofing and stress-reducing properties
of the paving geotextile layer. Thinner, heat-bonded geotextiles have also been used.
However, a significant variation in constructability and performance has been found between
different paving geotextiles.
Although lighter-weight (e.g., 3 to 4 oz/yd2 {100 to 135 g/m2}) geotextiles have been
previously used, both theory and a limited amount of field evidence indicate that geotextiles
with a greater mass per unit area (and a greater retention of asphalt), perform better than
lighter-weight geotextiles by further reducing stress at the tip of the underlying pavement
crack (Graf and Werner, 1993; Grzybowska et al., 1993; Walsh, 1993). As a result, the
2

current AASHTO specification requires a minimum of 4.1 oz/yd2 (140 g/m )


Heavier geotextiles are recommended where pavements have an extensive amount of cracks.
They have also been found to be more effective in cold regions. Numerical analysis indicates
2

that at some level of mass per unit area (e.g., > 6 to 7 oz/yd2 {200 g/m }), the bonding of the
overlay would be reduced due to shear on the geotextile (Grzybowska et al., 1993; Walsh,
1993).
For overlay design, the appropriate paving grade geotextile should be selected with
consideration given to pavement conditions, pavement deflection measurements, and the
overlay design traffic (EAL), as presented in Table 6-1.
6.3-9 Cost Considerations
The installed cost of the geotextile interlayer system includes the cost of the geotextile, the
additional tackcoat to saturate the geotextile, and installation. The design thickness of an AC
overlay with a geotextile interlayer should be determined as if the geotextile is not present.
The economic justification of geotextile use is then derived from:

An increase in pavement life, a decrease in pavement maintenance costs, and an


increase in pavement serviceability due to retardation and possible reduction of
reflection cracks.
An increased structural capacity due to drier base and subgrade materials.
A combination of the above two items.

FHWA NHI-07-092
Geosynthetics Engineering

6-13

Pavement Overlays
August 2008

Table 6-1
Paving Grade Geotextile Selection
Pavement Conditions
Rating1

Deflections
(mm)

Overlay Design Traffic


(EAL)

Lighter grade2
(~4.1 oz/yd2 {140 g/m2})

65 - 80

< 1.5

50,000

Medium grade
(~5 oz/yd2 {170 g/m2})

40 - 50

> 1.5

2,000,000

Class 3 (heaviest grade)


(~6 oz/yd2 {200 g/m2 })

20 30

> 1.5

> 2,000,000

Paving Grade Geotextile

1. From The Asphalt Institute (1983), and adopted from TRB Record No. 700:
65 - 80: Fairly good, slight longitudinal and alligator cracking. Few slightly rough and uneven.
40 - 50: Poor to fair, moderate longitudinal and alligator cracks. Surface is slightly rough and uneven.
20 - 30: Poor conditions with extensive alligator & longitudinal cracks. Surface is very rough and uneven.

Increased life of the overlay also lowers vehicle operating costs due to higher levels of
serviceability and lowers user delay costs due to future preventive and rehabilitative
maintenance interventions (Tighe et al., 2003 and Amini, 2005), which should also be
included in the economic analysis of these treatment.
The old pavement surface condition and overall installation play a very important role in the
performance of the paving geotextile. The deteriorated pavement should be repaired,
including filling joints and cracks and replacing sections with potholes and faults in their
base or subgrade. Under favorable conditions, reflection cracks can be retarded for
approximately 1 to 5 years as compared to the overlay without the paving grade geotextile.
The broad range is directly related to the load levels and magnitude of deformation at the
joint or crack.
The anticipated life improvement, under favorable conditions, is
approximately 100 to 200% that of an overlay of the same design thickness without a
geotextile. Favorable conditions for the use of a paving grade geotextile with pavement
repaving include:

The presence of fatigue-related pavement failure, evidenced by alligator cracks.


Pavement cracks no wider than 1/8 in. (3 mm) to prevent loss of tack coat into the
crack and correspondingly an unsaturated geotextile over the crack.
Maximum differential vertical movement at the joint or crack of less than 8 mil (0.2
mm).
The thickness of the new overlay designed to meet the structural requirements of the
pavement.

FHWA NHI-07-092
Geosynthetics Engineering

6-14

Pavement Overlays
August 2008

The economic benefit of the geotextile interlayer functioning as a moisture barrier is


currently not quantified in cost analyses. The effect of the geotextile on the quality of
drainage might be used to objectively estimate an increase in pavement structural capacity.
This increased capacity then can be used to estimate increased pavement life or to design a
thinner AC overlay.
These potential economic benefits can be combined for a particular project. Other cost
benefits, currently not quantified, include potential improvement of aesthetics and improved
ride quality. Alternatively, some engineers may reduce the overlay thickness based upon an
equivalent performance thickness to justify economics. Extensive research conducted by
Caltrans (Predoehl, 1990), implies that a geotextile interlayer is equivalent to 1.2 in (30 mm)
of asphalt concrete for relatively thin (i.e., 5 in. {120 mm}) overlays that are structurally
adequate. This equivalent value was confirmed in the study by Maxim Technologies (1997),
which reviewed the results of over 100 pavement sections (AC and PCC) on which
performance of the HMA overlay system with and without geotextiles was monitored. The
Maxim study also developed a pavement design model with consideration for both the
environmental and structural effects. Their analysis indicated a total average effect of 1.32 in
(33 mm) with 0.5 in (12.5 mm) environmental equivalent thickness benefit and 0.8 in. (20
mm) structural equivalent thickness benefit. A useful rule of thumb, based upon typical inplace costs, is that a geotextile interlayer is roughly equivalent to the cost of about 0.6 in (15
mm) of asphalt concrete. Cost of installed geotextile generally decreases with increased
quantities, simpler site geometry, and an experienced local installer.
Considerable insight into the economics of overlay design with geotextiles can be gained
from historic cost and performance data (Barksdale, 1991). The national and international
study by Maxim Technologies (1997) is just one example. A number of regional studies
have also been developed, many of these and the associated reports can be found on the
internet, including:

A regional study in Greenville County, South Carolina titled Study of Pavement


Maintenance Techniques used on Greenville County Maintained Roads (Sprague,
2005) in which overlay treatments on 370 roads were evaluated for a 6 year period of
time (available at: www.gmanow.com).

A synthesis of practice by the Mississippi Department of Transportation in


cooperation with FHWA titled Potential Applications of Paving Fabrics to Reduce
Reflective Cracking (Amini, 2005).
(available at: http://www.gomdot.com/research/pdf/PavFabr.pdf).

FHWA NHI-07-092
Geosynthetics Engineering

6-15

Pavement Overlays
August 2008

A synthesis of practice by the Texas Department of Transportation in cooperation


with FHWA titled Geosynthetics in Flexible and Rigid Pavement Overlay Systems
to Reduce Refection Cracking (Cleveland et al., 2002).
(available at: http://tti.tamu.edu/documents/1777-1.pdf).

A report on a test section in the county of San Diego, California where chip seals
were placed over geotextiles and found to have no reflection cracks after 17 years of
service. A life cycle cost analysis is provided showing the cost effectiveness of these
sections versus other treatments (Davis, 2005).
(available in the Transportation Research Circular Number E-C078 on Roadway
Pavement Preservation at: www.trb.org/publications/circulars/ec078.pdf).

A local study of geotextiles in chip seal treatments was performed by the county of
Sacramento Department of Transportation titled Chip Seal over Fabric Excelsior
Road in which geotextiles with two different binders under a variety of chip seals
were evaluated over a 6-year period of time.
(available at: www.aia-us.org/docs/SacramentoCountyChipOverFabricReport.pdf).

A final economic analysis issue is the probability of success. Geotextile interlayers, as well
as other rehabilitation techniques, are not always effective in improving pavement
performance. Therefore, an estimate of the probability of success should be included in all
economic analyses (Barksdale, 1991). The probability for success will obviously increase
with thoroughness of rehabilitation design, local experience with geotextile interlayers, and
thoroughness of construction inspection.

6.3-10 Specifications
The following example specification is a combination of the AASHTO M288 (2006)
geotextile material specification and its accompanying construction/installation guidelines.
The specification was based on the combined experience of the Texas and California
Departments of Transportation, which have had the greatest success using geotextiles in
pavement overlays.

FHWA NHI-07-092
Geosynthetics Engineering

6-16

Pavement Overlays
August 2008

PAVEMENT OVERLAY GEOTEXTILE SPECIFICATION


(after AASHTO M288, 2006)

1.

SCOPE

1.1

Description. This specification is applicable to the use of a paving fabric, saturated with
asphalt cement, between pavement layers. The function of the pavement geotextile is to act
as a waterproofing and stress relieving membrane within the pavement structure. This
specification is not intended to describe geotextile membrane systems specifically designed
for pavement joints and localized (spot) repairs.

2.

REFERENCED DOCUMENTS

2.1

AASHTO Standards
T88
Particle Size Analysis of Soils
T90
Determining the Plastic Limit and Plasticity Index of Soils
T99
The Moisture-Density Relationships of Soils Using a 5.5 lb (2.5 kg) Rammer and a
12 in. (305 mm) Drop

2.2

ASTM Standards
D 123 Standard Terminology Relating to Textiles
D 276 Test Methods for Identification of Fibers in Textiles
D 4354 Practice for Sampling of Geosynthetics for Testing
D 4439 Terminology for Geosynthetics
D 4632 Test Method for Grab Breaking Load and Elongation of Geotextiles
D 4759 Practice for Determining the Specification Conformance of Geosynthetics
D 4873 Guide for Identification, Storage, and Handling of Geotextiles
D 6140 Test Method for Determining the Asphalt Retention of Paving Fabrics

3.

PHYSICAL AND CHEMICAL REQUIREMENTS

3.1

Fibers used in the manufacture of paving fabrics and the threads used in joining paving
fabrics by sewing, shall consist of long chain synthetic polymers, composed of at least 95%
by weight polyolefins or polyesters. They shall be formed into a stable network such that the
filaments or yarns retain their dimensional stability relative to each other, including selvages.

3.2

Paving fabric Requirements. The paving fabric shall meet the requirements of following
Table. All numeric values in the following table represent minimum average roll values
(MARV) in the weaker principal direction (i.e., average test results of any roll in a lot
sampled for conformance or quality assurance testing shall meet or exceed the minimum
values).

FHWA NHI-07-092
Geosynthetics Engineering

6-17

Pavement Overlays
August 2008

Paving Fabric Property Requirementsa


Property

Test Method

Units

Requirements

Grab Strength

ASTM D 4632

lb (N)

100 (450)

Mass per Unit Area

ASTM D 3776

oz/yd2 (gm/m2)

4.1 (140)

Ultimate Elongation

ASTM D 4632

> 50

Asphalt Retention(1)

ASTM D 6140

gal/yd2 (l/m2)

(b,c)

Melting Point

ASTM D 276

F (C)

300 (150)

NOTES:
(a)
All numeric values represent MARV in the weaker principal direction.
(b)
Asphalt required to saturate paving fabric only. Asphalt retention must be provided in
manufacturer certification. Value does not indicate the asphalt application rate
required for construction. Refer to M288 for discussion of asphalt application rate.
(c)
Product asphalt retention property must meet the MARV provided by the manufacturers
certification.

4.

CERTIFICATION

4.1

The Contractor shall provide to the Engineer a certificate stating the name of the
manufacturer, product name, style number, chemical composition of the filaments or yarns
and other pertinent information to fully describe the paving fabric.

4.2

The Manufacturer is responsible for establishing and maintaining a quality control program to
assure compliance with the requirements of the specification. Documentation describing the
quality control program shall be made available upon request.

4.3

The Manufacturers certificate shall state that the furnished paving fabric meets MARV
requirements of the specification as evaluated under the Manufacturers quality control
program. A person having legal authority to bind the Manufacturer shall attest to the
certificate.

4.4

Either mislabeling or misrepresentation of materials shall be reason to reject those paving


fabric products.

5.

SAMPLING, TESTING, AND ACCEPTANCE

5.1

Paving fabrics shall be subject to sampling and testing to verify conformance with this
specification. Sampling shall be in accordance with the most current ASTM D 4354 using
the section titled, Procedure for Sampling for Purchasers Specification Conformation
Testing. In the absence of purchasers testing, verification may be based on manufacturers
certifications as a result of a testing by the manufacturer of quality assurance samples

FHWA NHI-07-092
Geosynthetics Engineering

6-18

Pavement Overlays
August 2008

obtained using he procedure for Sampling or Manufacturers Quality Assurance (MQA)


Testing. A lot size shall be considered to be the shipment quantity of the given product or a
truckload of the given product, whichever is smaller.
5.2

Testing shall be performed in accordance with the methods referenced in this specification for
the indicated application. The number of specimens to test per sample is specified by each
test method. Paving fabric product acceptance shall be based on ASTM D 4759. Product
acceptance is determined by comparing the average test results of all specimens within a
given sample to the specification MARV. Refer to ASTM D 4759 for more details regarding
paving fabric acceptance procedures.

6.

SHIPMENT AND STORAGE

6.1

Paving fabric labeling, shipment, and storage shall follow ASTM D 4873. Product labels
shall clearly show the manufacturer or supplier name, style number, and roll number. Each
shipping document shall include a notation certifying that the material is in accordance with
the manufacturers certificate.

6.2

Each paving fabric roll shall be wrapped with a material that will protect the paving fabric,
including the ends of the rolls, from damage due to shipment, water, sunlight, and
contaminants. The protective wrapping shall be maintained during periods of shipment and
storage.

6.3

During storage, paving fabric rolls shall be elevated off the ground and adequately covered to
protect them from the following: site construction damage, precipitation, extended ultraviolet
radiation including sunlight, chemicals that are strong acids or strong bases, flames including
welding sparks, temperatures in excess of 160F (71C), and any other environmental
condition that may damage the physical property values of the paving fabric.

7.

MATERIALS

7.1

Sealant. The sealant material used to impregnate and seal the paving fabric, as well as bond
it to both the base pavement and overlay, shall be a paving grade asphalt recommended by the
paving fabric manufacturer, and approved by the engineer.
a. Uncut asphalt cements are the preferred sealants; however, cationic and anionic emulsions
may be used provided the precautions outlined in M288 are followed. Cutbacks and
emulsions which contain solvents shall not be used.
b. The grade of asphalt cement specified for hot-mix design in each geographic location is
generally the most acceptable material.

FHWA NHI-07-092
Geosynthetics Engineering

6-19

Pavement Overlays
August 2008

7.2

Sand. Washed concrete sand may be spread over an asphalt saturated paving fabric to
facilitate movement of equipment during construction or to prevent tearing or delamination of
the paving fabric. Hot-mix broadcast in front of construction vehicle tire may also be used to
serve this purpose. If sand is applied, excess quantities shall be removed from the paving
fabric prior to placing the surface course.
a. Sand is not usually required. However, ambient temperatures are occasionally sufficiently
high to cause bleed-through of the asphalt sealant resulting in undesirable paving fabric
adhesion to construction vehicle tires.

8.

EQUIPMENT

8.1

The asphalt distributor shall be capable of spraying the asphalt sealant at the prescribed
uniform application rate. Not streaking, skipping, or dripping will be permitted. The
distributor shall also be equipped with a hand spray having a single nozzle and positive shutoff valve.

8.2

Mechanical or manual lay down equipment shall be capable of laying the paving fabric
smoothly.

8.3

The following miscellaneous equipment shall be provided: stiff bristle brooms or squeegees
to smooth the paving fabric; scissors or blades to cut the paving fabric; brushes for applying
asphalt sealant to paving fabric overlaps.

8.4

Pneumatic rolling equipment to smooth the paving fabric into the sealant, and sanding
equipment may be required for certain jobs. Rolling is especially required on jobs where thin
lifts or chip seals are being placed. Rolling helps ensure paving fabric bond to the adjoining
pavement layers in the absence of heat and weight associated with thick lifts of asphaltic
pavement.

9.

CONSTRUCTION

9.1

Neither the asphalt sealant nor the paving fabric shall be placed when weather conditions, in
the opinion of the Engineer, are not suitable. Air and pavement temperatures shall be
sufficient to allow the asphalt sealant to hold the paving fabric in place. For asphalt cements,
air temperature shall be 50F (10C) and rising. For asphalt emulsions, air temperature shall
be 60F (15C) and rising.

9.2

The surface on which the paving fabric is to be placed shall be reasonably free of dirt, water,
vegetation, or other debris. Cracks exceeding 0.1 in. (3 mm) in width shall be filled with a
suitable crack filler. Potholes shall be properly repaired as directed by the Engineer. Fillers
shall be allowed to cure prior to paving fabric placement.

FHWA NHI-07-092
Geosynthetics Engineering

6-20

Pavement Overlays
August 2008

9.3

The specified rate of asphalt sealant application must be sufficient to satisfy the asphalt
retention properties of the paving fabric, and bond the paving fabric and overlay to the old
pavement.
NOTE: When emulsions are used, the application rate must be increased
to offset water content of the emulsion.
a. Application of the sealant shall be by distributor spray bar, with hand spraying kept to a
minimum. Temperature of the asphalt sealant shall be sufficiently high to permit uniform
spray pattern. For asphalt cements the minimum temperature shall be 290F (145C). To
avoid damage to the paving fabric, however, the distributor tank temperature shall not exceed
320F (160C).
b. Spray patterns for asphalt emulsion are improved by heating. Temperatures in the 130F
to 160F (55C to 70C) range are desirable. A temperature of 160F (70C) shall not be
exceeded since higher temperatures may break the emulsion.
c. The target width of asphalt sealant application shall be the paving fabric width plus 6 in.
(150 mm). The asphalt sealant shall not be applied any farther in advance of paving fabric
placement than the distance the contractor can maintain free of traffic.
d. Asphalt spills shall be cleaned from the road surface to avoid flushing and paving fabric
movement.
e. When asphalt emulsions are used, the emulsion shall be cured prior to placing the paving
fabric and final wearing surface. This means essentially no moisture remaining.

9.4

The paving fabric shall be placed onto the asphalt sealant with minimum wrinkling prior to
the time the asphalt has cooled and lost tackiness. As directed by the Engineer, wrinkles or
folds in excess of 1 in. (25 mm) shall be slit and laid flat.
a. Brooming and/or pneumatic rolling will be required to maximize paving fabric contact
with the pavement surface.
b. Overlap of paving fabric joints shall be sufficient to ensure full closure of the joint, but
should not exceed 6 in. (150 mm). Transverse joints shall be lapped in the direction of
paving to prevent edge pickup by the paver. A second application of asphalt sealant to the
paving fabric overlaps will be required if in the judgement of the Engineer additional asphalt
sealant is needed to ensure proper bonding of the double paving fabric layer.
c. Removal and replacement of paving fabric that is damaged will be the responsibility of the
contractor.

FHWA NHI-07-092
Geosynthetics Engineering

6-21

Pavement Overlays
August 2008

NOTE: The problems associated with wrinkles are related to the


thickness of the asphalt lift being placed over the paving fabric. When
wrinkles are large enough to be folded over, there usually is not enough
asphalt available from the tack coat to satisfy the requirement of multiple
layers of paving fabric. Therefore, wrinkles should be slit and laid flat.
Sufficient asphalt sealant should be sprayed on the top of the paving
fabric to satisfy the requirement of the lapped paving fabric.
NOTE: In overlapping adjacent rolls of paving fabric it is desirable to
keep the lapped dimension as small as possible and still provide a
positive overlap. If the lapped dimension becomes too large, the
problem of inadequate tack to satisfy the two lifts of paving fabric and
the old pavement may occur. If this problem does occur then additional
asphaltic sealant should be added to the lapped areas. In the application
of the additional sealant, care should be taken not to apply too much
since an excess will cause flushing.
d. Trafficking the paving fabric will be permitted for emergency and construction vehicles
only.
9.5

Placement of the hot-mix overlay should closely follow paving fabric laydown. The
temperature of the mix shall not exceed 320F (160C). In the event asphalt bleeds through
the paving fabric causing construction problems before the overlay is placed, the affected
areas shall be blotted by spreading sand. To avoid movement of, or damage to the seal
coat saturated paving fabric, turning of the paver and other vehicles shall be gradual and kept
to a minimum.

9.6

Prior to placing a seal coat (or thin overlay such as an open-graded friction course), lightly
sand the paving fabric at a spread rate of 1.2 to 1.8 lb per square yard (0.65 to 1 kg per square
meter), and pneumatically roll the paving fabric tightly into the sealant.
ADVISORY It is recommended that for safety considerations,
trafficking of the paving fabric should not be allowed. However, if the
contracting agency elects to allow trafficking, the following verbiage is
recommended: If approved by the Engineer, the sealant saturated
paving fabric may be opened to traffic for 24 to 48 hours prior to
installing the surface course. Warning signs shall be placed which advise
the motorist that the surface may be slippery when wet. The signs shall
also post the appropriate safe speed. Excess sand shall be broomed from
the surface prior to placing the overlay. If, in the judgement of the
Engineer, the paving fabric lacks tackiness following exposure to traffic,
a light tack coat shall be applied prior to the overlay.

FHWA NHI-07-092
Geosynthetics Engineering

6-22

Pavement Overlays
August 2008

10.

METHOD OF MEASUREMENT
10.1 The paving fabric shall be measured by the number of square yards (square meters)
computed from the payment lines shown on the plans or from payment lines established in
writing by the Engineer. This excludes seam overlaps.
10.2 Asphalt sealant for the paving fabric will be measured by the liter

11.

BASIS OF PAYMENT
11.1 The accepted quantities of paving fabric shall be paid for per square yard (square meter)
in place.
11.2

The accepted quantities of asphalt sealant for the paving fabric will be paid for at the
contract unit price per gallon (liter) complete in place.

11.3 Payment will be made under:


Pay Item

Pay Unit

Pavement Overlay Paving fabric


Asphalt Sealant

Square Yard (Square Meter)


Gallon (Liter)

6.3-11 Field Inspection


Because proper construction procedures are essential for a successful AC overlay project
with geotextiles, good field inspection is very important. Prior to construction, the field
inspector should review the guidelines in Section 1.7. Most geotextile manufacturers and
suppliers will provide technical assistance during the initial stages of a fabric overlay project.
This assistance may be particularly beneficial to inexperienced inspectors and contractors.
One construction problem observed in some installations is the placement of insufficient tack
coat. Tack coat should be a separate pay item, and field inspection should monitor the
quantity of tack coat placed. Monitoring can be done by gauging or by placing a test coupon
(e.g., a square piece of geotextile or fiber board) and weighing it.
6.3-12 Recycling
AC overlays used with a geotextile can be recycled. The most common practice is to mill
down to just above the geotextile interlayer (i.e., within about in. {12.5 mm}). This
process maintains the benefits of the interlayer when the recycled overlay is replaced
FHWA NHI-07-092
Geosynthetics Engineering

6-23

Pavement Overlays
August 2008

(Marienfeld and Smiley, 1994). Alternatively, milling may include the geotextile interlayer.
A detailed study of recycling a nonwoven polypropylene geotextile (Christman, 1981)
concluded that overlay with geotextile interlayers does not pose any problem to the milling
operation. Chisel milling teeth rather than conical teeth and slower forward speed can be
used to produce the smallest geotextile pieces (Dykes, 1980, Cleveland et al., 2002). Proper
installation, especially the proper amount of tackcoat improves bonding and thus the milling
operation. Also, milling to a depth of at least in. (12.5 mm) beneath the geotextile is
recommended to prevent going in and out of the fabric. No apparent differences have been
noted in the properties of dryer-drum recycled mixtures (50-50 blend) containing or lacking
geotextiles. Texas DOT indicates that heater scarification can cause problems when a
geosynthetic is present (Cleveland, et al., 2002). Texas DOT also indicates that cold milling
does not usually present problems and a typical geotextile does not significantly affect
mixture properties. However, cold recycling has been reported to create problems with
equipment entanglement when the geotextile is not adequately milled. For cold recycling,
the FHWA Pavement Preservation Checklist for Cold In-place Asphalt Recycling
recommends that 90 percent of the geotextile pieces should be milled to a size of less than 2
in.2 (1200 mm2) and that the maximum length of any individual piece is 100 mm (4 in.)
(FHWA-IF-06-012, 2005). Milling rates may have to be adjusted and screens may have to
be added or removed in order to meet these requirements.

6.4

GEOGRIDS

6.4-1 Geogrid Functions


High-strength and high-stiffness polymer and fiberglass geogrids have been used for fullwidth and strip overlay applications as well as in new asphalt pavements. The geogrids in
this application primarily function as a reinforcing element, provided they are sufficiently
stiff (i.e., the elastic modulus of the geogrid must be greater than that of the surrounding
HMA {Lytton, 1989}).
6.4-2 Applications
The primary application of geogrids is as a stress resistance layer in HMA overlays to reduce
the development of reflection cracks from either old AC or PCC pavements. Research at the
University of Nottingham has indicated that geogrids can also reduce rutting in the HMA
overlay layer when placed over AC pavements (Brown, 2006). The research at the
University of Nottingham found that the geogrid increased the life to critical pavement rut
depth by a factor of three.

FHWA NHI-07-092
Geosynthetics Engineering

6-24

Pavement Overlays
August 2008

Geogrids have also been used in new construction to increase the fatigue life of AC
pavements and reduce rutting. The University of Nottingham has performed extensive
research on this application. They have developed the relationship between the maximum
tensile strain and the number of cycles to produce failure through cracking, which is
modified by the presence of a high tensile modulus polymeric geogrid as shown in Figure 6-3
(Brown, 2006). This figure was developed for use in mechanistic design and should be
verified through monitored field performance data before its extended use.
6.4-3 Design
General
Unfortunately, most information on these applications is proprietary and independent
published information on geogrid reinforcement in HMA overlays and pavement systems is
somewhat limited. Most of the research has been experimental work in the laboratory
supported by theoretical evaluation with notable work by West Virginia University (Kutuk
and Siriwardane, 1998) and the University of Nottingham (Brown et al., 2001). Both studies
demonstrated the ability of geogrids to arrest reflection crack propagation. Full scale field
performance is needed to verify these results. Design guidance and recommended property
values are not available. Research is still in progress to develop a mechanistic design method
for reinforced overlays which takes into account all the relevant parameters.

Vertical Subgrade Strain


(microstrain)

Load Applications (Millions) to Critical Cracking Conditions (msa)


Figure 6-3.

Relationship between the vertical compressive strain at the top of the subgrade
and the number of load applications in geogrid reinforced pavement (after
Brown, 2006).

FHWA NHI-07-092
Geosynthetics Engineering

6-25

Pavement Overlays
August 2008

Geogrid Selection
The key design parameter is the stiffness of the geogrid. However, guidance is not available
as to the minimum modulus requirements. Other important parameters include the shear
strength of interface, the geogrid location (at the bottom of the overlay and/or between HMA
layers, and installation of the geogrid). Brown (2006) notes that the interface shear strength
is important since any serious reduction in the continuity between layers on either side of the
reinforcement will reduce the structural integrity of the pavement. Brown further states that
a balance has to be reached between the provision of crack inhibiting reinforcement and the
need for an adequate shear strength since there will usually be some reduction in the latter.
Again, however, protocol for measuring the interface value has not been developed as a
standard of practice either through ASTM or AASHTO. The optimum installation location
does appear to be at the base of the overlay, with some additional improvements noted by
both West Virginia University and the University of Nottingham when two layers were used
(i.e., one at the base and the other at an intermediate level in the asphalt).
6.4-4 Installation
Geogrid installation follows the same placement as geotextiles. As with geotextiles in
overlays, high quality construction is required to achieve the benefits of the reinforcement.
To achieve proper interface conditions, care should again be taken to insure the proper type,
application, and application rate of asphalt sealant. The temperature of the asphalt sealant
and hot mix asphalt should also be monitored and controlled to resist damage to the geogrid.
6.4-5 Cost Considerations (Brown, 2006)
The cost of geogrid reinforcement is relatively high compared to geotextiles and adds
substantially to the cost of an overlay. However, the potential performance may provide
significant cost-benefit. Based on beam testing of three reflection crack treatments,
Caltabiano and Brunton (1991) developed the following relative cost comparison of various
overlay treatments.
Overlay
Standard asphalt
Polymer modified asphalt
Geotextile
Geogrid

Relative Life Relative Costs


1.0
-2.5
2.5
5.0
1.0
10.0
2.0

This information indicates that both geosynthetics were more effective than the polymer
modified asphalt.

FHWA NHI-07-092
Geosynthetics Engineering

6-26

Pavement Overlays
August 2008

6.4-6 Specifications
The following example geogrid specification is from the Texas Department of Transportation
(Cleveland et al., 2002), who have had good success using geosynthetics in pavement
overlays.
SPECIAL SPECIFICATION
31xx
REINFORCING GRID FOR JOINT REPAIR
1. Description. Install a reinforcing grid in accordance with the details shown on the plans to
prevent reflective cracking of transverse and longitudinal joints in asphaltic paving overlay
mixtures.
2. Materials.
(1) Reinforcing Grid. Provide grid meeting Table xxx, Reinforcing Grid for Joint Repair.
Use roll widths shown on the plans or as approved.
(2) Asphalt. Provide the grade of asphalt for tack coat as shown on the plans and in
accordance with Item 300, Asphalts, Oils, and Emulsions.
(3) Packaging Requirements. Ensure each roll of grid is packaged individually in a suitable
wrapper to help protect from damage due to ultra-violet light and moisture during normal
storage and handling.
(4) Storage Requirements. Store material in dry covered conditions free from dust. Store
vertically to avoid misshapen rolls.
(5) Identification Requirements. Label or tag such that the sample identification can be
read without opening the roll packaging. Label each roll with the manufacturers name,
job number, loom number, production date and shift, tare weight of packaging material,
width and length of grid on the roll, and net weight of the grid.
(6) Safety Precautions. Gloves are recommended to prevent contact with the material.
3

Equipment. For longitudinal full-width reinforcing grid, provide applicable equipment in


accordance with Item 316, Surface Treatments.

3. Construction. Apply reinforcing grid when air temperature is above 50F (122C) and rising. Do
not apply grid when air temperature is 60F (140C) and falling. In all cases, do not apply grid
when surface temperature is below 60F (140C). Do not apply when, according to the Engineer,
weather conditions are not suitable. Measure air temperature in the shade and away from artificial
heat. Cease reinforcing grid installation if the Engineer determines that weather conditions
prevent proper placement.
(1) Surface Preparation. Prepare the surface by cleaning off dirt, dust, or other debris. Set string
lines for alignment, if required. Remove existing raised pavement markers in accordance with
the plans. When shown on the plans, remove vegetation and blade pavement edges. Fill
cracks exceeding 1/8 inch (3 mm) in width with approved crack filler. Fill cracks exceeding 1
inch in width with a fined grained bituminous mixture or other approved material. Ensure
crack sealing material is flush with the existing pavement surface. Repair faulted cracks or
FHWA NHI-07-092
Geosynthetics Engineering

6-27

Pavement Overlays
August 2008

joints with vertical deformation greater than inch (13 mm) with a fine grained bituminous
mixture or other approved material. Ensure crack filler and patching materials are cured prior
to the placement of the level-up paving mixture.
(2) Asphalt Binder Application. Apply:
(a) with an asphalt distributor unless otherwise approved,
(b) at the rate shown on the plans or as directed,

(c) within 15F (122C) of the temperature selected by the Engineer not to exceed 320F,
(d) approximately 6 in. (150 mm) outside the reinforcing grid width,
(e) with paper or other approved material at the beginning and end of each shot to construct a
straight transverse joint and to prevent overlapping of the asphalt. Unless otherwise
approved, match longitudinal joints with the lane lines. The Engineer may require a string
line, if necessary, to keep joints straight with no overlapping. Do not contaminate asphalt
binder.
(3) Level-Up Paving Mixture. Place and compact a fine-grained paving mixture as a leveling
course in accordance with the specifications shown on the plans. Unless otherwise shown, the
compacted target lift thickness is between 3/4 and 1 in. (19 to 25 mm).
(4) Reinforcing Grid Placement. Unless otherwise directed, furnish the Engineer with
manufactures recommendations for reinforcing grid installation. If required by the Engineer,
ensure a manufacturers representative is present on site during the first three days of the grid
placement.
(a) Installation. Apply asphalt binder at the rate shown on the plans, unless otherwise
directed. Install grid either by hand or mechanical means under sufficient tension to
eliminate ripples and provide sufficient adhesion to avoid dislodging of the grid. Should
ripples occur, these must be removed by pulling the grid tight or in extreme cases, for
example, in tight radius, by cutting and laying flat. A sharp knife may be used for cutting.
Roll the grid surface with a rubber coated drum roller or pneumatic tire roller to seat grid.
Tires must be cleaned regularly with an approved asphalt-cleaning agent.
(b) Transverse Joints. Overlap transverse joints in the direction of the paver a minimum of
3 inches (75 mm) with the top layer in the direction of traffic. If required, apply
additional asphalt binder to secure overlapping grid layer.
(c) Longitudinal Joints. Overlap longitudinal joints a minimum of 1 in. (25 mm). If
required, apply additional asphalt binder to make secure overlapping grid layer.
After the rolling is completed, construction and emergency traffic may drive on the grid.
Minimize turning movements of paving machinery. Minimize braking from vehicles by
installing appropriate signs at intersections and driveways. Remove and patch damaged
sections. No payment will be made for repair work. All grid placed in a day shall be covered
with asphaltic concrete the same day, within permissible laying temperatures, and compacted
in accordance with applicable specifications as shown on the plans.
5. Measurement. The reinforcing grid will be measured by the linear foot (meter) of joint or crack
repaired or by the square yard (square meter) of the actual area complete in place. No allowance
will be made for overlapping at joints.

FHWA NHI-07-092
Geosynthetics Engineering

6-28

Pavement Overlays
August 2008

6. Payment. The work performed and materials furnished in accordance with this Item and
measured as provided under "Measurement" will be paid for at the unit price bid for "Reinforcing
Grid for Joint Repair" of the type specified and by the width for the linear foot (meter)
measurement. This price shall be full compensation for cleaning the existing pavement; for
furnishing, preparing, hauling and placing all materials; for all manipulation, including rolling,
and for all labor, tools, equipment and incidentals necessary to complete the work

6.5

GEOCOMPOSITES

Geogrid-geotextile composites are available; however, there is limited experience with these
products to date. The intent of such a composite is to have a material that installs similarly to
geotextile overlays. A properly installed composite should function as stress-relieving
interlayer, moisture barrier, and reinforcement (provided the geogrid has a high modulus).
6.5-1 Membrane and Composite Strips
General
A variety of commercially available, heavy-duty membrane strips are used over cracks and
joints of PCC pavements that are overlaid with AC. Typically, these materials are
composites of woven or nonwoven geotextiles and modified asphalt membranes. Materials
of single-layer geotextiles with rubber-asphalt membranes are typically used for strip
waterproofing. Materials of double-layer geotextiles that sandwich a modified asphalt
membrane are typically used to reduce and retard reflective cracking. Crack reduction
interlayers are typically 0.14 in. (3.5 mm) thick and are capable of maintaining 95% of their
thickness during installation and in-service use. Interlayer strips are typically 1 to 3 ft (0.3 to
1 m) wide, and usually weigh 47 to 94 oz/yd2 (1600 to 3300 g/m2 ) -- which is heavier than
the typical geotextile interlayer that weighs about 38 oz/yd2 (1300 g/m2) with asphalt
impregnation (Barksdale, 1991).
A composite that combines both stress relief and stress resistance components was developed
at the University of Illinois (Dempsey, 2002). The geocomposite is known as the interlayer
stress-absorbing composite (ISAC) system and was developed and evaluated for the purpose
of effectively alleviating or mitigating the problem of reflection cracking in an asphalt
concrete (AC) overlay. ISAC materials and performance properties were carefully selected
through comprehensive theoretical studies and laboratory evaluation programs. The ISAC
system consists of a low-stiffness geotextile as the bottom layer, a viscoelastic membrane
layer as the core, and a very high stiffness geotextile for the upper layer. In order to evaluate
the effectiveness of the ISAC layer to control reflective cracking, a laboratory pavement
section with an AC overlay placed on a jointed Portland cement concrete slab was
constructed and tested in an environmental chamber. A mechanical device was used to
FHWA NHI-07-092
Geosynthetics Engineering

6-29

Pavement Overlays
August 2008

simulate thermal strain in the slab, and the joint was opened and closed at an extremely slow
rate. The testing was conducted at 30F (-1C), and strain in the overlay was monitored using
a sensitive linear variable differential transducer device. The force required to pull and push
the slab was monitored using a load cell placed between the slab and a hydraulic ram.
Performance of the ISAC system was evaluated by comparing the cycles to failure of an
ISAC-treated overlay with the performance of a control section without ISAC and of test
sections containing two commercially available products. The base isolation properties of
the ISAC system were demonstrated in the laboratory evaluation studies, which indicated
that the ISAC system vastly outperformed the control section and the sections with the two
commercial products tested. Several years of field performance testing have shown that the
ISAC system is highly effective for mitigating reflective cracking in AC overlays used on
both airport and highway pavement systems (Dempsey, 2002).
Installation
The installation of heavy-duty membranes is relatively easy. Usually the manufacturer's
installation recommendations are followed because of the wide variation of products and
installation requirements. Single-, two-, or three-step installation processes are required for
the various products.
Advantages
Advantages of strips include:
Limited area of installation, and, therefore, less potential installation problems.
Factory-applied asphalt, and, therefore, less field variances.
Heavier weight, and possible function as a stress-relieving membrane interlayer of the
material.
The moderate amount of documented field performance data developed to date has been
summarized by Barksdale (1991). FHWA-NHI 131008 reports that several states have had
success with these treatments on both longitudinal and transverse joints and cracks and
provides the results of several studies.
Disadvantages
The cost of membrane and composite strips is relatively high (on the order of $4 to $15 per
square yard ($4.40 to $16.40 per square meter) installed and the cost-effectiveness of this
approach has not been established.
6.5-2 Specifications
The construction specifications for geocomposites are similar to the specifications presented
for geotextiles and geogrids, except that the geocomposites often come with an adhesive
binder and may not require a tack coat.
FHWA NHI-07-092
Geosynthetics Engineering

6-30

Pavement Overlays
August 2008

6.6

REFERENCES

Key references are noted in bold type.


AASHTO (2006). Standard Specifications for Geotextiles - M 288, Standard
Specifications for Transportation Materials and Methods of Sampling and
Testing, 26th Edition, American Association of State Transportation and
Highway Officials, Washington, D.C.
AASHTO (1993). AASHTO Guide for Design of Pavement Structures, American
Association of State Highway and Transportation Officials, Washington, D.C.
Aldrich, R.C. (1986). Evaluation of Asphalt Rubber and Engineering Fabrics as Pavement
Interlayers, Misc. Paper GL-86-34 (untraced series N-86), 7 p.
Amini, F. (2005). Potential Applications of Paving Fabrics to Reduce Reflective Cracking,
FHWA/MS-DOT-RD-174, performed in cooperation with the Mississippi
Department of Transportation, Jackson, Mississippi and U.S. Department of
Transportation, Federal Highway Administration, Washington D.C., 45 p.
Barksdale, R.D. (1991). Fabrics in Asphalt Overlays and Pavement Maintenance,
National Cooperative Highway Research Program Report 171, Transportation
Research Board, National Research Council, Washington. D.C., 72 p.
Brown, S.F. (2006). Geosynthetics in Asphalt Pavements, International Geosynthetic Society
Mini Lecture No. 18 and 19 GSA ID#132, http://www.geosynthetica.net/tech
_docs.asp.
Brown, N. R. (2003). Solution for Distressed Pavements and Crack Reflection,
Transportation Research Record, No. 1819, 2003, pp. 313-317.
Brown, S.F., Thom, N.H. and Saunders P. J. (2001). A Study of Grid Reinforced Asphalt to
Combat Reflection Cracking. Asphalt Paving Technology, Vol. 70, pp 543-570.
Caltabiano, M.A. and Brunton, J.M. (1991). Reflection Cracking in Asphalt Overlays,
Asphalt Paving Technology: Proceedings-Association of Asphalt Paving Technical
Sessions[C].USA: Assoc. of Asphalt Paving Technologists.
Christman, R. (1981). Material Properties and Equipment Capabilities Resulting From
Recycling Bituminous Concrete Pavements Containing Petromat, Pavement
Resource Managers, Newington, CT, 139 p.
Cleveland, G.S., Button, J.W. and Lytton, R.L. (2002). Geosynthetics in Flexible and Rigid
Pavement Overlay Systems to Reduce Reflection Cracking, FHWA/TX-02/1777-1,
performed in cooperation with the Texas Department of Transportation and the U.S.
FHWA NHI-07-092
Geosynthetics Engineering

6-31

Pavement Overlays
August 2008

Department of Transportation, Federal Highway Administration, Washington D.C.,


298 p.

Davis, L. (2005). Chip Sealing over Fabric in Borrego Springs, California, Roadway
Pavement Preservation 2005, Transportation Research Circular Number E-C078,
Transportation Research Board, Washington D.C., pp. 42-53.
Dempsey, B. J. (2002). Development and Performance of Interlayer Stress-Absorbing
Composite in Asphalt Concrete Overlays, Transportation Research Record 1809, pp.
175-183.
Dykes, J.W. (1980). Use of Fabric Interlayers to Retard Reflection Cracking, Proceedings,
AAPT, Volume 49, pp. 354-568.
FHWA (current edition). Pavement Rehabilitation Manual, Pavement Division, Office of
Highway Operations.
FHWA (2005). Pavement Preservation Checklist Series 12- Cold In-Place Asphalt Recycling
Application Checklist, FHWA Publication Number FHWA-IF-06-012, Federal
Highway Administration, Washington, D.C., November, 2005,
FHWA (1998). FHWA-NHI 131008, Techniques for Pavement Rehabilitation, Reference
Manual, NHI Training Course.
GMA (2007). Personal correspondence in the review of this manual.
Graf, B., and Werner. G. (1993). Design of Asphalt Overlay Fabric System Against
Reflective Cracking, Proceedings of the Second RILEM-Conference on Reflective
Cracking in Pavements, E & FN Spon, Liege, Belgium, pp. 159-168.
Gransberg, D. and James, D.M.B. (2005). Chip Seal Best Practices, National Cooperative
Highway Research Program, Synthesis of Highway Practice 342, Transportation
Research Board, National Academy Press, Washington, D.C., 111 p.
Grzybowska,W., Wojtowicz, J., and Fonkerko, L. (1993). Application of Geo-Synthetics to
Overlays in Cracow Region of Poland. Proceedings of the Second RILEMConference on Reflective Cracking in Pavements, E & FN Spon, Liege, Belgium, pp.
290298.
Jayawickrama, P. W. and Lytton, R. (1987). Methodology for Predicting Asphalt Concrete
Overlay Life against Reflective Cracking, Proceedings of the Sixth International
Conference on the Structural Design of Asphalt Pavements, Volume I.
Kutuk, B. and Siriwardane, H.J. (1998). Performance of Flexible Pavements Reinforced with
Geogrids, Department of Civil and Environmental Engineering, West Virginia
University, WVDOF RP 98/CFC 98-259, performed for the West Virginia
Department of Transportation, Division of Highways.
FHWA NHI-07-092
Geosynthetics Engineering

6-32

Pavement Overlays
August 2008

Lytton, R. L. (1989). Use of Geotextiles for Reinforcement and Strain Relief in Asphalt
Concrete, Geotextiles and Geomembranes, Vol. 8, 217-237.
Marienfeld, M.L. and Baker, T.L. (1999). Paving Fabric Interlayer as a Pavement Moisture
Barrier, Transportation Research Circular Number E-C006, Transportation Research
Board, Washington D.C., 24 p. http://gulliver.trb.org/publications/circulars/ec006
Marienfeld, M.L. and Smiley, D. (1994). Paving Fabrics: The Why and the How-To,
Geotechnical Fabrics Report, Vol. 12, No. 4, pp 24-29.
Maxim Technologies, Inc. (1997). Nonwoven Paving Fabrics Study Final Report,
Prepared for the Industrial Fabrics Association International - Geotextile
Division, available at: www.gma.now.com.
Monismith, C.L. and Long, F. (1999). Overlay Design for Cracked and Seated Portland
Cement Concrete (PCC) Pavement Interstate Route 710, Technical Memorandum
TM UCB PRC 99-3 prepared for the Long Life Pavement Task Force by the
Pavement Research Center, Institute for Transportation Studies, University of
California, Berkeley, 37 P.
Predoehl, N.H. (1990). Evaluation of Paving Fabric Test Installations in California - Final
Report, FHWA/CA/TL-90/02, Office of Transportation Materials and Research,
California Department of Transportation, Sacramento, CA.
Sprague, C.J. (2005). Study of Pavement Maintained Techniques Used on Greenville County
Maintained Roads, Phase 2 Report, submitted by TRI/Environmental, Inc. to the
Geosynthetic Materials Association, available at: www.gma.now.com
Tighe, S., Hass, R., and Ponniah, J. (2003). Life Cycle Cost Analysis of Mitigating Reflective
Cracking, Transportation Research Record, No. 1823, pp. 73-79.
Walsh, I.D. (1993). Thin Overlay to Concrete Carriageway to Minimize Reflective Cracking.
Proceedings of the Second RILEM-Conference on Reflective Cracking in Pavements,
E & FN Spon, Liege, Belgium.

FHWA NHI-07-092
Geosynthetics Engineering

6-33

Pavement Overlays
August 2008

FHWA NHI-07-092
Geosynthetics Engineering

6-34

Pavement Overlays
August 2008

7.0 REINFORCED EMBANKMENTS ON SOFT FOUNDATIONS

7.1

BACKGROUND

Embankments constructed on soft foundation soils have a tendency to spread laterally


because of horizontal earth pressures acting within the embankment. These earth pressures
cause horizontal shear stresses at the base of the embankment that must be resisted by the
foundation soil. If the foundation soil does not have adequate shear resistance, failure can
result. Properly designed horizontal layers of high-strength geotextiles or geogrids can
provide reinforcement, which increase stability and prevent such failures. Both materials can
be used equally well, provided they have the requisite design properties. There are some
differences in how they are installed, especially with respect to seaming and field
workability. Also, at some very soft sites, especially where there is no root mat or vegetative
layer, geogrids may require a lightweight geotextile separator to provide filtration and
prevent contamination of the first lift if it is an open-graded or similar type soil. A
lightweight geotextile is not required beneath the first lift if it is sand, which meets soil
filtration criteria.
The reinforcement may also reduce horizontal and vertical displacements of the underlying
soil and thus reduce differential settlement. It should be noted that the reinforcement will not
reduce the magnitude of long-term consolidation or secondary settlement of the embankment.
The use of reinforcement in embankment construction may allow for:
an increase in the design factor of safety;
an increase in the height of the embankment;
a reduction or elimination of stabilizing side berms;
a reduction in embankment displacements during construction, thus reducing fill
requirements; and/or
an improvement in embankment performance due to increased uniformity of postconstruction settlement.
This chapter assumes that all the common foundation treatment alternatives for the
stabilization of embankments on soft or problem foundation soils have been carefully
considered during the preliminary design phase. Holtz (1989) discusses these treatment
alternatives and provides guidance about when embankment reinforcement is feasible. In
some situations, the most economical final design may be some combination of a
conventional foundation treatment alternative together with geosynthetic reinforcement.
Examples include preloading and stage construction with prefabricated (wick) vertical drains,
FHWA NHI-07-092
Geosynthetic Engineering

7 1

7 Reinforced Embankments
August 2008

the use of stabilizing berms, lightweight fill or column supported embankments - each used
with geosynthetic reinforcement at the base of the embankment. In addition to the
information in Chapter 2 on prefabricated drains and Section 7.12 of this chapter on column
supported embankments, FHWA NHI-06-020, Ground Improvement Methods Reference
Manual Volume II (Elias et al., 2006) provides detailed information on prefabricated
vertical drains, column supported embankments, and lightweight fill technologies.

7.2

APPLICATIONS

Reinforced embankments over weak foundations typically fall into one of two situations construction over uniform deposits, and construction over local anomalies (Bonaparte, Holtz,
and Giroud, 1985). The more common application is embankments, dikes, or levees
constructed over very soft, saturated silt, clay, or peat layers (Figure 7-1). In this situation,
the reinforcement is usually placed with its strong direction perpendicular to the centerline of
the embankment, and plane strain conditions are assumed to prevail. Additional
reinforcement with its strong direction oriented parallel to the centerline may also be required
at the ends of the embankment.
The second reinforced embankment situation includes foundations below the embankment
that are locally weak or contain voids. These zones or voids may be caused by sinkholes,
thawing ice (thermokarsts), old streambeds, or pockets of silt, clay, or peat (Figure 7-1). In
this application, the role of the reinforcement is to bridge over the weak zones or voids, and
tensile reinforcement may be required in more than one direction. Thus, the strong direction
of the reinforcing must be placed in proper orientation with respect to the embankment
centerline (Bonaparte and Christopher, 1987).
Geotextiles may also be used as separators for displacement-type embankment construction
(Holtz, 1989) and as a stabilization layer to allow for embankment construction (see
Chapter 5). In this application, the geotextile does not provide any reinforcement but only
acts as a separator to maintain the integrity of the embankment as it displaces the subgrade
soils. In this case, geotextile design is based upon constructability and survivability, and a
high elongation material may be selected. Prefabricated geocomposite drains may also be
placed as a drainage layer at the base of the embankment to allow for pore pressure
dissipation and consolidation as an alternate to using clean, free draining granular fill for the
first lift.

FHWA NHI-07-092
Geosynthetic Engineering

7 2

7 Reinforced Embankments
August 2008

Figure 7-1. Reinforced embankment applications (after Bonaparte and Christopher, 1987).
Biaxial geogrids may also be used as a stabilization layer for embankment construction. This
stabilization geogrid may provide reinforcement strength in the embankments longitudinal
direction (see Step 9 in Sections 7.3-2 and 7.3-3). A lightweight geotextile filter, if needed,
can be used in conjunction with the geogrid.

7.3

DESIGN GUIDELINES FOR REINFORCED EMBANKMENTS ON SOFT


SOILS

7.3-1 Design Considerations


As with ordinary embankments on soft soils, the basic design approach for reinforced
embankments is to design against failure. The ways in which embankments constructed on
soft foundations can fail have been described by Terzaghi and Peck (1967); Haliburton,
Anglin and Lawmaster (1978 a and b); Fowler (1981); Christopher and Holtz (1985); and
Koerner (1990), among others. Figure 7-2 shows unsatisfactory behavior that can occur in
reinforced embankments. The three possible modes of failure indicate the types of stability
analyses that are required. In addition, settlement of the embankment and potential creep of
FHWA NHI-07-092
Geosynthetic Engineering

7 3

7 Reinforced Embankments
August 2008

the reinforcement must be considered, although creep is only a factor if the creep rate in the
reinforcement is greater than the strength gain occurring in the foundation due to
consolidation. Because the most critical condition for embankment stability is at the end of
construction, the reinforcement only has to function until the foundation soils gain sufficient
strength to support the embankment.

Figure 7-2. Reinforced embankments failure modes (after Haliburton et al., 1978b).

FHWA NHI-07-092
Geosynthetic Engineering

7 4

7 Reinforced Embankments
August 2008

The calculations required for stability and settlement utilize conventional geotechnical design
procedures modified only for the presence of the reinforcement.
The stability of an embankment over soft soil is usually determined by the total stress
method of analysis, which is conservative since the analysis generally assumes that no
strength gain occurs in the compressible soil. The stability analyses presented in this text
uses the total stress approach, because it is simple and appropriate for reinforcement design
(Holtz, 1989).
It is always possible to calculate stability in terms of the effective stresses using the effective
stress shear strength parameters. However, this calculation requires an accurate estimate of
the field pore pressures to be made during the project design phase. Additionally, highquality, undisturbed samples of the foundation soils must be obtained and Ko consolidatedundrained triaxial tests conducted in order to obtain the required design soil parameters.
Because the prediction of in-situ pore pressures in advance of construction is not easy, it is
essential that field pore pressure measurements using high quality piezometers be made
during construction to control the rate of embankment filling. Preloading and staged
embankment construction are discussed in detail by Ladd (1991). Note that by taking into
account the strength gain that occurs with controlled rate (e.g. staged) embankment
construction, lower strength and therefore lower cost reinforcement can be utilized.
However; the time required for construction may be significantly increased and the costs of
the site investigation, laboratory testing, design analyses, field instrumentation, and
inspection are also greater.
The total stress design steps and methodology are detailed in the following section.
[Note: The subjects of site investigation and laboratory testing, soil shear strength
determination, and field instrumentation are addressed in detail in the following FHWA
references: NHI-01-031 Subsurface Investigations - Geotechnical Site Characterization
(NHI course No. 132031 reference manual{Mayne et al., 2002}); IF-02-034 Geotechnical
Engineering Circular No. 5 Evaluation of Soil and Rock Properties (Sabatini, et al., 2002);
NHI-06-088 Soils and Foundations Workshop (NHI course No. 132012 reference manual
{Samtani and Nowatzki, 2006}); and HI-98-034 Geotechnical Instrumentation (NHI course
No. 132041 reference manual {Dunnicliff, 1988}).]
7.3-2 Design Steps
The following is a step-by-step procedure for design of reinforced embankments. Additional
comments on each step can be found in Section 7.3-3.
FHWA NHI-07-092
Geosynthetic Engineering

7 5

7 Reinforced Embankments
August 2008

STEP 1.

Define embankment dimensions and loading conditions.

A.

Embankment height, H

B.

Embankment length

C.

Width of crest

D.

Side slopes, b/H

E.

External loads
1.
surcharges
2.
temporary (traffic) loads
3.
dynamic loads

F.

Environmental considerations
1.
frost action
2.
shrinkage and swelling
3.
drainage, erosion, and scour

G.

Embankment construction rate


1.
project constraints
2.
anticipated or planned rate of construction

STEP 2.

Establish the soil profile and determine the engineering properties of the
foundation soil.

A.

From a subsurface soils investigation, determine


1.
subsurface stratigraphy and soil profile
2.
groundwater table (location, fluctuation)

B.

Engineering properties of the subsoils


1.
Undrained shear strength, cu, for end of construction
2.
Drained shear strength parameters, c' and ', for long-term conditions
3.
Consolidation parameters (Cc, Cr, cv, p')
4.
Chemical and biological factors that may be detrimental to the
reinforcement

C.

Variation of properties with depth and areal extent

FHWA NHI-07-092
Geosynthetic Engineering

7 6

7 Reinforced Embankments
August 2008

STEP 3.

Obtain engineering properties of embankment fill materials.

A.

Classification properties

B.

Moisture-density relationships

C.

Shear strength properties

D.

Chemical and biological factors that may be detrimental to the reinforcement

STEP 4.

Establish minimum appropriate factors of safety and operational settlement


criteria for the embankment. Suggested minimum factors of safety are as
follows.

A.

Bearing capacity:
Overall bearing capacity: 2.0
Local bearing capacity (i.e., lateral squeeze type failure): 1.3 to 2.0

B.

Global (rotational) shear stability at the end of construction: 1.3

C.

Internal shear stability, long-term: 1.5

D.

Lateral spreading (sliding): 1.5

E.

Dynamic loading: 1.1

F.

Settlement criteria: dependent upon project requirements

STEP 5.
A.

Check bearing capacity.


When the thickness of the soft soil is much greater than the width of the
embankment, use classical bearing capacity theory:
qult = fill H = cu Nc

[7-1]

where Nc, the bearing capacity factor, is usually taken as 5.14 -- the value for
a strip footing on a cohesive soil of constant undrained shear strength, cu, with
depth. This approach may underestimate the bearing capacity of reinforced
embankments, as discussed in Section 7.3-3.
FHWA NHI-07-092
Geosynthetic Engineering

7 7

7 Reinforced Embankments
August 2008

B.

STEP 6.

When the soft soil is of limited depth, perform a lateral squeeze analysis
(Section 7.3-3).
Check rotational shear stability.

Perform a rotational slip surface analysis on the unreinforced embankment and


foundation to determine the critical failure surface and the factor of safety against
local shear instability.
A.

If the calculated factor of safety is greater than the minimum required, then
reinforcement is not needed. Check lateral embankment spreading (Step 7).

B.

If the factor of safety is less than the required minimum, then calculate the
required reinforcement strength, Tg, to provide an adequate factor of safety
using Figure 7-3 or alternative solutions (Section 7.3-3), where:

Tg =
STEP 7.

FS (M D ) M R
R cos( )

Check lateral spreading (sliding) stability.

Perform a lateral spreading or sliding wedge stability analysis (Figure 7-4).

FS =

Fresisting
Fdriving

1 H b tan
b tan f
f
= 2
=
1 K H2
Ka H
2 a

A.

If the calculated factor of safety is greater than the minimum required, then
reinforcement is not needed for this failure mode possibility.

B.

If the factor of safety is inadequate, then determine the lateral spreading


strength of reinforcement, Tls, required -- see Figure 7-4b. Soil/geosynthetic
cohesion, Ca, should be based on undrained direct shear tests on the
soil/geosynthetic interface and assumed equal to 0 for extremely soft soils and
low embankments. A cohesion value should be included with placement of
the second and subsequent fills in staged embankment construction.

FHWA NHI-07-092
Geosynthetic Engineering

7 8

7 Reinforced Embankments
August 2008

FS =

2 (b c a + Tls )
Ka H 2

where:
b =
H =
Ka =
' =
=
sg =
cu =
ca =

length of embankment side slope


height of embankment
coefficient of lateral earth pressure for embankment fill soil
friction angle of embankment soil
unit weight of embankment soil
embankment soil to geosynthetic interface friction angle
cohesion (total stress) of foundation soil
adhesion of foundation soil to geosynthetic reinforcement
(Assume ca = 0 for 1st stage loading on extremely soft soils.)

In absence of test data, the value of tan sg may conservatively be taken as 2/3
tan . In absence of test data, the value of ca should be assumed to be 0.

C.

Check sliding above the reinforcement. See Figure 7-4a.

FS =

STEP 8.

b tan sg
Ka H

Establish tolerable geosynthetic deformation requirements and calculate the


required reinforcement modulus, J, based on wide width (ASTM D 4595)
tensile testing.
Reinforcement Modulus:

J = Tls /geosynthetic

[7-2]

Recommendations for strain limits, based on type of fill soil materials and for
construction over peats, are:
Cohesionless soil fills:
Cohesive soil fills:
Peat foundations:

FHWA NHI-07-092
Geosynthetic Engineering

7 9

geosynthetic = 5 to 10%
geosynthetic = 2%
geosynthetic = 2 to 10%

[7-3]
[7-4]
[7-5]

7 Reinforced Embankments
August 2008

Figure 7-3.

Reinforcement required to provide rotational stability (a) Christopher and


Holtz (1985) after Wager (1981); (b) Bonaparte and Christopher (1987) for
the case in which the reinforcement does not increase soil strength.

FHWA NHI-07-092
Geosynthetic Engineering

7 10

7 Reinforced Embankments
August 2008

Figure 7-4.

Reinforcement required to limit lateral embankment spreading


(a) embankment sliding on reinforcement; (b) rupture of reinforcement and
embankment sliding on foundation soil (Bonaparte and Christopher, 1987).

FHWA NHI-07-092
Geosynthetic Engineering

7 11

7 Reinforced Embankments
August 2008

STEP 9.

Establish geosynthetic strength requirements in the embankments


longitudinal direction (i.e., direction of the embankment alignment).

A.

Check bearing capacity and rotational slope stability at the ends of the
embankment (Steps 5 and 6).

B.

Use strength and elongation determined from Steps 7 and 8 to control


embankment spreading during construction and to control bending following
construction.

C.

As the strength of the seams transverse to the embankment alignment control


strength requirements, geosynthetic seam strength requirements are the higher
of the strengths determined from Steps 9.A or 9.B.

STEP 10.

Establish geosynthetic properties (Section 7.4).

A.

Design strengths and modulus are based on the ASTM D 4595 wide width
tensile test. This test standard permits definition of tensile modulus in terms
of: (i) initial tensile modulus; (ii) offset tensile modulus; or (iii) secant tensile
modulus. Furthermore, the secant modulus may be defined between any two
strain points. Geosynthetic modulus for design of embankments should be
determined using a secant modulus, defined with the zero strain point and
design strain limit (i.e., 2 to 10%) point.

B.

Geotextile seam strength is quantified with the ASTM D 4884 test method,
and is equal to the strength required in the embankments longitudinal
direction. Geogrid overlap strength, for longitudinal direction strength, is
quantified with pullout testing (ASTM D 6706).

C.

Soil-geosynthetic friction, sg, based on ASTM D 5321 with on-site soils. For
preliminary estimates, assume sg = 2/3; for final design, testing is
recommended.

D.

Geotextile stiffness based on site conditions and experience. See Sect. 7.4-5.

E.

Select survivability and constructability requirements for the geosynthetic


based on site conditions, backfill materials, and equipment, using Tables 7-1,
7-2, and 7-3.

FHWA NHI-07-092
Geosynthetic Engineering

7 12

7 Reinforced Embankments
August 2008

STEP 11.

Estimate magnitude and rate of embankment settlement.

Use conventional geotechnical procedures and practices for this step.


STEP 12.

Establish construction sequence and procedures.

See Section 7.8.


STEP 13.

Establish construction observation requirements.

See Sections 7.8 and 7.9.


STEP 14.

Hold preconstruction meetings.

Consider a partnering type contract with a disputes resolution board.


STEP 15.

Observe construction and build with confidence (if the procedures outlined in
these guidelines are followed!)

7.3-3 Comments on the Design Procedure


STEPS 1 and 2 need no further elaboration.
STEP 3. Obtain embankment fill properties.
Follow traditional geotechnical practice, except that the first few lifts of fill material
just above the geosynthetic should be free-draining granular materials. This
requirement provides the best frictional interaction between the geosynthetic and fill,
as well as providing a drainage layer for excess pore water to dissipate from the
underlying soils. Other fill materials may be used above this layer as long as the
strain compatibility of the geosynthetic is evaluated with respect to the backfill
materials (Step 8).
When a fill is placed on soft ground, the main driving force is from the weight of the
embankment itself. It may be advantageous to use a lightweight fill material to
reduce the driving forces, thereby increasing the overall global stability of the fill.
The reduction in driving force will depend upon the type of lightweight fill material
used. The geotechnical properties of various types of lightweight fill materials are
discussed in detail in FWHA NHI-06-019 Ground Improvement Methods Reference
FHWA NHI-07-092
Geosynthetic Engineering

7 13

7 Reinforced Embankments
August 2008

Manual Volume I (Elias et al., 2006). A secondary benefit of the use of lightweight
fill material is the reduction in settlement under loading. The amount of settlement
will be reduced proportionately to the reduction in load.
STEP 4. Establish design factors of safety.
The minimum factors of safety previously stated are recommended for projects with
modern state-of-the-practice geotechnical site investigations and laboratory testing.
Those factors may be adjusted depending on the method of analysis, type and use of
facility being designed, the known conditions of the subsurface, the quality of the
samples and soils testing, the cost of failure, the probability of extreme events
occurring, and the engineer's previous experience on similar projects and sites. In
short, all of the uncertainties in loads, analyses, and soil properties influence the
choice of appropriate factors of safety. Typical factors of safety for unreinforced
embankments also seem to be appropriate for reinforced embankments.
When the calculated factor of safety is greater than 1 but less than the minimum
allowable factor of safety for design, say 1.3 or 1.5, then the geosynthetic provides an
additional factor of safety or a second line of defense against failure. On the other
hand, when the calculated factor of safety for the unreinforced embankment is
significantly less than 1, the geosynthetic reinforcement is the difference between
success and failure. In this latter case, construction considerations (Section 7.8)
become crucial to the project success.
Maximum tolerable post-construction settlement and embankment deformations,
which depend on project requirements, must also be established.
STEP 5. Check overall bearing capacity.
Overall Bearing
Reinforcement does not increase the overall bearing capacity of the foundation soil.
If the foundation soil cannot support the weight of the embankment, then the
embankment cannot be built. Thus, the overall bearing capacity of the entire
embankment must be satisfactory before considering any possible reinforcement. As
such, the vertical stress due to the embankment can be treated as an average stress
over the entire width of the embankment, similar to a semi-rigid mat foundation.
The bearing capacity can be calculated using classical soil mechanics methods
(Terzaghi and Peck, 1967; Vesic, 1975; Perloff and Baron, 1976; and U.S. Navy,
FHWA NHI-07-092
Geosynthetic Engineering

7 14

7 Reinforced Embankments
August 2008

1986), which use limiting equilibrium-type analyses for strip footings, assuming
logarithmic spiral failure surfaces on an infinitely deep foundation. These analyses
are not appropriate if the thickness of the underlying soft deposit is small compared to
the width of the embankment. In this case, high lateral stresses in the confined soft
stratum beneath the embankment could lead to a lateral squeeze-type failure. Use of
reinforced soils slopes (Chapter 8) or of mechanically stabilized earth walls (Chapter
9) can lead to high lateral stresses in underlying soft foundation soils. See following
discussion for guidance on assessing this failure mechanism.
In a review of 40 reinforced embankment case histories, Humphrey and Holtz (1986)
and Humphrey (1987) found that in many cases, the failure height predicted by
classical bearing capacity theory was significantly less than the actual constructed
height, especially if high strength geotextiles and geogrids were used as the
reinforcement. Figure 7-5 shows the embankment height versus average undrained
shear strength of the foundation. Significantly, four embankments failed at heights of
6.6 ft. (2 m) greater than predicted by Equation 7-1 (line B in Figure 7-5). The two
reinforced embankments that failed below line B were either on peat or underreinforced (Humphrey, 1987). It appears that in many cases, the reinforcement
enhances the beneficial effect the following factors have on stability:
limited thickness or increasing strength with depth of the soft foundation soils
(Rowe and Soderman, 1987 a and b; Jewell, 1988);
the dry crust (Humphrey and Holtz, 1989);
flat embankment side slopes (e.g., Humphrey and Holtz, 1987); or
dissipation of excess pore pressures during construction.
If the factor of safety for bearing capacity is sufficient, then continue with the next
step. If not, consider increasing the embankment's width, flattening the slopes,
adding toe berms, or improving the foundation soils by using stage construction and
drainage enhancement or other alternatives, such as relocating the alignment or
placing the roadway on an elevated structure.

FHWA NHI-07-092
Geosynthetic Engineering

7 15

7 Reinforced Embankments
August 2008

Figure 7-5. Embankment height versus undrained shear strength of foundation; line A:
classical bearing capacity theory (Eq. 7-1); line B: line A + 6.6 ft. (2 m) (after Humphrey,
1987). 1 m = 3.3 ft.
Lateral Squeeze
High lateral stresses in a confined soft stratum beneath an embankment could lead to
a lateral squeeze-type failure. Lateral squeeze-type failure of the foundation should
be anticipated if fill x Hfill > 3cu, (see FHWA Soils and Foundation Manual, FHWA
NHI-06-088 {Samtani and Nowatzki, 2006}) and a weak soil layer exists beneath the
embankment to a depth that is less that the width of the embankment. The shear
forces developed under the embankment should be compared to the corresponding
shear strength of the soil. Approaches discussed by Jrgenson (1934), Silvestri
(1983), and Bonaparte, Holtz and Giroud (1985), Rowe and Soderman (1987a), Hird
and Jewell (1990), and Humphrey and Rowe (1991) are appropriate. The designer
should be aware that the analysis for lateral squeeze is only approximate, and no
single method is completely accepted by geotechnical engineers at present. When the
depth of the soft layer, DS, is greater than the base width of the embankment, general
global bearing capacity and overall stability will govern the design.

FHWA NHI-07-092
Geosynthetic Engineering

7 16

7 Reinforced Embankments
August 2008

The approach by Silvestri (1983) is presented and demonstrated below for lateral
squeeze failure at the toe of an embankment side slope. If a weak soil layer exists
beneath the embankment to a limited depth DS which is less than the width of the
slope b' (see Figure 7-6), the factor of safety against failure by squeezing may be
calculated from:

FS squeezing =

2 cu
4.14 cu
+
1.3
Ds (tan )
H

[7-6]

where:
=
=
Ds =
H =
cu =

angle of slope.
unit weight of soil in slope.
depth of soft soil beneath slope base of the
embankment.
height of slope.
undrained shear strength of soft soil beneath slope.

Caution is advised and rigorous analysis (e.g, numerical modeling and/or extensive
subsurface investigation with careful evaluation of cu) should be performed when FS
< 2. For factors of safety below 2, cu should be confirmed through rigorous
laboratory testing on undisturbed samples direct simple shear, evaluation of over
consolidation ratio (e.g. Ladd, 1991), or triaxial compression with pore pressure
measurements and/or field vane shear tests. Careful monitoring during construction
will be required with piezometers, surface survey monuments (both within and
outside the toe of the embankment), and inclinometers installed for construction
control.

FS =
Figure 7-6.

2 cu
4.14 c u
+
Ds (tan )
H

Local bearing failure (lateral squeeze).

FHWA NHI-07-092
Geosynthetic Engineering

7 17

7 Reinforced Embankments
August 2008

If the foundation soils are cohesive and limited to a depth of less than the base width
of the embankment, then local stability should be evaluated. As an example, assume
that the foundation soils had an undrained shear strength of 340 psf (16 kPa) and
extended to a depth of 10 ft (3 m) at which point the granular soils were encountered,
and the embankment fill unit weight is 120 lb/ft3 (18.8 kN/m3). Constructing even a
13 ft (4 m) high embankment with a 2H:1V side slope would create a problem in
accordance with equation 7-6 as follows.
FS squeezing =

2 cu
4.14 c u
+

D s (tan )
H

FS squeezing =

2 (170 psf )
4.14 (170 psf )
+
= 1.02
3
(120 lb / ft )(10 ft )(tan 26.6 ) 13 ft (120 lb / ft 3 )

Since FSsqueezing is lower than the recommended 1.3, the stability conditions must be
improved. This could be accomplished by either reducing the slope angle, use of
lightweight embankment fill, or by placing a surcharge at the toe (which effectively
reduces the slope angle). In addition, if the resulting factor of safety is less than 2,
refinement of the analysis should be considered as previously discussed (i.e., careful
evaluation of cu, consider performing numerical modeling, and install instrumentation
for construction control).
STEP 6. Check rotational shear stability.
The next step is to calculate the factor of safety against a circular failure through the
embankment and foundation using classical limiting equilibrium-type stability
analyses. If the factor of safety does not meet the minimum design requirements
(Step 4), then the reinforcing tensile force required to increase the factor of safety to
an acceptable level must be estimated.
This is done by assuming that the reinforcement acts as a stabilizing tensile force at
its intersection with the slip surface being considered. The reinforcement thus
provides the additional resisting moment required to obtain the minimum required
factor of safety. The analysis is shown in Figure 7-3.
The analysis consists of determining the most critical failure surface(s) using
conventional limiting equilibrium analysis methods. For each critical sliding surface,
the driving moment (MD) and soil resisting moment (MR) are determined as shown in
FHWA NHI-07-092
Geosynthetic Engineering

7 18

7 Reinforced Embankments
August 2008

Figure 7-3a. The additional resisting moment MR to provide the required factor of
safety is calculated as shown in Figure 7-3b. Then one or more layers of geotextiles
or geogrids with sufficient tensile strength at tolerable strains (Step 7) are added at the
base of the embankment to provide the required additional resisting moment. If
multiple layers are used, they must be separated by a granular layer and they must
have compatible stress-strain properties (e.g., the same type of reinforcement must be
used for each layer).
A number of procedures have been proposed for determining the required additional
reinforcement, and these are summarized by Christopher and Holtz (1985), Bonaparte
and Christopher (1987), Holtz (1990), and Humphrey and Rowe (1991). The basic
difference in the approaches is in the assumption of the reinforcement force
orientation at the location of the critical slip surface (the angle in Figures 7-3a and
7-3b). It is conservative to assume that the reinforcing force acts horizontally at the
location of the reinforcement ( = 0). In this case, the additional reinforcing moment
is equal to the required geosynthetic strength, Tg, times the vertical distance, y, from
the plane of the reinforcement to the center of rotation, or:
MR = Tg y

[7-6a]

as determined for the most critical failure surface, shown in Figure 7-3a. This
approach is conservative because it neglects any possible reinforcement reorientation
along the alignment of the failure surface, as well as any confining effect of the
reinforcement.
A less-conservative approach assumes that the reinforcement bends due to local
displacements of the foundation soils at the onset of failure, with the maximum
possible reorientation located tangent to the slip surface ( = 2 in Figure 7-3b). In
this case,
MR = Tg [R cos (2 - $)]
[7-6b]
where,
2
=
angle from horizontal to tangent line as shown in Figure 7-3.
Limited field evidence indicates that it is actually somewhere in between the
horizontal and tangential (Bonaparte and Christopher, 1987) depending on the
foundation soils, the depth of soft soil from the original ground line in relation to the
width of the embankment (D/B ratio), and the stiffness of the reinforcement. Based
on the minimal information available, the following suggestions are provided for
selecting the orientation:
FHWA NHI-07-092
Geosynthetic Engineering

7 19

7 Reinforced Embankments
August 2008

=0

for brittle, strain-sensitive foundations soils (e.g., leached


marine clays) or where a crust layer is considered in the
analysis for increased support;
for D/B < 0.4 and moderate to highly compressible soils (e.g.,
= 2/2
soft clays, peats);
for D/B 0.4 highly compressible soils (e.g., soft clays, peats);
=2
and reinforcement with high elongation potential (,design $
10%), and large tolerable deformations; and
=0
when in any doubt!
Other approaches, as discussed by Bonaparte and Christopher (1987), require a more
rigorous analysis of the foundation soils deformation characteristics and the
reinforcement strength compatibility.
In each method, the depth of the critical failure surface must be relatively shallow,
i.e., y in Figure 7-3a must be large, otherwise the geosynthetic contribution toward
increasing the resisting moment will be small. On the other hand, Jewell (1988) notes
that shallow slip surfaces tend to underestimate the driving force in the embankment,
and both he and Leshchinsky (1987) have suggested methods to address this problem.
STEP 7. Check lateral spreading (sliding) stability.
A simplified analysis for calculating the reinforcement required to limit lateral
embankment spreading is illustrated in Figure 7-4. For unreinforced as well as
reinforced embankments, the driving forces result from the lateral earth pressures
developed within the embankment and which must, for equilibrium, be transferred to
the foundation by shearing stresses (Holtz, 1990). Instability occurs in the
embankment when either:
1.
the embankment slides on the reinforcement (Figure 7-4a); or
2.
the reinforcement fails in tension and the embankment slides on the
foundation soil (Figure 7-4b).
In the latter case, the shearing resistance of the foundation soils just below the
embankment is insufficient to maintain equilibrium. Thus, in both cases, the
reinforcement must have sufficient friction to resist sliding on the reinforcement
plane, and the geosynthetic tensile strength must be sufficient to resist rupture as the
potential sliding surface passes through the reinforcement.
The forces involved in the analysis of embankment spreading are shown in Figure 7-4
for the two cases above. The lateral earth pressures, usually assumed to be active, are
FHWA NHI-07-092
Geosynthetic Engineering

7 20

7 Reinforced Embankments
August 2008

a maximum at the crest of the embankment. The factor of safety against embankment
spreading is found from the ratio of the resisting forces to the actuating (driving)
forces. The recommended factor of safety against sliding is 1.5 (Step 4). If the
required soil-geosynthetic friction angle is greater than that reasonably achieved with
the reinforcement, embankment soils and subgrade, then the embankment slopes must
be flattened or berms must be added. Sliding resistance can be increased by the soil
improvement techniques mentioned above. Generally, however, there is sufficient
frictional resistance between geotextiles and geogrids commonly used for
reinforcement and granular fill. If this is the case, then the resultant lateral earth
pressures must be resisted by the tension in the reinforcement.
In the case where an MSE or RSS structure is founded at the end of the embankment
(but not supporting a bridge structure) the length b may be taken as the reinforcement
length, L, of the MSE or RSS structure. An MSE or RSS structure should only be
included at the end of an embankment after the foundation soil has been adequately
improved (i.e., through surcharging) to support such structures or other ground
improvement techniques are employed, such as stabilization berms, lightweight fill,
etc.
STEP 8. Establish tolerable deformation requirements for the geosynthetic.
Excessive deformation of the embankment and its reinforcement may limit its
serviceability and impair its function, even if total collapse does not occur. Thus, an
analysis to establish deformation limits of the reinforcement must be performed. The
most common way to limit deformations is to limit the allowable strain in the
geosynthetic. This is done because the geosynthetic tensile forces required to prevent
failure by lateral spreading are not developed without some strain, and some lateral
movement must be expected. Thus, geosynthetic modulus is used to control lateral
spreading (Step 7). The distribution of strain in the geosynthetic is assumed to vary
linearly from zero at the toe to a maximum value beneath the crest of the
embankment. This is consistent with the development of lateral earth pressures
beneath the slopes of the embankment.
For the assumed linear strain distribution, the maximum strain in the geosynthetic
will be equal to twice the average strain in the embankment. Fowler and Haliburton
(1980) and Fowler (1981) found that an average lateral spreading of 5% was
reasonable, both from a construction and geosynthetic property standpoint. If 5% is
the average strain, then the maximum expected strain would be 10%, and the
geosynthetic modulus would be determined at 10% strain (Equation 7-3). However,
FHWA NHI-07-092
Geosynthetic Engineering

7 21

7 Reinforced Embankments
August 2008

it has been suggested that a modulus at 10% strain would be too large, and that
smaller maximum values at, say 2 to 5%, are more appropriate. Additional discussion
of geosynthetic deformation is given in Christopher and Holtz (1985 and 1989),
Bonaparte, Holtz and Giroud (1985), Rowe and Mylleville (1989 and 1990), and
Humphrey and Rowe (1991).
If cohesive soils are used in the embankment, then the modulus should be determined
at 2% strain to reduce the possibility of embankment cracking (Equation 7-4). Of
course, if embankment cracking is not a concern, then these limiting reinforcement
strain values could be increased. Keep in mind, however, that if cracking occurs, no
resistance to sliding is provided. Further, the cracks could fill with water, which
would add to the driving forces.

STEP 9. Establish geosynthetic strength requirements in the longitudinal direction.


Most embankments are relatively long but narrow in shape. Thus, during
construction, stresses are imposed on the geosynthetic in the longitudinal direction,
i.e., along the direction of the centerline. Reinforcement may be also required for
loadings that occur at bridge abutments, and due to differential settlements and
embankment bending, especially over nonuniform foundation conditions and at the
edges of soft soil deposit.
Because both sliding and rotational failures are possible, analysis procedures
discussed in Steps 6 and 7 should be applied, but in the direction along the alignment
of the embankment. This determines the longitudinal strength requirements of the
geosynthetic. Because the usual placement of the geosynthetic is in strips
perpendicular to the centerline, the longitudinal stability will be controlled by the
strength of the transverse seams.
STEP 10. Establish geosynthetic properties.
See Section 7.4 for a determining the required properties of the geosynthetic.
STEP 11. Estimate magnitude and rate of embankment settlement.
Although not part of the stability analyses, both the magnitude and rate of settlement
of the embankment should be considered in any reinforcement design. There is some
evidence from finite element studies that differential settlements may be reduced
FHWA NHI-07-092
Geosynthetic Engineering

7 22

7 Reinforced Embankments
August 2008

somewhat by the presence of geosynthetic reinforcement.


Long-term or
consolidation settlements are not influenced by the geosynthetic, since
compressibility of the foundation soils is not altered by the reinforcement, although
the stress distribution may be somewhat different. Present recommendations provide
for reinforcement design as outlined in Steps 6 - 10 above. Then use conventional
geotechnical methods to estimate immediate, consolidation, and secondary
settlements, as if the embankment was unreinforced (Christopher and Holtz, 1985).
Possible creep of reinforced embankments on soft foundations should be considered
in terms of the geosynthetic creep rate versus the consolidation rate and strength gain
of the foundation. If the foundation soil consolidates and gains strength at a rate
faster than (or equal to) the rate the geosynthetic loses strength due to creep, there is
no problem. Many soft soils such as peats, silts and clays with sand lenses have high
permeability, therefore, they gain strength rapidly, but each case should be analyzed
individually.
Time required for settlement can be substantially decreased with foundation drains.
Consolidation of soft ground using vertical drains is a technique used since the 1920s.
Today, the most common method is the use of wick drains, which can best be
described as prefabricated vertical drains (PVDs), since drainage is via pressure, and
not by wicking. PVDs are used to accelerate consolidation of soft saturated
compressible soils under load.
The most common use of PVDs is to accelerate
consolidation for approach embankments at bridges or other embankment
construction over soft soils, where the total post construction settlement is not
acceptable.
When PVDs are used to accelerate settlement, the subsoil must meet the following
criteria:
Moderate to high compressibility.
Low permeability.
Full saturation.
Final embankment loads must exceed maximum past pressure.
Secondary consolidation must not be a major concern.
Low-to-moderate shear strength.
The evaluation, design, cost, specification, and construction with PVDs are discussed
in detail in FWHA NHI-06-019 Ground Improvement Methods Reference Manual
Volume I (Elias et al., 2006). Filtration of the PVD geotextile should be evaluated
following the guidelines in Chapter 2 of this manual.
FHWA NHI-07-092
Geosynthetic Engineering

7 23

7 Reinforced Embankments
August 2008

STEP 12. Establish construction sequence and procedures.


The importance of proper construction procedures for geosynthetic reinforced
embankments on very soft foundations cannot be over emphasized. A specific
construction sequence is usually required to avoid failures during construction.
See Section 7.8 for details on site preparation, special construction equipment,
geosynthetic placement procedures, seaming techniques, and fill placement and
compaction procedures.
STEP 13. Establish construction observation requirements
See Sections 7.8 and 7.9.
A.

B.

Instrumentation. As a minimum, install piezometers, settlement points, and


surface survey monuments. Also consider inclinometers to observe lateral
movement with depth.
Note that the purpose of the instrumentation in soft ground reinforcement
projects is not for research but to verify design assumptions and to control
and, usually, expedite construction.
Geosynthetic inspection. Be sure field personnel understand:
geosynthetic submittal for acceptance prior to installation;
testing requirements;
fill placement procedures; and
seam integrity verification.

STEP 14. Hold preconstruction meetings


It has been our experience that the more potential contractors know about the overall
project, the site conditions, and the assumptions and expectations of the designers, the
more realistically they can bid; and, the project is more successful. Prebid and
preconstruction information meetings with contractors have been very successful in
establishing a good, professional working relationship between owner, design
engineer, and contractor. Partnering type contracts and a disputes resolution board
can also be used to reduce problems, claims, and litigation.

FHWA NHI-07-092
Geosynthetic Engineering

7 24

7 Reinforced Embankments
August 2008

STEP 15. Observe construction


Inspection should be performed by a trained and knowledgeable inspector, and good
documentation of construction should be maintained.

7.4

SELECTION OF GEOSYNTHETIC AND FILL PROPERTIES

Once the design strength requirements have been established, the appropriate geosynthetic
must be selected. In addition to its tensile and frictional properties, drainage requirements,
construction conditions, and environmental factors must also be considered. Geosynthetic
properties required for reinforcement applications are given in Table 7-1. The selection of
appropriate fill materials is also an important aspect of the design. When possible, granular
fill is preferred, especially for the first few lifts above the geosynthetic.

Table 7-1. Geosynthetic Properties Required for Reinforcement Applications.

Property1

Criteria and Parameter


Design Requirements:
a. Mechanical
Tensile strength
Tensile modulus
Seam strength
Tension creep
Soil-geosynthetic friction
b. Hydraulic
Piping resistance
Permeability

Wide width strength


Wide width strength
Wide width strength
Tension creep
Soil-geosynthetic friction angle
Apparent opening size
Permeability

Constructability Requirements:
Tensile strength
Puncture resistance
Tear resistance

Grab strength
Puncture resistance
Trapezoidal tear

Longevity:
UV stability (if exposed)
Soil compatibility (where required)

UV resistance
Chemical; Biological

NOTE: 1. See Table 1-3 for specific test procedures.

FHWA NHI-07-092
Geosynthetic Engineering

7 25

7 Reinforced Embankments
August 2008

7.4-1 Geotextile and Geogrid Strength Requirements


The most important mechanical properties are the tensile strength and modulus of the
reinforcement, seam strength, soil-geosynthetic friction, and system creep resistance.

The tensile strength and modulus values should preferably be determined by an in-soil tensile
test. From research by McGown, Andrawes, and Kabir (1982) and others, we know that insoil properties of many geosynthetics are markedly different than those from tests conducted
in air. However, in-soil tests are not yet routine nor standardized, and the test proposed test
methods need additional work. The practical alternate is to conservatively use a
representative (i.e., wide strip) tensile test as a measure of the in-soil strength. This point is
discussed by Christopher and Holtz (1985) and Bonaparte, Holtz, and Giroud (1985).
Therefore, strength and modulus are based on testing of wide specimens. ASTM D 4595,
Standard Test Method for Tensile Properties of Geotextiles by the Wide-Width Strip Method,
is used for geotextiles, and ASTM D 6637 Standard Test Method for Determining Tensile
Properties of Geogrids by the Single or Multi-Rib Tensile Method with Method B or C (wide
specimen) is used for geogrids. These test standards permits definition of tensile modulus in
terms of: (i) initial tensile modulus; (ii) offset tensile modulus; or (iii) secant tensile
modulus. Furthermore, the secant modulus may be defined between any two strain points.
Geosynthetic modulus for design of embankments should be determined using a secant
modulus, defined with the zero strain point and design strain limit (i.e., 2 to 10%) point.
The following minimum criteria for tensile strength of geosynthetics are recommended.
1.
For ordinary cases, determine the design tensile strength Td (the larger of Tg
and Tls) and the required secant modulus at 2 to 10% strain.
2.
The ultimate tensile strength Tult obviously must be greater that the design
tensile strength, Td. Note that Tg includes an inherent safety factor against
overload and sudden failure that is equal to the rotational stability safety
factor. The tensile strength requirements should be increased to account for
installation damage, depending on the severity of the conditions.
3.
The strain of the reinforcement at failure should be at least 1.5 times the
secant modulus strain to avoid brittle failure. For exceptionally soft
foundations where the reinforcement will be subjected to very large tensile
stresses during construction, the geosynthetic must have either sufficient
strength to support the embankment itself, or the reinforcement and the
embankment must be allowed to deform. In this case, an elongation at rupture
of up to 50% may be acceptable. In either case, high tensile strength
geosynthetics and special construction procedures (Section 7.8) are required.
4.
If there is a possibility of tension cracks forming in the embankment or high
FHWA NHI-07-092
Geosynthetic Engineering

7 26

7 Reinforced Embankments
August 2008

5.

6.

strain levels occurring during construction (such as might occur, for example,
with cohesive embankments), the lateral spreading strength, Tls, at 2% strain
should be required.
The required lateral spreading strength, Tls, should be increased to account for
creep and installation damage as the creep potential of the geosynthetic
depends on the creep potential of the foundation. If significant creep is
expected in the foundation, the creep potential of the geosynthetic at design
stresses should be evaluated, recognizing that strength gains in the foundation
will reduce the creep potential. Installation damage potential will depend on
the severity of the conditions.
Strength requirements must be evaluated and specified for both the machine
and cross machine directions of the geosynthetic. Usually, the seam strength
controls the cross machine geosynthetic strength requirements.

Depending on the strength requirements, geosynthetic availability, and seam efficiency, more
than one layer of reinforcement may be necessary to obtain the required tensile strength. If
multiple layers are used, a granular layer of 8 to 12 in. (200 to 300 mm) must be placed
between each successive geosynthetic layer or the layers must be mechanically connected
(e.g., sewn) together. Also, the geosynthetics must be strain compatible; that is, the same
type of geosynthetic should be used for each layer.
For soil-geosynthetic friction values, either direct shear or pullout tests should be utilized. If
test values are not available, Bell (1980) recommends that for sand embankments, the soilgeosynthetic friction angle is from 2/3 up to the full of the sand. Since these early
recommendations, a number of direct shear and pullout tests have been performed on both
geogrids and geotextiles and the recommendations still apply. It is recommended that in the
absence of tests, a soil-geosynthetic friction angle of 2/3 should be conservatively used for
granular fill placed directly on the geosynthetic. For clay soils, friction tests are definitely
warranted and should be performed under all circumstances.
The creep properties of geosynthetics in reinforced soil systems are not well established. Insoil creep tests are possible but are far from routine today. For design, it is recommended
that the working stress be kept much lower than the creep limit of the geosynthetic. Values
of 40 to 60% of the ultimate stress are typically satisfactory for this purpose. Live loads
versus dead loads also must be taken into account. Short-term live loadings are much less
detrimental in terms of creep than sustained dead loads. And finally, as discussed in Section
7.3-3 Step 11, the relative rates of deformation of the geosynthetic versus the consolidation
and strength gain of the foundation soil must be considered. In most cases, creep is not an
issue in reinforced embankment stability.
FHWA NHI-07-092
Geosynthetic Engineering

7 27

7 Reinforced Embankments
August 2008

7.4-2 Drainage Requirements


The geosynthetic must allow for free vertical drainage of the foundation soils to reduce pore
pressure buildup below the embankment. Pertinent geosynthetic hydraulic properties are
piping resistance and permeability (Table 7-1). It is recommended that the permeability of
the geosynthetic be at least 10 times that of the underlying soil. Permeability values could be
based on consolidation tests and taken at initial load levels to simulate initial placement of
fill. The opening size should be selected based on the requirements of Section 2.3. The
opening size should be a maximum to reduce the risk of clogging, while still providing
retention of the underlying soil.
7.4-3 Environmental Considerations
For most embankment reinforcement situations, geosynthetics have a high resistance to
chemical and biological attack; therefore, chemical and biological compatibility is usually
not a concern. However, in unusual situations such as very low (i.e., < 3) or very high (i.e., >
9) pH soils, or other unusual chemical environments -- such as in industrial areas or near
mine or other waste dumps -- the chemical compatibility of the polymer(s) in the
geosynthetic should be checked to assure it will retain the design strength at least until the
underlying subsoil is strong enough to support the structure without reinforcement.
7.4-4 Constructability (Survivability) Requirements
In addition to the design strength requirements, the geotextile or geogrid must also have
sufficient strength to survive construction. If the geotextile is ripped, punctured, or torn
during construction, support strength for the embankment structure will be reduced and
failure could result. Constructability property requirements are listed in Table 7-1. (These
are also called survivability requirements.) Tables 7-2 and 7-3 were developed by
Haliburton, Lawmaster, and McGuffey (1982) specifically for reinforced embankment
construction with varying subgrade conditions, construction equipment, and lift thicknesses
(see also Christopher and Holtz, 1985). The specific property values are provided in Table 74 and Table 7-5. The high and moderate class conditions are taken directly from
survivability tables in Chapter 5 for road construction (e.g., Table 5-3 and 5-4 from
AASHTO M-288 Specification (2006) for geotextiles and Table 5-5 for geogrids) and are
equivalent to Class 1 and Class 2 geosynthetics, respectively. The very high class requires
greater strength than the requirements in Chapter 5 due to the possibility of constructing
embankments on uncleared subgrade, which is a much harsher condition than anticipated for
roads. For all critical applications, high to very high survivability geotextiles and geogrids
are recommended. As the construction of the first lift of the embankment is analogous to
construction of a temporary haul road, survivability requirements discussed in Section 5.9 are
also appropriate here.

FHWA NHI-07-092
Geosynthetic Engineering

7 28

7 Reinforced Embankments
August 2008

Table 7-2. Required Degree of Geosynthetic Survivability as a Function


of Subgrade Conditions And Construction Equipment.
Construction Equipment and
6 to 12 in. (150 to 300 mm) Cover Material
Initial Lift Thickness
Low Ground
Pressure
Equipment
(< 4 psi)
{< 30 kPa}

SUBGRADE CONDITIONS

{>30 kPa, < 60 kPa}

High Ground
Pressure
Equipment
(> 8 psi)
{>60 kPa}

Medium Ground
Pressure
Equipment
(> 4 psi, < 8 psi)

Subgrade has been cleared of all obstacles except


grass, weeds, leaves, and fine wood debris. Surface is
smooth and level, and shallow depressions and humps
do not exceed 6 in. (150 mm) in depth and height. All
larger depressions are filled. Alternatively, a smooth
working table may be placed.

Moderate/
Low

Moderate

High

Subgrade has been cleared of obstacles larger than


small- to moderate-sized tree limbs and rocks. Tree
trunks and stumps should be removed or covered with
a partial working table. Depressions and humps should
not exceed 18 in. (450 mm) in depth and height.
Larger depressions should be filled.

Moderate

High

Very High

High

Very High

Not
Recommended

Minimal site preparation is required. Trees may be


felled, delimbed, and left in place. Stumps should be
cut to project not more than ~6 in. (150 mm) above
subgrade. Geosynthetic may be draped directly over
the tree trunks, stumps, large depressions and humps,
holes, stream channels, and large boulders. Items
should be removed only if, where placed, the
Geosynthetic and cover material over them will distort
the finished road surface.

NOTES:
1. Recommendations are for 6 to 12 in. (150 to 300 mm) initial thickness. For other initial lift thickness:
12 to 18 in. (300 to 450 mm): Reduce survivability requirement one level
18 to 24 in. (450 to 600 mm): Reduce survivability requirement two levels
> 24 in. (> 600 mm): Reduce survivability requirement three levels
2. For special construction techniques such as prerutting, increase survivability requirement one level.
3. Placement of excessive initial cover material thickness may cause bearing failure of soft subgrades.
4. Note that equipment used for embankment construction (even High Ground Pressure equipment) have
significantly lower ground contact pressures than equipment used for roadway construction (Table 5-2).

FHWA NHI-07-092
Geosynthetic Engineering

7 29

7 Reinforced Embankments
August 2008

Table 7-3. Required Degree of Geosynthetic Survivability as a


Function of Cover Material and Construction Equipment.

COVER MATERIAL
Fine sand to + 2 in.
(50 mm) diameter
gravel, rounded to
subangular

Coarse aggregate with


diameter up to one-half
proposed lift thickness,
may be angular

Some to most aggregate


with diameter greater than
one-half proposed lift
thickness, angular and
sharp-edged, few fines

Low ground
pressure equipment
(4 psi)
{30 kPa}

Moderate/Low

Moderate

High

Medium ground
pressure equipment
(> 4 psi, < 8 psi)
{>30 kPa,<60 kPa}

Moderate

High

Very High

Medium ground
pressure equipment
(> 4 psi, < 8 psi)
{>30 kPa,<60 kPa}

Moderate/Low

Moderate

High

High ground
pressure equipment
(> 8 psi)
{>60 kPa}

Moderate

High

Very High

18 to 24 in.
(450 to 600 mm)
Initial Lift
Thickness

High ground
pressure equipment
(> 8 psi)
{>60 kPa}

Moderate/Low

Moderate

High

> 24 in.
(> 600 mm)
Initial Lift
Thickness

High ground
pressure equipment
(> 8 psi)
{>60 kPa}

Moderate/Low

Moderate/Low

Moderate

CONSTRUCTION

6 to 12 in.
(150 to 300 mm)
Initial Lift
Thickness

12 to 18 in.
(300 to 450 mm)
Initial Lift
Thickness

NOTES:
1. For special construction techniques such as prerutting, increase geosynthetic survivability requirement one
level.
2. Placement of excessive initial cover material thickness may cause bearing failure of soft subgrades.
3. Note that equipment used for embankment construction (even High Ground Pressure equipment) have
significantly lower ground contact pressures than equipment used for roadway construction (Table 5-2).

FHWA NHI-07-092
Geosynthetic Engineering

7 30

7 Reinforced Embankments
August 2008

Table 7-4. Minimum Geotextile Property Requirements1,2,3


for Geotextile Survivability (after AASHTO, 2006)
ASTM
Test
Method

Units

Grab Strength

D 4632

Tear Strength
Puncture Strength

Property

Required Degree of Geotextile Survivability


Very High

High

Moderate

(see Note 4)

1400

1100

D 4533

(see Note 4)

500

400

D 6241

(see Note 4)

2750

2200

NOTES:
1. Acceptance of geotextile material shall be based on ASTM D 4759.
2. Acceptance shall be based upon testing of either conformance samples obtained using Procedure A of ASTM
D 4354, or based on manufacturers certifications and testing of quality assurance samples obtained using
Procedure B of ASTM D 4354.
3. Minimum; use value in weaker principal direction. All numerical values represent minimum average roll
value (i.e., test results from any sampled roll in a lot shall meet or exceed the minimum values in the table).
Lot samples according to ASTM D 4354.
4. Recommend survivability of candidate Very High survivability geotextile(s) be demonstrated on a
field/project basis or the use of a High survivability geotextile as a sacrificial layer.
CONVERSION: 1 N = 0.225 lbf

7.4-5 Stiffness and Workability


For extremely soft soil conditions, geosynthetic stiffness or workability may be an important
consideration. The workability of a geosynthetic is its ability to support workers during
initial placement and sewing operations and to support construction equipment during the
first lift placement. Workability is generally related to geosynthetic stiffness; however,
stiffness evaluation techniques and correlations with field workability are very poor (Tan,
1990). The workability guidelines based on subgrade CBR (Christopher and Holtz, 1985)
are satisfactory for CBR > 1.0. For very soft subgrades, much stiffer geosynthetics are
required. Other aspects of field workability such as water absorption, bulk density, and
fastening method (i.e., geotextile sewn seam or geogrid overlap) should also be considered,
especially on very soft sites.

FHWA NHI-07-092
Geosynthetic Engineering

7 31

7 Reinforced Embankments
August 2008

Table 7-5. Geogrid Survivability Property Requirements1,2,3

Property

Test Method

Units

Requirement
Geogrid Class4

SURVIVABILITY

CLASS 15

CLASS 2

Ultimate Multi-Rib
Tensile Strength

ASTM D 6637

kN/m

18

12

Junction Strength6

GSI GRI GG2

110

110

Ultraviolet Stability
(Retained Strength)

ASTM D 4355

50% after 500 hours of exposure

OPENING CHARACTERISTICS
Opening Size

Direct measure

mm

Opening Size > D50 of aggregate above geogrid

Separation

ASTM D 422

mm

D85 of aggregate above geogrid < 5 D85 subgrade


Other wise use separation geotextile with geogrid

NOTES:
1. Acceptance of geogrid material shall be based on ASTM D 4759.
2. Acceptance shall be based upon testing of either conformance samples obtained using Procedure A of
ASTM D 4354, or based on manufacturers certifications and testing of quality assurance samples obtained
using Procedure B of ASTM D 4354.
3. Minimum; use value in stronger principal direction for ultimate multi-rib tensile and retained strengths, and
use value in weaker principal direction for junction strength. All numerical values represent minimum
average roll value (i.e., test results from any sampled roll in a lot shall meet or exceed the minimum values
in the table). Lot samples according to ASTM D 4354.
4. Class 1 is considered a High survivability geogrid and Class 2 as a Moderate survivability geogrid.
Recommend survivability of candidate Very High survivability geogrid(s) be demonstrated on a
field/project basis or the use of a High survivability geogrid as a sacrificial layer in conditions requiring
Very High survivability.
5. Default geogrid selection. The engineer may specify a Class 2 geogrid for moderate survivability
conditions, see Table 5-2.
6. Junction strength requirements have not been fully supported by data, and until such data is established,
manufacturers shall submit data from full scale installation damage tests in accordance with ASTM D 5818
documenting integrity of junctions. For soft soil applications, a minimum of 6 in. (150 mm) of cover
aggregate shall be placed over the geogrid and a loaded dump truck used to traverse the section a minimum
number of passes to achieve 4 in. (100 mm) of rutting. A photographic record of the geogrid after
exhumation shall be provided, which clearly shows that junctions have not been displaced or otherwise
damaged during the installation process.

FHWA NHI-07-092
Geosynthetic Engineering

7 32

7 Reinforced Embankments
August 2008

7.4-6 Fill Considerations


The first lift of fill material just above the geosynthetic should be free-draining granular
materials. This requirement provides the best frictional interaction between the geosynthetic
and fill, as well as a drainage layer for excess pore water dissipation of the underlying soils.
Other lower permeability, (preferably granular) fill materials may be used above this layer as
long as the strain compatibility of the geosynthetic is evaluated with respect to the backfill
material, as discussed in Section 7.3-3, Step 8.

Most reinforcement analyses assume that the fill material is granular. In fact, in the past the
use of cohesive soils together with geosynthetic reinforcement has been discouraged. This
may be an unrealistic restriction, although there are problems with placing and compacting
cohesive earth fills on especially soft subsoils. Furthermore, the frictional resistance between
geosynthetics and cohesive soils is problematic. It may be possible to use composite
embankments. Cohesionless fill could be used for the first 18 to 36 in. (0.5 to 1 m); then the
rest of the embankment could be constructed to grade with locally available materials.

7.5

DESIGN EXAMPLE

DEFINITION OF DESIGN EXAMPLE

Project Description: A 4-lane highway is to be constructed over a peat bog. Alignment and
anticipated settlement require construction of an embankment with an
average height of 6.5 ft. See project cross section figure.

Type of Structure: embankment supporting a permanent paved road

Type of Application: geosynthetic reinforcement

Alternatives:

i) excavate and replace - wetlands do not allow;


ii) lightweight fill - high cost;
iii) stone columns - soils too soft;
iv) drainage and surcharge - yes; or
v) very flat (8H:1V) slope - right-of-way restriction

GIVEN DATA

Geometry

as shown in project cross section figure

Geosynthetic

geotextile (a geogrid also may be used for this example problem; however,
this example represents an actual case history where a geotextile was used)

FHWA NHI-07-092
Geosynthetic Engineering

7 33

7 Reinforced Embankments
August 2008

Soils

Stability

Stability analyses of the unreinforced embankment were conducted with the STABL
computer program. The most critical condition for embankments on soft soils is endof-construction case; therefore, UU (unconsolidated, undrained) soil shear strength
values are used in analyses.

Results of the analyses:


a. With 4:1 side slopes and sand/gravel fill ( = 138 lb/ft3), FS 0.72.
b. Since FS was substantially less than 1 for 4H:1V slopes, flatter slopes were
evaluated, even though additional right-of-way would be required. With 8:1 side
slopes and sand/gravel fill ( = 138 lb/ft3), a FS 0.87 was computed.
c. Light-weight fill ( = 100 lb/ft3) was also considered, with it, the FS varied
between 0.90 to 1.15

Transportation Department required safety factors are:


Fsmin > 1.5 for long-term conditions
FSallow 1.3 for short-term conditions

subsurface exploration indicates cu = 100 psf in weakest areas


soft soils are underlain by firmer soils of cu = 500 psf
embankment fill soil will be sands and gravel
lightweight fill costs $250,000 more than sand/gravel

Project Cross Section


REQUIRED
Design geotextile reinforcement to provide a stable embankment.

DEFINE
A. Geotextile function(s):
FHWA NHI-07-092
Geosynthetic Engineering

7 34

7 Reinforced Embankments
August 2008

B. Geotextile properties required:


C. Geotextile specification:
SOLUTION
A. Geotextile function(s):
Primary
reinforcement (for short-term conditions)
Secondary - separation and filtration
B. Geotextile properties required:
tensile characteristics
interface shear strength
survivability
apparent opening size (AOS)

DESIGN
Design embankment with geotextile reinforcement to meet short-term stability requirements.

STEP 1.DEFINE DIMENSIONS AND LOADING CONDITIONS


See project cross section figure.

STEP 2.SUBSURFACE CONDITIONS AND PROPERTIES


Undrained shear strength provided in given data. Design for end-of-construction. Long-term
design with drained shear strength parameters not covered within this example.

STEP 3.EMBANKMENT FILL PROPERTIES


sand and gravel, with
(m = 138 lb/ft3
N' = 35E

STEP 4.ESTABLISH DESIGN REQUIREMENTS


-Transportation Department required safety factors are:
FSmin > 1.5 for long-term conditions
FSmin 1.3 for short-term conditions

FHWA NHI-07-092
Geosynthetic Engineering

7 35

7 Reinforced Embankments
August 2008

- settlement
Primary consolidation must be completed prior to paving roadway.
A total fill height of 6.5 ft is anticipated to reach design elevation. This height
includes the additional fill material thickness to compensate for anticipated
settlements.
STEP 5.CHECK OVERALL BEARING CAPACITY
Recommended minimum safety factor (section 7.3-2) is 2.
A. Overall bearing capacity of soil, ignoring footing size is
qult = c Nc
qult = 100 psf x 5.14 = 514 psf
Considering depth of embedment (i.e., shearing will have to occur through the embankment
for a bearing capacity failure) the bearing capacity is more accurately computed (see
Meyerhof) as follows.
Nc = 4.14 + 0.5 B/D where, B = the base width of the embankment (~ 100 ft),
and
D = the average depth of the soft soil (~ 15 ft)
Nc = 4.14 + 0.5 (100 ft / 15 ft) = 7.5
qult = 100 psf x 7.5 = 750 psf
maximum load, Pmax = m H
w/o a geotextile Pmax = 138 lb/ft3 x 6.5 ft = 900 psf
implies FS = 750 / 900 = 0.83

NO GOOD

with a geotextile, and assuming that the geotextile will result in an even distribution of the
embankment load over the width of the geotextile (i.e., account for the slopes at the
embankment edges),
Pavg = AE m / B
where, A = cross section area of embankment, and
B = base width of the embankment
Pavg = {[ (100 ft + 50 ft) 6.5 ft] 138 lb/ft3 } / 100 ft
Pavg = 672 psf < qult worst case

Safety Factor Marginal

Add berms to increase bearing capacity. Berms, 10 ft wide, can be added within the existing
right-of-way, increasing the base width to 120 ft. With this increase in width,
Nc = 4.14 + 0.5 (120 ft / 15 ft) = 8.14
qult = 100 psf x 8.1 = 814 psf
and,
FHWA NHI-07-092
Geosynthetic Engineering

7 36

7 Reinforced Embankments
August 2008

Pavg = 672 psf (100 / 120) = 560 psf


FS = 814 psf / 560 psf = 1.45

Safety Factor O.K.

B. Lateral squeeze
From FHWA Foundation Manual (Cheney and Chassie, 1993) If fill x Hfill > 3c, then lateral squeeze of the foundation soil can occur. Since Pmax =
900 psf is much greater than 3c, even considering the crust layer (c = 200 psf), a rigorous
lateral squeeze analysis was performed using the method by Jrgeson (1934). In this method,
the lateral stress beneath the toe of the embankment is determined through charts or finite
element analysis and compared to the shear strength of the soil. This method indicated a
safety factor of approximately 1 for the 100 ft base width. Adding the berm and extending
the reinforcement to the toe of the berm decreases the potential for lateral squeeze as the
lateral stress is reduced at the toe of the berm. The berms increased FSSQUEEZZE to greater
than 1.5.
Also, comparing the reinforced design with Figure 7-5 indicates that the reinforced
structure should be stable.
STEP 6.PERFORM ROTATIONAL SHEAR STABILITY ANALYSIS
Recommended minimum safety factor at end of construction (section 7.3-2) is 1.3.
The critical unreinforced failure surface is found through rotational stability methods. For
this project, STABL4M was used and the critical, unreinforced surface FS = 0.72. As the
soil supporting the embankment was highly compressible peat, the reinforcement was
assumed to rotate such that = (Figure 7-3 and Eq. 7-4b). Thus,

FS req =

Tg =
therefore,

M R + Tg R
MD

1.3

1 .3 M D M R
R

Tg 18,000 lb/ft

Feasible - yes. Geosynthetics are available which exceed this strength requirement,
especially if multiple layers are used. For this project, an installation damage factor of
approximately equal to 1.0, and 2 layers were used:
Bottom:
6,000 lb/ft
Top:
12,000 lb/ft
The use of 2 layers allowed the lower cost bottom material to be used over the full
embankment plus berm width, while the higher strength and more expensive geotextile was
only placed under the embankment section where it was required.
FHWA NHI-07-092
Geosynthetic Engineering

7 37

7 Reinforced Embankments
August 2008

STEP 7.CHECK LATERAL SPREADING (SLIDING) STABILITY


Recommended minimum safety factor (section 7.3-2) is 1.5.
A. from Figure 7-4b:
T = FS x PA = FS x 0.5 Ka (m H2
T = 1.5 (0.5) [tan2 (45 - 35/2)] (138 lb/ft3) (6.5 ft)2
T = 1185 lb/ft
Use Reduction Factors (RF) = 3 for creep and 1 installation damage
therefore, Tls = 3560 lb/ft
Tls < Tg, therefore Tdesign = Tg = 18,000 kN/m
B. check sliding:

FS =

FS =

b tan sg
Ka H

26 ft tan 23
0.27 6.5 ft

FS > 6, OK

STEP 8.ESTABLISH TOLERABLE DEFORMATION (LIMIT STRAIN) REQUIREMENTS


For cohesionless sand and gravel over deformable peat use = 10%

STEP 9.EVALUATE GEOSYNTHETIC


DIRECTION

STRENGTH

REQUIRED

IN

LONGITUDINAL

From Step 7,
use TL = Tls = 53 kN/m for reinforcement and seams in the cross machine (X-MD) direction

STEP 10.ESTABLISH GEOSYNTHETIC PROPERTIES


A. Design strength and elongation based upon ASTM D 4595
Ultimate tensile strength
Td1 = Tult $ 6,000 lb/ft in MD - Layer 1
Td2 = Tult $ 12,000 lb/ft in MD - Layer 2
Tult $ 3,560 lb/ft in X-MD - both layers
FHWA NHI-07-092
Geosynthetic Engineering

7 38

7 Reinforced Embankments
August 2008

Reinforcement Modulus, J
J = Tls / 0.10 = 3,560 lb/ft for limit strain of 10%
J $ 35,600 lb/ft - MD and X-MD, both directions
B. seam strength
Tseam $ 3,560 lb/ft

with controlled fill placement

C. soil-geosynthetic adhesion
from testing, per ASTM D 5321, Nsg $ 23
D. geotextile stiffness based upon site conditions and experience
E. survivability and constructability requirements
Assume:1. medium ground pressure equipment
2. 12 in. first lift
3. uncleared subgrade
Use a Very High Survivability geotextile (from Tables 7-2 and 7-3). Therefore, from Table 7-4, the
survivability of candidate geotextile reinforcements shall be demonstrated on a field/project basis or a
High survivability geotextile, meeting the minimum average roll values listed below, may be used
as a sacrificial layer.

Property
Grab Strength
Tear Resistance
Puncture Strength

ASTM
Test Method
D 4632
D 4533
D 6241

Minimum
Strength
1400 N
(315 lbs)
500 N
(110 lbs)
2750 N
(620 lbs)

Drainage and filtration requirements Need grain size distribution of subgrade soils
Determine:maximum AOS for retention
minimum kg> ks
minimum AOS for clogging resistance

Complete Steps 11 through 15 to finish design.


STEP 11.PERFORM SETTLEMENT ANALYSIS
STEP 12.ESTABLISH CONSTRUCTION SEQUENCE REQUIREMENTS
STEP 13.ESTABLISH CONSTRUCTION OBSERVATION REQUIREMENTS
STEP 14.HOLD PRECONSTRUCTION MEETING
STEP 15.OBSERVE CONSTRUCTION

FHWA NHI-07-092
Geosynthetic Engineering

7 39

7 Reinforced Embankments
August 2008

7.6

SPECIFICATIONS

Because the reinforcement requirements for soft-ground embankment construction will be


project and site specific, standard specifications, which include suggested geosynthetic
properties, are not appropriate, and special provisions or a separate project specification must
be used. The following examples, one for a geotextile reinforcement and another for geogrid
reinforcement include most of the items that should be considered in a reinforced
embankment project.
HIGH STRENGTH GEOTEXTILE FOR EMBANKMENT REINFORCEMENT
(from Washington Department of Transportation, October 27, 1997)

Description
This work shall consist of furnishing and placing construction geotextile in accordance with the
details shown in the plans, these specifications, or as directed by the Engineer.
Materials
Geotextile and Thread for Sewing
The material shall be a woven geotextile consisting only of long chain polymeric filaments or
yarns formed into a stable network such that the filaments or yarns retain their position relative to
each other during handling, placement, and design service life. At least 95 percent by mass of the
of the material shall be polyolefins or polyesters. The material shall be free from defects or tears.
The geotextile shall be free of any treatment or coating which might adversely alter its hydraulic
or physical properties after installation. The geotextile shall conform to the properties as
indicated in Table 1.
Thread used shall be high strength polypropylene, polyester, or Kevlar thread. Nylon threads will
not be allowed.
Geotextile Approval
Source Approval
The Contractor shall submit to the Engineer the following information regarding each
geotextile proposed for use:
Manufacturer's name and current address,
Full Product name,
Geotextile structure, including fiber/yarn type, and
Geotextile polymer type(s).
If the geotextile source has not been previously evaluated, a sample of each proposed geotextile
shall be submitted to the Olympia Service Center Materials Laboratory in Tumwater for
evaluation. After the sample and required information for each geotextile type have arrived at the
Olympia Service Center Materials Laboratory in Tumwater, a maximum of 14 calendar days will
be required for this testing. Source approval will be based on conformance to the applicable
values from Table 1. Source approval shall not be the basis of acceptance of specific lots of
material unless the lot sampled can be clearly identified, and the number of samples tested and
approved meet the requirements of WSDOT Test Method 914.
FHWA NHI-07-092
Geosynthetic Engineering

7 40

7 Reinforced Embankments
August 2008

Geotextile Properties
Table 1. Properties for high strength geotextile for embankment reinforcement.
Test Method1

Geotextile Property Requirements2

AOS
Water Permittivity
Tensile Strength, min.
in machine direction
Tensile Strength, min.
in x-machine direction
Secant Modulus at 5% strain

ASTM D4751
ASTM D4491

0.84 mm max. (#20 sieve)


0.02/sec. min.

ASTM D4595

(to be based on project specific design)

ASTM D4595

(to be based on project specific design)

ASTM D4595

(to be based on project specific design)

Seam Breaking Strength

ASTM D4884

(to be based on project specific design)

Puncture Resistance
Tear Strength, min.
in machine and
x-machine direction
Ultraviolet (UV) Radiation
Stability

ASTM D4833

330 N min.

ASTM D4533

330 N min.

ASTM D4355

50% Strength Retained min., after 500


Hrs in weatherometer

Property

The test procedures are essentially in conformance with the most recently approved ASTM
geotextile test procedures, except geotextile sampling and specimen conditioning, which are in
accordance with WSDOT Test Methods 914 an 915, respectively. Copies of these test methods
are available at the Olympia Service Center Materials Laboratory in Tumwater, Washington.
2
All geotextile properties listed above are minimum average roll values (i.e., the test result for
any sampled roll in a lot shall meet or exceed the values listed).
Geotextile Samples for Source Approval
Each sample shall have minimum dimensions of 1.5 meters by the full roll width of the
geotextile. A minimum of 6 square meters of geotextile shall be submitted to the Engineer
for testing. The geotextile machine direction shall be marked clearly on each sample
submitted for testing. The machine direction is defined as the direction perpendicular to the
axis of the geotextile roll.
The geotextile samples shall be cut from the geotextile roll with scissors, sharp knife, or other
suitable method which produces a smooth geotextile edge and does not cause geotextile
ripping or tearing. The samples shall not be taken from the outer wrap of the geotextile nor
the inner wrap of the core.
Acceptance Samples
Samples will be randomly taken by the Engineer at the job site to confirm that the geotextile
meets the property values specified.
Approval will be based on testing of samples from each lot. A "lot" shall be defined for the
purposes of this specification as all geotextile rolls within the consignment (i.e., all rolls sent
to the project site) which were produced by the same manufacturer during a continuous
period of production at the same manufacturing plant and have the same product name. After
the samples and manufacturer's certificate of compliance have arrived at the Olympia Service
FHWA NHI-07-092
Geosynthetic Engineering

7 41

7 Reinforced Embankments
August 2008

Center Materials Laboratory in Tumwater, a maximum of 14 calendar days will be required


for this testing. If the results of the testing show that a geotextile lot, as defined, does not
meet the properties required in Table 1, the roll or rolls which were sampled will be rejected.
Two additional rolls for each roll tested which failed from the lot previously tested will then
be selected at random by the Engineer for sampling and retesting. If the retesting shows that
any of the additional rolls tested do not meet the required properties, the entire lot will be
rejected. If the test results from all the rolls retested meet the required properties, the entire
lot minus the roll(s) which failed will be accepted. All geotextile which has defects,
deterioration, or damage, as determined by the Engineer, will also be rejected. All rejected
geotextile shall be replaced at no expense to the Contracting Agency.
Certificate of Compliance
The Contractor shall provide a manufacturer's certificate of compliance to the Engineer which
includes the following information about each geotextile roll to be used:
Manufacturer's name and current address,
Full product name,
Geotextile structure, including fiber/yarn type,
Geotextile polymer type(s),
Geotextile roll number, and
Certified test results.
Approval Of Seams
If the geotextile seams are to be sewn in the field, the Contractor shall provide a section of
sewn seam which can be sampled by the Engineer before the geotextile is installed.
The seam sewn for sampling shall be sewn using the same equipment and procedures as will
be used to sew the production seams. The seam sewn for sampling must be at least 2 meters
in length. If the seams are sewn in the factory, the Engineer will obtain samples of the
factory seam at random from any of the rolls to be used. The seam assembly description shall
be submitted by the Contractor to the Engineer and will be included with the seam sample
obtained for testing. This description shall include the seam type, stitch type, sewing thread
type(s), and stitch density.
Construction Requirements
Geotextile Roll Identification, Storage, and Handling
Geotextile roll identification, storage, and handling shall be in conformance to ASTM D 4873.
During periods of shipment and storage, the geotextile shall be stored off the ground. The
geotextile shall be covered at all times during shipment and storage such that it is fully protected
from ultraviolet radiation including sunlight, site construction damage, precipitation, chemicals
that are strong and acids or strong bases, flames including welding sparks, temperatures in excess
of 70o C, and any other environmental condition that may damage the physical property values of
the geotextile.
Preparation and Placement of the Geotextile Reinforcement
The area to be covered by the geotextile shall be graded to a smooth, uniform condition free from
ruts, potholes, and protruding objects such as rocks or sticks. The Contractor may construct a
working platform, up to 0.6 meters in thickness, in lieu of grading the existing ground surface. A
working platform is required where stumps or other protruding objects which cannot be removed
without excessively disturbing the subgrade are present. All stumps shall be cut flush with the
ground surface and covered with at least 150 mm of fill before placement of the first geotextile
FHWA NHI-07-092
Geosynthetic Engineering

7 42

7 Reinforced Embankments
August 2008

layer. The geotextile shall be spread immediately ahead of the covering operation. The
geotextile shall be laid with the machine direction perpendicular or parallel to centerline as shown
in Plans. Perpendicular and parallel directions shall alternate. All seams shall be sewn. Seams to
connect the geotextile strips end to end will not be allowed, as shown in the Plans. The geotextile
shall not be left exposed to sunlight during installation for a total of more than 14 calendar days.
The geotextile shall be laid smooth without excessive wrinkles. Under no circumstances shall the
geotextile be dragged through mud or over sharp objects which could damage the geotextile. The
cover material shall be placed on the geotextile in such a manner that a minimum of 200 mm of
material will be between the equipment tires or tracks and the geotextile at all times.
Construction vehicles shall be limited in size and weight such that rutting in the initial lift above
the geotextile is not greater than 75 mm deep, to prevent overstressing the geotextile. Turning of
vehicles on the first lift above the geotextile will not be permitted. Compaction of the first lift
above the geotextile shall be limited to routing of placement and spreading equipment only. No
vibratory compaction will be allowed on the first lift.
Small soil piles or the manufacturers recommended method shall be used as needed to hold the
geotextile in place until the specified cover material is placed.
Should the geotextile be torn or punctured or the sewn joints disturbed, as evidenced by visible
geotextile damage, subgrade pumping, intrusion, or roadbed distortion, the backfill around the
damaged or displaced area shall be removed and the damaged area repaired or replaced by the
Contractor at no expense to the Contracting Agency. The repair shall consist of a patch of the
same type of geotextile placed over the damaged area. The patch shall be sewn at all edges.
If geotextile seams are to be sewn in the field or at the factory, the seams shall consist of two
parallel rows of stitching, or shall consist of a J-seam, Type Ssn-1, using a single row of stitching.
The two rows of stitching shall be 25 mm apart with a tolerance of plus or minus 13 mm and
shall not cross, except for restitching. The stitching shall be a lock-type stitch. The minimum
seam allowance, i.e., the minimum distance from the geotextile edge to the stitch line nearest to
that edge, shall be 40 mm if a flat or prayer seam, Type SSa-2, is used. The minimum seam
allowance for all other seam types shall be 25 mm. The seam, stitch type, and the equipment
used to perform the stitching shall be as recommended by the manufacturer of the geotextile and
as approved by the Engineer.
The seams shall be sewn in such a manner that the seam can be inspected readily by the Engineer
or his representative. The seam strength will be tested and shall meet the requirements stated in
this Specification.
Embankment construction shall be kept symmetrical at all times to prevent localized bearing
capacity failures beneath the embankment or lateral tipping or sliding of the embankment. Any
fill placed directly on the geotextile shall be spread immediately. Stockpiling of fill on the
geotextile will not be allowed.
The embankment shall be compacted using Method B of Section 2-03.3(14)C. Vibratory or
sheepsfoot rollers shall not be used to compact the fill until at least 0.5 meters of fill is covering
the bottom geotextile layer and until at least 0.3 meters of fill is covering each subsequent
geotextile layer above the bottom layer.
The geotextile shall be pretensioned during installation using either Method 1 or Method 2 as
described herein. The method selected will depend on whether or not a mudwave forms during
placement of the first one or two lifts. If a mudwave forms as fill is pushed onto the first layer of
FHWA NHI-07-092
Geosynthetic Engineering

7 43

7 Reinforced Embankments
August 2008

geotextile, Method 1 shall be used. Method 1 shall continue to be used until the mudwave ceases
to form as fill is placed and spread. Once mudwave formation ceases, Method 2 shall be used
until the uppermost geotextile layer is covered with a minimum of 0.3 meters of fill. These
special construction methods are not needed for fill construction above this level. If a mudwave
does not form as fill is pushed onto the first layer of geotextile, then Method 2 shall be used
initially and until the uppermost geotextile layer is covered with at least 0.3 meters of fill.
Method 1
After the working platform, if needed, has been constructed, the first layer of geotextile shall
be laid in continuous transverse strips and the joints sewn together. The geotextile shall be
stretched manually to ensure that no wrinkles are present in the geotextile. The fill shall be
end-dumped and spread from the edge of the geotextile. The fill shall first be placed along
the outside edges of the geotextile to form access roads. These access roads will serve three
purposes: to lock the edges of the geotextile in place, to contain the mudwave, and to provide
access as needed to place fill in the center of the embankment. These access roads shall be
approximately 5 meters wide. The access roads at the edges of the geotextile shall have a
minimum height of 0.6 meters when completed. Once the access roads are approximately 15
meters in length, fill shall be kept ahead of the filling operation, and the access roads shall be
kept approximately 15 meters ahead of this filling operation as shown in the Plans. Keeping
the mudwave ahead of this filling operation and keeping the edges of the geotextile from
moving by use of the access roads will effectively pre-tension the geotextile. The geotextile
shall be laid out no more than 6 meters ahead of the end of the access roads at any time to
prevent overstressing of the geotextile seams.
Method 2
After the working platform, if needed, has been constructed, the first layer of geotextile shall
be laid and sewn as in Method 1. The first lift of material shall be spread from the edge of
the geotextile, keeping the center of the advancing fill lift ahead of the outside edges of the
lift as shown in the Plans. The geotextile shall be manually pulled taut prior to fill placement.
Embankment construction shall continue in this manner for subsequent lifts until the
uppermost geotextile layer is completely covered with 0.3 meters of compacted fill.
Measurement
High strength geotextile for embankment reinforcement will be measured by the square meter for
the ground surface area actually covered.
Payment
The unit contract price per square meter for High Strength Geotextile For Embankment
Reinforcement, shall be full pay to complete the work as specified.

7.7 COST CONSIDERATIONS

The cost analysis for a geosynthetic reinforced embankment includes:


1. Geosynthetic cost: including purchase price, factory prefabrication, and shipping.
2. Site preparation: including clearing and grubbing, and working table preparation.
FHWA NHI-07-092
Geosynthetic Engineering

7 44

7 Reinforced Embankments
August 2008

3.

4.

Geosynthetic placement: related to field workability (see Christopher and Holtz,


1989),
a) with no working table, or
b) with a working table.
Fill material: including purchasing, hauling, dumping, compaction, allowance for
additional fill due to embankment subsidence. (NOTE: Use free-draining granular
fill for the lifts adjacent to geosynthetic to provide good adherence and drainage.)

7.8 CONSTRUCTION PROCEDURES

The construction procedures for reinforced embankments on soft foundations are extremely
important. Improper fill placement procedures can lead to geosynthetic damage, nonuniform
settlements, and even embankment failure. By the use of low ground pressure equipment, a
properly selected geosynthetic, and proper procedures for placement of the fill, these
problems can essentially be eliminated. Essential construction details are outlined below.
The Washington State DOT Special Provision (see Section 7.6) provides additional details.
A. Prepare subgrade:
1. Cut trees and stumps flush with ground surface.
2. Do not remove or disturb root or meadow mat.
3. Leave small vegetative cover, such as grass and reeds, in place.
4. For undulating sites or areas where there are many stumps and fallen trees,
consider a working table for placement of the reinforcement. In this case, a
lower strength sacrificial geosynthetic designed only for constructability can be
used to construct and support the working table.
B. Geosynthetic placement procedures:
1. Orient the geosynthetic with the machine direction perpendicular to the
embankment alignment. No seams should be allowed parallel to the alignment.
Therefore,
The geosynthetic rolls should be shipped in unseamed machine direction
lengths equal to one or more multiples of the embankment design base
width.
The geosynthetic should be manufactured with the largest machine
width possible.
These widths should be factory-sewn to provide the largest width
compatible with shipping and field handling.

FHWA NHI-07-092
Geosynthetic Engineering

7 45

7 Reinforced Embankments
August 2008

2.

Unroll the geosynthetic as smoothly as possible transverse to the alignment.


(Do not drag it.)

3.

Geotextiles should be sewn as required with all seams up and every stitch
inspected. Geogrids may be joined to hold adjacent rolls together or maintain
overlaps by ties, clamps, cables, etc.

4.

The geosynthetic should be manually pulled taut to remove wrinkles. Weights


(sand bags, tires, etc.) or pins may be required to prevent lifting by wind.
Before covering, the Engineer should examine the geosynthetic for holes, rips,
tears, etc. Defects, if any, should be repaired by.
Large defects, should be replaced by cutting along the panel seam and
sewing in a new panel.
Smaller defects, can be cut out and a new panel resewn into that section,
if possible.
Defects less than 6 in. (150 mm), can be overlapped a minimum of 3
feet (1 m) or more in all directions from the defective area. (Additional
overlap may be required, depending on the geosynthetic-to-geosynthetic
friction angle).

5.

NOTE: If a weak link exists in the geosynthetic, either through a defective


seam or tear, the system will tell the engineer about it in a dramatic way -spectacular failure! (Holtz, 1990)
C. Fill placement, spreading, and compaction procedures:
1.

Construction sequence for extremely soft foundations (when a mudwave forms)


is shown in Figure 7-7.
a. End-dump fill along edges of geosynthetic to form toe berms or access
roads.
Use trucks and equipment compatible with constructability
design assumptions (Table 7-1).
End-dump on the previously placed fill; do not dump directly
on the geosynthetic.
Limit height of dumped piles, e.g., to less than 3 feet (1 m)
above the geosynthetic layer, to avoid a local bearing failure.
Spread piles immediately to avoid local depressions.
Use lightweight dozers and/or front-end loaders to spread the
fill.

FHWA NHI-07-092
Geosynthetic Engineering

7 46

7 Reinforced Embankments
August 2008

b.

c.

d.
e.

2.

Toe berms should extend one to two panel widths ahead of the
remainder of the embankment fill placement.
After constructing the toe berms, spread fill in the area between the toe
berms.
Placement should be parallel to the alignment and symmetrical
from the toe berm inward toward the center to maintain a Ushaped leading edge (concave outward) to contain the
mudwave (Figure 7-8).
Traffic on the first lift should be parallel to the embankment
alignment; no turning of construction equipment should be allowed.
Construction vehicles should be limited in size and weight to
limit initial lift rutting to 3 in. (75 mm). If rut depths exceed 3
in. (75 mm), decrease the construction vehicle size and/or
weight.
The first lift should be compacted only by tracking in place with
dozers or end-loaders.
Once the embankment is at least 24 in. (600 mm) above the original
ground, subsequent lifts can be compacted with a smooth drum
vibratory roller or other suitable compactor. If localized liquefied
conditions occur, the vibrator should be turned off and the weight of
the drum alone should be used for compaction. Other types of
compaction equipment also can be used for nongranular fill.

After placement, the geosynthetic should be covered within 48 hours.


For less severe foundation conditions (i.e., when no mudwave forms):
a. Place the geosynthetic with no wrinkles or folds; if necessary,
manually pull it taut prior to fill placement.
b. Place fill symmetrically from the center outward in an inverted U
(convex outward) construction process, as shown in Figure 7-9. Use
fill placement to maintain tension in the geosynthetic.
c. Minimize pile heights to avoid localized depressions.
d. Limit construction vehicle size and weight so initial lift rutting is no
greater than 3 in. (75 mm).
e. Smooth-drum or rubber-tired rollers may be considered for
compaction of first lift; however, do not overcompact. If weaving or
localized quick conditions are observed, the first lift should be
compacted by tracking with construction equipment.

FHWA NHI-07-092
Geosynthetic Engineering

7 47

7 Reinforced Embankments
August 2008

D. Construction monitoring:
1. Monitoring should include piezometers to indicate the magnitude of excess
pore pressure developed during construction. If excessive pore pressures are
observed, construction should be halted until the pressures drop to a
predetermined safe value.
2.

Settlement plates should be installed at the geosynthetic level to monitor


settlement during construction and to adjust fill requirements appropriately.

3.

Inclinometers should be considered at the embankment toes to monitor lateral


displacement.

Photographs of reinforced embankment construction are shown in Figure 7-10.

SEQUENCE OF CONSTRUCTION
1.
2.
3.
4.
5.

6.

LAY GEOSYNTHETIC IN CONTINUOUS TRAVERSE STRIPS, SEW STRIPS TOGETHER.


END DUMP ACCESS ROADS.
CONSTRUCT OUTSIDE SECTIONS TO ANCHOR GEOSYNTHETIC.
CONSTRUCT OUTSIDE SECTION TO SET GEOSYNTHETIC.
CONSTRUCT INTERIOR SECTIONS TO TENSION GEOSYNTHETIC.
CONSTRUCT FINAL CENTER SECTION

Figure 7-7. Construction sequence for geosynthetic reinforced embankments for extremely
weak foundations (from Haliburton, Douglas and Fowler, 1977).

FHWA NHI-07-092
Geosynthetic Engineering

7 48

7 Reinforced Embankments
August 2008

Figure 7-8. Placement of fill between toe berms on extremely soft foundations (CBR < 1)
with a mud wave anticipated.

FHWA NHI-07-092
Geosynthetic Engineering

7 49

7 Reinforced Embankments
August 2008

Figure 7-9. Fill placement to tension geotextile on moderate ground conditions; moderate
subgrade (CBR > 1); no mud wave.

FHWA NHI-07-092
Geosynthetic Engineering

7 50

7 Reinforced Embankments
August 2008

(a)

(b)

(c)

Figure 7-10. Reinforced embankment construction; a) geosynthetic placement;


b) fill dumping; and c) fill spreading.
FHWA NHI-07-092
Geosynthetic Engineering

7 51

7 Reinforced Embankments
August 2008

7.9 INSPECTION

Since implemented construction procedures are crucial to the success of reinforced


embankments on very soft foundations, competent and professional construction inspection
is absolutely essential. Field personnel must be properly trained to observe every phase of
the construction and to ensure that (1) the specified material is delivered to the project, (2)
the geosynthetic is not damaged during construction, and (3) the specified sequence of
construction operations are explicitly followed. Field personnel should review the checklist
in Section 1.7.

7.10 REINFORCED EMBANKMENTS FOR ROADWAY WIDENING

Special considerations are required for widening of existing roadway embankments founded
on soft foundations. Construction sequencing of fill placement, connection of the
geosynthetic to the existing embankment, and settlements of both the existing and new fills
must be addressed by the design engineer. Analytical techniques for geosynthetic
reinforcement requirements are the same as those discussed in Section 7.3.
Two example roadway widening cross sections are illustrated in Figure 7-11. The addition
of a vehicle lane on either side of an existing roadway (Figure 7-11a) is feasible if the traffic
can be detoured during construction. In this case, the reinforcement may be placed
continuously across the existing embankment and beneath the two new outer fill sections.
Placing both new lanes to one side of the embankment (Figure 7-11b) may allow for
maintaining one lane of traffic flow during construction. With the new fill placed to one side
of the existing embankment, the anchorage of the geosynthetic into the existing embankment
becomes an important design step.
Both the new fill sections and the existing fill sections will most likely settle during and after
fill placement, although the amount of settlement will be greater for the new fill sections.
The existing fills settle because of the influence of the new, adjacent fill loads on their
foundation soils. The amount of settlements is a function of the foundation soils and amount
of load (fill height). When fill is placed to one side of an embankment (Figure 7-11b) the
pavement may need substantial maintenance during construction and until settlements are
nearly complete. Alternatively, light-weight fill could be used to reduce the settlement of the
new fill and existing sections.

FHWA NHI-07-092
Geosynthetic Engineering

7 52

7 Reinforced Embankments
August 2008

Note that the sections in Figure 7-11 do not indicate a geosynthetic reinforcement layer
beneath the existing embankment section. Typically, the reinforcement for the embankment
widening section would be designed assuming no contribution of existing section
geosynthetic in reinforcing the new and combined sections. Therefore, connection of the
new reinforcement to any existing reinforcement is normally not required.
For soft subgrades, where a mud wave is anticipated, construction should be parallel to the
alignment with the outside fill placed in advance of the fill adjacent to the existing
embankment. For firm subgrades, with no mudwave, fill may be placed outward,
perpendicular to the alignment.

(a)

(b)
Figure 7-11. Reinforced embankment construction for roadway widening; a) fill placement
on both sides of existing embankment; b) fill placement on one side of the
existing fill.

FHWA NHI-07-092
Geosynthetic Engineering

7 53

7 Reinforced Embankments
August 2008

7.11 REINFORCEMENT OF EMBANKMENTS COVERING LARGE AREAS

Special considerations are required for constructing large reinforced areas, such as parking
lots, toll plazas, storage yards for maintenance materials and equipment, and construction
pads. Loads are more biaxial than conventional highway embankments, and design strengths
and strain considerations must be the same in all directions. Analytical techniques for
geosynthetic reinforcement requirements are the same as those discussed in Section 7.3.
Because geosynthetic strength requirements will be the same in both directions, including
across the seams, special seaming techniques must often be considered to meet required
strength requirements. Ends of rolls may also require butt seaming. In this case, rolls of
different lengths should be used to stagger the butt seams. Two layers of fabric should be
considered, with the bottom layer seams laid in one direction, and the top layer seams laid
perpendicular to the bottom layer. The layers should be separated by a minimum lift
thickness, usually 12 in. (300 mm), soil layer.
For extremely soft subgrades, the construction sequence must be well planned to
accommodate the formation and movement of mudwaves. Uncontained mudwaves moving
outside of the construction can create stability problems at the edges of the embankment. It
may be desirable to construct the fill in parallel embankment sections, then connect the
embankments to cover the entire area. Another method staggers the embankment load by
constructing a wide, low embankment with a higher embankment in the center. The outside
low embankments are constructed first and act as berms for the center construction. Next, an
adjacent low embankment is constructed from the outside into the existing embankment; then
the central high embankment is spread over the internal adjacent low embankment. Other
construction schemes can be considered depending on the specific design requirements. In
all cases, a perimeter berm system is necessary to contain the mudwave.

7.12 COLUMN SUPPORTED EMBANKMENTS

An alternate approach of embankment construction on soft soils may be used when time
constraints are critical to the success of the project. Column supported embankments (CSE)
with a geosynthetic reinforced load transfer platform are designed to transfer the load of the
embankment through the soft compressible soil layer to a firm foundation, thus eliminating
the construction wait time for dissipation of pore water pressures and minimizing settlement
of the foundation soils. This technology was first used in Sweden in 1971, and has been used
successfully on projects in the U.S. since 1994.

FHWA NHI-07-092
Geosynthetic Engineering

7 54

7 Reinforced Embankments
August 2008

The load from the embankment must be effectively transferred to the columns to prevent
punching of the columns through the embankment fill causing differential settlement at the
surface of the embankment. If the columns are placed close enough together, soil arching
will occur and the load will be transferred to the columns. A conventional CSE is where
the columns are spaced relatively close together, and some battered columns are used at the
sides of the embankment to prevent lateral spreading. In order to minimize the number of
columns required to support the embankment and increase the efficiency of the design, a
geosynthetic reinforced load transfer platform (LTP) may be used. The load transfer platform
consists of one or more layers of geosynthetic reinforcement placed between the top of the
columns and the bottom of the embankment. A CSE with geosynthetic reinforcement is
schematically shown in Figure 7-12.
The key advantage to CSE is that construction may proceed rapidly in one stage. One major
benefit of CSE technology is that it is not limited to any one-column type. Where the
infrastructure precludes high-vibration techniques, the type of column used for the CSE
system may be selected to minimize or eliminate the potential for vibrations. If contaminated
soils are anticipated at a site, the column type may be selected so that there are no spoils from
the installation process. The designer has the flexibility of selection of the most appropriate
column for the project. Total and differential settlement of the embankment may be
drastically reduced when using CSE over conventional approaches. A potential disadvantage
of CSE is often initial construction cost when compared to other solutions. However, if the
time savings when using CSE technology is included in the economic analysis, the cost may
be far less than other solutions.
Design procedures and recommendations are presented in FHWA NHI-06-020, Ground
Improvement Methods Volume II (Elias et al., 2006) the reference manual used with the
3-day NHI Ground Improvement Course #132034. There are two basic design approaches.
One approach models the geosynthetic as a catenary and assumes: one layer of geosynthetic
reinforcement is used; soil arch forms in the embankment; and the reinforcement is deformed
during loading. The other approach uses a beam theory model and assumes: a minimum of
three layers of geosynthetic reinforcement; vertical spacing between reinforcements of 8 to
18 in. (20 0 450 mm); granular fill platform thickness > one-half the clear span between
columns; and soil arch fully develops within the height of platform.
Applications where CSE technology is appropriate for transportation include
embankment stabilization
roadway widening
bridge approach fill stabilization
bridge abutment and other foundation support
FHWA NHI-07-092
Geosynthetic Engineering

7 55

7 Reinforced Embankments
August 2008

A considerable amount of highway widening and reconstruction work will be required in


future years. Some of this work will involve building additional lanes immediately adjacent
to existing highways constructed on moderate to high fills over soft cohesive soils, such as
those found in wetland areas. For this application, differential settlement between the
existing and new construction is an important consideration, in addition to embankment
stability. Support of the new fill on CSE offers a viable design alternative to conventional
construction.
CSE may be used whenever an embankment must be constructed on soft compressible soil.
To date, the technology has been limited to embankment heights in the range of 33 feet (10
m). CSE technology reduces post construction settlements of the embankment surface to
typically less than 2 to 4 in. (50 to 100 mm). A generalized summary of the factors that
should be considered when assessing the feasibility of utilizing CSE technology on a project
is presented in FHWA NHI-06-020, Ground Improvement Methods Reference Manual
Volume II.

Figure 7-12. Column supported embankment with geosynthetic reinforcement.

FHWA NHI-07-092
Geosynthetic Engineering

7 56

7 Reinforced Embankments
August 2008

7.13 REFERENCES

AASHTO, Standard Specifications for Geotextiles - M 288 (2006). Standard Specifications


for Transportation Materials and Methods of Sampling and Testing, 26th Edition,
American Association of State Transportation and Highway Officials, Washington,
D.C.,
ASTM, Annual Books of ASTM Standards, (2006). Volume 4.13 Geosynthetics, American
International, West Conshohocken, Pennsylvania.
Bell, J.R. (1980). Design Criteria for Selected Geotextile Installations, Proceedings of the
1st Canadian Symposium on Geotextiles, pp. 35-37.
Bonaparte, R. and Christopher, B.R. (1987). Design and Construction of Reinforced
Embankments Over Weak Foundations, Proceedings of the Symposium on Reinforced
Layered Systems, Transportation Research Record 1153, Transportation Research Board,
Washington, D.C., pp. 26-39.
Bonaparte, R., Holtz, R.R. and Giroud, J.P. (1985). Soil Reinforcement Design Using
Geotextiles and Geogrids, Geotextile Testing and The Design Engineer, J.E. Fluet, Jr.,
Editor, ASTM STP 952, 1987, Proceedings of a Symposium held in Los Angeles, CA,
July 1985, pp. 69-118.
Cheney, R.S. and Chassie, R.G. (1993). Soils and Foundations Workshop Manual, HI-88099, 395 p.
Christopher, B.R. and Holtz, R.D. (1985). Geotextile Engineering Manual, FHWA-TS86/203, 1044 p.
Christopher, B.R. and Holtz, R.D. (1989). Geotextile Design and Construction Guidelines,
FHWA-HI-90-001, 297 p.
Dunnicliff, J. (1998). Geotechnical Instrumentation; FHWA HI-98-034; NHI course No.
132041 reference manual; 238 pp.
Elias, V., Welsh, J., Warren, J., Lukas, R., Collin, J.G. and Berg, R.R. (2006). Ground
Improvement Methods; FHWA NHI-06-019 Volume I and NHI-06-020 Volume II; NHI
course No. 132034 reference manual; 536 p and 520 p.
Fowler, J. (1981). Design, Construction and Analysis of Fabric-Reinforced Embankment
Test Section at Pinto Pass, Mobile, Alabama, Technical Report EL-81-7, USAE
Waterways Experiment Station, 238 p.
Fowler, J. and Haliburton, T.A. (1980). Design and Construction of Fabric Reinforced
Embankments, The Use of Geotextiles for Soil Improvement, Preprint 80-177, ASCE
Convention, pp. 89-118.
FHWA NHI-07-092
Geosynthetic Engineering

7 57

7 Reinforced Embankments
August 2008

Haliburton T.A., Lawmaster, J.D. and McGuffey, V.E. (1982). Use of Engineering Fabrics
in Transportation Related Applications, Final Report Under Contract No. DTFH61-80C-0094.
Haliburton, T.A., Anglin, C.C. and Lawmaster, J.D. (1978a). Testing of Geotechnical Fabric
for Use as Reinforcement, Geotechnical Testing Journal, American Society for Testing
and Materials, Vol. 1, No. 4, pp. 203-212.
Haliburton, T.A., Anglin, C.C. and Lawmaster, J.D. (1978b). Selection of Geotechnical
Fabrics for Embankment Reinforcement, Report to U.S. Army Engineer District, Mobile,
Oklahoma State University, Stillwater, 138p.
Haliburton, T.A., Douglas, P.A. and Fowler, J. (1977). Feasibility of Pinto Island as a LongTerm Dredged Material Disposal Site, Miscellaneous Paper, D-77-3, U.S. Army
Waterways Experiment Station.
Hird, C.C. and Jewell, R.A. (1990). Theory of Reinforced Embankments, Reinforced
Embankments - Theory and Practice, Shercliff, D.A., Ed., Thomas Telford Ltd., London,
UK, pp. 117-142.
Holtz, R.D. (1990). Design and Construction of Geosynthetically Reinforced Embankments
on Very Soft Soils, State-of-the-Art Paper, Session 5, Performance of Reinforced Soil
Structure, Proceedings of the International Reinforced Soil Conference, Glasgow,
British Geotechnical Society, pp. 391-402.
Holtz, R.D. (1989). Treatment of Problem Foundations for Highway Embankments,
Synthesis of Highway Practice 147, National Cooperative Highway Research Program,
Transportation Research Board, Washington, D.C., 72p.
Humphrey, D.N. and Rowe, R.K. (1991). Design of Reinforced Embankments - Recent
Developments in the State of the Art, Geotechnical Engineering Congress 1991, McLean,
F., Campbell, D.A. and Harris, D.W., Eds., ASCE Geotechnical Special Publication No.
27, Vol. 2, June, pp. 1006-1020.
Humphrey, D.N. and Holtz, R.D. (1989). Effects of a Surface Crust on Reinforced
Embankment Design, Proceedings of Geosynthetics '89, Industrial Fabrics Association
International, St. Paul, MN, Vol. 1, pp. 136-147.
Humphrey, D.N. (1987). Discussion of Current Design Methods by R.M. Koerner, B-L Hwu
and M.H. Wayne, Geotextiles and Geomembranes, Vol. 6, No. 1, pp. 89-92.
Humphrey, D.N. and Holtz, R.D. (1987). Use of Reinforcement for Embankment Widening,
Proceedings of Geosynthetics '87, Industrial Fabrics Association International, St. Paul,
MN, Vol. 1, pp. 278-288.

FHWA NHI-07-092
Geosynthetic Engineering

7 58

7 Reinforced Embankments
August 2008

Humphrey, D.N. and Holtz, R.D. (1986). Reinforced Embankments - A Review of Case
Histories, Geotextiles and Geomembranes, Vol. 4, No. 2, pp.129-144.
Jewell, R.A. (1988). The Mechanics of Reinforced Embankments on Soft Soils, Geotextiles
and Geomembranes, Vol. 7, No. 4, pp.237-273.
Jrgenson, L. (1934). The Shearing Resistance of Soils, Journal of the Boston Society of
Civil Engineers. Also in Contribution to Soil Mechanics, 1925-1940, BSCE, pp. 134217.
Koerner, R.M., Editor (1990). The Seaming of Geosynthetics, Special Issue, Geotextiles and
Geomembranes, Vol. 9, Nos. 4-6, pp. 281-564.
Ladd, C.C. (1991). Stability Evaluation During Staged Construction, 22nd Terzaghi Lecture,
Journal of Geotechnical Engineering, American Society of Civil Engineers, Vol. 117,
No. 4, pp. 537-615.
Leshchinsky, D. (1987). Short-Term Stability of Reinforced Embankment over Clayey
Foundation, Soils and Foundations, The Japanese Society of Soil Mechanics and
Foundation Engineering, Vol. 27, No. 3, pp. 43-57.
Mayne, P.W., Christopher, B.R. and DeJong, J. (2002). Subsurface Investigations
Geotechnical Site Characterization, FHWA NHI-01-031, NHI course No. 132031
reference manual, 300 pp.
McGown, A., Andrawes, K.Z., and Kabir, M.H. (1982). Load-Extension Testing of
Geotextiles Confined in Soil, Proceedings of the Second International Conference on
Geotextiles, Las Vegas, Vol. 3, pp. 793-798.
Perloff, W.H. and Baron, W. (1976). Soil Mechanics: Principles and Applications, Ronald,
745 p.
Rowe, R.K. and Mylleville, B.L.J. (1990). Implications of Adopting an Allowable
Geosynthetic Strain in Estimating Stability, Proceedings of the 4th International
Conference on Geotextiles, Geomembranes, and Related Products, The Hague, Vol. 1,
pp. 131-136.
Rowe, R.K. and Mylleville, B.L.J. (1989). Consideration of Strain in the Design of
Reinforced Embankments, Proceedings of Geosynthetics '89, Industrial Fabrics
Association International, St. Paul, MN, Vol. 1, pp. 124-135.
Rowe, R.K. and Soderman, K.L. (1987a). Reinforcement of Embankments on Soils Whose
Strength Increases With Depth, Proceedings of Geosynthetics '87, Industrial Fabrics
Association International, St. Paul, MN, Vol. 1, pp. 266-277.

FHWA NHI-07-092
Geosynthetic Engineering

7 59

7 Reinforced Embankments
August 2008

Rowe, R.K. and Soderman, K.L. (1987b). Stabilization of Very Soft Soils Using High
Strength Geosynthetics: The Role of Finite Element Analyses, Geotextiles and
Geomembranes, Vol. 6, No. 1, pp. 53-80.
Sabatini, P.J., Bachus, R.C., Mayne, P.W., Schneider, J.A. and Zettler, T.E. (2002). GEC
No. 5 Evaluation of Soil and Rock Properties, Geotechnical Engineering Circular No.
5, FHWA IF-02-034, 385 pp.
Samtani, N.C. and Nowatzki, E.A. (2006). Soils and Foundations Workshop Reference
Manual, FHWA NHI-06-088, NHI course No. 132012 reference manual.
Silvestri, V. (1983). The Bearing Capacity of Dykes and Fills Founded on Soft Soils of
Limited Thickness, Canadian Geotechnical Journal, Vol. 20, No. 3, pp. 428-436.
Tan, S.L. (1990).
Stress-Deflection Characteristics of Soft Soils Overlain with
Geosynthetics, MSCE Thesis, University of Washington, 146 p.
Terzaghi, K. and Peck, R.B. (1967). Soil Mechanics in Engineering Practice, 2nd Edition,
John Wiley & Sons, New York, 729 p.
U.S. Department of the Navy (1986). Foundations and Earth Structures, Design Manual 7.2,
Naval Facilities Engineering Command, Alexandria, VA, (can be downloaded from
http://www.geotechlinks.com)
Vesic, A.A. (1975). Bearing Capacity of Shallow Foundations, Chapter 3 in Foundation
Engineering Handbook, Winterkorn and Fang, Editors, Van Nostrand Reinhold, pp. 121147.
Washington State Department of Transportation (1997).
Embankment Reinforcement.

High Strength Geotextile for

Wager, O. (1981). Building of a Site Road over a Bog at Kilanda, Alvsborg County, Sweden
in Preparation for Erection of Three 400kV Power Lines, Report to the Swedish State
Power Board, AB Fodervvnader, Bor, Sweden, 16p.

FHWA NHI-07-092
Geosynthetic Engineering

7 60

7 Reinforced Embankments
August 2008

8.0 REINFORCED SLOPES

8.1

BACKGROUND

Even if foundation conditions are satisfactory, slopes may be unstable at the desired slope
angle. For new construction, the cost of fill, right-of-way, and other considerations may
make a steeper slope desirable. Existing slopes, natural or manmade, may also be unstable,
as is painfully obvious when they fail. As shown in Figure 8-1, multiple layers of geogrids
or geotextiles may be placed in an earthfill slope during construction or reconstruction to
reinforce the soil and provide increased slope stability. Reinforced soil slopes (RSS) are a
form of mechanically stabilized earth that incorporates planar reinforcing elements in
constructed earth-sloped structures with face inclinations less than 70 (FHWA RD-89-043
{Christopher et al., 1989}). MSE structures with face inclinations of 70 to 90 are classified
as walls. These are addressed in Chapter 9.
In this chapter, analysis of the reinforcement and construction details required to provide a
safe slope will be reviewed. The design method included in this chapter was first developed
in the early 1980s for landslide repair in northern California. This approach has been
validated by the thousands of reinforced soil slopes constructed over the past two decades
and through results of an extensive FHWA research program on reinforced soil structures as
detailed in Reinforced Soil Structures Volume I - Design and Construction Guidelines and
Volume II - Summary of Research and Systems Information (Christopher et al., 1989).
Contracting options and guideline specifications are included from FHWA-SA-93-025
Guidelines for Design, Specification, and Contracting of Geosynthetic Mechanically
Stabilized Earth Slopes on Firm Foundations (Berg, 1993). The guidelines within this
chapter are consistent with those detailed in FHWA NHI-00-043 Mechanically Stabilized
Earth Walls and Reinforced Soil Slopes, Design and Construction Guidelines (Elias et al.,
2001), the reference manual for NHI courses 132042 and 132043.
8.2

APPLICATIONS

Geosynthetics are primarily used as slope reinforcement for construction of slopes to angles
steeper than those constructed with the fill material being used, as illustrated in Figure 8-1a.
Geosynthetics used in this manner can provide significant project economy by:
creating usable land space at the crest or toe of the reinforced slope;
reducing the volume of fill required;
allowing the use of less-than-high-quality fill; and
eliminating the expense of facing elements required on MSE walls.
FHWA NHI-07-092
Geosynthetic Engineering

8 1

8 Reinforced Slopes
August 2008

Figure 8-1.

Use of geosynthetics in engineered slopes: (a) to increase stability of a slope;


and (b) to provide improved compaction and surficial stability at edge of
slopes (after Berg et al., 1990).

FHWA NHI-07-092
Geosynthetic Engineering

8 2

8 Reinforced Slopes
August 2008

Applications which highlight some of these advantages, illustrated in Figure 8-2, include:
construction of new highway embankments;
construction alternative to retaining walls;
widening of existing highway embankments; and
repair of failed slopes.
Other applications of reinforced soil slopes include:
steepening end slopes of approach embankments and decreasing bridge spans;
temporary road widening for detours;
steepening slopes to decrease length of box culverts;
upstream/downstream face stability and increased height of dams;
construction of permanent levees and temporary flood control structures; and
embankment construction with wet, fine-grained soils

Figure 8-2.

Applications of RSSs: (a) construction of new embankments; (b) alternative


to retaining walls; (c) widening existing embankments; and (d) repair of
landslides (after Tensar, 1987).

FHWA NHI-07-092
Geosynthetic Engineering

8 3

8 Reinforced Slopes
August 2008

The design of reinforcement for safe, steep slopes requires rigorous analysis. The design of
reinforcement for these applications is critical because reinforcement failure results in slope
failure. To date, several thousand reinforced slope structures have been successfully
constructed at various slope face angles. The tallest structure constructed in the U.S. to date,
is a 1H:1V reinforced slope 242 feet (74 m) high (Yeager Airport {Geosynthetics, 2008}).
A second purpose of geosynthetics placed at the edges of a compacted fill slope is to provide
lateral resistance during compaction (Iwasaki and Watanabe, 1978) and surficial stability
(Thielen and Collin, 1993). The increased lateral resistance allows for increased compacted
soil density over that normally achieved and provides increased lateral confinement for soil
at the face. Even modest amounts of reinforcement in compacted slopes have been found to
prevent sloughing and reduce slope erosion. Edge reinforcement also allows compaction
equipment to operate safely near the edge of the slope. Further compaction improvements
have been found in cohesive soils using geosynthetics with in-plane drainage capabilities
(e.g., nonwoven geotextiles) which allow for rapid pore pressure dissipation in the
compacted soil (Zornberg and Mitchell, 1992).
Design for the compaction improvement application is simple. Place a geogrid or geotextile
that will survive construction at every lift or every other lift in a continuous plane along the
edge of the slope. Only narrow strips, about 4 to 6 ft (1.2 to 2 m) in width, at 12 to 20 in.
(0.3 to 0.5 m) vertical spacing are required. No reinforcement design is required if the
overall slope is found to be safe without reinforcement. Where the slope angle approaches
the angle of repose of the soil, a face stability analysis should be performed using the method
presented in Section 8.3. Where reinforcement is required by analysis, the narrow strip
geosynthetic may be considered as a secondary reinforcement used to improve compaction
and to stabilize the slope face between primary layers.

8.3

DESIGN GUIDELINES FOR REINFORCED SLOPES

8.3-1 Design Concepts


The overall design requirements for reinforced slopes are similar to those for unreinforced
slopes: the factor of safety must be adequate for both the short-term and long-term
conditions and for all possible modes of failure.
Permanent, critical reinforced structures should be designed using comprehensive slope
stability analyses. A structure may be considered permanent if its design life is greater than 3
years. An application is considered critical if there is mobilized tension in the reinforcement
for the life of the structure, if reinforcement failure results in failure of the structure, or if the
FHWA NHI-07-092
Geosynthetic Engineering

8 4

8 Reinforced Slopes
August 2008

consequences of failure include personal injury or significant property damage (Bonaparte


and Berg, 1987). A RSS is typically not considered critical if the safety factor against
instability of the same unreinforced slope is greater than 1.1, and the reinforcement is used to
increase the safety factor.
Failure modes of reinforced slopes (Berg et al., 1989) include:
1. internal, where the failure plane passes through the reinforcing elements;
2. external, where the failure surface passes behind and underneath the reinforced mass;
and
3. compound, where the failure surface passes partially behind and partially through the
reinforced soil mass.
In optimally designed slopes, the stability safety factor will be approximately equal in two or
all three modes.
Reinforced slopes are currently analyzed using modified versions of the classical limit
equilibrium slope stability methods. A circular or wedge-type potential failure surface is
assumed, and the relationship between driving and resisting forces or moments determines
the slope's factor of safety. Based on their tensile capacity and orientation, reinforcement
layers intersecting the potential failure surface are generally, and in this manual, assumed to
increase the resisting moment or force (see Section 8.3-4 for alternate assumptions). The
tensile capacity of a reinforcement layer is the minimum of its allowable pullout resistance
behind, or in front of, the potential failure surface and/or its long-term design tensile strength,
whichever is smaller. A wide variety of potential failure surfaces must be considered,
including deep-seated surfaces that may pass partially through or behind the reinforced zone.
Detailed design of reinforced slopes is performed by determining the factor of safety with
sequentially modified reinforcement layouts until the target factor of safety is achieved.
8.3-2 Design of Reinforced Slopes
The steps for design of a reinforced soil slope are:

STEP 1.

Establish the geometric, loading, and performance requirements for design.

STEP 2.

Determine the subsurface stratigraphy and the engineering properties of the


in-situ soils.

FHWA NHI-07-092
Geosynthetic Engineering

8 5

8 Reinforced Slopes
August 2008

STEP 3.

Determine the engineering properties of the available fill soils.

STEP 4.

Evaluate design parameters for the reinforcement (design reinforcement


strength, durability criteria, soil-reinforcement interaction).

STEP 5.

Determine the factor of safety of the unreinforced slope.

STEP 6.

Design reinforcement to provide stable slope.


Method A - Direct reinforcement design
Method B - Trial reinforcement layout analysis

STEP 7.

Select slope face treatment.

STEP 8.

Check external stability.

STEP 9.

Check seismic stability.

STEP 10. Evaluate requirements for subsurface and surface water control.

STEP 11. Develop specifications and contract documents.


Details required for each step, along with equations for analysis, are presented in section 8.33. The procedure in section 8.3-3 assumes that the slope will be constructed on a stable
foundation (i.e., a circular or wedge-shaped failure surface through the foundation is not
critical and local bearing support is clearly adequate). The user is referred to Chapter 7 for
use of reinforcement in embankments over weak foundation soils.
For slide repair applications, it is also very important that solutions address the cause of
original failure. Make sure that the new reinforced soil slope will not have the same
problems. If water table or erratic water flows exist, particular attention must be paid to
drainage. In natural soils, it is also necessary to identify any weak seams that could affect
stability.
FHWA NHI-07-092
Geosynthetic Engineering

8 6

8 Reinforced Slopes
August 2008

8.3-3 Reinforced Slope Design Guidelines


The following provides design procedure details for reinforced soil slopes.
STEP 1.

Establish the geometric, load, and performance requirements for design


(Figure 8-3).

Geometric and load requirements:


a. Slope height, H
b. Slope angle,
c. External (surcharge) loads:
Surcharge load, q
Temporary live load, q
Design Seismic acceleration, Am (see Division 1A, AASHTO Standard
Specifications for Highway Bridges (2002))
Traffic barrier load see article 2.7 of AASHTO Standard Specifications for
Highway Bridges (2002) and AASHTO Roadside Design Guide
(1989)
Performance requirements:
a. External stability and settlement
Horizontal sliding of the MSE mass along its base, FS > 1.3
Deep-seated (overall stability), FS > 1.3
Local bearing failure (lateral squeeze), FS > 1.3
Dynamic loading: FS > 1.1
Settlement post construction magnitude and time rate based on project
requirements.
b. Compound failure modes (for planes passing behind and through the reinforced
mass)
Compound failure surfaces, FS > 1.3
c. Internal stability
Internal failure surfaces, FS > 1.3

FHWA NHI-07-092
Geosynthetic Engineering

8 7

8 Reinforced Slopes
August 2008

Figure 8-3.

Requirements for design of a reinforced slope.

STEP 2. Determine the engineering properties of the in-situ soils in the slope.
a. Determine the foundation and retained soil (i.e., soil beneath and behind
reinforced zone) profiles along the alignment (every 100 to 200 ft (30 to 60 m),
depending on the homogeneity of the subsurface profile) deep enough to evaluate
a potential deep-seated failure (recommended exploration depth is twice the
height of the slope or to refusal).
b. Determine the foundation and retained fill soil strength parameters (cu, Nu and c',
N'); unit weight (wet and dry); and foundation consolidation parameters (Cc, Cr, cv
and N'p) for each layer.
c. Locate the groundwater table, dw, and piezometric surfaces (especially important
if water will exit slope).
d. For slope and landslide repairs, identify the cause of instability and locate the
previous failure surface.

STEP 3.

Determine properties of reinforced fill and, if different, the fill behind the
reinforced zone.
See recommendations in Section 8.4-1.

a. Gradation and plasticity index


b. Compaction characteristics and placement requirements
FHWA NHI-07-092
Geosynthetic Engineering

8 8

8 Reinforced Slopes
August 2008

c. Shear strength parameters, cu, Nu and c', N'


d. Chemical composition of soil, pH

STEP 4.

Evaluate design parameters for the reinforcement.


See recommendations in section 8.4-1 and Appendix H.

a. Allowable geosynthetic strength, Tal = ultimate strength (Tult) ) reduction factors


(RF) for creep, installation damage, and durability
b. Pullout Resistance: Use FS > 1.5 for granular soils and FS > 2 for cohesive soils,
with a minimum embedment length beyond the failure plane (Le) equal to 3 feet
(1 m).

STEP 5.
a.

Check unreinforced stability.

Evaluate unreinforced stability to determine: if reinforcement is required; critical


nature of the design (i.e., unreinforced FS < or > 1); potential deep-seated failure
problems; and the preliminary extent of the reinforced zone.
Perform a stability analysis using conventional stability methods (see FHWA
Soils and Foundations Reference Manual {Samtani and Notwatski, 2006)}) to
determine safety factors and driving moments for potential failure surfaces.
Use circular arc and/or sliding wedge methods, as appropriate, and consider
failures through the toe, through the face (at several elevations), and deep seated
below the toe. Failure surface exit points should be defined within each of the
potential failure zones.
A number of stability analysis computer programs are available for rapid evaluation
(e.g., the STABL family of programs developed at Purdue University and FHWA
developed ReSSA program). In all cases, you should perform a few calculations by
hand to verify reasonability of the computer program.

b.

Determine the size of the critical zone to be reinforced.


Examine the full range of potential failure surfaces found to have:
Unreinforced safety factor FSU < Required safety factor FSR
Plot all of these surfaces on the slope's cross-section.
The surfaces that just meet the target factor of safety roughly envelope the limits
of the critical zone to be reinforced, as illustrated in Figure 8-4.

FHWA NHI-07-092
Geosynthetic Engineering

8 9

8 Reinforced Slopes
August 2008

Figure 8-4.

c.

Critical zone defined by rotational and sliding surface that meet the
required safety factor.

Critical failure surfaces extending below the toe of the slope are indications of deep
foundation and edge bearing capacity problems that must be addressed prior to
completion of design (see Step 8). For such cases, a more-extensive foundation
analysis is warranted. Geosynthetics may be used to reinforce the base of the
embankment and to construct toe berms for improved embankment stability, as
reviewed in Chapter 7. Other foundation improvement measures should be
considered (see FHWA NHI-06-019 and FHWA NHI-06-020, Ground Improvement
Methods {Elias et al., 2006}).

STEP 6.

Design reinforcement to provide for a stable slope.

Several approaches are available for the design of slope reinforcement, many of which are
contained in Christopher and Holtz (FHWA-TS-86/203,1985). Two methods are presented
in this section. The first method uses a direct design approach to obtain the reinforcing
requirements. The second method, analyzes trial reinforcement layouts.
Method A - Direct design approach.
The first method, presented in Figure 8-5 for a rotational slip surface, uses any
conventional slope stability computer program, and the steps necessary to manually
calculate the reinforcement requirements. This design approach can accommodate
fairly complex conditions depending on the analytical method used (e.g., Bishop,
Janbu, etc.).
FHWA NHI-07-092
Geosynthetic Engineering

8 10

8 Reinforced Slopes
August 2008

The assumed orientation of the reinforcement tensile force influences the calculated
slope safety factor.
In a conservative approach, the deformability of the
reinforcements is not taken into account; therefore, the tensile forces per unit width of
reinforcement, Tr, are always assumed to be horizontal to the reinforcements, as
illustrated in Figure 8-5. However, close to failure, the reinforcements may elongate
along the failure surface, and an inclination from the horizontal can be considered.
Tensile force direction is therefore dependent on the extensibility of the
reinforcements used, and for continuous (i.e., 100% coverage) extensible
geosynthetic reinforcement, a T inclination tangent to the sliding surface is
recommended. For discontinuous strips (i.e., < 100% coverage) of geosynthetic
reinforcement, a horizontal orientation should be conservatively assumed.
Judgment and experience in selection of the most appropriate design is required. The
following steps are necessary:
a. Calculate the total reinforcement tension per unit width of slope, TS, required
to obtain the target minimum factor of safety for each potential failure circle
inside the critical zone (Step 5) that extends through or below the toe of the
slope (see Figure 8-5). Use the following equation:
TS = (FS R FSU )

MD
D

where:
TS = sum of required tensile force per unit width of reinforcement
(considering rupture and pullout) in all reinforcement layers
intersecting the failure surface
MD = driving moment about the center of the failure circle
D = the moment arm of TS about the center of failure circle
= radius of circle R for continuous, sheet type geosynthetic
reinforcement (i.e., assumed to act tangentially to the circle)
= vertical distance, Y, to centroid of TS for discrete element, strip type
reinforcement (Assume H/3 above slope base for preliminary
calculations (i.e., assumed to act in a horizontal plane intersecting the
failure surface at H/3 above the slope base).
FSR = target minimum slope safety factor which is applied to both the soil
and reinforcement
FSU = unreinforced slope safety factor

FHWA NHI-07-092
Geosynthetic Engineering

8 11

8 Reinforced Slopes
August 2008

Factor of safety of unreinforced slope:

FS u =
where: W
LSP
q
Jf

=
=
=
=

R esisting Moment ( M R ) LSP f R


=
Driving Moment ( M D )
Wx + q d

weight of sliding earth mass


length of slip plane
surcharge
shear strength of soil

Factor of safety of reinforced slope:

FS = FS u +
where: Ts
D
D
D

=
=
=
=

Figure 8-5.

Ts D
MD

sum of available tensile force per width of reinforcement for all reinforcement layers
moment arm of Ts about the center of rotation
R for continuous geosynthetic reinforcement
Y for discontinuous, strip type geosynthetic reinforcement

Rotational shear approach to determine required strength of reinforcement.

FHWA NHI-07-092
Geosynthetic Engineering

8 12

8 Reinforced Slopes
August 2008

TS-MAX, the largest TS calculated establishes the total required design tension.
NOTE: The minimum safety factor usually does not control the location
of TS-MAX, the most critical surface is the surface requiring the largest
magnitude of reinforcement.

b. Determine the total required design tension per unit width of slope, TS-MAX, using
the charts in Figure 8-6 and compare results with Step 6a. If substantially
different, check the validity of the charts based on the limiting assumptions listed
in the figure and recheck Steps 5 and 6a.
Figure 8-6 provides a method for quickly checking the computer-generated
results. The charts are based upon simplified analysis methods of two-part and
one-part wedge-type failure surfaces and are limited by the assumptions noted on
the figure. Note that Figure 8-6 is not intended to be a single design tool. Other
design charts are also available from Jewell et al. (1984); Jewell (1990); Werner
and Resl (1986); Ruegger (1986); and Leshchinsky and Boedeker (1989). Several
computer programs are also available for analyzing a slope with given
reinforcement and can also be used as a check. Judgment in selection of other
appropriate design methods (ex., most conservative or experience) is required.
c. Determine the distribution of reinforcement:

For low slopes (H < 20 ft (6 m)) assume a uniform reinforcement distribution


and use TS-MAX to determine spacing or the required tension Tmax requirements
for each reinforcement layer.

For high slopes (H > 20 ft (6 m)), divide the slope into two (top and bottom)
or three (top, middle, and bottom) reinforcement zones of equal height and use
a factored TS-MAX in each zone for spacing or design tension requirements.
The total required tension in each zone is found from:

For two zones:


TBottom = TS-MAX
TTop = TS-MAX
For three zones:
TBottom = TS-MAX
TMiddle = a TS-MAX
TTop = 1/6 TS-MAX
The force is assumed to be uniformly distributed over the entire zone.
FHWA NHI-07-092
Geosynthetic Engineering

8 13

8 Reinforced Slopes
August 2008

a. Reinforcement force coefficient

b. Reinforcement length ratio

CHART PROCEDURE:
1) Determine force coefficient K from figure above, where Nr = friction angle of reinforced fill:

tan r
FS R

f = tan 1
2) Determine:
where:

TS-MAX = 0.5 K (r (H)2


H' = H + q/(r
q = a uniform load

3) Determine the required reinforcement length at the top LT and bottom LB of the slope from the figure above.
LIMITING ASSUMPTIONS
Extensible reinforcement.
Slopes constructed with uniform, cohesionless soil, c = 0).
No pore pressures within slope.
Competent, level foundation soils.
No seismic forces.
Uniform surcharge nor greater than 0.2 (r H.
Relatively high soil/reinforcement interface friction angle, Nsg = 0.9 Nr (may not be appropriate for some
geotextiles).

Figure 8-6. Sliding wedge approach to determine the coefficient of earth pressure K
(after Schmertmann et al., 1987).
NOTE: Charts The Tensar Corporation.
FHWA NHI-07-092
Geosynthetic Engineering

8 14

8 Reinforced Slopes
August 2008

d. Determine reinforcement vertical spacing Sv or the maximum design tension Tmax


requirements.
For each zone, calculate the design tension, Tmax, requirements for each reinforcing layer
based on an assumed Sv or, if the allowable reinforcement strength is known, calculate
the minimum vertical spacing and number of reinforcing layers N required for each zone
based on:
Tmax =

where:
Rc

Sv
Tzone
Ta
Hzone
N

Tzone S v Tzone
=
Ta Rc
H zone
N

= percent coverage of reinforcement, in plan view, which equals


the width of the reinforcement divided by the horizontal
spacing (Rc = 1 for continuous sheets)
= vertical spacing of reinforcement; should be multiples of
compaction layer thickness for ease of construction
= maximum reinforcement tension required for each zone; Tzone
equals TS-MAX for low slopes (H < 20 ft (6 m))
= Tal
= height of zone, and is equal to Ttop, Tmiddle, and TBottom for high
slopes (H > 20 ft (6 m))
= number of reinforcement layers

Short (4 to 6 ft (1.2 to 2 m)) lengths of intermediate reinforcement layers can be used


to maintain a maximum vertical spacing of 24 in. (600 mm) or less for face stability
and compaction quality (Figure 8-7).
-

For slopes less than 1H:1V, closer spaced reinforcements (i.e., every lift or
every other lift, but no greater than 16 in. (400 mm)) generally preclude
having to wrap the face. Wrapped faces are usually required for steeper
slopes to prevent face sloughing. Alternative vertical spacings can be used to
prevent face sloughing, but in these cases a face stability analysis should be
performed using the method presented in this section or by evaluating the face
as an infinite slope (Thielen and Collin, 1993) using:

FHWA NHI-07-092
Geosynthetic Engineering

8 15

8 Reinforced Slopes
August 2008

F .S . =

c' H + ( g w ) H z cos 2 tan ' + F g (cos sin + sin 2 tan ' )

g H z cos sin

where:

c
N
(g
(w
z
H
$
Fg

=
=
=
=
=

effective cohesion
effective friction angle
saturated unit weight of soil
unit weight of water
vertical depth to failure plane defined by the depth of
saturation
= vertical slope height
= slope angle
= summation of geosynthetic resisting force.

Fg is computed with only the tangential force component. The available


resisting force at each geosynthetic layer is limited by (i) long-term allowable
strength, (ii) long-term pullout resistance of the geosynthetic in the slide mass,
and (iii) long-term pullout resistance of the geosynthetic behind the slide
mass.

Intermediate reinforcement should be placed in continuous layers and need


not be as strong as the primary reinforcement, but in any case, all
reinforcements should be strong enough to survive installation (e.g., see
Tables 5-1 and 5-2) and provide localized tensile reinforcement to the surficial
soils.

Figure 8-7 Spacing and embedding requirements for slope reinforcement showing
primary and intermediate reinforcement layout.
FHWA NHI-07-092
Geosynthetic Engineering

8 16

8 Reinforced Slopes
August 2008

If the interface friction angle of the intermediate reinforcement sr is less than


that of the primary reinforcement r, then sr should be used in the analysis for
the portion of the failure surface intersecting the reinforced soil zone.

e. Check partial slope heights to ensure that the rule-of-thumb reinforcement force
distribution is adequate for critical or complex structures. Recalculate TS using the
equation in Step 6a to determine potential failure of partial slope heights. Checks
may be performed at heights where reinforcement vertical spacing, strengths, and/or
lengths change. Alternatively, checks may be performed directly above each layer of
primary reinforcement.
f. Determine the reinforcement lengths required.
The embedment length, Le, of each reinforcement layer beyond the most critical
sliding surface found in Step 6a (i.e., circle found for TS-MAX) must be sufficient to
provide adequate pullout resistance based on:
Le =

Tmax FS
F * 'v 2 Rc C

where:
Tmax = maximum required reinforcement tension
FS
= 1.5 for granular soils and 2 for cohesive soils, with a minimum
embedment length, Le = 3 ft (1 m).
Le C = the total surface area per unit width of the reinforcement in the
resistive zone behind the failure surface
Le
= the embedment or adherence length in the resisting zone behind the
failure surface
C
= the reinforcement effective unit perimeter; e.g., C = 2 for strips,
grids, and sheets
F*
= the pullout resistance (or friction-bearing-interaction) factor

= a scale effect correction factor to account for a non linear stress


reduction over the embedded length of highly extensible
reinforcements, based on laboratory data (generally 0.6 to 1.0 for
geosynthetic reinforcements, see Elias et al., 2001).
= the effective vertical stress at the soil-reinforcement interfaces
'v
Rc
= reinforcement coverage ratio = width or reinforcement divided by
the center-to-center horizontal reinforcement spacing = 1 for
continuous reinforcement.

FHWA NHI-07-092
Geosynthetic Engineering

8 17

8 Reinforced Slopes
August 2008

Minimum value of Le is 3 ft (1 m). For cohesive soils, check Le for both short- and
long-term pullout conditions. For long-term design, use N'r with cr = 0. For shortterm evaluation, conservatively use Nr and cr = 0 from consolidated-undrained triaxial
tests, or run pullout tests.

Plot the reinforcement lengths as obtained from the pullout evaluation on a slope
cross section containing the rough limits of the critical zone determined in Step 5, see
Figure 8-8 (note, that this example uses a uniform reinforcement spacing).
- The length required for sliding stability at the base will generally control the
length of the lower reinforcement levels.
- Lower layer lengths must extend to the limits of the critical zone as shown in
Figure 8-8. Longer reinforcements may be required to resolve deep seated failure
problems (see Step 8).
- Upper levels of reinforcement may not have to extend to the limits of the critical
zone provided sufficient reinforcement exists in the lower levels to provide FSR
for all circles within the critical zone as shown in Figure 8-8.

Figure 8-8.

Developing reinforcement lengths.

FHWA NHI-07-092
Geosynthetic Engineering

8 18

8 Reinforced Slopes
August 2008

Check that the sum of the reinforcement forces passing though each failure surface is
greater than Ts, from Step 6a, required for that surface.
- Only count reinforcement that extends 3 ft (1 m) beyond the surface to
account for pullout resistance.
- If the available reinforcement resistance is not sufficient, increase the length
of the reinforcement not passing through the surface or increase the strength
of lower level reinforcement.

Simplify the layout by lengthening some reinforcement layers to create two or three
sections of equal reinforcement length for ease of construction and inspection.

Reinforcement layers generally do not need to extend to the limits of the critical zone,
except for the lowest levels of each reinforcement section.

Check the obtained length using Chart B in Figure 8-6. Note: Le is already included
in the total length, Lt and LB from Chart B.

g. Check design lengths of complex designs.

When checking a design that has zones of different reinforcement length, lower zones
may be over-reinforced to provide reduced lengths of upper reinforcement levels.

In evaluating the length requirements for such cases, the reinforcement pullout
stability must be carefully checked in each zone for the critical surfaces exiting at the
base of each length zone.

STEP 6.

Method B - Trial reinforcement analysis.


Another way to design reinforcement for a stable slope is to develop a trial layout
of reinforcement and analyze the reinforced slope with a computer program, such
as the ReSSA program developed for FHWA (see Elias et al. {2001}) for more
details and guidelines for using this program). Layout includes number, length,
design strength, and vertical distribution of the geosynthetic reinforcement. The
charts presented in Figure 8-6 provide a method for generating a preliminary
layout. Note that these charts were developed with the specific assumptions noted
on the figure.
Analyze the RSS with trial geosynthetic reinforcement layouts. The most
economical reinforcement layout will be one which results in approximately

FHWA NHI-07-092
Geosynthetic Engineering

8 19

8 Reinforced Slopes
August 2008

equal, but greater than the minimum required, stability safety factor for internal,
external, and compound failure planes. A contour plot of lowest safety factor
values about the trial failure circle centroids is recommended to map and locate
the minimum safety factors for the three modes of failure.
External stability analysis in Step 8 will then include an evaluation of local
bearing capacity, foundation settlement, and dynamic stability.
STEP 7.

Select slope face treatment.

Slope facing requirements will depend on soil type, slope angle and the reinforcement
spacing as shown in Table 8-1.
If slope facing is required to prevent sloughing (i.e., slope angle is greater than Nsoil)
or erosion, several options are available. Sufficient reinforcement lengths could be
provided for wrapped faced structures. A face wrap may not be required for slopes
up to approximately 1H:1V. In this case, the reinforcement can be simply extended
to the face. For this option, a facing treatment as detailed in Section 8.5 Treatment of
Outward Face, should be applied at sufficient intervals during construction to prevent
face erosion. For wrapped or no wrap construction, the reinforcement should be
maintained at close spacing (i.e., every lift or every other lift but no greater than 16
in. (400 mm)). For armored, hard faced systems the maximum spacing should be no
greater than 32 in. (800 mm). A positive frictional or mechanical connection should
be provided between the reinforcement and armored type facing systems.
The following procedures are recommended for wrapping the face.
-

Turn up reinforcement at the face of the slope and return the reinforcement a
minimum of 3 ft (1 m) into the embankment below the next reinforcement
layer.

For steep slopes (> 1H:1V), temporary formwork (e.g., slip formed with
removable boards or left-in-place welded wire mesh) should be required to
support the face during construction.

If vertical reinforcement spacings of 18 in. (450 mm) or greater are used,


permanent welded wire mesh or gabion basket type facings should be
considered.

For geogrids, a fine mesh screen or geotextile may be required at the face to
retain backfill materials.

FHWA NHI-07-092
Geosynthetic Engineering

8 20

8 Reinforced Slopes
August 2008

Table 8-1. RSS Slope Facing Options (modified after Collin, 1996).
Slope Face Angle
and Soil Type

Type of Facing
When Geosynthetic is Not Wrapped at Face

Vegetated Face1
> 45o
(> ~1H:1V)
All Soil Types
35o to 45o
(~ 1.4H:1V to 1H:1V)
Clean Sands (SP)3
Rounded Gravel (GP)
o

Hard Facing2

Not Recommended

Gabions

Gabions
Not Recommended

Soil-Cement
Sod

35 to 45
(~ 1.4H:1V to 1H:1V)
Silts (ML)
Sandy Silts (ML)
Silty Sands (SM)

Bioreinforcement

Soil-Cement

Drainage
Composites4

Stone Veneer

Gabions

Sod

35o to 45o
(~ 1.4H:1V to 1H:1V)
Clayey Sands (SC)
Well graded sands &
gravels (SW & GW)

Temporary
Erosion Blanket
w/ Seed or Sod

25o to 35o
(~ 2H:1V to 1.4H:1V)
All Soil Types
Notes:
1.
2.
3.
4.

STEP 8.

Permanent
Erosion Mat
w/ Seed or Sod
Temporary
Erosion Blanket
w/ Seed or Sod
Permanent
Erosion Mat
w/ Seed or Sod

When Geosynthetic is Wrapped at Face

Vegetated Face1

Hard Facing2

Sod

Wire Baskets

Permanent Erosion
Blanket w/ seed

Permanent
Erosion Blanket
w/ seed

Permanent
Erosion Blanket
w/ seed

Stone
Shotcrete
Wire Baskets
Stone
Shotcrete
Wire Baskets
Stone
Shotcrete

Hard Facing
Not Needed

Geosynthetic
Wrap Generally
Not Needed

Geosynthetic
Wrap Not
Needed

Hard Facing
Not Needed

Geosynthetic
Wrap Not
Needed

Geosynthetic
Wrap Not
Needed

Vertical spacing of reinforcement (primary/secondary) shall be no greater than 16 in. (400 mm) with primary
reinforcements spaced no greater than 32 in. (800 mm) when secondary reinforcement is used.
Vertical spacing of primary reinforcement shall be no greater than 32 in. (800 mm).
Unified Soil Classification
Geosynthetic or natural horizontal drainage layers to intercept and drain the saturated soil at the face of the
slope.

Check external stability.

The external stability of a reinforced soil mass depends on the soil mass's ability to act as a
stable block and withstand all external loads without failure. Failure possibilities include
sliding, deep-seated overall instability, local bearing capacity failure at the toe (lateral
squeeze-type failure), as well as compound failures initiating internally and externally
through the short- and long-term conditions.
FHWA NHI-07-092
Geosynthetic Engineering

8 21

8 Reinforced Slopes
August 2008

a. Sliding resistance.
Evaluate the width of the reinforced soil mass at any level to resist sliding along
the reinforcement. A wedge type failure surface defined by the limits of the
reinforcement (the length of the reinforcement at the depth of evaluation defined
in step 5). The analysis can best be performed using a computerized method
which takes into account all soil strata and interface friction values.
A simple analysis using a sliding block method can be performed as a check. In
this method, an active wedge is assumed at the back of the reinforced soil mass
with the back of the wedge extending up at an angle of 45 + /2. Using this
assumption, the driving force is equal to the active earth pressure and the resisting
force is the frictional resistance provided by the weakest layer, either the
reinforced soil, the foundation soil or the soil-reinforcement interface. The
following relationships are then used:
Resisting Force = FS x Sliding Force
(W + Pa sin b) tan min = FS Pa cos b
with:
W =
W =
Pa =
where:
L
H
FS
PA
min

(r, (b
b

FHWA NHI-07-092
Geosynthetic Engineering

L2 (r (tan $r)
for L < H
2
[LH - H /(2tan$)] ((r)
for L > H
2
(b H Ka
= length of bottom reinforcing layer in each zone where there is a
reinforcement length change
= height of slope
= factor of safety for sliding (>1.5)
= active earth pressure
= minimum angle of shearing friction either between reinforced
soil and geosynthetic or the friction angle of the foundation
soil
= slope angle
= unit weight of the reinforced and retained backfill, respectively
= friction angle of retained fill (Note: If geotextile filter or
geocomposite drains are placed continuously on the
backslope, then Nb should be set equal to the interface
friction angle between the geosynthetic and the retained
fill).

8 22

8 Reinforced Slopes
August 2008

b. Deep-seated global stability.


As a check, potential deep-seated failure surfaces behind the reinforced soil mass
should be reevaluated. The analysis performed in Step 5 should provide this
information. However, as a check, classical rotational slope stability methods
such as simplified Bishop (1955), Morgenstern and Price (1965), Spencer (1981),
or others, may be used. Appropriate computer programs also may be used (see
FHWA Soils & Foundations Reference Manual {Samtani and Nowatzki, 2006}).
c. Local bearing failure at the toe (lateral squeeze).
Consideration must be given to the bearing capacity at the toe of the slope. High
lateral stresses in a confined soft stratum beneath the embankment could lead to a
lateral squeeze-type failure. This must be analyzed if the slope is on a soft not
firm foundation.
If a weak foundation soil layer exists beneath the reinforced slope to a limited
depth DS which is less than the width of the slope b', the factor of safety against
failure by squeezing may be calculated from (Silvestri, 1983):
FS squeezing =

2 cu
4.14 c u
+
1 .3
D s tan
H

where:
=
=
Ds =
H =
cu =

angle of slope
unit weight of soil in slope
depth of soft foundation soil beneath the base of the slope
height of slope
undrained shear strength of soft soil beneath slope

Caution is advised and rigorous analysis (e.g, numerical modeling) should be


performed when FS < 2. This approach is somewhat conservative as it does not
provide any influence from the reinforcement. When the depth of the soft layer,
DS, is greater than the base width of the slope, b', general slope stability will
govern the design.
d. Foundation settlement.
The magnitude of foundation settlement should be determined using ordinary
geotechnical engineering procedures (see Samtani and Nowatzki, 2006). If the
FHWA NHI-07-092
Geosynthetic Engineering

8 23

8 Reinforced Slopes
August 2008

calculated settlement exceeds project requirements, then foundation soils must be


improved.
STEP 9.

Check seismic stability.


Perform a pseudo-static type analysis using a seismic ground coefficient A,
obtained from local building code and a design seismic acceleration Am equal to
Am = A/2. Reinforced soil slopes are more flexible structures than rigid walls or
MSE walls. As such, Am can be taken as A/2 as allowed by AASHTO (2002)
Standard Specifications for Highway Bridges (Division 1A-Seismic Design, 6.4.3
Abutments.) (Elias et al., 2001)
F.S. dynamic 1.1
In the pseudo-static method, seismic stability is determined by adding a horizontal
and/or vertical force at the centroid of each slice to the moment equilibrium
equation. The additional force is equal to the seismic coefficient times the total
weight of the sliding mass. It is assumed that this force has no influence on the
normal force and resisting moment, so that only the driving moment is affected.
The liquefaction potential of the foundation soil should also be evaluated.
For critical projects in areas of potentially high seismic risk, a complete dynamic
analysis (i.e., stability and deformation analyses with reduced soil strengths and
site specific acceleration) should be performed. A psuedo-static analysis is
generally considered not suitable if the strength of the soil is reduced by more
than 15% by the cyclic loading (Duncan and Wright, 2005; Kramer, 1996). The
Unified Methodology for Seismic Stability and Deformation Analysis of slopes
presented in FHWA GEC No. 3 (Kavazanjian et al., 1997) recommends a
Newmark deformation analysis is the computed factor of safety (with this
particular methodology) is less than 1.0. Duncan and Wright (2005) provide a
summary of suggested methods for performing psuedo-static screening analyses.

STEP 10. Evaluate requirements for subsurface and surface water control.
a. Subsurface water control.
Uncontrolled subsurface water seepage can decrease slope stability and ultimately
result in slope failure. Hydrostatic forces on the rear of the reinforced mass and
uncontrolled seepage into the reinforced mass will decrease stability. Seepage
FHWA NHI-07-092
Geosynthetic Engineering

8 24

8 Reinforced Slopes
August 2008

through the mass can reduce pullout capacity of the geosynthetic and create
erosion at the face. Consider the water source and the permeability of the natural
and fill soils through which water must flow when designing subsurface water
drainage features. Subsurface water can be controlled by using free draining
reinforced fill (i.e., fill with only a few percent fines and seepage analysis to
confirm adequate drainage), a drainage system at the rear of and possible beneath
the reinforced mass, and/or use of composite geosynthetic reinforcements with inplane drainage capacity (Soong and Koerner, 1999). Design consideration for
drainage systems to control subsurface water include:

Design of subsurface water drainage features should address flow rate, filtration,
placement, and outlet details.

Drains are typically placed at the rear of and/or beneath the reinforced mass.
Geocomposite drainage systems or conventional granular blanket and trench drains
could be used.

Lateral spacing of outlet is dictated by site geometry, estimated flow, and existing
agency standards. Outlet design should address long-term performance and
maintenance requirements.

The design of geocomposite drainage materials is addressed in Chapter 2.

Slope stability analyses should account for interface shear strength along a
geocomposite drain. The geocomposite/soil interface will most likely have a friction
value that is lower than that of the soil. Thus, a potential failure surface may be
induced along the interface.

Geotextiles reinforcements (primary and intermediate layers) must be more


permeable than the reinforced fill material to prevent a hydraulic build-up above the
geotextile layers during precipitation.
Special emphasis on the design and construction of subsurface drainage features
is recommended for structures where drainage is critical for maintaining slope
stability. Redundancy in the drainage system is also recommended in these
cases.

FHWA NHI-07-092
Geosynthetic Engineering

8 25

8 Reinforced Slopes
August 2008

b. Surface water runoff.


Slope stability can be threatened by erosion due to surface water runoff. Erosion rills
and gullies can lead to surface sloughing and possibly deep-seated failure surfaces.
Erosion control and revegetation measures must, therefore, be an integral part of all
reinforced slope system designs and specifications.
Surface water runoff should be collected above the reinforced slope and channeled or
piped below the base of the slope. Standard agency drainage details should be
utilized.
If not otherwise protected, reinforced slopes should be vegetated immediately (or
soon) after construction to prevent or minimize erosion due to rainfall and runoff on
the face. Vegetation requirements will vary by geographic and climatic conditions
and are therefore project-specific. Geosynthetic reinforced slopes are inherently
difficult sites to establish and maintain vegetative cover due to these steep slopes.
The steepness of the slope limits the amount of water absorbed by the soil before
runoff occurs. Once vegetation is established on the face, it must be maintained to
ensure long-term survival.
A synthetic (permanent) erosion control mat that is stabilized against ultraviolet light
and is inert to naturally occurring soil-born chemicals and bacteria may be required
with seeding. The erosion control mat serves three functions: 1) to protect the bare
soil face against erosion until vegetation is established, 2) to reduce runoff velocity
for increased water absorption by the soil, thereby promoting long-term survival of
the vegetative cover, and 3) to reinforce the root system of the vegetative cover.
Maintenance of vegetation will still be required.
A permanent synthetic mat may not be required in applications characterized by
flatter slopes (less than 1:1), low height slopes, and/or moderate runoff. In these
cases, a temporary (degradable) erosion blanket may be specified to protect the slope
face and promote growth until vegetative cover is firmly established. Refer to
Chapter 4 for design of erosion mats and blankets.
Erosion control mats and blankets vary widely in type, cost, and more importantly
applicability to project conditions. Slope protection should not be left to the
Contractor's or vendors discretion. See Section 8.5 for additional guidance on
slope face treatment.

FHWA NHI-07-092
Geosynthetic Engineering

8 26

8 Reinforced Slopes
August 2008

STEP 11. Develop specifications and contract documents.


Specifications are discussed in Section 8.9.
8.3-4 Computer Assisted Design
The ideal method for reinforced slope design is to use a conventional slope stability computer
program that has been modified to account for the stabilizing effect of reinforcement. Such
programs should account for reinforcement strength and pullout capacity, compute reinforced
and unreinforced safety factors automatically, and have some searching routine to help locate
critical surfaces. The method may also include the confinement effects of the reinforcement
on the shear strength of the soil in the vicinity of the reinforcement.
A generic program ReSSA developed by FHWA performs both reinforcement design of
simple structures and evaluation of simple and complex geometries and soil stratigraphy.
The ReSSA Version 1.0 was distributed to state DOTs. Newer version (3.0 as of 2007) of
ReSSA with enhanced features is available, for purchase, at www.geogprograms.com.
Several other reinforced slope programs are commercially available. These programs
generally do not design the reinforcement but allow for an evaluation of a given
reinforcement layout.
An iterative approach then follows to optimize either the
reinforcement strength or layout. Some of the programs are limited to simple soil profiles
and simple reinforcement layouts. Additionally, external stability evaluation may be limited
to specific soil and reinforcement conditions and a single mode of failure. In some cases, the
programs are reinforcement-specific.
With computerized analyses, the actual factor of safety value FS is dependent upon how the
specific program accounts for the reinforcement tension in the moment equilibrium equation.
The method of analysis in this manual and in FHWAs ReSSA program, as well as many
others, assume the reinforcement force as contributing to the resisting moment, i.e.:
FS R =

where,

FHWA NHI-07-092
Geosynthetic Engineering

FSR =
MR =
MD =

M R + TS R
MD

the stability factor of safety required


resisting moment provided by the strength of the soil
driving moment about the center of the failure circle

8 27

8 Reinforced Slopes
August 2008

TS =

sum of required tensile force per unit width of


reinforcement (considering rupture and pullout) in all
reinforcement layers intersecting the failure surface
the moment arm of TS about the center of failure circle

With this assumption, the FSR is applied to both the soil and the reinforcement as part
of the analysis. The sum of required tensile force, Ts, is calculated with long-term
reinforcement strengths, Tal (see Section 8.4-3). The long-term reinforcement
strength is equal to the ultimate tensile divided by reduction factors to account for
creep, installation damage, and durability, i.e., TULT / (RFCR x RFID x RFD).
Some computer programs use an assumption that the reinforcement force is a
negative driving component, thus the factor of safety value is computed as:
FS R =

MR
M D TS R

With this assumption, the stability factor of safety, FSR, is not applied to TS.
Therefore, the sum of the required tensile force, Ts, should be computed with the
long tensile strength Tal divided by the required safety factor (i.e., target stability
factor of safety, FSR). This provides an appropriate factor of safety for uncertainty
in the reinforcement material strength. Likewise, the method used to develop design
charts should be carefully evaluated to determine FS used to obtain the sum of the
required tensile force.

8.4

MATERIAL PROPERTIES

8.4-1 Reinforced Slope Systems


Reinforced soil systems consist of planar reinforcements arranged in horizontal planes in the
fill soil to resist outward movement of the reinforced soil mass. Facing treatments ranging
from vegetation to flexible armor systems are applied to prevent raveling and sloughing of
the face. These systems are generic in nature and can incorporate any of a variety of
reinforcements and facing systems. This section provides the material properties required for
design.
8.4-2 Soils
Any soil meeting the requirements for embankment construction can be used in a reinforced
slope system. From a reinforcement point of view alone, even lower-quality soil than that
FHWA NHI-07-092
Geosynthetic Engineering

8 28

8 Reinforced Slopes
August 2008

conventionally used in unreinforced slope construction could be used; however, a higherquality material offers less durability concerns and is easier to handle, place, and compact,
which tends to speed up construction. Therefore, the following guidelines are provided as
recommended backfill requirements for reinforced engineered slopes.
Gradation (Elias et al., 2001):
engineered slopes are:

Recommended backfill requirements for reinforced

Sieve Size
-in. (20 mm)*
No. 4 (4.75 mm)
No. 40 (0.425 mm)
No. 200 (0.075 mm)

Percent Passing
100 - 75
100 - 20
0 - 60
0 - 50

* The maximum size can be increased up to 4-in. (100 mm)


provided field tests have been, or will be, performed to evaluate
potential strength reduction due to installation damage (see
Appendix H). In any case, geosynthetic strength reduction factors
for site damage should be checked in relation to particle size and
angularity of the larger particles.

Plasticity Index (PI) < 20 (AASHTO T-90)


Soundness: Magnesium sulfate soundness loss less than 30% after 4 cycles,
based on AASHTO T 104 or equivalent sodium sulfate soundness of less than
15 percent after 5 cycles.
Definition of total and effective stress shear strength properties becomes more important as
the percentage passing the No.200 (0.075 mm) sieve increases. Likewise, drainage and
filtration design are more critical. Fill materials outside of these gradation and plasticity
index requirements have been used successfully (Christopher et al., 1990; Hayden et al.,
1991). However, soils outside of the recommended gradation range should be carefully
evaluated and monitored.
Chemical Composition (Elias et al., 2001): The chemical composition of the reinforced fill
and retained soils should be assessed for effect on reinforcement durability (primarily pH and
oxidation agents). Some of the soil environments posing potential concern when using
geosynthetics are listed in Appendix H. Use of polyester geosynthetics should be limited to
soils with 3 < pH < 9; and use of polyolefins (polypropylene and polyethylene) should be
limited to soils with pH > 3. Soil pH should be determined in accordance with AASHTO T289.
FHWA NHI-07-092
Geosynthetic Engineering

8 29

8 Reinforced Slopes
August 2008

Compaction (Christopher et al., 1989): Soil fill shall be compacted to 95% of optimum dry
density ((d) and + or - 2 % of the optimum moisture content, wopt, according to AASHTO T99. Cohesive soils should be compacted in 6 to 8 in. (150 to 200 mm) compacted lifts, and
granular soils in 8 to 12 in. (200 to 300 mm) compacted lifts.
Shear Strength: Peak shear strength parameters determined using direct shear or
consolidated-drained (CD) triaxial tests should be used in the analysis (Christopher et al.,
1989). Effective stress strength parameters should be used for granular soils with less than
15% passing the No.200 (0.075 mm) sieve.
For all other soils, peak effective stress and total stress strength parameters should be
determined. These parameters should be used in the analyses to check stability for the
immediately-after-construction and long-term cases. Use CD direct shear tests (sheared
slowly enough for adequate sample drainage), or consolidated-undrained (CU) triaxial tests
with pore water pressures measured for determination of effective stress parameters. Use CU
direct shear or triaxial tests for determination of total stress parameters.
Shear strength testing is recommended. However, use of assumed shear values based on
Agency guidelines and experience may be acceptable for some projects. Verification of site
soil type(s) should be completed following excavation or identification of borrow pit, as
applicable.
Unit Weights: Dry unit weight for compaction control, moist unit weight for analyses, and
saturated unit weight for analyses (where applicable) should be determined for the fill soil.
8.4-3 Geosynthetic Reinforcement
Geosynthetic design strength must be determined by testing and analysis methods that
account for long-term interaction (e.g., grid/soil stress transfer) and durability of the all
geosynthetic components. Geogrids transfer stress to the soil through passive soil resistance
on the grid's transverse members and through friction between the soil and the geogrid's
horizontal surfaces (Mitchell and Villet, 1987). Geotextiles transfer stress to the soil through
friction.
An inherent advantage of geosynthetics is their longevity in fairly aggressive soil conditions.
The anticipated half-life of some geosynthetics in normal soil environments is in excess of
1,000 years. However, as with steel reinforcements, strength characteristics must be adjusted
to account for potential degradation in the specific environmental conditions, even in
relatively neutral soils. Questionable soil environments are listed in Appendix H.

FHWA NHI-07-092
Geosynthetic Engineering

8 30

8 Reinforced Slopes
August 2008

Tensile Strengths: Long-term tensile strength (Tal) of the geosynthetic shall be determined
using a partial factor of safety approach (Bonaparte and Berg, 1987). Reduction factors are
used to account for installation damage, chemical and biological conditions and to control
potential creep deformation of the polymer. Where applicable, a reduction is also applied for
seams and connections. The total reduction factor is based upon the mathematical product of
these factors. The long-term tensile strength, Tal, thus can be obtained from:
Tal =

Tult
RF

with RF equal to the product of all applicable reduction factors:


RF = RFCR xRFID xRFD
where:
= long-term geosynthetic tensile strength,(lb/ft {kN/m});
Tal
Tult
= ultimate geosynthetic tensile strength, based upon MARV, (lb/ft {kN/m});
RFCR = creep reduction factor, ratio of Tultlot (ultimate strength of roll used for
creep testing specimens) to creep-limiting strength, (dimensionless);
RFID = installation damage reduction factor, (dimensionless); and
RFD = durability reduction factor for chemical and biological degradation,
(dimensionless).
RF values for durable geosynthetics in non-aggressive, granular soil environments range
from about 2.5 to 7. Appendix H suggests that a default value RF = 7 may be used for
routine, non-critical structures which meet the soil, geosynthetic and structural limitations
listed in the appendix. However, as indicated by the range of RF values, there is a potential
to significantly reduce the reinforcing requirements and the corresponding cost of the
structure by obtaining a reduced RF from test data.
The procedure presented above and detailed in Appendix H is derived from FHWA SA-96071 (Elias and Christopher, 1997), FHWA SA-93-025 (Berg, 1993), and the AASHTO Task
Force 27 (1990) guidelines for geosynthetic reinforced soil retaining walls.
For RSS structures, the FS value will be dependent upon the analysis tools utilized by the
designer. With computerized analyses, the FS value is dependent upon how the specific
program accounts for the reinforcement tension is computing a stability factor of safety, as
discussed in Section 8.3-4.
FHWA NHI-07-092
Geosynthetic Engineering

8 31

8 Reinforced Slopes
August 2008

Soil-Reinforcement Interaction: Two types of soil-reinforcement interaction coefficients


or interface shear strengths must be determined for design: pullout coefficient, and interface
friction coefficient (Task Force 27 Report, 1990). Pullout coefficients are used in stability
analyses to compute mobilized tensile force at the front and tail of each reinforcement layer.
Interface friction coefficients are used to check factors of safety against outward sliding of
the entire reinforced mass.
Detailed procedures for quantifying interface friction and pullout interaction properties are
presented in Appendix H. The ultimate pullout resistance, Pr, of the reinforcement per unit
width of reinforcement is given by:

Pr = 2 C F C " C Fv C Le

where:
= pullout resistance per unit width of reinforcement
Pr
Le C 2 = the total surface area per unit width of the reinforcement in the
resistance zone behind the failure surface
= the embedment or adherence length in the resisting zone behind the
Le
failure surface

F
= the pullout resistance (or friction-bearing-interaction) factor
"
= a scale effect correction factor
Fv
= the effective vertical stress at the soil-reinforcement interfaces
For preliminary design in the absence of specific geosynthetic test data, and for standard
backfill materials with the exception of uniform sands (i.e., coefficient of uniformity, Cu < 4),
it is acceptable to use conservative default values for F* and as shown in Table 8-2. The soil
friction angle is normally established by testing, though a lower bound value of 28 degrees is
often used.
Table 8-2 Default Values For F* and " Pullout Factors
Reinforcement Type

Default F*

Default "

Geogrid

0.67 tan N

0.8

Geotextile

0.67 tan N

0.6

FHWA NHI-07-092
Geosynthetic Engineering

8 32

8 Reinforced Slopes
August 2008

8.5

TREATMENT OF OUTER FACE (Elias et al., 2001)

a.

Grass Type Vegetation


Stability of a slope can be threatened by erosion due to surface water runoff. Erosion
control and revegetation measures must, therefore, be an integral part of all reinforced
slope system designs and specifications. If not otherwise protected, reinforced slopes
should be vegetated after construction to prevent or minimize erosion due to rainfall
and runoff on the face. Vegetation requirements will vary by geographic and climatic
conditions and are, therefore, project specific.
For the unwrapped face (the soil surface exposed), erosion control measures are
necessary to prevent unraveling and sloughing of the face. A wrapped face helps
reduce erosion problems; however, treatments are still required on the face to shade
the geosynthetic and prevent ultraviolet light exposure that will degrade the
geosynthetic over time. In either case, conventional vegetated facing treatments
generally rely on low growth, grass type vegetation with more costly flexible armor
occasionally used where vegetation cannot be established. Geosynthetic reinforced
slopes can be difficult sites to establish and maintain grass type vegetative cover due
to the steep grades that can be achieved. The steepness of the grade limits the amount
of water absorbed by the soil before runoff occurs. Although root penetration should
not affect the reinforcement, the reinforcement will most likely restrict root growth.
This can have an adverse influence on the growth of some plants. Grass is also
frequently ineffective where slopes are impacted by waterways.
A synthetic (permanent) erosion control mat is normally used to improve the
performance of grass cover. This mat must also be stabilized against ultra-violet light
and should be inert to naturally occurring soil-born chemicals and bacteria. The
erosion control mat serves to: 1) protect the bare soil face against erosion until the
vegetation is established; 2) assist in reducing runoff velocity for increased water
absorption by the soil, thus promoting long-term survival of the vegetative cover; and
3) reinforce the surficial root system of the vegetative cover.
Once vegetation is established on the face, it must be protected to ensure long-term
survival. Maintenance issues, such as mowing, must also be carefully considered.
The shorter, weaker root structure of most grasses may not provide adequate
reinforcement and erosion protection. Grass is highly susceptible to fire, which can
also destroy the synthetic erosion control mat. Downdrag from snow loads or upland
slides may also strip matting and vegetation off the slope face. The low erosion

FHWA NHI-07-092
Geosynthetic Engineering

8 33

8 Reinforced Slopes
August 2008

tolerance combined with other factors previously mentioned creates a need to


evaluate revegetation measures as an integral part of the design. Slope face
protection should not be left to the construction contractor or vendor's discretion.
Guidance should be obtained from maintenance and regional landscaping groups in
the selection of the most appropriate low maintenance vegetation.
b.

Soil Bioengineering (Woody Vegetation)


An alternative to low growth, grass type vegetation is the use of soil bioengineering
methods to establish hardier, woody type vegetation in the face of the slope (Sotir and
Christopher, 2000). Soil bioengineering uses living vegetation purposely arranged
and imbedded in the ground to prevent shallow mass movement and surficial erosion.
However, the use of soil bioengineering in itself is limited to stable slope masses.
Combining this highly erosive-resistant system with geosynthetic reinforcement
produces a very durable, low maintenance structure with exceptional aesthetic and
environmental qualities.
Appropriately applied, soil bioengineering offers a cost-effective and attractive
approach for stabilizing slopes against erosion and shallow mass movement,
capitalizing on the benefits and advantages that vegetation offers. The value of
vegetation in civil engineering and the role woody vegetation plays in the
stabilization of slopes has gained considerable recognition in recent years (Gray and
Sotir, 1995). Woody vegetation improves the hydrology and mechanical stability of
slopes through root reinforcement and surface protection. The use of deeply-installed
and rooted woody plant materials, purposely arranged and imbedded during slope
construction offers:

Immediate erosion control for slopes; stream, and shoreline;

Improved face stability through mechanical reinforcement by roots;

Reduced maintenance costs, with less need to return to revegetate or cut grass;

Modification of soil moisture regimes through improved drainage and


depletion of soil moisture and increase of soil suction by root uptake and
transpiration;

Enhanced wildlife habitat and ecological diversity; and

Improved aesthetic quality and naturalization.


The biological and mechanical elements must be analyzed and designed to work
together in an integrated and complementary manner to achieve the required project
goals. In addition to using engineering principles to analyze and design the slope
stabilization systems, plant science and horticulture are needed to select and establish

FHWA NHI-07-092
Geosynthetic Engineering

8 34

8 Reinforced Slopes
August 2008

the appropriate vegetation for root reinforcement, erosion control, aesthetics and the
environment. Numerous areas of expertise must integrate to provide the knowledge
and awareness required for success. RSS systems require knowledge of the
mechanisms involving mass and surficial stability of slopes. Likewise when the
vegetative aspects are appropriate to serve as reinforcements and drains, an
understanding of the hydraulic and mechanical effects of slope vegetation is
necessary.
Vegetation Selection: The vegetation used in the vegetated reinforced soil slope
(VRSS) system is typically in the form of live woody branch cuttings from species
that root adventitiously or from, bare root and/or container plants. Plant materials
may be selected for a variety of tolerances including: drought, salt, flooding, fire,
deposition, and shade. They may be chosen for their environmental wildlife value,
water cleansing capabilities, flower, branch and leaf color or fruits. Other interests
for selection may include size, form, rate of growth rooting characteristics and ease of
propagation. Time of year for construction of a VRSS system also plays a critical roll
in plant selection.
Vegetation Placement: The decision to use natives, naturalized or ornamental species
is also an important consideration. The plant materials are placed on the frontal
section of the formed terraces. Typically 6 to 12 in. (150 to 300 mm), protrudes
beyond the constructed terrace edge or finished face, and 1.5 to 10 ft (0.5 to 3 m) of
the live branch cuttings when used are embedded in the reinforced backfill behind, or
as in the case of rooted plants, are placed 1 to 3 ft (0.3 to 1 m) in the backfill. The
process of plant installation is best and least expensive when it occurs simultaneously
with the conventional construction activities, but may be incorporated later if
necessary.
Vegetation Development: Typically soil bioengineering VRSS systems offer
immediate results from the surface erosion control structural/mechanical and
hydraulic perspectives. Over time, (generally within the first year) they develop
substantial top and root growth further enhancing those benefits, as well as providing
aesthetic and environmental values.
Design Issues: There are several agronomic and geotechnical design issues that must
be considered, especially in relation to selection of geosynthetic reinforcement and
type of vegetation. Considerations include root and top growth potential. The root
growth potential consideration is important when face reinforcement enhancement is
required. This will require a review of the vertical spacing based on the anticipated
FHWA NHI-07-092
Geosynthetic Engineering

8 35

8 Reinforced Slopes
August 2008

root growth for the specific type of plant. In addition to spacing, the selected type of
reinforcements is also important. Open-mesh geogrid-type reinforcements, for
example, are excellent as the roots will grow through the grid and further "knit" the
system together. On the other hand, geocomposites, providing both reinforcement and
lateral drainage, offer enhanced water and oxygen opportunities for the healthy
development of the woody vegetation. Dependent upon the species selected, aspect,
climatic conditions, soils etc., dense woody vegetation can provide ultraviolet light
protection within the first growing season and maintain the cover thereafter.
In arid regions, geosynthetics that will promote moisture movement into the slope
such as non-woven geotextiles or geocomposites may be preferred. Likewise, the
need for water and nutrients in the slope to sustain and promote vegetative growth
must be balanced against the desire to remove water so as to reduce hydrostatic
pressures. Plants can be installed to promote drainage toward geosynthetic drainage
net composites placed at the back of the reinforced soil section.
Organic matter is not required; however, a medium that provides nourishment for
plant growth and development is necessary. As mentioned earlier, the agronomic
needs must be balanced with the geotechnical requirements, but these are typically
compatible. For both, a well-drained backfill is needed. The plants also require
sufficient fines to provide moisture and nutrients while this may be a limitation, under
most circumstances, some slight modifications in the specifications to allow for some
non-plastic fines in the backfill in the selected frontal zone offers a simple solution to
this problem.
While many plants can be installed throughout the year, the most cost effective,
highest rate of survival and best overall performance and function occurs when
construction is planned around the dormant season for the plants, or just prior to the
rainy season. This may require some specific construction coordination in relation to
the placement of fill, and in some cases may preclude the use of a VRSS structure.
c.

Armored
Where vegetation cannot be established or for aesthetic reasons, a permanent facing
may also be used. Permanent facing such as gunite or emulsified asphalt may also be
applied to a RSS slope face to provide long-term ultra-violet protection, if the
geosynthetic UV resistance is not adequate for the life of the structure.

FHWA NHI-07-092
Geosynthetic Engineering

8 36

8 Reinforced Slopes
August 2008

8.6

Welded wire mesh, gabions, or geocells may be used to facilitate face


construction and provide permanent facing systems.
Other armored facing elements may include riprap, stone veneer, articulating
modular units, or fabric-formed concrete.
Structural elements facing elements (see MSE walls) may also be used, especially
if discrete reinforcing elements such as metallic strips are used. These facing
elements may include prefabricated concrete slabs, modular precast blocks, or
precast slabs.

PRELIMINARY DESIGN AND COST EXAMPLE

EVALUATION AND COST ESTIMATE EXAMPLE


A 0.6 mile long, 16 feet high, 2.5H:1V side slope road embankment in a suburban area is to be
widened by one lane. At least a 20-foot width extension is required to allow for the additional lane
plus shoulder improvements. Several options are being considered.
1.

Simply extend the slope of the embankment.

2.

Construct a 8-foot high concrete, cantilever retaining wall at the toe of the slope, extend the
alignment, and slope down at 2.5H:1V to the wall. (Of course a geosynthetically reinforced
retaining wall should also be considered, but that's covered in the next chapter.)

3.

Construct a 1H:1V reinforced soil slope up from the toe of the existing slope, which will add
25 feet to the alignment, enough for future widening, if required.

A guardrail is required for all options and is not included in the cost comparisons.
Option 1
The first alternative will require 12 yd3 fill per lineal foot of embankment length. The fill is locally
available with some hauling required and has an estimated in-place cost of $6/yd3 (about $3.65 per
ton). The cost of the 20-foot right of way is $1.40/ft2, for a cost of $28 per foot of embankment
length. Finally, hydroseeding and mulching will cost approximately $0.07 per square foot of face, or
approximately $3 lineal foot of embankment. Thus the total cost of embankment will be $103 for the
full height per foot length of embankment or $6.50/ft2 of vertical face. There will also be a project
delay while the additional right of way is obtained.
Option 2
Based on previous projects in this area, the concrete retaining wall is estimated to cost $37/ft2 of
vertical wall face including backfill. Thus, the 8 ft high wall will cost $300 per foot length of
FHWA NHI-07-092
Geosynthetic Engineering

8 37

8 Reinforced Slopes
August 2008

embankment. This leads to a cost of $18.50/ft2 of vertical embankment per foot length of structure.
In addition 7 yd3 of fill will be required to construct the sloped portion, adding $43 per foot of
embankment, or $2.70/ft2 of vertical face, to the cost. Hydroseeding and mulching of the slope will
add about a $0.10/ft2 of vertical face to the cost. Thus, the total cost of this option is estimated at
$21.50/ft2 of vertical embankment face.
This option will require an additional 2 weeks of
construction time to allow the concrete to cure. On some projects, additional costs can be incurred
due to the delay plus additional traffic control and highway personnel required for inspection during
removal of the forms. Since this project was part of a larger project, such delays were not considered.
Option 3
This option will require a preliminary design to determine the quantity of reinforcement.
STEP 1.
a.
b.
c.

Slope description
H = 16 ft
= 45o
q = 250 psf (for pavement section) + 2% road grade
Performance requirements

a. External Stability:
Sliding Stability: FSmin = 1.3
Overall slope stability and deep seated: FSmin = 1.3
Dynamic loading: no requirement
Settlement: analysis required
b. Compound Failure: FSmin = 1.3
c. Internal Stability: FSmin = 1.3

STEP 2. Engineering properties of foundation soils.


a. Review of soil borings from the original embankment construction indicates foundation
soils consisting of stiff to very stiff, low-plasticity silty clay with interbedded seams
of sand and gravel. The soils tend to increase in density and strength with depth.
b. (= 120 lb/ft3, Topt = 15% UU = 2,000 psf, N'= 28o, c'= 0
c. At the time of the borings, dw = 6.5 ft below the original ground surface.
d. Not applicable
STEP 3.

Properties of reinforced and embankment fill


(The existing embankment fill is a clayey sand and gravel). For preliminary evaluation,
the properties of the embankment fill are assumed for the reinforced section as follows:

FHWA NHI-07-092
Geosynthetic Engineering

8 38

8 Reinforced Slopes
August 2008

a. Sieve Size
Percent Passing
1-in. (100 mm)
100%
3/4-in. (20 mm)
99%
No.4 (4.75 mm)
63%
No.40 (0.425 mm)
45%
No.200 (0.075 mm)
25%
PI (of fines) = 10
Gravel is competent
pH = 7.5
b. (r = 135 lb/ft3, Topt = 15%
c. N' = 33o, c' = 0
d. Soil is relatively inert

STEP 4.

Design parameters for reinforcement


For preliminary analysis use default values.
a. Tal = Tult/Rf
b. FSpo = 1.5

STEP 5. Check unreinforced stability


Using STABL5M, the minimum unreinforced factor of safety was 0.68 with the critical zone
defined by the target factor of safety FSR as shown in Design Figure A.

STEP 6.

Calculate TS for the FSR

Option A. From the computer runs, obtain FSU, MD and R for each failure surface within the
critical zone, and calculate TS as follows. (NOTE: With minor code modification, this could
easily be done as part of the computer analysis, as is done in the FHWA program RSS.)

TS = (1.3 F .SU )

a.

MD
R

Evaluating all of the surfaces in the critical zone indicates maximum TS-MAX = 3,400 lb/ft
for FSU = 0.89 as shown in Design Figure B.
b. Checking TS MAX by using Figure 8-6:

FHWA NHI-07-092
Geosynthetic Engineering

8 39

8 Reinforced Slopes
August 2008

tan r
FS R

f = tan 1

tan 33
= tan 1
= 26.5
1 .3

From Figure 8-6, K . 0.14


and,
H'

= H + q/(r + 0.1 m (for 2% road grade)


= 16 ft + (250 lb/ft2 ) 135 lb/ft3) + 0.3 ft = 18.2 ft

then,
TS-MAX = 0.5KrHl
= 0.5(0.14)(135 lb/ft3)(18.2 ft)2
= 3,130 lb/ft
The evaluation using Figure 8-6 appears to be in good agreement with the computer
analysis.
c. Determine the distribution of reinforcement.
Since H < 20 ft, use a uniform spacing. Due to the cohesive nature of the backfill,
maximum compaction lifts of 8 in. are recommended.
d. A reinforcement spacing, Sv, equal to 16 in. will be used. Therefore, N = 16 ft/1.25 ft =
12.8. Use 12 layers with the bottom layer placed after the first lift of embankment fill.

Td

Tmax 3,400 lb / ft
=
= 283 lb ft
N
12

(NOTE: Other reinforcement options such as using short secondary reinforcements at


every lift with spacing and strength increased for primary reinforcements could be
considered, and should be evaluated, for selecting the most cost-effective option for final
design.)
e. Since this is a simple structure, rechecking Ts above each layer of reinforcement is not
performed.
f. For preliminary analysis, the critical zone found in the computer analysis (Figure A) can
be used to define the reinforcement limits. This is especially true for this problem, since
the factor of safety for sliding (FSsliding > 1.3) is greater than the internal stability
requirement (FSinternal > 1.3); thus, the sliding failure surface well encompasses the most
critical reinforcement surface.
FHWA NHI-07-092
Geosynthetic Engineering

8 40

8 Reinforced Slopes
August 2008

As measured at the bottom and top of the sliding surface in Figure A, the required lengths
of reinforcement are:
Lbottom = 17.5 ft
Ltop
= 9.5 ft

Check length of embedment beyond the critical surface Le and factor of safety against
pullout.
Since the most critical location for pullout is the reinforcement near the top of the slope
(depth Z = 8 in.), subtract the distance from the critical surface to the face of the slope in
Figure B from Ltop. This gives Le at top = 4.3 ft.
Assuming the most conservative assumption for pullout factors F* and " from Section 8.4
and Appendix H gives F* = 0.67 tanN and " = 0.6. Therefore,

FS PO =

Le F v C 4.3 (0.67 tan 33 )(0.6 ) 0.67 ft x 135 lb / ft 3 + 250 lb / ft 2 (2 )


=
Tmax
283 lb / ft

FSPO = 2.7 > 1.5 required

Check the length requirement using Figure 8-6.


For LB

tan 28
= 22.2
1.3

f = tan 1
From Figure 8-6:
thus,
LB

LB/H' = 0.96
= 18.2 ft (0.96) = 17.5 ft

For LT

tan 33
f = tan 1
= 26.5
1.3
From Figure 8-6:
thus,

LT

LT/H = 0.52
= 18.2 ft (0.52) = 9.5 ft

Using Figure 8-6, the evaluation again appears to be in good agreement with the computer
analysis.
FHWA NHI-07-092
Geosynthetic Engineering

8 41

8 Reinforced Slopes
August 2008

g. This is a simple structure and additional evaluation of design lengths is not required.
Option B.

Since this is a preliminary analysis and a fairly simple problem, Figure 8-6 or
any number of proprietary computer programs, can be used to rapidly
evaluate Ts and Td.

In summary, 12 layers of reinforcement are required with a design strength, Td, of 283 lb/ft
and an average length of 13.5 ft over the full height of embankment. This would result in a
total of 18 yd2 reinforcement per foot length of embankment or 1.1 yd2 per vertical foot of
height. Adding 10% to 15% for overlaps and overages results in an anticipated reinforcement
volume of 1.3 yd2 per vertical embankment face. Based on the cost information in Appendix
G, reinforcement with an allowable strength Ta $ 283 lb/ft would cost approximately $1.00 to
$1.50/yd2. Assuming $0.50 yd2 for handling and placement, the in-place cost of
reinforcement would be approximately $2.50/ft2 of vertical embankment face.
Approximately 7.5 yd3 of additional backfill would be required for this option, adding
$2.80/ft2 to the cost of this option. In addition, overexcavation and backfill of existing
embankment material will be required to allow for reinforcement placement. Assuming
$1.50/yd3 for overexcavation and replacement will add approximately $0.40/ft2 of vertical
face. Erosion protection for the face will also add a cost of $0.45/ft2 of vertical face. Thus,
the total estimated cost for this option totals approximately $6.15/ft2 of vertical embankment
face.
Option 3 provides a slightly lower cost than Option 1 plus it does not require additional right-of-way.

Figure A

Unreinforced stability analysis.

FHWA NHI-07-092
Geosynthetic Engineering

8 42

8 Reinforced Slopes
August 2008

Figure B

8.7

Surface requiring maximum reinforcement (i.e., most critical reinforced surface).

COST CONSIDERATIONS

As with any other reinforcement application, an appropriate benefit-to-cost ratio analysis


should be completed to determine if the steeper slope with reinforcement is justified
economically over the flatter slope with increased right-of-way and materials costs, etc. In
some cases, however, the height of the embankment will be controlled by grade
requirements, and the slope might as well be as steep as possible. With respect to economy,
the factors to consider are:
cut or fill earthwork quantities;
size of slope area;
average height of slope area;
angle of slope;
cost of nonselect versus select backfills;
erosion protection requirements;
cost and availability of right-of-way needed;
complicated horizontal and vertical alignment changes;
safety equipment (guardrails, fences, etc.)
need for temporary excavation support systems;
cost of structures running through the embankment (e.g., box culvert);
maintenance of traffic during construction; and
aesthetics.
Figure 8-9 provides a rapid first order assessment of cost for comparing a flatter unreinforced
slope with a steeper reinforced slope.
FHWA NHI-07-092
Geosynthetic Engineering

8 43

8 Reinforced Slopes
August 2008

Figure 8-9.

8.8

Cost evaluation of reinforced soil slopes.

IMPLEMENTATION

The recent availability of many new geosynthetic reinforcement materials as well as


drainage and erosion control products requires Engineers to consider many alternatives
before preparing contract bid documents so that proven, cost-effective materials can be
chosen. Reinforced soil slopes may be contracted using two different approaches. Slope
structures can be contracted on the basis of (Berg and DiMaggio, 1994):
In-house (Agency) design with geosynthetic reinforcement, drainage details, erosion
measures, and construction execution specified.
System or end-result approach using approved systems, similar to mechanically
stabilized earth (MSE) walls, with lines and grades noted on the drawings.
For either approach, the following assumptions should be considered:
Geosynthetic reinforced slope systems can successfully compete with select
embankment fill requirements, other landslide stabilization techniques, and
unreinforced embankment slopes in urban areas.
Value engineering proposals are allowed, based on Agency standard procedures.
Geosynthetic reinforced slope systems submitted for use in a value engineering
proposal should have previous approval.
Though they may incorporate proprietary materials, reinforced slope systems are nonproprietary and may be bid competitively with geosynthetic reinforcement material
FHWA NHI-07-092
Geosynthetic Engineering

8 44

8 Reinforced Slopes
August 2008

alternatives. Geosynthetic reinforcement design parameters must be based upon


documentation that is provided by the manufacturer, submitted and approved by the
Agency, or based upon default partial safety values as described in Section 8.3 and
Appendix H.
Designers contemplating the use of reinforced slope systems should offer the same
degree of involvement to all suppliers who can accomplish the project objectives.
Geosynthetic reinforcement material specifications and special provisions for
reinforced slope systems should require suppliers to provide a qualified and
experienced representative at the site, for a minimum of three days, to assist the
contractor and Agency inspectors at the start of construction. If there is more than
one slope on a project, then this criterion should apply to construction of the initial
slope only. From then on, the representative should be available on an as-needed
basis, as requested by the Agency Engineer, during construction of the remainder of
the slope(s).
An in-house design approach and an end result approach to reinforced slope
solicitation are included in this document. Some user agencies prefer one approach to
the other, or a mixed use of approaches depending on the critical natrue of the slope
structures. Both approaches are acceptable if properly implemented. Each
approach has advantages and disadvantages.

Any proprietary material should undergo an Agency review prior to inclusion as an


alternate offered either during the design (in-house) or construction (value engineering
or end result) phase.
It is highly recommended that each Transportation Agency develop documented procedures
for the following.
Review and approval of geosynthetic soil reinforcing materials, both longterm tensile strength and soil-geosynthetic interaction properties.
Review and approval of drainage composite materials.
Review and approval of erosion control materials.
Review and approval of geosynthetic reinforced slope systems and suppliers
(geosynthetic soil reinforcing materials, drainage composites, and erosion
control materials).
In-house design of reinforced slopes.
Contracting for outside design and supply of reinforced slope systems.
In house designs are generally prepared for on a project-specific basis. However, standard
designs may be developed and used as agency in-house designs, as discussed in Sec. 8.9-3.

FHWA NHI-07-092
Geosynthetic Engineering

8 45

8 Reinforced Slopes
August 2008

8.9 SPECIFICATIONS AND CONTRACTING APPROACH


This section includes:
requirements for the method specification of reinforcement materials based on inhouse design; and
a guideline for the (performance) specification of a reinforced slope system based on
line and grade conceptual plans developed by the agency.
Conceptual plans must contain the geometric, geotechnical, and project-specific design
information. A framework is provided for developing an Agency specification or special
provision. Generic pullout and allowable tensile strength values, based upon preapproved
geotextiles or geogrids should be inserted into the corresponding Tables 8-3 and 8-4.
Reinforced slopes may be designed in-house or by a geosynthetic reinforcement supplier
based upon a performance specification. Geosynthetic reinforcement, erosion measures, and
drainage details must be specified when contracting based on an in-house design. Permanent
geosynthetic erosion mats are usually required on steeper and taller slopes, and degradable
blankets may be sufficient for flatter and shorter slopes. Specifications for these materials
should follow the recommendations in Chapter 4. Erosion control material from an Agencyapproved products list could be used in lieu of a specification. If required, geosynthetic
drainage composite specifications from Chapter 2 should also be included. Drainage design,
required flow capacities, and geotextile filtration criteria are project specific and require
specific evaluation for a project. Thus, the authors recommend that drainage features be
required in all walls unless the engineer determines such feature is, or features are, not
required for a specific project or structure. Line and grade drawings, erosion control
design, and a system specification are required for the second option. Alternatives in
specified materials or systems are possible during design and post-award periods.
All feasible, cost-effective material alternates should be seriously considered for in-house
design projects:
1. Alternate materials during the design phase - Consultants and the Agency should
consider all approved materials and shall provide generic specifications for
geosynthetic reinforcement and other materials used in the system.
2. Value engineering (VE) may be applied by the Contractor for the use of material
alternates. However, this approach will require soil-geosynthetic interaction
evaluation and possible testing. Proposals incorporating unapproved materials, not
submitted for review prior to a project's design phase, should not be approved for
VE proposals.

FHWA NHI-07-092
Geosynthetic Engineering

8 46

8 Reinforced Slopes
August 2008

8.9-1 Specification for Geosynthetic Soil Reinforcement (Material and construction


specification for an Agency Design).
1. DESCRIPTION:
Work shall consist of furnishing geosynthetic soil reinforcement and all equipment necessary to
install the geogrid in construction of reinforced soil slopes.
2. GEOSYNTHETIC REINFORCEMENT MATERIAL:
2.1 The specific geosynthetic reinforcement material and supplier shall be preapproved by the
Agency as outlined in the Agency's reinforced slope policy.
2.2 The geosynthetic reinforcement shall consist of a geogrid or geotextile that can develop
sufficient mechanical interaction with the surrounding soil or rock. The geosynthetic
reinforcement structure shall be dimensionally stable and able to retain its geometry under
construction stresses and shall have high resistance to damage during construction, to
ultraviolet degradation, and to all forms of chemical and biological degradation encountered
in the soil being reinforced.
2.3 The geosynthetics shall have an Allowable Strength (Tal) and Pullout Resistance, for the soil
type(s) indicated, as listed in Table 8-3 for geotextiles and/or Table 8-4 for geogrids.
2.4 The geosynthetics shall meet respective durability property requirements listed in Table 8-5.
2.5 The permeability of a geotextile reinforcement shall be greater than the permeability of the
reinforced fill soil.
2.6 Certification:
The contractor shall submit a manufacturer's certification that the
geosynthetics supplied meet the respective index criteria set when geosynthetic was approved
by the Agency, measured in full accordance with all test methods and standards specified. In
case of dispute over validity of values, the Engineer can require the Contractor to supply test
data from an Agency-approved laboratory to support the certified values submitted.
2.7 Quality Assurance/Index Properties: Testing procedures for measuring design properties
require elaborate equipment, tedious set up procedures and long durations for testing. These
tests are inappropriate for quality assurance (QA) testing of geosynthetic reinforcements
received on site. In lieu of these tests for design properties, a series of index criteria may be
established for QA testing. These index criteria include mechanical and geometric properties
that directly impact the design strength and soil interaction behavior of geosynthetics. It is
likely that each family of products will have varying index properties and QC/QA test
methods. QA testing should measure the respective index criteria set when geosynthetic was
FHWA NHI-07-092
Geosynthetic Engineering

8 47

8 Reinforced Slopes
August 2008

approved by the Agency. Minimum average roll values, per ASTM D 4759, shall be used for
conformance.
3. CONSTRUCTION:
3.1 Delivery, Storage, and Handling:
Follow requirements set forth under materials
specifications for geosynthetic reinforcement, drainage composite, and geosynthetic erosion
mat.
3.2 On-Site Representative: Geosynthetic reinforcement material suppliers shall provide a
qualified and experienced representative on site, for a minimum of three days, to assist the
Contractor and Agency inspectors at the start of construction. If there is more than one slope
on a project then this criteria will apply to construction of the initial slope only. The
representative shall also be available on an as-needed basis, as requested by the Agency
Engineer, during construction of the remaining slope(s).
3.3 Site Excavation: All areas immediately beneath the installation area for the geosynthetic
reinforcement shall be properly prepared as detailed on the plans, specified elsewhere within
the specifications, or directed by the Engineer. Subgrade surface shall be level, free from
deleterious materials, loose, or otherwise unsuitable soils. Prior to placement of geosynthetic
reinforcement, subgrade shall be proofrolled to provide a uniform and firm surface. Any soft
areas, as determined by the Owner's Engineer, shall be excavated and replaced with suitable
compacted soils. Foundation surface shall be inspected and approved by the Owner's
Geotechnical Engineer prior to fill placement. Benching the backcut into competent soil is
recommended to improve stability.
3.4 Geosynthetic Placement: The geosynthetic reinforcement shall be installed in accordance
with the manufacturer's recommendations. The geosynthetic reinforcement shall be placed
within the layers of the compacted soil as shown on the plans or as directed.
The geosynthetic reinforcement shall be placed in continuous, longitudinal strips in the
direction of main reinforcement. However, if the Contractor is unable to complete a required
length with a single continuous length of geogrid, a joint may be made with the Engineer's
approval. Only one joint per length of geogrid shall be allowed. This joint shall be made for
the full width of the strip by using a similar material with similar strength. Joints in geogrid
reinforcement shall be pulled and held taut during fill placement. Joints shall not be used
with geotextiles.
Adjacent strips, in the case of 100% coverage in plan view, need not be overlapped. The
minimum horizontal coverage is 50%, with horizontal spacings between reinforcement no
greater than 3 ft (1 m). Horizontal coverage of less than 100% shall not be allowed unless
specifically detailed in the construction drawings.
FHWA NHI-07-092
Geosynthetic Engineering

8 48

8 Reinforced Slopes
August 2008

Adjacent rolls of geosynthetic reinforcement shall be overlapped or mechanically connected


where exposed in a wrap-around face system, as applicable.
Place only that amount of geosynthetic reinforcement required for immediately pending work
to prevent undue damage. After a layer of geosynthetic reinforcement has been placed, the
next succeeding layer of soil shall be placed and compacted as appropriate. After the
specified soil layer has been placed, the next geosynthetic reinforcement layer shall be
installed. The process shall be repeated for each subsequent layer of geosynthetic
reinforcement and soil.
Geosynthetic reinforcement shall be placed to lay flat and pulled tight prior to backfilling.
After a layer of geosynthetic reinforcement has been placed, suitable means, such as pins or
small piles of soil, shall be used to hold the geosynthetic reinforcement in position until the
subsequent soil layer can be placed. Under no circumstances shall a track-type vehicle be
allowed on the geosynthetic reinforcement before at least 6 in. (150 mm) of soil has been
placed.
During construction, the surface of the fill should be kept approximately horizontal.
Geosynthetic reinforcement shall be placed directly on the compacted horizontal fill surface.
Geosynthetic reinforcements are to be placed within 3 in. (75 mm) of the design elevations
and extend the length as shown on the elevation view, unless otherwise directed by the
Owner's Engineer. Correct orientation of the geosynthetic reinforcement shall be verified by
the Contractor.
3.5 Fill Placement: Fill shall be compacted as specified by project specifications or to at least 95
percent of the maximum density determined in accordance with AASHTO T-99, whichever is
greater.
Density testing shall be made every lift of soil placement or as otherwise specified by the
Owner's Engineer or contract documents.
Backfill shall be placed, spread, and compacted in such a manner to minimize the
development of wrinkles and/or displacement of the geosynthetic reinforcement.
Fill shall be placed in 12-in. (300 mm) maximum lift thickness where heavy compaction
equipment is to be used, and 6-in. (150 mm) maximum uncompacted lift thickness where
hand-operated equipment is used.
Backfill shall be graded away from the slope crest and rolled at the end of each workday to
prevent ponding of water on the surface of the reinforced soil mass.

FHWA NHI-07-092
Geosynthetic Engineering

8 49

8 Reinforced Slopes
August 2008

Tracked construction equipment shall not be operated directly upon the geosynthetic
reinforcement. A minimum fill thickness of 6 in. (150 mm) is required prior to operation of
tracked vehicles over the geosynthetic reinforcement. Turning of tracked vehicles should be
kept to a minimum to prevent tracks from displacing the fill and the geosynthetic
reinforcement.
If recommended by the geogrid manufacturer and approved by the Engineer, rubber-tired
equipment may pass over the geogrid reinforcement at slow speeds, less than 10 mph.
Sudden braking and sharp turning shall be avoided.
3.6 Erosion Control Material Installation:
installation notes.
3.7 Geosynthetic Drainage Composite:
Specification for installation notes.

See Erosion Control Material Specification for

See Geocomposite Drainage Composite Material

3.8 Final Slope Geometry Verification: Contractor shall confirm that as-built slope geometries
conform to approximate geometries shown on construction drawings.
4. METHOD OF MEASUREMENT:
Measurement of geosynthetic reinforcement is on a square-yard basis and is computed on the
total area of geosynthetic reinforcement shown on the construction drawings, exclusive of the
area of geosynthetics used in any overlaps. Overlaps are an incidental item.
5. BASIS OF PAYMENT:
5.1 The accepted quantities of geosynthetic reinforcement by type will be paid for per-squareyard in-place.
5.2 Payment will be made under:
Pay Item
Geogrid Soil Reinforcement - Type I
Geogrid Soil Reinforcement - Type II
Geogrid Soil Reinforcement - Type III
or
Geotextile Soil Reinforcement - Type I
Geotextile Soil Reinforcement - Type II
Geotextile Soil Reinforcement - Type III
Material Supplier Representative
(exceeding 3 days)
FHWA NHI-07-092
Geosynthetic Engineering

8 50

Pay Unit
square yard
square yard
square yard
square yard
square yard
square yard
person-day

8 Reinforced Slopes
August 2008

Table 8-3
Allowable Geotextile Strength with Various Soil Typesa

Geotextile

Ultimate Strength
TULT
(lb/ft)a

Long-Term
Allowable Strength
Tal (lb/ft)

Minimum
Permeability, k
(or Permittivity, R)

ASTM D 4595c

FHWAd

ASTM D 4491

For use with


these fillsb

Note (e)

GW-GM

Note (e)

SW-SM-SC

Note (e)

GW-GM

Note (e)

SW-SM-SC

NOTES:
a) Based upon minimum average roll values.
b) Per Unified Soil Classification System.
c) ASTM D 4595 shall be with an 8-in. width specimen, or a 4-in. specimen width with correlation to an 8-in.
width. Correlation methodology shall be submitted to, and is subject to acceptance by, the Engineer.
d) The Long Term Allowable Tensile Strength shall be determined by applying appropriate reduction factors to
the Ultimate Tensile Strength of the geogrid to account for installation damage, survivability, creep,
durability and degradation. A 75-year design life shall be used in determining the long term allowable
tensile strength. The FHWA methodology (FHWA NHI-00-043 {Elias et al., 2001}) shall be used for this
computation. Proposed strength and reduction factors are subject to approval. Minimum durability
reduction factors by soil pH per Table 8-5. Minimum installation damage factor is 1.10. Minimum creep
reduction factor for polypropylene geotextiles is ___; unless lower value has been recommended in an
approved evaluation report (e.g., HITEC MSE Wall Evaluation Report), in which case the minimum is equal
to the evaluation report recommended value. Minimum creep reduction factor for polyester geotextiles is
___, unless a lower value has been recommended in an approved evaluation report (e.g., HITEC MSE Wall
Evaluation Report), in which case the minimum is equal to the evaluation report recommended value.
e) Equal to or greater than the permeability of the fill soil.

FHWA NHI-07-092
Geosynthetic Engineering

8 51

8 Reinforced Slopes
August 2008

Table 8-4
Allowable Geogrid Strength with Various Soil Typesa

Geogrid

Ultimate Strength
TULT (lb/ft)a

Long-Term
Allowable Strength
Tal (lb/ft)

ASTM D 6637c

FHWAd

For use with


these fillsb

GW-GM

SW-SM-SC

GW-GM

SW-SM-SC

NOTES:
a) Based upon minimum average roll values.
b) Per Unified Soil Classification System.
c) ASTM D 6637 provides an option of three testing methods. The option used to define the ultimate strength
shall be the same as used to define Tultlot in the creep reduction factor calculation (see Appendix H).
d) The Long Term Allowable Tensile Strength shall be determined by applying appropriate reduction factors to
the Ultimate Tensile Strength of the geogrid to account for installation damage, survivability, creep,
durability and degradation. A 75-year design life shall be used in determining the long term allowable
tensile strength. The FHWA methodology (FHWA NHI-00-043 {Elias et al., 2001}) shall be used for this
computation. Proposed strength and reduction factors are subject to approval. Minimum durability
reduction factors by soil pH per Table 8-5. Minimum installation damage factor is 1.10. Minimum creep
reduction factor for polyethylene geogrids is ___; unless lower value has been recommended in an approved
evaluation report (e.g., HITEC MSE Wall Evaluation Report), in which case the minimum is equal to the
evaluation report recommended value. Minimum creep reduction factor for polyester geogrids is ___, unless
a lower value has been recommended in an approved evaluation report (e.g., HITEC MSE Wall Evaluation
Report), in which case the minimum is equal to the evaluation report recommended value.

FHWA NHI-07-092
Geosynthetic Engineering

8 52

8 Reinforced Slopes
August 2008

Table 8-5
Durability Reduction Factors by Product and Soil Fill pH
Minimum Reduction
Geosynthetic
Soil Fill pH
Factor, RFD
5 < pH < 8
1.6

PET Geotextiles
w/ Mn < 20,000 & 40 < CEG < 50

PET Coated Geogrids and PET Geotextiles


w/ Mn > 25,000 & CEG < 30
HDPE and PP Geogrids

3 < pH < 5
8 < pH < 9

2.0

5 < pH < 8

1.15

3 < pH < 5
8 < pH < 9

1.3

3 < pH

1.10

NOTES:
a. Mn = number average molecular weight
b. CEG = carboxyl end group
c. Use of geosynthetics outside of the indicated pH or molecular property range (PET) requires specific
product testing.
d. See FHWA NHI-00-044 (Elias, 2001) for detailed discussions on geosynthetic durability.

8.9-2

Specification for Geosynthetic Reinforced Soil Slope System (Vendor/Contractor


Design) (after Berg, 1993)

1.0 DESCRIPTION
Work shall consist of design, furnishing materials, and construction of geosynthetic reinforced
soil slope structure. Supply of geosynthetic reinforcement, drainage composite, and erosion
control materials, and site assistance are all to be furnished by the slope system supplier.

2.0 REINFORCED SLOPE SYSTEM


2.1 Acceptable Suppliers The following suppliers can provide an agency approved system:
(a) ___________________________
(b) ___________________________
(c) ___________________________
2.2 Materials: Only those geosynthetic reinforcement, drainage composite, and erosion mat
materials approved by the Agency prior to advertisement for bids on the project under
consideration shall be utilized in the slope construction. Geogrid Soil Reinforcement,
Geotextile Soil Reinforcement, Drainage Composite, and Geosynthetic Erosion Mat
materials are specified under respective material specifications.
FHWA NHI-07-092
Geosynthetic Engineering

8 53

8 Reinforced Slopes
August 2008

2.3 Design Submittal: The Contractor shall submit six (6) sets of detailed design calculations,
construction drawings, and shop drawings for approval within thirty (30) days of
authorization to proceed and at least sixty (60) days prior to the beginning of reinforced
slope construction. The calculations and drawings shall be prepared and sealed by a
professional engineer, licensed in the State. Submittal shall conform to Agency
requirements for steepened reinforced soil systems.
2.4 Material Submittals: The Contractor shall submit six (6) sets of manufacturer's certification
that indicate the geosynthetic soil reinforcement, drainage composite, and geosynthetic
erosion mat meet the requirements set forth in the respective material specifications, for
approval at least sixty (60) days prior to the start of reinforced slope construction.

3.0 CONSTRUCTION
(Should follow the specification details in Section 8.9-1)

4.0 METHOD OF MEASUREMENT


Measurement of Geosynthetic Reinforced Soil Slope Systems is on a vertical square foot basis.
Payment shall cover reinforced slope design, and supply and installation of geosynthetic soil
reinforcement, reinforced soil fill, drainage composite, and geosynthetic erosion mat.
Excavation of any suitable materials and replacement with select fill, as directed by the
Engineer, shall be paid under a separate pay item.
Quantities of reinforced soil slope system as shown on the plans may be increased or decreased
at the direction of the Engineer based on construction procedures and actual site conditions.

5.0 BASIS OF PAYMENT


5.1 The accepted quantities of geosynthetic reinforced soil slope system will be paid for per
vertical square-foot in-place.
5.2 Payment will be made under:

Pay Item
Geosynthetic Slope System
Material Supplier Representative

FHWA NHI-07-092
Geosynthetic Engineering

8 54

Pay Unit
vertical square-foot
person-day

8 Reinforced Slopes
August 2008

8.10

INSTALLATION PROCEDURES

Reinforcement layers are easily incorporated between the compacted lifts of fill. Therefore,
construction of reinforced slopes is very similar to normal embankment construction. The
following is the usual construction sequence:
A. Site preparation.
Clear and grub site.
Remove all slide debris (if a slope reinstatement project).
Prepare level subgrade for placement of first level of reinforcing.
Proofroll subgrade at the base of the slope with roller or rubber-tired vehicle.
B. Place the first reinforcing layer.
Reinforcement should be placed with the principal strength direction
perpendicular to the face of slope.
Secure reinforcement with retaining pins to prevent movement during fill
placement.
A minimum overlap of 6 in. (150 mm) is recommended along the edges
perpendicular to the slope for wrapped-face structures. Alternatively, with
geogrid reinforcement, the edges may be clipped or tied together. When
geosynthetics are not required for face support, no overlap is required and
edges should be butted.
C. Place backfill on reinforcement.
Place fill to required lift thickness on the reinforcement using a front-end
loader operating on previously placed fill or natural ground.
Maintain a minimum of 6 in. (150 mm) between reinforcement and wheels
of construction equipment. This requirement may be waived for rubber-tired
equipment provided that field trials, including geosynthetic strength tests,
have demonstrated that anticipated traffic conditions will not damage the
specific geosynthetic reinforcement.
Compact with a vibratory roller or plate-type compactor for granular
materials, or a rubber-tired vehicle for cohesive materials.
When placing and compacting the backfill material, avoid any deformation
or movement of the reinforcement.
Use lightweight compaction equipment near the slope face to help maintain
face alignment.
D. Compaction control.
Provide close control on the water content and backfill density. It should be
compacted at least 95% of the standard AASHTO T-99 or ASTM D 698
maximum density within 2% of optimum moisture.
FHWA NHI-07-092
Geosynthetic Engineering

8 55

8 Reinforced Slopes
August 2008

If the backfill is a coarse aggregate, then a relative density or a method type


compaction specification should be used.
E. Face construction.
As indicated in the design section (8.3-3), a face wrap generally is not required for
slopes up to 1H:1V, if the reinforcement is maintained at close spacing (i.e., every
lift or every other lift, but no greater than 16 in. (400 mm)). In this case the
reinforcement can be simply extended to the face. For this option, a facing
treatment should be applied to prevent erosion during and after construction (e.g., a
geosynthetic erosion control mat and vegetation). If slope facing is required to
prevent sloughing (i.e., slope angle $ is greater than Nsoil) or erosion, sufficient
reinforcement lengths could be provided for a wrapped-face structure. The
following procedures are recommended for wrapping the face.
Turn up reinforcement at the face of the slope and return the reinforcement a
minimum of 3 to 4 ft (1 to 1.2 m) into the embankment below the next
reinforcement layer (see Figure 8-10).
For steep slopes, formwork is required to support the face during
construction and permanent welded wire mesh or gabion faces should be
used where reinforcement spacings of greater than 18 in.(0.5 m), are used.
For geogrids, a fine mesh screen or geotextile may be required at the face to
retain backfill materials.
F. Continue with additional reinforcing materials and backfill.
NOTE: If drainage layers are required, they should be constructed directly
behind or on the sides of the reinforced section.
Several construction photos from reinforced slope projects are shown in Figure 8-11.

FHWA NHI-07-092
Geosynthetic Engineering

8 56

8 Reinforced Slopes
August 2008

Figure 8-10. Construction of reinforced slopes.

FHWA NHI-07-092
Geosynthetic Engineering

8 57

8 Reinforced Slopes
August 2008

(a)

(c)
(b)
Figure 8-11. Reinforced slope construction: a) geogrid and fill placement; b) soil fill erosion
control mat facing; and c) finished, vegetated 1:1 slope.

FHWA NHI-07-092
Geosynthetic Engineering

8 58

8 Reinforced Slopes
August 2008

8.11

FIELD INSPECTION

As with all geosynthetic construction, and especially with critical structures such as
reinforced slopes, competent and professional field inspection is absolutely essential for
successful construction. Field personnel must be properly trained to observe and document
every phase of the construction (see NHI course #132080 Inspection of MSEW and RSS).
They must make sure that the specified material is delivered to the project, that the
geosynthetic is not damaged during construction, and that the specified sequence of
construction operations are explicitly followed. Field personnel should review the checklist
items in Section 1.7. Other important details for RSS that should be monitored and
documented include:
C Drainage system components and installation
C Vertical spacing of primary and secondary reinforcements
C Length of primary and secondary reinforcements as well as wrap-facing return
C Construction of the slope face, facing alignment and application of the facing
treatment to minimize geosynthetic exposure to ultraviolet light.

8.12

STANDARD DESIGNS

RSS structures are customarily designed on a project-specific basis. Most agencies use a
line-and-grade contracting approach, thus the contractor selected RSS vendor provides the
detailed design after contract bid and award. This approach works well. However, standard
designs can be developed and implemented by an agency for RSS structures.
Use of standard designs for RSS structures offers the following advantages over a line-andgrade approach:
C Agency is more responsible for design details and integrating slope design with other
components.
C Pre evaluation and approval of materials and material combinations, as opposed to
evaluating contractor submittal post bid.
C Economy of agency design versus vendor design/stamping of small reinforced slopes.
C Agency makes design decisions versus vendors making design decisions.
C More equitable bid environment as agency is responsible for design details, and
vendors are not making varying assumptions.
C Filters out substandard work, systems and designs with associated approved product
lists.

FHWA NHI-07-092
Geosynthetic Engineering

8 59

8 Reinforced Slopes
August 2008

The Minnesota Department of Transportation (Mn/DOT) recently developed and


implemented (in-house) standardized RSS designs (Berg, 2000). The use of these standard
designs is limited by geometric, subsurface and economic constraints. Structures outside of
these constraints should be designed on a project-specific basis. The general approach used
in developing these standards could be followed by other agencies to develop their own,
agency-specific standard designs.
Standardized designs require generic designs and generic materials. Generic designs require
definition of slope geometry and surcharge loads, soil reinforcement strength, structure
height limit, and slope facing treatment. As an example, the Mn/DOT standard designs
address two geometric and surcharge loadings, two reinforced soil fills, and can be used for
slopes up to 26.2 feet (8 m) in height. Three reinforcement long-term strengths, Tal, of 700,
1050 and 1400 plf (10, 15 and 20 kN/m) are used in the standard designs, though a structure
must use the same reinforcement throughout its height and length.
Generic material properties used definitions of shear strength and unit weight of the
reinforced fill, retained backfill and foundation soils applicable to the agencys specifications
and regional geology. Definition of generic material properties requires the development of
approved product lists for soil reinforcements and face erosion control materials. A standard
face treatment is provided, however, it is footnoted with Develop site-specific
recommendations for highly shaded areas, highly visible urban applications, or in sensitive
areas.
An example design cross section and reinforcement layout table from the Mn/DOT standard
designs is presented in Figure 8-12. Note that the Mn/DOT standard designs are not directly
applicable to, nor should they be used by, other agencies.

FHWA NHI-07-092
Geosynthetic Engineering

8 60

8 Reinforced Slopes
August 2008

Figure 8-12. Example of standard RSS design (Mn/DOT (Berg, 2000)).


FHWA NHI-07-092
Geosynthetic Engineering

8 61

8 Reinforced Slopes
August 2008

8.13

REFERENCES

References quoted within this section are listed below. The FHWANHI-00-043 manual
(Elias et al., 2001) reference is a comprehensive guideline specifically addressing reinforced
slopes in transportation applications. It is a key reference for design, specification, and
contracting. This and other key references are noted in bold type.

AASHTO (2002). Standard Specifications for Highway Bridges, Seventeenth Edition,


American Association of State Transportation and Highway Officials, Washington,
D.C.
AASHTO (1990). Design Guidelines for Use of Extensible Reinforcements (Geosynthetic)
for Mechanically Stabilized Earth Walls in Permanent Applications, Task Force 27
Report -In Situ Soil Improvement Techniques, American Association of State
Transportation and Highway Officials, Washington, D.C.
AASHTO (1989). Roadside Design Guide, American Association of State Transportation
and Highway Officials, Washington, D.C.
ASTM Test Methods - see Appendix E.
Berg, R.R. (2000). Minnesota Department of Transportation Standard MSEW and RSS
Designs, Project Report Volume I, Summary Report, Minnesota Department of
Transportation, Oakdale, MN, 98 p.
Berg, R.R. and DiMaggio, J.D. (1994). U.S. Guidelines for Reinforced Slopes in
Transportation Applications, Proceedings of the Fifth International Conference on
Geotextiles, Geomembranes and Related Products, Vol. 1, Singapore, September, pp.
233-236.
Berg, R.R. (1993). Guidelines for Design, Specification, & Contracting of Geosynthetic
Mechanically Stabilized Earth Slopes on Firm Foundations, Federal Highway
Administration, FHWA-SA-93-025, 87 p.
Berg, R.R., Anderson, R.P., Race, R.J., and Chouery-Curtis, V.E. (1990). Reinforced Soil
Highway Slopes, Transportation Research Record No. 1288, Geotechnical Engineering,
Transportation Research Board, Washington, D.C., pp. 99-108.
Berg, R.R., Chouery-Curtis, V.E. and Watson, C.H. (1989). Critical Failure Planes in
Analysis of Reinforced Slopes, Proceedings of Geosynthetics '89, Volume 1, San Diego,
CA, February.
Bishop, A.W. (1955). The Use of the Slip Circle in the Stability Analysis of Slopes,
Geotechnique, Volume 5, Number 1.

FHWA NHI-07-092
Geosynthetic Engineering

8 62

8 Reinforced Slopes
August 2008

Bonaparte, R. and Berg, R.R. (1987). Long-Term Allowable Tension for Geosynthetic
Reinforcement, Proceedings of Geosynthetics '87 Conference, Volume 1, New Orleans,
LA, pp. 181-192.
Christopher, B.R. and Holtz, R.D. (1985). Geotextile Engineering Manual, Federal Highway
Administration, FHWA-TS-86/203, 1044 p.
Christopher, B.R., Gill, S.A., Giroud, J.P., Juran, I. Scholsser, F., Mitchell, J.K. and
Dunnicliff, J. (1989). Reinforced Soil Structures, Volume I. Design and Construction
Guidelines and Volume II Summary of Research and Systems Information, Federal
Highway Administration, FHWA-RD-89-043, 287 p.
Collin, J.G. (1996). Controlling Surficial Stability Problems on Reinforced Steepened Slopes,
Geotechnical Fabrics Report, IFAI.
Duncan, J.M. and Wright, S.G. (2005). Soil Strength and Slope Stability, John Wiley &
Sons, Inc., Hoboken, N.J., 297 p.
Elias, V., Welsh, J., Warren, J., Lukas, R.,. Collin, J.G. and Berg, R.R. (2006). Ground
Improvement Methods, Federal Highway Adminstration, FHWA NHI-06-019 (Vol. I)
and FHWA NHI-06-020 (Vol. II), 536 (Vol. I) and 520 (Vol. II) p.

Elias, V., Christopher, B.R. and Berg, R.R. (2001). Mechanically Stabilized Earth
Walls and Reinforced Soil Slopes, Design & Construction Guidelines, Federal
Highway Administration, FHWA-NHI-00-043, 418 p.
Elias, V. (2001). Corrosion/Degradation of Soil Reinforcements for Mechanically Stabilized
Earth Walls and Reinforced Soil Slopes, Federal Highway Administration, FHWA NHI00-044, 94 p.
Elias, V. and Christopher, B.R. (1997). Mechanically Stabilized Earth Walls and Reinforced
Soil Slopes, Design & Construction Guidelines, Federal Highway Administration,
FHWA-SA-96-071, 371 p.
Geosynthetics (2008). Project Showcase 2007 International Achievement Awards,
Geosynthetics, Vol 26, No. 1, February/March, IFAI, pp 8-9.
Gray, D.H. and Sotir, R. (1995). Biotechnical and Soil Bioengineering Slope Stabilization, A
Practical Guide for Erosion Control, John Wiley & Sons, New York, NY.
Hayden, R.F., Schmertmann, G.R., Qedan, B.Q., and McGuire, M.S. (1991). High Clay
Embankment Over Cannon Creek Constructed With Geogrid Reinforcement,
Proceedings of Geosynthetics '91, Volume 2, Atlanta, GA, pp. 799-822.

FHWA NHI-07-092
Geosynthetic Engineering

8 63

8 Reinforced Slopes
August 2008

Iwasaki, K. and Watanabi, S. (1978). Reinforcement of Highway Embankments in Japan,


Proceedings of the Symposium on Earth Reinforcement, ASCE, Pittsburgh, PA, pp. 473500.
Jewell, R.A., Paine, N. and Woods, R.I. (1984). Design Methods for Steep Reinforced
Embankments, Proceedings of the Symposium on Polymer Grid Reinforcement, Institute
of Civil Engineering, London, U.K., pp. 18-30.
Jewell, R.A. (1990). Revised Design Charts for Steep Reinforced Slopes, Reinforced
Embankments: Theory and Practice in the British Isles, Thomas Telford, London, U.K.
Kavazanjian, Jr., E., Matasovi, N., Hadj-Hamou, T., Sabatini, P.J. (1997). Geotechnical
Engineering Circular No. 3, Design Guidance: Geotechnical Earthquake Engineering
for Highways, Volume I Design Principles, Federal Highway Administration, FHWA
SA-97-076, 186 p.
Kramer, S.L. (1996). Geotechnical Earthquake Engineering, Prentice Hall, Upper Saddle
River, N.J., 653 p.
Leshchinsky, D. and Boedeker, R.H. (1989). Geosynthetic Reinforced Soil Structures,
Journal of Geotechnical Engineering, ASCE, Volume 115, Number 10, pp. 1459-1478.
Mitchell, J.K. and Villet, W.C.B. (1987). Reinforcement of Earth Slopes and Embankments,
NCHRP Report No. 290, Transportation Research Board, Washington, D.C.
Morgenstern, N. and Price, V.E. (1965). The Analysis of the Stability of General Slip
Surfaces, Geotechnique, Volume 15, Number 1, pp. 79-93.
Ruegger, R. (1986). Geotextile Reinforced Soil Structures on which Vegetation can be
Established, Proceedings of the 3rd International Conference on Geotextiles, Vienna,
Austria, Volume II, pp. 453-458.
Samtani, N.C. and Nowatzki, E.A. (2006). Soils and Foundations Reference Manual, Federal
Highway Administration, FHWA NHI-06-088 (Vol. I) and FHWA NHI-06-089 (Vol.
II), 462 (Vol. I) and (Vol. II) 594 p.
Schmertmann, G.R., Chouery-Curtis, V.E., Johnson, R.D. and Bonaparte, R. (1987). Design
Charts for Geogrid-Reinforced Soil Slopes, Proceedings of Geosynthetics '87, New
Orleans, LA, Volume 1, pp. 108-120.
Silvestri, V. (1983). The Bearing Capacity of Dykes and Fills Founded on Soft Soils of
Limited Thickness, Canadian Geotechnical Journal, Vol. 20, No. 3, pp. 428-436.
Soong, T. and Koerner, R.M. (1999). Geosynthetic Reinforced and Geocomposite Drained
Retaining Walls Utilizing Low Permeability Backfill Soils, GRI Report #24,
Geosynthetic Research Institute, Folsom, PA, 140 p.
FHWA NHI-07-092
Geosynthetic Engineering

8 64

8 Reinforced Slopes
August 2008

Sotir, R.B. and Christopher, B.R. (2000). Soil Bioengineering and Geosynthetics for Slope
Stabilization, Proceedings, Geosynthetics Asia 2000, Selangor Daruf, Ehsan Malaysia.
Spencer, E. (1981). Slip Circles and Critical Shear Planes, Journal of the Geotechnical
Engineering Division, ASCE, Volume 107, Number GT7, pp. 929-942.
The Tensar Corporation (1987). Slope Reinforcement, Brochure, 10 p.
Thielen, D.L. and Collin, J.G. (1993). Geogrid Reinforcement for Surficial Stability of
Slopes, Proceedings of Geosynthetics '93, Vancouver, B.C., Volume 1, pp. 229-244.
Werner, G. and Resl, S. (1986). Stability Mechanisms in Geotextile Reinforced EarthStructures, Proceedings of the 3rd International Conference on Geotextiles, Vienna,
Austria, Volume II, pp. 465-470.
Zornberg, J.G. and J.K. Mitchell (1992). Poorly Draining Backfills for Reinforced Soil
Structures - A State of the Art Review, Geotechnical Research Report No. UCB/GT/9210, Department of Civil Engineering, University of California, Berkeley, 101 p.

FHWA NHI-07-092
Geosynthetic Engineering

8 65

8 Reinforced Slopes
August 2008

FHWA NHI-07-092
Geosynthetic Engineering

8 66

8 Reinforced Slopes
August 2008

9.0 MECHANICALLY STABILIZED EARTH RETAINING


WALLS AND ABUTMENTS

The purpose of this chapter is to review the variety of available geosynthetic MSE wall types,
discuss their typical use, consider the advantages of geosynthetic MSE walls, and detail how
they are designed, specified, and constructed. Detailed design, contracting, and construction
guidelines are provided in FHWA NHI-00-043 Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes (Elias et al., 2001) (the reference manual for NHI courses 132042
and132043); AASHTO Standard Specifications for Highway Bridges (ASD) (2002), and the
most recent edition of AASHTO LRFD Bridge Design Specifications (2007, with interims).
The design guidelines presented within this chapter follow the allowable strength design
(ASD), as addressed in FHWA NHI-00-043 (Elias et al., 2001) and AASHTO (2002).
LRFD-based design of MSE walls (AASTHO, 2007) is summarized in section 9.13.

9.1 BACKGROUND
Retaining walls in transportation engineering are quite common. They are required where a
slope is uneconomical or not technically feasible. When selecting a retaining wall type,
mechanically stabilized earth (MSE) walls should always be considered. MSE (i.e.,
reinforced soil) walls are basically comprised of some type of reinforcing element in the soil
fill to help resist lateral earth pressures. When compared with conventional retaining wall
systems, there are often significant advantages to MSE retaining walls. MSE walls are very
cost effective, especially for walls in fill embankment cross sections. Furthermore, these
systems are more flexible than conventional earth retaining walls such as reinforced concrete
cantilever or gravity walls. Therefore, they are very suitable for sites with poor foundations
and for seismically active areas.
The modern invention of reinforcing the soil in fill applications was developed by Vidal in
France in the mid-1960s. The Vidal system, called Reinforced Earth, uses metal strips for
reinforcement, as shown schematically in Figure 9-1. The design and construction of
Reinforced Earth walls is quite well established, and thousands have been successfully
built worldwide in the last 35 years. During this time, other similar reinforcing systems, both
proprietary and nonproprietary, utilizing different types of metallic reinforcement have been
developed (e.g., VSL, Hilfiker, etc.; see Mitchell and Villet, 1987, and Christopher et al.,
1989).

FHWA NHI-07-092
Geosynthetic Engineering

9 1

9 MSE Walls
August 2008

Figure 9-1.

Component parts of a Reinforced Earth wall (Lee et al., 1973).

The use of geogrids or geotextiles rather than metallic strips, shown conceptually in Figure 92, is really a further development of the Reinforced Earth concept. Geosynthetics offer a
viable and often very economical alternative to metallic reinforcement for both permanent
and temporary walls, especially under certain environmental conditions. Reinforcing with
geosynthetics has been used since 1977 (Bell et al., 1975). Today the use in transportation
walls is quite common in many states. In the U.S., maximum heights of geosynthetic
reinforced walls constructed to date are about 60 ft (18 m), whereas metallic reinforced walls
have exceeded 150 ft (46 m) in height. A significant benefit of using geosynthetics is the
wide variety of wall facings available, resulting in greater aesthetic options. Metallic
reinforcement is typically used with articulated precast concrete panels. Alternate facing
systems for geosynthetic reinforced walls are discussed in Section 9.3.

FHWA NHI-07-092
Geosynthetic Engineering

9 2

9 MSE Walls
August 2008

Figure 9-2. Reinforced retaining wall systems using geosynthetics: (a) with wrap- around
geosynthetic facing, (b) with modular block masonry units, and (c) with fullheight (propped) precast panels.

9.2

APPLICATIONS

MSE structures, including those reinforced with geosynthetics, should be considered as costeffective alternates for all applications where conventional gravity, cast-in-place concrete
cantilever, bin-type, or metallic reinforced soil retaining walls are specified. This includes
bridge abutments as well as locations where conventional earthen embankments cannot be
constructed due to right-of-way restrictions (another alternative is a reinforced slope, see
Chapter 8). Conceptually, geosynthetic MSE walls can be used for any fill wall situation and
for low- to moderate-height cut-wall situations. Similar to other MSE wall types, the
relatively wide wall base width required for geosynthetic walls typically precludes their use
in tall cut situations. Figure 9-3 shows several completed geosynthetic reinforced retaining
walls.
Geosynthetic MSE walls are generally less expensive than conventional earth retaining
systems. Using geogrids or geotextiles as reinforcement has been found to be 30 to 50% less
expensive than other reinforced soil construction with concrete facing panels, especially for
small- to medium-sized projects (Allen and Holtz, 1991). They may be most cost-effective
in temporary or detour construction, and in low-volume road construction (e.g., national
forests and parks).

FHWA NHI-07-092
Geosynthetic Engineering

9 3

9 MSE Walls
August 2008

(a)

(b)

(c)

(d)

Figure 9-3. Examples of geosynthetic MSE walls: a. full-height panels, geogrid


reinforced wall, Arizona; b. modular block wall units, geogrid-reinforced wall,
Minnesota; c. modular block wall units, geogrid-reinforced bridge abutment,
Colorado; and d. temporary geotextile wrap around wall, Washington.

FHWA NHI-07-092
Geosynthetic Engineering

9 4

9 MSE Walls
August 2008

Due to their greater flexibility, MSE walls offer significant technical and cost advantages
over conventional gravity or reinforced concrete cantilever walls at sites with poor
foundations and/or slope conditions. These sites commonly require costly additional
construction procedures, such as deep foundations, excavation and replacement, or other
foundation soil improvement techniques.
The level of confidence needed for the design of a geosynthetic wall depends on the
criticality of the project (Carroll and Richardson, 1986). The criticality depends on the
design life, maximum wall height, the soil environment, risk of loss of life, and impact to the
public and to other structures if failure occurs. Assessment of criticality is rather subjective,
and sound engineering judgement is required. The Engineer or regulatory authority should
determine the critical nature of a given application. Design, as summarized and discussed
within this chapter, assumes that walls are classified as permanent, critical structures.
Of course, the method could be conservatively used to design temporary and other noncritical structures.

9.3 DESCRIPTION OF MSE WALLS


9.3-1 Soil Reinforcements
Geosynthetic MSE walls may utilize geogrids or geotextiles as soil reinforcing elements.
However, the prevalent material used in highway walls today is geogrid reinforcement. This
trend is driven both by needs of transportation agencies and by geosynthetic manufacturers
and suppliers of packaged wall systems. One of these needs enhanced aesthetics of the
completed wall is obviously controlled by the facing used. As discussed below, the facing
can dictate a preferred type of geosynthetic reinforcement.
9.3-2 Facings
A significant advantage of geosynthetic MSE walls over other earth retaining structures is the
variety of facings that can be used and the resulting aesthetic options that can be provided.
Descriptions of various facings are provided below. Some examples are illustrated in Figure
9-4.
Modular Block Wall (MBW) Units are the most common facing currently used for
geosynthetic MSE wall construction. These facing elements are also known as modular
block wall units and as concrete masonry unit (CMU). They are popular because of their
aesthetic appeal, widespread availability, and relative low cost (Berg, 1991). A broken
block, or natural stone-like, finish is the most popular MBW unit face finish.

FHWA NHI-07-092
Geosynthetic Engineering

9 5

9 MSE Walls
August 2008

Figure 9-4. Possible geosynthetic MSE wall facings: (a) geosynthetic wrapped face temporary wall; (b) geosynthetic facing protected by shotcrete; (c) full-height
precast concrete (propped) panels; and (d) modular concrete units.
MBW units are relatively small, squat concrete units, specially designed and manufactured
for retaining wall applications. The units are typically manufactured by a dry casting process
and weigh 35 to 110 lbs (15 to 50 kg) each, with 75 to 110 lbs (35 to 50 kg) units routinely
used for highway works. The nominal depth (dimension perpendicular to wall face) of
MBW units usually ranges between 12 to 20 inches (0.3 to 0.5 m). These large units can
provide significant contribution to stability, particularly for low- to moderate-height gravity
and MSE walls. But, for MSE wall design they are treated simply as a facing and their
FHWA NHI-07-092
Geosynthetic Engineering

9 6

9 MSE Walls
August 2008

contribution to stability ignored (except possibly for sliding stability, see Section 9.4). Unit
height is typically 8 in. (200 mm), but can vary between 4 to 8 in. (100 to 200 mm) for the
various manufacturers. Exposed face length typically ranges between 8 to 16 in. (200 to 600
mm). The durability of MBW units in freeze-thaw environments and when exposed to
roadway deicing salts is a concern. See Sections 9.9-1 and 9.9-2 for discussion of this issue.
MBW units may be manufactured solid or with cores. Full height cores are typically filled
with aggregate during erection. Units are normally dry-stacked (i.e., without mortar) in a
running bond configuration. Vertically adjacent units may be connected with shear pins, lips,
or other alignment aids.
The vertical connection mechanism between MBW units also contributes to the connection
strength between the geosynthetic reinforcement and the MBW units. Connection strength
must be addressed in design, and often controls the maximum allowable tensile load in a
given layer of reinforcement. Therefore, the reinforcement design strength and vertical
spacing of layers is specific to the particular combination of MBW unit and geosynthetic
reinforcement utilized.
Geogrids, both stiff and flexible, are the common reinforcing elements of MBW unit-faced
MSE walls in highway applications. Geotextiles have been used in walls with MBW unit
facings, but to a limited extent. A detailed description of MBW units, and design with these
units, is presented by NCMA (1997) and is summarized by Bathurst and Simac (1994).
Wrap-Around facings are commonly used: i) for temporary structures; ii) for walls that will
be subject to significant post-construction settlement; iii) where aesthetic requirements are
low; and/or iv) where post-construction facings are applied for protection and aesthetics. The
geosynthetic facing may be left exposed for temporary walls, as illustrated in Figure 9-5(a), if
the geosynthetic is stabilized against ultraviolet light degradation. A consistent vertical
spacing of reinforcement, and therefore wrap height, of 12 to 18 in (0.3 to 0.45 m) is
typically used. A sprayed concrete facing is usually applied to permanent walls to provide
protection against ultraviolet exposure, potential vandalism, and possible fire. Precast
concrete or wood panels may also be attached after construction.
Geotextiles are commonly used in wrap-around-faced MSE walls. With the proper
ultraviolet light stabilizer, these structures perform satisfactorily for a few years. They
should be covered by a permanent facing for longer-term applications. Geogrids are also
used for wrap facings, though a geotextile, an erosion control blanket, or sod is required to
retain fill soil. Alternatively, rock or gravel can be used in the wrap area and a filter placed
between the stone and fill soil. Secondary, biaxial geogrid can be used to provide the face
FHWA NHI-07-092
Geosynthetic Engineering

9 7

9 MSE Walls
August 2008

wrap for the primary, uniaxial soil reinforcing elements for geogrids. With the proper
ultraviolet light stabilizer specified, geogrids can be left uncovered for a number of years;
reportedly for design lives of 50 years or more for heavy, stiff geogrids (Wrigley, 1987).
Welded-Wire Facing is popular for construction of both temporary and permanent walls.
For temporary construction walls that are eventually buried in-place or dismantled, L-shaped
welded-wire mesh, with 12 24 in (300 600 mm) vertical and horizontal legs are normally
used. Such walls are commonly used for staged lane construction on grade separation
projects. Geosynthetic, or steel, soil reinforcements are used in these walls. Galvanized steel
is used for permanent walls, and non-galvanized or black steel is used for these temporary
walls. An example permanent wall is illustrated in Figure 9-5(b).
Permanent welded-wire facing is also used with geosynthetic, or steel, soil reinforcement.
The facing may be L-shaped welded wire mesh or woven steel gabions. The steel is
galvanized and should be designed with consideration of corrosion loss over the life to the
structure. Stone fill in the zone of the steel facing is often used with these walls. A
geotextile filter is typically used between the back face of the gabion baskets or stone fill and
the reinforced wall fill soil to prevent soil from piping through the stones. Aesthetically, the
walls constructed with L-shaped welded wire mesh have a look similar to stone-fill gabions.
Soil and nutrients can also be blended with the gravel to promote growth of vegetation and
create green walls, which is a very popular facing in Europe, as shown in Figure 9-5(c).

Figure 9-5(a). Geotextile temporary wrap-around wall.


FHWA NHI-07-092
Geosynthetic Engineering

9 8

9 MSE Walls
August 2008

Figure 9-5(b). Galvanized, WWM faced geogrid reinforced soil wall.

Figure 9-5(c). A green welded wire faced, geosynthetic reinforced soil wall.
FHWA NHI-07-092
Geosynthetic Engineering

9 9

9 MSE Walls
August 2008

Segmental Precast Concrete Panels, similar to panels used to face metallic MSE systems,
are used to face geosynthetic reinforced MSE walls. However, stiff, polyethylene geogrids
are exclusively used for precast concrete panel faces, where tabs of the geogrid are cast into
the concrete and field-attached to the soil reinforcing geogrid layers. Flexible, polyester
geogrids are not used because casting them into wet concrete would expose the geogrids to a
high-alkaline environment.
Full-Height Concrete Panels are also used to face geosynthetic MSE walls. These are used
in only a few states, and only where aesthetics of full-height panels are specifically desired.
Similar to the segmental precast concrete panels, stiff, polyethylene geogrids are exclusively
used for precast concrete panel faces where tabs are cast into the panel.
Timber facings are commonly used for geosynthetic MSE walls. Timber-faced walls are
normally used for low- to moderate-height structures, landscaping, or maintenance
construction. Geotextiles and geogrids are used with timber facings.

9.4 DESIGN GUIDELINES FOR MSE WALLS


9.4-1 Approaches and Models
A number of approaches to geotextile and geogrid reinforced retaining wall design have been
proposed, and these are summarized by Christopher and Holtz (1985), Mitchell and Villet
(1987), Christopher et al. (1989), and Claybourn and Wu (1993). The most commonly used
method is classical Rankine earth pressure theory combined with tensile-resistant tie-backs,
in which the reinforcement extends beyond an assumed Rankine failure plane. Figure 9-6
shows a MBW unit-faced system and the model typically analyzed. Because this design
approach was first proposed by Steward et al. (1977) of the U.S. Forest Service, it is often
referred to as the Forest Service or tie-back wedge method.
The simplified coherent gravity method (Elias et al., 2001 and AASHTO, 2002) is
recommended for internal design of MSE walls for transportation works. This
simplified approach was developed so that iterative design procedures are avoided and by
practical considerations of some of the complex refinements of the available methods (i.e.,
the coherent gravity method (AASHTO, 1996, with 1997 interims) and the structural
stiffness method {Christopher et al., 1989}). For geosynthetics, it is essentially the same as
the tie-back wedge method with a few minor changes as noted in the following paragraphs.

FHWA NHI-07-092
Geosynthetic Engineering

9 10

9 MSE Walls
August 2008

Figure 9-6.

Actual geosynthetic reinforced soil wall in contrast to the design model.

The simplified method uses an assumed Rankine failure surface. The lateral earth pressure
coefficient, K, is determined by applying a multiplier to the active earth pressure coefficient.
The active earth pressure coefficient is determined using a Coulomb earth pressure
relationship, assuming no wall friction and a horizontal backslope ( angle equal to zero).
For a vertical wall the earth pressure, therefore, reduces to the Rankine active earth pressure
equation. For wall face batters equal to or greater than 8 degrees, a simplified form of the
Coulomb equation can be used, as discussed in section 9.4-3.
The basic approach for internal stability is a limiting equilibrium analysis, with consideration
of the reinforced soil mass's possible failure modes as given in Table 9-1. These failure
modes are analogous to those of metallic reinforced MSE walls.
As with conventional retaining structures, overall (external) stability and wall settlement
must also be satisfactory. In fact, external stability considerations (i.e., sliding) generally
control the length of the reinforcement required.
A popular design method for modular block wall (MBW) unit faced, geosynthetic reinforced
soil walls is the National Concrete Masonry Association design procedure (NCMA, 1997).
This method is well documented and has an accompanying generic design and analysis
computer program. Although similar in many ways to the FHWA/AASHTO procedure,
FHWA NHI-07-092
Geosynthetic Engineering

9 11

9 MSE Walls
August 2008

there are some significant differences leading to a design that is not in compliance with
FHWA/AASHTO requirements. A consistent approach to design is recommended for
highway works, particularly where vendor designs are used. Therefore, the NCMA (1997)
design method is not recommended for general use in design of transportation works.

Table 9-1
Internal Failure Modes and Required Properties for MSE Walls
Failure Mode for
Geosynthetic Reinforcement

Failure Mode for


Metallic Reinforcement

Property Required

Geogrid or geotextile rupture

Strips or meshes break

Tensile strength

Geogrid or geotextile pullout

Strips or meshes pullout

Soil-reinforcement
interaction (passive
resistance, frictional
resistance)

Excessive creep of geogrid or


geotextile

N/A

Creep resistance

Connection Failure by
rupture of geosynthetic or
pullout from face unit

Strips or meshes break, or


strips or meshes pullout

Tensile failure or pullout

9.4-2 Design Steps


The following is a step-by-step procedure for the design of geosynthetic reinforced walls.
STEP 1.

Establish design limits, scope of project, and external loads (Figure 9-7).

A. Wall height, H
B. Wall length
C. Face batter angle
D. External loads:
1. Temporary concentrated live loads, q
2. Uniform surcharge loads, q
3. Seismic loads, Am
FHWA NHI-07-092
Geosynthetic Engineering

9 12

9 MSE Walls
August 2008

E. Site topography
1. Wall toe slope
2. Wall backfill slope
3. Surface water drainage patterns
F. Type of facing and connections:
1. Modular block wall units, timbers, segmental precast panels, etc.
4. Full-height concrete panels
5. Welded-wire
6. Wrapped
G. Vertical spacing requirements, Si, based on facing connections, stability during
construction, lift thickness, and placement considerations (e.g., maximum s = 18
inches (0.5 m) for geotextile- and geogrid-wrapped faced walls), and
reinforcement strength.
H. Environmental conditions such as frost action, scour, shrinkage and swelling,
drainage, seepage, rainfall runoff, chemical nature of backfill and seepage water
(e.g., pH range, hydrolysis potential, chlorides, sulfates, chemical solvents, diesel
fuel, other hydrocarbons, etc.), etc.
I. Design and service life periods

STEP 2.

Determine engineering properties of foundation soil (Figure 9-7).

A. Determine the soil profile below the base of the wall; exploration depth should be at
least twice the height of the wall or to refusal. Borings should be spaced at least
every 100 to 150 ft (30 to 45 m) along the wall's alignment at the front and at the back
of the reinforced soil section.
B. Determine the foundation soil strength parameters (cu, Nu, c', and N'), unit weight ((),
and consolidation parameters (Cc, Cr, cv and F'p) for each foundation stratum.
C. Establish location of groundwater table. Check criticality of drainage behind and
beneath the wall.

FHWA NHI-07-092
Geosynthetic Engineering

9 13

9 MSE Walls
August 2008

For vertical walls

= 45 +

r l

2
For walls with face batter > 8 or more from the vertical,
tan ( ) =

tan ( ) + tan ( )[tan ( ) + cot ( + 90 )][1 + tan ( + 90 ) cot ( + 90 )]


1 + tan ( + 90 )[tan ( ) + cot ( + 90 )]

where * = $

Figure 9-7.

Geometric and loading characteristics of geosynthetic MSE walls.

FHWA NHI-07-092
Geosynthetic Engineering

9 14

9 MSE Walls
August 2008

STEP 3.

Determine properties of both the reinforced fill and retained backfill soils (see
Section 9.6 for recommended soil fill requirements).

A. Water content, gradation and plasticity. Note that soils with appreciable fines (silts
and clays) are not recommended for MSE walls.
B. Compaction characteristics (maximum dry unit weight, (d, and optimum water
content, wopt, or relative density).
C. Coefficient of permeability, k, to evaluate drainage requirements.
D. Angle of internal friction, N'.
E. pH, oxidation agents, etc. (For a discussion of chemical and biological characteristics
of the backfill that could affect geosynthetic durability, see Section 9.6.)

STEP 4.

Establish design factors of safety (the values below are recommended minimums;
local codes may require greater values) and performance criteria.

A. External stability:
1. Sliding: FS > 1.5
2. Bearing capacity: FS > 2.5
(AASHTO 5.8.3 notes A lessor FS, of 2.0, could be used if justified by a
geotechnical analysis.)
3. Deep-seated (overall) stability: FS > 1.3
4. Settlement: Maximum allowable total and differential based on performance
requirements of the project.
5. Seismic Stability: F.S. > 75% of static F.S. (All failure modes)
B. Internal stability:
1. Determine the allowable long-term tensile strength, Tal, of the reinforcement; see
Appendix H.
Remember to consider connection strength between the
reinforcement and facing element, which may limit the reinforcement's design
tensile strength value.
2. Determine the long-term design strength, Ta, of the reinforcement, where:
Ta = Tal / FS
with a minimum FS against internal stability failure of 1.5 normally used.
FHWA NHI-07-092
Geosynthetic Engineering

9 15

9 MSE Walls
August 2008

3. Pullout resistance: FS > 1.5; minimum embedment length is 3 ft (1 m). Use FS >
1.1 for seismic pullout.

STEP 5.

Determine preliminary wall dimensions.

A. For analyzing a first trial section, assume a reinforced section length of L = 0.7H for a
level backfill and a length of L 0.85H with a sloping backfill.
B. Determine wall embedment depth.
1. Minimum embedment depth, H1, at the front of the wall (Figure 9-7):
Minimum H1
Slope in Front of Wall
horizontal (walls)
H/20
horizontal (abutments)
H/10
3H:1V
H/10
2H:1V
H/7
3H:2V
H/5
A minimum horizontal bench of 4 ft (1.2 m) wide should be provided in front
of walls founded on slopes. In any case, the minimum H1 is 18 in (0.5 m).
2. Consider possible frost action, shrinkage and swelling potential of foundation
soils, and seismic activity. Check bearing capacity and global stability.
Adjust embedment depth and/or bench width as needed.

STEP 6.

Develop the internal and external lateral earth pressure diagrams for the
reinforced section. It is recommended that computations for external stability be
made assuming the reinforced soil mass and facing to be a rigid body, and for
internal stability using the Simplified method (Elias et al., 2001; AASHTO,
2002).

A. Consider the internal stresses from the reinforced soil fill and dead load and live load
surcharges.
B. Consider the external stresses from the retained backfill plus dead load and live load
surcharges. External live loads are ignored when they increase stability and are
applied when they decrease stability.
FHWA NHI-07-092
Geosynthetic Engineering

9 16

9 MSE Walls
August 2008

C. Evaluate seepage forces and determine drainage requirements to assure design


assumptions.
D. Combine earth, surcharge, and live load pressure diagrams into a composite diagram
for the internal and external design.

STEP 7.

Check external wall stability.

A. Sliding resistance. Check with and without surcharge.


B. Bearing capacity of the foundation.
C. Deep-seated (overall) stability.
D. Seismic analysis.

STEP 8.

Settlement Analysis

A. Estimate total, differential along the alignment, and differential from back to front
settlements of the reinforced section using conventional settlement analyses.
B. Compare estimated differential settlement along the wall alignment to distortion
limits of potential facings.

STEP 9.

Calculate the maximum horizontal stress at each level of reinforcement.

A. Determine, at each reinforcement level, the vertical stress distribution due to


reinforced fill weight and the uniform surcharge and resultant external forces.
B. Determine, at each reinforcement level, the additional vertical stress due to any
concentrated surcharges.
C. Calculate the horizontal stresses, Fh, using the lateral earth pressure diagram from
Step 6.

FHWA NHI-07-092
Geosynthetic Engineering

9 17

9 MSE Walls
August 2008

STEP 10. Check internal stability and determine reinforcement requirements.


Use the lateral earth pressure diagrams developed in Step 6 for the reinforced section.
A. Calculate the maximum tension Tmax in each reinforcement layer per unit width of
wall based on the vertical spacing, Sv of reinforcing layers to resist the internal lateral
pressures.
B. Determine the long-term design strength Ta > Tmax, where Ta is equal to the long-term
allowable strength Tal, defined in Appendix H, divided by the FS against internal
failure selected in Step 4.
C. Determine the long-term connection design strength Tac and compare to Ta. If Tac is
lower, then Tac > Tmax. See Section 9.6 for Tac equations. Note that for some systems
(ex., MBW facings) Tac will be a function of overburden pressure and thus varies with
distance down from top of wall.
D. Check the local stability of MBW units, timber, or concrete panels that are used for
the wall facing. If a wrap-around face is used, determine overlap length, Lo, for the
folded portion of the geosynthetic at the face using pullout capacity.
E. Check length of the reinforcement, Le, required to develop pullout resistance beyond
the Rankine failure wedge. A minimum Le = 3 ft (1 m) is recommended.

STEP 11. Design internal and surficial drainage system components.

STEP 12. Prepare plans and specifications.

9.4-3 Comments on the Design Procedure


Again, for additional design details refer to the Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes Design and Construction Guidelines (Elias et al., 2001) and to the
AASHTO Standard Specification for Bridges (2002).

STEPS 1 Establish design limits, scope of project, and external loads.

FHWA NHI-07-092
Geosynthetic Engineering

9 18

9 MSE Walls
August 2008

Step 1 A-E and H-I need no further elaboration, however, there are several considerations
in selecting the facing system (step 1F), several of which will affect the vertical spacing
of reinforcements (Step 1G).

Modular Block Walls


For modular concrete facing blocks (MBW), sufficient inter-unit shear capacity must
be available, and the maximum spacing between reinforcement layers shall be limited
to twice the front to back width, Wu, of the modular concrete facing unit or 32 inches
(0.8 m) whichever is less. The maximum facing height above the uppermost
reinforcement layer and the maximum depth of facing below the bottom
reinforcement layer should be limited to the width, Wu, of the modular concrete
facing unit used.
The inter-unit shear capacity as obtained by testing (ASTM D 6916) at the
appropriate normal load should exceed the horizontal earth pressure at the facing by a
Factor of Safety of 2.
For seismic performance categories "C" or higher (AASHTO Division 1A), facing
connections in modular block faced walls (MBW) shall not be fully dependent on
frictional resistance between the backfill reinforcement and facing blocks. Shear
resisting devices between the facing blocks and soil reinforcement such as shear keys,
pins, etc. shall be used. For connections partially or fully dependent on friction
between the facing blocks and the soil reinforcement, the long-term connection
strength Tac, should be reduced to 80 percent of its static value. Further, the blocks
above the uppermost layer soil reinforcement layer must be secured against toppling
under all seismic events.

Welded Wire Walls (Elias et al., 2001)


Welded wire or similar facing panels shall be designed in a manner which prevents
the occurrence of excessive bulging as backfill behind the facing elements
compresses due to compaction stresses, self weight of the backfill or lack of section
modulus. Bulging at the face between soil reinforcement elements in both the
horizontal and vertical direction should be limited to 1 to 2 in. (25 to 50 mm) as
measured from the theoretical wall line. This may be accomplished by requiring the
placement of a nominal 2 ft (600 mm) deep zone of rockfill or cobbles directly behind
the facing, using baskets with a vertical height 18 in. (450 mm), decreasing the
vertical spacing between continuous reinforcements and vertically and horizontally
for discontinuous reinforcements, increasing the section modulus of the facing
material and by providing sufficient overlap between adjacent facing panels. In
FHWA NHI-07-092
Geosynthetic Engineering

9 19

9 MSE Walls
August 2008

addition, the reinforcements must not be restrained and have the ability to slide
vertically with respect to the facing material. Furthermore, the top of the flexible
facing panel at the top of the wall shall be attached to a soil reinforcement layer to
provide stability to the top facing panel.

Wrapped Faced Walls (after Elias et al., 2001)


As previously indicated, a consistent vertical spacing of reinforcement on the order of
every lift or every other lift should be used with a maximum spacing of 18 in (0.45
m). Geosynthetic facing elements should not be left exposed to sunlight (specifically
ultraviolet radiation) for permanent walls. If geosynthetic facing elements must be
left exposed permanently to sunlight, the geosynthetic shall be stabilized to be
resistant to ultraviolet radiation. Furthermore, product specific test data should be
provided which can be extrapolated to the intended design life and which proves that
the product will be capable of performing as intended in an exposed environment.
Alternately a protective facing shall be constructed in addition (e.g., concrete,
shotcrete, etc.).
STEP 2. needs no elaboration
STEP 3. Determine reinforced fill and retained backfill properties.
Requirements for reinforced fill are presented in Section 9.6-1. There are no specific
requirements for the retained backfill soils stated in either the AASHTO or FHWA
guidelines. However, retained backfill soils generally should meet state agency
embankment fill soil requirements. The engineering properties of these two separate fill
materials has a significant influence on the design of the reinforced soil volume.
The moist unit weights, (m, of the reinforced fill and retained backfill soils can be
determined from the standard Proctor test (AASHTO T-99) or alternatively, from a
vibratory-type relative density test. The angles of internal friction, N', should be
consistent with the respective design value of unit weight. Conservative estimates can be
made for granular materials, or alternatively for major projects, this soil property can be
determined by drained direct shear (ASTM D 3080) or triaxial tests.
Conventional compaction control density measurements should be performed for fills
where a majority of the material passes a -in. (20 mm) sieve. For coarse, gravelly
backfills, use either relative density for compaction control or a method-type compaction
specification for fill placement. The latter is appropriate if the fill contains more than
30% of -in. (20 mm) or larger materials.

FHWA NHI-07-092
Geosynthetic Engineering

9 20

9 MSE Walls
August 2008

STEP 4. Establish design factors of safety.


Minimum recommended factors of safety for the various potential internal, external, and
global failures are based on AASHTO and FHWA guidelines. Determine if values higher
than the minimums should be used for a particular project or structure or agency.
STEP 5. Determine preliminary wall dimensions, including wall embedment depth.
Since the design process is trial and error, it is necessary to initially analyze a set of
assumed trial wall dimensions. The recommended minimum value of L > 0.7H is a good
place to start. Surcharges and sloping fills will likely increase the reinforcement length
requirements. For low walls a minimum length reinforcement length of 8 ft (2.4 m) is
required. This minimum length shall provide sufficient depth for placement and
compaction of the wall fill behind low walls. AASHTO LRFD Bridge Design
Specifications (2007) also states a soil reinforcement of L > 0.7 H, and notes the
minimum requirement in accompanying commentary as follows. In general, a minimum
reinforcement length of 8.0 ft., regardless of wall height, has been recommended based
on historical practice, primarily due to size limitations of conventional spreading and
compaction equipment. Shorter minimum reinforcement lengths, on the order of 6.0 ft.,
but no less than 70 percent of the wall height, can be considered if smaller compaction
equipment is used, facing panel alignment can be maintained, and minimum requirements
for wall external stability are met.
For walls founded on rock or competent foundation soil (foundation materials which will
exhibit minimal post construction settlements), FHWA NHI-00-043 (Elias et al., 2001)
allows shorter reinforcements to be used at the base (minimum base length Lbase 0.4H
8 ft {2.4 m}), provided compensating lengths are added in the central and upper portions
of the wall. Special design procedures are required as covered in FHWA NHI-00-043.
Minimum base width requirements and design guidelines for back-to-back walls are also
covered in FHWA NHI-00-043.
Unless the foundation is on rock, a minimum embedment depth is required to provide
adequate bearing capacity and to provide for environmental considerations such as frost
action, shrinkage and swelling clays, or earthquakes. The recommendations given earlier
under Step 5 are conservative. Frost or moisture sensitive soils could always be removed
and replaced to reduce embedment requirements.
Embedment of the wall also helps resist the lateral earth pressure exerted by the
reinforced fill through passive resistance at the toe. This resistance is neglected for
design purposes because it may not always be there. Construction sequence, possible
scour, or future excavation at the front of the wall may eliminate it.
FHWA NHI-07-092
Geosynthetic Engineering

9 21

9 MSE Walls
August 2008

STEP 6.

Develop lateral earth pressure diagrams for both the reinforced fill and retained
backfill.

A. Determine the internal lateral stresses H at any level z from the weight of the
reinforced fill (z using the properties as determined in Step 3, plus any uniform
surcharges q, concentrated surcharges )v, live loads )q or any concentrated stresses
from horizontal surcharges )h.
H = K v + h
where: K = 1.0 x Ka = Ka for geosynthetics; and
v = (r z + q + )q + )v
Various approaches for considering the lateral earth pressures due to distributed
surcharges, concentrated surcharges, and live loads are discussed by Christopher and
Holtz (1985), Christopher et al. (1989), and Elias et al. (2001). Terzaghi and Peck
(1967), Wu (1975), Perloff and Baron (1976), the U.S. Forest Service (Steward, et al.,
1977), Simac et al. (1993), and the U.S. Navy DM 7.01 (1986) also provide suitable
methods.
The increment of vertical stress due to concentrated vertical loads v is evaluated
using a 2V:1H pyramidal distribution by AASHTO (2002) and FHWA (Elias et al.,
2001).
For calculating the vertical stress in the reinforced section of walls supporting a
backslope on angle $, use:

v = r z + 0.5L(tan ) r
For internal stability, the active earth pressure coefficient, Ka, should be determined
using the Coulomb method, but assuming no wall friction and that the backslope
angle, , is always equal to zero. Thus, for a near-vertical face batter, the Coulomb
equation simplifies mathematically to the simplest form of the Rankine equation:
Ka = tan2 (45E - N/2)

[9-1]

FHWA (Elias et al., 2001) recommends that for walls with face batter > 8 the
following simplified form of the Coulomb equation be used:
FHWA NHI-07-092
Geosynthetic Engineering

9 22

9 MSE Walls
August 2008

KA =

sin 2 ( + )
sin
sin 1 +
sin

[9-2]

where: 2 is the face inclination clockwise from the horizontal (see Figure 9-7).
AASHTO 5.8.4.1 (2002) notes that the simplied Coulomb equation can be used if
the wall face is battered.
Determine the appropriate lateral earth pressure distribution diagram for the design
height of the retaining wall.
In conventional retaining wall design, active earth pressure conditions (earth pressure
coefficient = Ka) are normally assumed. There may be some situations, however,
where the wall is prevented from moving (examples include abutments of rigid frame
bridges; walls on bedrock), and at-rest earth pressure conditions (Ko), or even greater
pressures due to compaction, are appropriate.
Ko may be estimated from the Jaky (1948) relationship:
Ko = 1 - sin Ncv
where Ncv =

[9-3]

constant volume friction angle.

B. Consider the external lateral stresses from the retained fill plus any distributed,
concentrated surcharges, or live loads.
Using the retained backfill properties as determined in Step 3, calculate the lateral
earth pressure coefficient and develop the external stability lateral earth pressure
diagram for the wall. This pressure acts along the height, measured from bottom of
wall to top of finished grade, at a vertical line at the back of the reinforced soil mass.
The lateral earth pressure coefficient, Ka, for external stability may be computed with
the Rankine earth pressure equation (9-1) if the backslope angle, $, is equal to zero
and the face batter is near-vertical. The following equation should be used for
sloping fill surcharges on walls with near-vertical face batters:
FHWA NHI-07-092
Geosynthetic Engineering

9 23

9 MSE Walls
August 2008

K a = cos

cos cos 2 cos

[9-4]

cos + cos cos


2

FHWA (Elias et al., 2001) recommends that for face batters > 8E, the coefficient of
earth pressure for external stability can be calculated with the general Coulomb case
as:
Ka =

[9-5]

sin 2 ( + )

sin ( + )sin ( )
sin sin ( )1 +

sin ( )sin ( + )

where 2 is the face inclination from horizontal, and $ is the surcharge slope angle.
The wall friction angle is assumed to be equal to a maximum of $, but < 2/3 N.
AASHTO 5.8.2 (2002) notes that the general Coulomb equation should be used with
no minimum batter value, and with equal to $.
See AASHTO (2002) or FHWA (Elias et al., 2001) guidelines for broken back slope
surcharge conditions.
C. Evaluate seepage forces and determine drainage requirements to assure design
assumptions.
The design procedures and equations presented within this chapter are based upon the
assumption that the wall soils are drained, and remain drained throughout the service
life of the structure. Therefore potential subsurface and surface water sources must
be evaluated and drainage features accordingly designed in Step 10 to assure that
these design assumptions are correct as discussed in Section 9.4-4.
D. Develop the composite pressure diagram:
The earth pressure and surcharge pressure diagrams are combined to develop a
composite earth pressure diagram that is used for design. See the standard references
for procedures on locating the resultant forces.

FHWA NHI-07-092
Geosynthetic Engineering

9 24

9 MSE Walls
August 2008

STEP 7.

Check external wall stability.

As with conventional retaining wall design, the overall stability of a geosynthetic MSE
wall must be satisfactory. External stability failure modes of sliding and bearing capacity
are evaluated by assuming that the reinforced soil mass acts as a rigid body, although in
reality the wall system is really quite flexible. It must resist the earth pressure imposed
by the backfill which is retained by the reinforced mass and any surcharge loads.
Potential external modes of failure to be considered are:
sliding of the wall;
limiting the location of the resultant of all forces (eccentricity);
bearing capacity of the wall foundation; and
stability of the slope created by the wall (both external and compound failure planes
see Chapter 8).
These failure modes and methods of design against them are discussed by Christopher
and Holtz (1985, 1989), Mitchell and Villett (1987), Christopher et al. (1989), and Elias
et al. (2001).
The potential for sliding along the base is checked by equating the external horizontal
forces with the shear stress at the base of the wall. Sliding must be evaluated with respect
to the minimum frictional resistance provided by either the reinforced soil, Nr, the
foundation soil, Nf, or the soil-reinforcement friction angle Nsg, as measured by interface
shear tests. Often, external stability, particularly sliding, controls the length of
reinforcement required. Reinforcement layers at the base of the wall are considerably
longer than required by internal earth pressure considerations alone. Generally,
reinforcement layers of the same length are used throughout the entire height of the wall.
The factor of safety against sliding should be at least 1.5.
Design for bearing capacity follows the same procedures as those outlined for an ordinary
shallow foundation. The entire reinforced soil mass is assumed to act as a footing.
Because there is a horizontal earth pressure component in addition to the vertical
gravitational component, the resultant is inclined and should pass through the middle
third of the foundation to insure there is no uplift (tension) in the base of this assumed
rigid mass. Provided the resultant location meets this criterion, an overturning stability
analysis is not necessary.

FHWA NHI-07-092
Geosynthetic Engineering

9 25

9 MSE Walls
August 2008

The bearing pressure (acting upon the foundation soil) shall be computed using a
Meyerhof-type distribution, that results in a uniform base pressure distribution over an
effective base width. The effective base width, B, is equal to:
B = L 2e
where:

[9-6]

L = the reinforcement length; and


e = the eccentricity of the resultant vertical forces at the base of the wall.

Allowable bearing pressure (what the foundation soil can support) or appropriate bearing
capacity factors must be used as in conventional geotechnical practice. The width of the
footing for ultimate bearing capacity calculations is equal to the length of the
reinforcement at the foundation level (AASHTO, 2002; Elias et al., 2001). The ultimate
bearing capacity, qult, is determined using classical soil mechanics:
qult = cf Nc + 0.5 r L N

[9-7]

For relatively thick MSE facing elements (e.g., large MBW units) it may be desirable to
include the facing dimensions and weight in bearing capacity calculations (AASHTO,
2002).
Note that ground factors (see Munfakh et al., 2001 or Samtani and Nowatzki, 2006) have
to be added to Equation 9-7 for conditions where a wall is founded upon a slope. The
dimensionless bearing capacity coefficients Nc and N, and ground factors can be readily
obtained from AASHTO (2002) or foundation engineering textbooks.
Due to the flexibility of MSE walls, the factor of safety for bearing capacity is lower than
normally used for stiffer reinforced concrete cantilever or gravity structures. Generally,
the factor of safety must be at least 2.5 with respect to the ultimate bearing capacity. A
lessor factor of safety of 2.0 could be used if justified by a geotechnical analysis that
calculates settlement and determines it to be acceptable (Elias et al. 2001). Use of the
lower factor of safety should be supported by both undrained and drained (effective
stress) parameters for cohesive soils, to permit evaluation of both long-term and shortterm conditions and settlement calculations should be based on consolidation tests on
cohesive soils or modulus determinations from appropriate field tests (e.g.,
pressuremeter, dilatometer, etc.) for granular soils.

FHWA NHI-07-092
Geosynthetic Engineering

9 26

9 MSE Walls
August 2008

The potential for local shear in the foundation soil should be checked if the MSE wall is
constructed on or adjacent to weak soils. To prevent large horizontal movements of the
wall structure on weak cohesive soils:
H < 3 c

[9-8]

where:

H
c

=
=
=

unit weight of wall fill soil


wall height
undrained shear strength of soft soil beneath the wall

Ground improvement of the foundation soils is likely required if adequate support


conditions are not available.
Other foundation design considerations include environmental factors such as frost
action, drainage, shrinkage or swelling of the foundation soils, and potential seismic
activity at the site. Each of these items must be checked to ensure adequate wall
performance is maintained throughout the wall's design and service life.
Overall slope stability typically requires a factor of safety of at least 1.3 for long-term
conditions. Note that the reinforced mass should not be considered as a rigid body for
overall slope stability analyses. Slope stability analysis methods that model the
reinforced fill and reinforcement as discrete elements should be used, as presented in
Chapter 8.
In seismically active areas, the reinforced wall and facing system, if any, must be stable
during earthquakes. Seismic stability is discussed in Section 9.4-4.

STEP 8.

Estimate settlement of the reinforced section.

Potential settlement of the wall structure should be assessed and conventional settlement
analyses for shallow foundations carried out to ensure that immediate, consolidation, and
secondary settlements of the wall are less than the performance requirements of the
project, if appropriate. Both total and differential settlements along the wall length
should be considered. For specific procedures, consult standard textbooks on foundation
engineering.

FHWA NHI-07-092
Geosynthetic Engineering

9 27

9 MSE Walls
August 2008

The amount of total and of differential settlements will affect the aesthetics of the
completed wall structure. The amount of anticipated differential settlement must be
considered when evaluating wall facing options. Tolerable differential settlement for
MBW units is on the order of 1 in 200. Geosynthetic wrapped-face and welded-wire
faced walls are much more deformable and can tolerate significant differential settlement
(on the order of 1 in 50, or greater). Walls with full height precast concrete panels should
be limited to differential settlements of 1 in 500. The limitations of differential
settlement for segmental precast concrete panel faced walls varies with the width of the
joint between panels, and ranges from 1 in 100 to 1 in 300 for typical joint widths (Elias
et al., 2001).

STEP 9. Calculate the maximum horizontal stress at each level of reinforcement.


Calculate, at each reinforcement level, the horizontal stress, Fh, along the potential active
earth pressure failure surface, as shown inclined at the angle R in Figure 9-7. From
Rankine earth pressure theory, R is inclined at 45 + Nr'/2 for a vertical wall, where Nr' is
the internal friction angle appropriate for the reinforced soil section. This Rankine failure
line should be used for geosynthetic reinforced walls with a face batter of less than 8
(FHWA (Elias et al., 2001)) or less than 10 (AASHTO, 2002). A failure wedge defined
with Coulomb earth pressure theory should be assumed for walls with greater batter. Use
the moist unit weight of the reinforced backfill plus, if present, uniform and concentrated
surcharge loads.
Use Ka and the lateral earth pressure diagram, as discussed in Step 6.
STEP 10. Check internal stability and determine reinforcement requirements.
Use the lateral earth pressure diagrams developed in Step 6 for the reinforced section.
A. Determine vertical spacings, Sv, of the geosynthetic reinforcing layers and the
required strength of the reinforcement, Tmax, required at each level to resist the
internal lateral pressures.
The required tensile strength, Tmax, of the geosynthetic is controlled by the vertical
spacing of the layers of the reinforcing, and it is obtained from:
Tmax = Sv FH

FHWA NHI-07-092
Geosynthetic Engineering

9 28

[9-9]

9 MSE Walls
August 2008

where:
=
Sv
FH

(distance to reinforcing layer above + distance to reinforcing layer


below)
horizontal earth pressure at middle of the layer

Vertical spacings should be based on multiples of the compacted fill lift thickness.
From Equation 9-9, it is obvious that greater vertical spacing between the horizontal
layers is possible if stronger geosynthetics are used. (NOTE: Vertical spacing may
be governed by the connection strength between the reinforcement and facing.) This
may reduce the cost of the reinforcement, as well as increase the fill placement rate to
some extent. Typical reinforcement spacing for MSE walls varies between 8 to 32 in.
(200 mm to 0.8 m) for geogrids and rigid facings, and between 8 to 12 in. (200 to 300
mm) for geosynthetic wrap walls. For MBW-faced wall, the maximum vertical
reinforcement spacing should be limited to twice the width (front to back) of MBW
units or 32 in. (0.8 m), whichever is less (AASHTO, 2002). Elias et al. (2001) notes
that the block-width reinforcement vertical spacing limitation is applicable to MBW
units deriving their connection capacity by friction.
B. Determine the length, Le, of geosynthetic reinforcement required to develop pullout
resistance beyond the Rankine (or Coulomb, for walls with batter equal to or greater
than 8 or 10) failure wedge.
This design step is necessary to calculate embedment length, Le, behind the assumed
failure plane (Figure 9-7). The angle of the assumed failure plane is taken to be the
Rankine failure angle (i.e., 45E + Nr/2), or Coulomb failure angle for walls with
batters equal to or greater than 8 (Elias et al., 2001) or 10 (AASHTO, 2002). Also,
this plane is usually assumed to initiate from the toe of the wall (At the back or soil
side of the facing) and proceed upward at that angle. This assumption results in
conservative embedment lengths. The formula for the embedment length, Le, is:
Le

where:
Ti
(r
z
Rc
FHWA NHI-07-092
Geosynthetic Engineering

Ti
(FS )
2 r zRc F

[9-10]

= computed tensile load in the geosynthetic;


= unit weight of backfill (reinforced section);
= depth of the layer being designed;
= reinforcement coverage ratio;

9 29

9 MSE Walls
August 2008

F*
"
FS

= coefficient of pullout interaction between soil and geosynthetic;


= scale correction factor;
= factor of safety against pullout failure.

For preliminary design in absence of specific geosynthetic test data, and for standard
backfill materials, with the exception of uniform sands (i.e., coefficient of uniformity,
Cu < 4), it is acceptable to use conservative default values for F* and as shown in
Table 9-2. The soil friction angle is normally established by testing, though a lower
bound value of 28 degrees is often used.

Table 9-2
Default Values For F* and " Pullout Factors
Reinforcement Type

Default F*

Default "

Geogrid

0.67 tan N

0.8

Geotextile

0.67 tan N

0.6

For wrap-around walls, the overlap length, Lo, must be long enough to transfer
stresses from the lower portion to the longer layer above it. The equation for
geosynthetic overlap length, Lo, is:
Lo =

Ti
(FS )
2 r zRc F

[9-11]

Again, a minimum value of approximately 3 ft (1 m) is recommended for Lo to insure


adequate anchorage of reinforcement layers.
STEP 11.

Design internal and surficial drainage system components.

Drainage requirements are discussed in the next section, Section 9-4.


STEP 12.

Prepare plans and specifications.

Specifications are discussed in Section 9.9.

FHWA NHI-07-092
Geosynthetic Engineering

9 30

9 MSE Walls
August 2008

9.4-4 Drainage
The design procedures and equations presented within this chapter are based upon the
assumption that the wall soils are drained, and remain drained throughout the service life of
the structure. Thus, wall design, detailing, specification, and construction must address
potential sources of subsurface water and potential effect on stability of the wall structure.
Potential sources of subsurface water are groundwater and surface water infiltration.
Subsurface water in the foundation soils will decrease allowable bearing pressures and may
increase anticipated wall settlements. A phreatic surface in the retained backfill will act as
destabilizing force and decrease the external stability of the MSE structure. While a phreatic
surface within the reinforced wall fill increase lateral pressures and decreases the internal
stability of a wall structure. Water moving through an MSE wall can also pipe, or erode,
soils from one zone to another and create paths for additional water flow and soil erosion; if
filtration criteria between zones is not addressed in design.
The primary component of an MSE wall is soil. Water has a profound effect on this primary
component of soil, as it can both decrease the soil shear strength (i.e., resistance) and
increase destabilizing forces (i.e., load). Thus, the authors recommend that drainage
features be required in all walls unless the engineer determines such feature is, or
features are, not required for a specific project or structure. For example, drainage is
usually not required when free draining reinforced fill (i.e., less than 3 to 5% non plastic
fines) is used, however, situations were the wall is influenced by tide or stream fluctuations
could require the use of rapidly draining backfill such as shot rock or open graded coarse
gravel for reinforced fill (i.e., AASHTO No. 57 stone) (Elias et al., 2001). Drains are also
highly recommended for hillside and cut construction to collect and divert groundwater,
including perched water, from the reinforced soil mass. Typical drainage features may
include a drainage blanket along side hill backcuts, a drain and pipe collection/discharge
system at the reinforced wall fill and retained backfill interface, a drainage blanket beneath
the reinforced wall fill, and a column of gravel fill behind the MSE facing (for collection and
discharge of surface water infiltration). Example details are shown in Figures 9-8 and 9-9.
Pavement structures above walls should, when possible, be sloped and positively drained
away from the wall. When surface drainage cannot be accommodated and high infiltration is
anticipated (ex., snow melt) a geomembrane can be placed beneath the pavement system at
the top of the wall and sloped to a back drain.
MSE walls can be designed for water loads, if needed. Standard soil mechanics principles
should be used to determine the effect of phreatic surface on wall loads. FHWA NHI-00-043
(Elias et al., 2001) requires that a minimum differential hydrostatic pressure equal to 3.3 ft
FHWA NHI-07-092
Geosynthetic Engineering

9 31

9 MSE Walls
August 2008

(1.0 m) be applied to all MSE walls along rivers and canals and that the effective unit
weights be used for internal and external stability calculations at levels just below the
equivalent surface of the pressure head line. Additional hydrostatic head should be applied
where water level fluctuations are anticipated unless rapidly drainage backfill is used as
discussed in the previous paragraph. A design guide with drainage design/detailing and with
equations for computing loads and resistance with water loads has been developed for MBW
unit faced MSE walls by NCMA (Collin et al., 2002).

9.4-5 Seismic Design (Allen and Holtz, 1991)


In seismically active areas, an analysis of the geosynthetic MSE wall stability under seismic
conditions must be performed. Seismic analyses can range from a simple pseudo-static
analysis to a complete dynamic soil-structure interaction analysis such as might be performed
on earth dams and other critical structures.
The generally conservative pseudo-static Mononabe-Okabe analysis is recommended for
geosynthetic MSE walls and ground accelerations < 0.29g, in the AASHTO (2002) and
FHWA (Elias et al., 2001) guidelines. This analysis correctly includes the horizontal inertial
forces for internal seismic resistance, as well as the pseudo-static thrust imposed by the
retained fill on the reinforced section. A detailed lateral deformation analysis is
recommended when anticipated ground accelerations are greater than 0.29 g (AASHTO,
2002).

Figure 9-8.

Example MSE wall drainage blanket detail (Elias et al., 2001).

FHWA NHI-07-092
Geosynthetic Engineering

9 32

9 MSE Walls
August 2008

Figure 9-9.

Drainage details for MBW faced, MSE wall (NCMA, 1997).

Because of their inherent flexibility, properly designed and constructed geosynthetic walls
are probably better able to resist seismic loadings (than other types of walls), but high walls
in earthquake-prone regions should be checked. The facing connections must also resist the
inertial force of the wall fascia that can occur during the design seismic event. Stress buildFHWA NHI-07-092
Geosynthetic Engineering

9 33

9 MSE Walls
August 2008

up behind the face, resulting from strain incompatibility between a relatively stiff facing
system and the extensible geosynthetic reinforcement must also be resisted by facing
connections. Additional research is needed to evaluate the effect of seismic forces on
geosynthetic walls with stiff facings.

9.5

LATERAL DISPLACEMENT

Lateral displacement of the wall face occurs primarily during construction, although some
also can occur due to post construction surcharge loads. Post-construction deformations can
also occur due to structure settlement. As noted by Christopher et al. (1989), there is no
standard method for evaluating the overall lateral displacement of reinforced soil walls.
The major factors influencing lateral displacements during construction include compaction
intensity, reinforcement to soil stiffness ratio (i.e., the modulus and the area of reinforcement
as compared to the modulus and area of the soil), reinforcement length, slack in
reinforcement connections at the wall face, and deformability of the facing system. An
empirical relationship for estimating relative lateral displacements during construction of
walls with granular backfills is presented in AASHTO (2002) and FHWA (Elias et al., 2001)
guidelines. The relationship was developed from finite element analyses, small-scale model
tests, and very limited field evidence from 20 ft (6 m) high test walls. Note that as L/H
decreases, the lateral deformation increases. The procedure has been found to provide good
predictions of wall face movement on several monitored MSE wall projects including wall
face movements of a 41 ft (12.6 m) high geotextile wall where predicted movements were
slightly greater than those observed (Holtz et al., 1991).
Two major factors influencing lateral displacements compaction intensity and slack in the
reinforcement at the wall face are contractor controlled. Therefore, geosynthetic MSE wall
construction specifications should state acceptable horizontal and vertical erected face
tolerances.

9.6

MATERIAL PROPERTIES

9.6-1 Reinforced Wall Fill Soil


Gradation: All soil fill material used in the structure volume shall be reasonably free from
organic or other deleterious materials and shall conform to the limits presented in Table 9-3.

FHWA NHI-07-092
Geosynthetic Engineering

9 34

9 MSE Walls
August 2008

Table 9-3
MSE Soil Fill Requirements
Sieve Size

Percent Passing

-inch1 (19 mm1)


No. 40 (0.425 mm)
No. 200 (0.075 mm)

100
0 60
0 - 15

Plasticity Index (PI) < 6 (AASHTO T-90)


Soundness: magnesium sulfate soundness loss < 30% after 4 cycles
NOTE:
1.
The maximum size can be increased up to 2 in. (100 mm), provided tests
have been or will be performed to evaluate geosynthetic strength reduction due
to installation damage (see Appendix H).

Chemical Composition (Elias et al., 2001; Elias, 2000): The chemical composition of the
fill and retained soils should be assessed for effect on durability of reinforcement (pH,
oxidation agents, etc.). Some soil environments posing potential concern when using
geosynthetics are listed in Appendix H. It is recommended that application of polyester
based geosynthetics be limited to soils with a pH range between 3 and 9. Polyolefin based
geosynthetics (i.e., polyethylene and polypropylene) should be limited to use with soils of pH
>3.
Compaction (Elias et al., 2001): A minimum density of 95 percent of AASHTO T-99
maximum value is recommended for retaining walls, and 100 percent of T-99 is
recommended for abutments and walls supporting structural foundations. Soil fill shall be
placed and compacted at or within + or - 2 percentage points of optimum moisture content,
wopt, according to AASHTO T-99. If the reinforced fill is free draining with less than 5
percent passing a No. 200 (0.075 mm) sieve, water content of the fill may be within + or - 3
percentage points of the optimum. Compacted lift height of 8 to 12 in. (200 to 300 mm) is
recommended for granular soils.
A small single or double drum walk-behind vibratory roller or vibratory plate compactor
should be used within 3 feet (1 m) of the wall face. Within this 3-foot (1 m) zone, quality
control should be maintained by a methods specification, such as three passes of light drum
compactor at a maximum lift height. Construction of a test pad to demonstrate proposed
methods will achieve the minimum required soil density (95% or 100% of T-99) to determine
the maximum lift height is recommended. Consideration should be given to reduced lift
FHWA NHI-07-092
Geosynthetic Engineering

9 35

9 MSE Walls
August 2008

thickness requirements near the face to decrease the required compactive effort. Note that
excessive compactive effort or use of too heavy of equipment near the face could result in
excessive face movement, and overstressing of reinforcement layers.
Compaction control testing of the reinforced backfill should be performed on a regular basis
during the entire construction project. A minimum frequency of one test with the reinforced
zone per every 5 feet (1.5 m) of wall height for every 100 feet (30 m) of wall is
recommended. Inconsistent compaction and undercompaction caused by insufficient
compactive effort will lead to gross misalignments and settlement problems, and should not
be permitted.

Shear Strength: Peak shear strength parameters should be used in the analysis (Christopher,
et al., 1989). Parameters should be determined using direct shear and triaxial tests.
Shear strength testing is recommended. However, use of assumed shear values based on
Agency guidelines and experience may be acceptable for some projects. Verification of site
soil type(s) should be completed following excavation or identification of borrow pit, as
applicable.

Unit Weights: Dry unit weight for compaction control, moist unit weight for analyses, and
saturated unit weight for analyses (where applicable) should be determined for the fill soil.
The unit weight value of should be consistent with the design angle of internal friction, N.
9.6-2 Geosynthetic Reinforcement
Geosynthetic reinforcement systems consist of geogrid or geotextile materials arranged in
horizontal planes in the backfill to resist outward movement of the reinforced soil mass.
Geogrids transfer stress to the soil through passive soil resistance on grid transverse members
and through friction between the soil and the geogrid's horizontal surfaces (Mitchell and
Villet, 1987). Geotextiles transfer stress to the soil through friction. Geosynthetic design
strength must be determined by testing and analysis methods that account for the long-term
geosynthetic-soil stress transfer and durability of the full geosynthetic structure. Long-term
soil stress transfer is characterized by the geosynthetic's ability to sustain long-term load inservice without excessive creep strains. Durability factors include site damage, chemical
degradation, and biological degradation. These factors may cause deterioration of either the
geosynthetic's tensile elements or the geosynthetic structure's geosynthetic/soil stress transfer
mechanism.
An inherent advantage of geosynthetics is their longevity in fairly aggressive soil conditions.
The anticipated half-life of some geosynthetics in normal soil environments is in excess of
FHWA NHI-07-092
Geosynthetic Engineering

9 36

9 MSE Walls
August 2008

1,000 years. However, as with steel reinforcements, strength characteristics must be adjusted
to account for potential degradation in the specific environmental conditions, even in
relatively neutral soils. Questionable soil environments are listed in Appendix H.

Allowable Tensile Strength: Allowable tensile strength (Ta) of the geosynthetic shall be
determined using a partial factor of safety approach (Bonaparte and Berg, 1987). Reduction
factors are used to account for installation damage, chemical and biological conditions and to
control potential creep deformation of the polymer. Where applicable, a reduction is also
applied for seams and connections. The total reduction factor is based upon the
mathematical product of these factors. The long-term tensile strength, Tal, thus can be
obtained from:
Tal =

Tult
RF

[9-12]

with RF equal to the product of all applicable reduction factors:


RF = RFCR RFID RFD

[9-13]

where:
Tal
Tult
RFCR
RFID
RFD

= long-term tensile strength,(lb/ft {kN/m});


= ultimate geosynthetic tensile strength, based upon MARV, (lb/ft
{kN/m});
= creep reduction factor, ratio of Tult to creep-limiting strength,
(dimensionless);
= installation damage reduction factor, (dimensionless); and
= durability reduction factor for chemical and biological degradation,
(dimensionless).

RF values for durable geosynthetics in non-aggressive, granular soil environments range


from 3 to 7. Appendix H suggests that a default value RF = 7 may be used for routine, noncritical structures which meet the soil, geosynthetic and structural limitations listed in the
appendix. However, as indicated by the range of RF values, there is a potential to
significantly reducing the reinforcing requirements and the corresponding cost of the
structure by obtaining a reduced RF from test data.
The procedure presented above and detailed in Appendix H is derived from Elias et al.
(2001), Elias and Christopher (1997), Berg (1993), the Task Force 27 (AASHTO, 1990)
guidelines for geosynthetic reinforced soil retaining walls, the Geosynthetic Research
Institute's Methods GG4a and GG4b - Standard Practice for Determination of the Long Term
FHWA NHI-07-092
Geosynthetic Engineering

9 37

9 MSE Walls
August 2008

Design Strength of Geogrids (1990, 1991), and the Geosynthetic Research Institute's Method
GT7 - Standard Practice for Determination of the Long Term Design Strength of Geotextiles
(1992).
Additionally, the following factors should be considered. The long-term strength determined
by dividing the ultimate strength by RF does not include an overall factor of safety to account
for variation from design assumptions (e.g., heavier loads than assumed, construction
placement, fill consistency, etc.). A safety factor is applied to the reinforcement when
designing MSE structures to quantify a safe allowable strength. Thus the allowable strength
of a geosynthetic for MSE applications can be defined as:
Ta =

Tal
FS

[9-14]

where:
Ta
FS

=
=

allowable geosynthetic tensile strength,(lb/ft {kN/m}); and


overall factor of safety to account for uncertainties in the geometry of
the structure, fill properties, reinforcement properties, and externally
applied loads.

For permanent, MSE wall structures, a minimum factor of safety, FS, of 1.5 is recommended.
Of course, the FS value should be dependent upon the specifics of each project.

Connection Strength: The design (or factored allowable) strength, Ta, may be limited by
the strength of the connection between the reinforcement and wall facing, Tac. There are
three primary types of connections used with geosynthetic reinforced MSE walls. They are:
C (Full) mechanical connection;
C Partial or full frictional connection; and
C Soil embedment (i.e., pullout resistance).
The original mechanical connection is a bodkin connection of polyethylene geogrid soil
reinforcement to a tab of polyethylene geogrid cast into a concrete panel. This is illustrated
in Figure 9-10. A few MBW unit systems have (full) mechanical connections with the
geogrid soil reinforcement; an example is shown in Figure 9-11. Another possible (though
not common) mechanical connection is sewn geotextile seam. Note that a (full) mechanical
connection strength should be independent of overburden pressure. Also note that polyester
geogrids and geotextiles should not be cast into concrete for connections, due to potential
chemical degradation (Elias et al., 2001). Other types of geotextiles also are not cast into
concrete for connections due to fabrication and field connection requirements.

FHWA NHI-07-092
Geosynthetic Engineering

9 38

9 MSE Walls
August 2008

Figure 9-10. Polyethylene geogrid bodkin connection detail.

Figure

9-11.

FHWA NHI-07-092
Geosynthetic Engineering

Example

MBW

mechanical

9 39

connection

(HITEC,

2003).

9 MSE Walls
August 2008

Full Mechanical Connections: For (full) mechanical connections, the maximum connection
strength as developed by testing reduced for long-term environmental aging, creep and
divided by a factor of safety (of at least 1.5 for permanent structures) is:

Tac

Tult CRu
RF D RFCR FS

[9-15]

where:
Tac =
CRu =
FS =
Tult, RFD, RFCR =

allowable, design connection strength


Reduction factor to account for reduced ultimate strength
resulting from the connection
factor of safety against connection failure, minimum of 1.5
as previously defined, see Equation 9-12.

The reduced ultimate connection strength based upon connection/seam strength CRu as
determined from ASTM D 4884 for seams is computed as:
CRu =

where:
Tultc =
Tlot

[9-16]

Tultc
Tlot

connection load per unit reinforcement width which results in rupture


of the reinforcement; and
the ultimate wide width tensile strength (ex., ASTM D 4595) for the
reinforcement material lot used for the connection strength testing.

.
Note the same test method and conditions must be used to define Tultc and Tlot.
Partial or Full Frictional Connections: The typical partial or full frictional connection is a
geogrid or geotextile placed horizontally between stacked MBW units. Compression from
the unit weight of the units and infill in the units above the geosynthetic creates friction,
while the shape of the units and/or inserts, which extend into or through the geosynthetic to
clamp or fasten the material, may act to create a mechanical component. The majority of
MBW unit faced wall systems employ a partial or full frictional connection. Though, as
previously noted, a few use a (full) mechanical connection.
The connection strength for partial or full frictional connections is based on the lower of two
values the pullout capacity of the connection and the long-term rupture strength of the
connection (AASHTO, 2002).

FHWA NHI-07-092
Geosynthetic Engineering

9 40

9 MSE Walls
August 2008

Pullout: For partial or full frictional connections that fail by pullout, the maximum
connection strength as developed by testing reduced for long-term environmental aging and
divided by a factor of safety (of at least 1.5 for permanent structures) is
Tac

[9-17]

Tult x CR s
RFD x FS

where the reduction factor to account for reduced strength due to connection pullout, CRs, is
equal to:
CRs =

where:
Tsc

Tsc
Tlot

[9-18]

peak load per unit reinforcement width in the connection test at a


specified confining pressure where pullout is the mode of failure.

Rupture Failure: For partial or full frictional connections that fail by long-term tensile
rupture, the maximum connection strength should be defined following the laboratory testing
and interpretation procedure defined in Appendix A of FHWA NHI-00-043 (Elias et al.,
2001). The maximum connection strength as developed by testing reduced for long-term
environmental aging and divided by a factor of safety (of at least 1.5 for permanent
structures) for partial or full frictional connections that fail by long-term tensile rupture, is
[9-19]

T x CRcr
Tac ult
RFD x FS

where the connection strength reduction factor resulting from long-term testing, CRcr, is
equal to:
T
[9-20]
CRcr = crc
Tlot
where Tcrc is the extrapolated (75 to 100 year) connection test strength.
Note that the environment at the connection may not be the same as the environment within
the MSE mass. Therefore, the long-term environmental aging factor, RFD, may be
significantly different than that used in computing the allowable reinforcement strength, Ta.

FHWA NHI-07-092
Geosynthetic Engineering

9 41

9 MSE Walls
August 2008

The connection strength as developed above is a function of normal pressure that is


developed by the weight of the units. Thus, it will vary from a minimum in the upper portion
of the structure to a maximum near the bottom of the structure for walls with no batter.
Further, since many MBW walls are constructed with a front batter, the column weight above
the base of the wall or above any other interface may not correspond to the weight of the
facing units above the reference elevation. The concept is known as the hinge height (Simac
et al., 1993). Hence, for walls with a batter (AASHTO, 2002) or with a batter greater than 8
(Elias et al., 2001), the normal stress is limited to the lesser of the hinge height, Hh, or the
height of the wall above the interface. This vertical pressure range should be used in
developing CRcr and CRs.
For wall facings constructed of geosynthetics anchored in soil (e.g., wrapped faced walls,
concrete facings with horizontal bars for wrapping the geosynthetics, gabions and welded
wire), the maximum and design connection strength should be determined following soilinteraction pullout calculation procedures. These procedures follow. Note that plastic ties or
hog rings should be used with gabions and welded wire to provide a positive connection and
maintain the position of the geosynthetic during construction; however, the ties are usually
not considered to provide additional connection strength.

Soil-Reinforcement Interaction: Two types of soil-reinforcement interaction coefficients


or interface shear strengths must be determined for design: pullout coefficient, and interface
friction coefficient (AASHTO, 1990). Pullout coefficients are used in stability analyses to
compute mobilized tensile force at the tail of each reinforcement layer. Interface friction
coefficients are used to check factors of safety against outward sliding of the entire
reinforced mass.
Detailed procedures for quantifying interface friction and pullout interaction properties are
presented in Appendix H. The ultimate pullout resistance, Pr, of the reinforcement per unit
width of reinforcement is given by:
Pr = 2 C F C " C Fv C Rc C Le

[9-21]

where:
Le C 2 = the total surface area per unit width of the reinforcement in the
resistance zone behind the failure surface
Le
= the embedment or adherence length in the resisting zone behind the
failure surface

= the pullout resistance (or friction-bearing-interaction) factor


F
FHWA NHI-07-092
Geosynthetic Engineering

9 42

9 MSE Walls
August 2008

"
Fv
Rc

= a scale effect correction factor


= the effective vertical stress at the soil-reinforcement interfaces
= reinforcement coverage ratio

Default values for F* and " are presented in Table 9-2.

9.7

COST CONSIDERATIONS

FHWA NHI-00-043 (Elias et al., 2001) indicates that the typical reinforcing cost is 20 to 30
percent of the total cost of an MSEW wall and depends on the face construction cost. For
example, at the FHWA-Colorado Department of Highways Glenwood Canyon wrapped
faced geotextile test walls with a shotcrete finish (Bell et al., 1983), the cost of the geotextile
was only about 25% of the wall's total cost. As can be seen in the cost example in the next
section, the geosynthetic soil reinforcement is approximately 15% to 20% of the total inplace cost of highway MSE walls with MBW unit facing and select granular soil fill.
Therefore, some conservatism on geosynthetic strength or on vertical spacing is not
necessarily excessively expensive. A major part of the Glenwood Canyon costs involved the
hauling and placement of backfill, as well as shotcrete facing. In some situations, especially
where contractors are unfamiliar with geosynthetic reinforcement, artificially high unit costs
have been placed on bid items such as the shotcrete facing, which effectively has made the
reinforced soil wall uneconomical. A cost comparison for reinforced versus other types of
retaining walls is presented in Figure 9-12. For low walls, geosynthetics are usually less
expensive than conventional walls and metallic MSE wall systems. At the time of its
construction, the Rainier Avenue wall was (shown on the figure) the highest geotextile wall
ever constructed (Allen, et al., 1992). It was unusually economical, partially because, as a
temporary structure, no special facing was used. Permanent facing on a wall of that height
would have increased its cost by $4.65/ft2 ($50/m2) or more.
Other factors impacting cost comparison include site preparation; facing cost, especially if
precast panels or other special treatments are required; special drainage required behind the
backfill; instrumentation; etc.

FHWA NHI-07-092
Geosynthetic Engineering

9 43

9 MSE Walls
August 2008

Figure 9-12. Cost comparison of reinforced systems (after Koerner et al., 1998).
(1 m = 3.3 ft; 1 m2 = 10.8 ft2)

9.8

COST ESTIMATE EXAMPLES

9.8-1 Geogrid, MBW Unit-Faced Wall


A preliminary cost estimate for an MSE wall is needed to assess its viability on a particular
project. Therefore, a rough design is required to estimate fill and soil reinforcement
quantities. The project's scope is not fully defined, and several assumptions will be required.
STEP 1.

Wall description

The wall will be approximately 650-foot long, and varies in exposed wall height from 13 to
20 feet. A gradual slope of 5H:1V will be above the wall. The wall will have a nominal
(e.g., < 3E) batter. Seismic loading can be ignored.
An MBW unit facing will be specified. The geosynthetic will be a geogrid. Reinforced wall
fill will be imported. Wall fill soils are not aggressive and pose no specific durability
concerns.

FHWA NHI-07-092
Geosynthetic Engineering

9 44

9 MSE Walls
August 2008

STEP 2.

Foundation soil

Wall will be founded on a competent foundation, well above the estimated water table. The
in-situ soils are silty sands, and an effective friction angle of 30 can be assumed for
conceptual design. A series of soil borings along the proposed wall alignment will be
completed prior to final design.

STEP 3.

Reinforced fill and retained backfill properties

A well-graded gravely sand, with in. maximum size, will be specified as wall fill, as it is
locally available at a cost of approximately $5.00 per ton delivered to site. An effective angle
of friction of 34E and a unit weight of 125 lb/ft3 can be assumed. The fill is nonaggressive,
and a minimum durability partial safety factor can be used.
The retained backfill will be on-site silty sand embankment material. An effective angle of
friction of 30E and a unit weight of 120 lb/ft3 can be assumed.

STEP 4.

Establish design factors of safety.

For external stability, use minimums of:


sliding
1.5
bearing capacity
2.5
overall stability
1.3

For internal stability, use minimums of:


FS = 1.5 against reinforcement failure
FS = 1.5 against pullout failure

STEP 5.

Determine preliminary wall dimensions.

Average exposed wall height is approximately 16.5 ft. An embedment depth of 1.5 feet
should be added to the exposed height. Total design height is 18 ft.
Assume L/H ratio of approximately 0.7. Use an L = 0.7 (18 ft) = 12.6 ft, use 13 ft.

FHWA NHI-07-092
Geosynthetic Engineering

9 45

9 MSE Walls
August 2008

STEP 6.

Develop earth pressure diagrams.

FH = Ka Fv
At the base of the wall,
Fv = (r H + 0.5 L (tan $) (b= 125 lb/ft3 (18 ft) + 0.5 (13 ft)(0.2) 125 lb/ft3 = 2,410 lb/ft2
For internal stability Ka = tan2(45 - N/2) = 0.28
For external stability Ka = tan2(45 - N/2) = 0.33

STEP 7.

Check external stability.

By observation and experience (must be checked in final design phase), it is assumed that the
L/H ratio of 0.7 will provide adequate external safety factors for the project conditions.

STEP 8.

Estimate settlement.

Due to sand type foundation and experience in this area, settlement is not a problem.

STEP 9.

Calculate horizontal stress at each layer of reinforcement.

Not required for conceptual design; see next step.


STEP 10. Check internal stability and determine reinforcement requirements.
The maximum lateral stress, FH, to be resisted by the geogrid is at the bottom of the wall and
is equal to:
FH = 0.283 (2,410 lb/ft2) = 682 lb/ft2
Assume 100% geogrid coverage in plan view. Assume a geogrid spacing and calculate Tmax
and Ta per Step 10 (i.e., use a maximum spacing of 2.0 ft to match block height intervals).
Assume vertical spacing of 2.0 ft, which is about one geogrid every three blocks. Therefore,
9 layers of geogrid will be used. The required strength of the lowest geogrid is therefore
equal to:
Tmax = 2.0 ft ( 682 lb/ft2) = 1,256 lb/ft < Ta

FHWA NHI-07-092
Geosynthetic Engineering

9 46

9 MSE Walls
August 2008

A geogrid with a long-term allowable strength of 1,400 lb/ft will be used. Using Ta = 1,400
lb/ft Y Tal = 2,100 lb/ft with FS = 1.5 for the bottom layers of geogrid. The top layers could
be reduced by about . Therefore, use 5 geogrid layers at the bottom at Tal = 2,100 lb/ft and
4 geogrid layers at the top at Tal = 1,400 lb/ft.

STEP 11.

Design internal and surficial drainage system components.

COST ESTIMATE:
Material Costs:
Leveling Pad - 650 ft ($3 / ft) = $2,000
Reinforced wall fill 650 ft (13 ft) (18 ft) (125 lb/ft3) = 9,500 ton
9,500 ton ($5.00 / ton) = $ 47,500 say $49,000 with overfill
Geogrid soil reinforcement 5 layers (13 ft) (650 ft) = 4,700 yd2 of 2,100 lb/ft; and
4 layers (13 ft) (650 ft) = 3,800 yd2 of 1,400 lb/ft
From the range presented in Appendix H, assume material costs, delivered to site, of $
3.00 / yd2 and $ 2.00 / yd2. Therefore, cost is 4,700 yd2 ($3.00 / yd2) + 3,800 yd2 ($2.00 /
yd2) = $21,700
MBW face units From local market, MBW units range in cost from $4.50 to $7.00 / ft2
Assume a cost of $6.50 / ft2
650 ft (18 ft) ($6.50 / ft2) = $64,400
Gravel fill within or behind the MBW units Assume 0.3 m thickness required. Assume a cost of $6.00 per compacted yd3.
650 ft (18 ft) (1 ft) ($6.00 / yd3) = $2,600
Subsurface drain behind the reinforced soil mass, and outlets Assume a cost of $5.00 per lineal foot of wall.
650 ft ($5.00 / lft) = $3,200
Engineering and Testing Costs:
A line-and-grade specification will be used. Based upon previous projects, assume cost of
design engineering, soil testing, and site assistance will be approximately $1 per ft2.
650 m (18 ft) ($1 / ft2) = $11,700

FHWA NHI-07-092
Geosynthetic Engineering

9 47

9 MSE Walls
August 2008

Installation Costs:
Based upon previous bids, assume cost to install will be approximately $5 per ft2
650 ft (18 ft) ($5 / ft2) = $58,500
TOTAL ESTIMATED COST:
Materials + Engineering/Site Assistance + Installation =
($2,000 + $49,000 + $21,700 + $64,400 + $2,600 + $3,200) + $11,700 + $58,500
= $213,500
{Check: This is equal to an installed cost of $18.25 / ft2, which is reasonable based upon past
experience (see Figure 9-12).} Based upon this cost estimate, the geosynthetic MSE wall
option is the most economical for this project. Therefore, it is recommended that final design
proceed using a geosynthetic MSE wall. Note that estimate does not include site preparation,
placement of random backfill, or final completion items (e.g., seeding, railings).

9.8-2 Geotextile Wrap Wall


A preliminary cost estimate for a temporary MSE wall is needed to assess its viability on a
particular project. Therefore, a rough design is required to estimate fill and soil
reinforcement quantities. The project scope is not fully defined, and several assumptions will
be required.

STEP 1.

Wall description.

The temporary wall will be approximately 165 ft long and approximately 26 feet high. A flat
slope and no traffic will be above the wall. Seismic loading can be ignored.
A wrap-around facing will be used and an ultraviolet-stabilized geotextile specified. Thus, a
gunite or other type of protective facing for this temporary structure will not be required.
Wall fill soils are not aggressive and pose no specific durability concerns.

STEP 2.

Foundation soil.

Wall will be founded on a competent foundation that overlies a soft compressible layer of
soil. Details of foundation bearing capacity and global stability do not have to be addressed
for this conceptual cost estimate, but will be addressed in final design. The in-situ soils are
silts, and an effective friction angle of 28E can be assumed for conceptual design.

FHWA NHI-07-092
Geosynthetic Engineering

9 48

9 MSE Walls
August 2008

STEP 3.

Reinforced fill and retained backfill properties.

A well-graded sand will be specified as wall fill, as it is locally available at a cost of


approximately $1.75 per ton delivered to site. An effective angle of friction of 32E and a unit
weight of 125 lb/ft3 can be assumed. The fill is non-aggressive, and a minimum durability
partial safety factor can be used.
The retained backfill will be on site silt embankment material. An effective angle of friction
of 28E and a unit weight of 120 lb/ft3 can be assumed.

STEP 4.

Establish design factors of safety.

For external stability, use minimums of:


sliding
1.5
bearing capacity
2.5
overall stability
1.3
For internal stability, use minimums of:
FS = 1.5 against reinforcement failure
FS = 1.5 against pullout failure

STEP 5.

Determine preliminary wall dimensions.

Average exposed wall height is approximately 26 ft. An embedment distance of 1.5 ft (the
minimum recommended) should be added to the exposed height. Design height is 27.5 ft.
Assume L/H ratio of approximately 0.7. Use an L = 0.7 (27.5 ft ) = 19.25 ft, use 20 ft.

STEP 6.

Develop earth pressure diagrams.

Fv at base of wall = (r H = 125 lb/ft3 (27.5 ft) = 3,438 lb/ft2


For internal stability Ka = tan2(45 - N/2) = 0.307
For external stability Ka = tan2(45 - N/2) = 0.361

FHWA NHI-07-092
Geosynthetic Engineering

9 49

9 MSE Walls
August 2008

STEP 7.

Check external stability.

By observation and experience, it is assumed that the L/H ratio of 0.7 will provide adequate
external safety factors for the project conditions. {This assumption will checked in final
design.} Therefore, use L = 20 ft.

STEP 8.

Estimate settlement.

Again, by observation and experience, settlement is likely not a problem for these project
conditions. Settlement will be quantified during final design.

STEP 9.

Calculate horizontal stress at each layer of reinforcement.

Not required for conceptual design; see next step.

STEP 10. Check internal stability and determine reinforcement requirements.


Lateral load to be resisted by the geotextile is equal to:
Ka ( H2 = (0.307) (125 lb/ft3) (27.5 ft)2 = 14,510 lb/ft
Assuming 100% geotextile coverage in plan view, the geotextiles must safely carry 14,500
lb/ft per unit width of wall. Assume a geotextile with a long-term allowable strength of 1,400
lb/ft will be used. The safe design strength of the geotextile is therefore equal to:
Ta = Tal / FS = 1,400 / 1.5 = 933 lb/ft
The approximate number of geotextile layers required is equal to:
14,500 / 933 = 15.5
Round this number up and add an additional layer for conceptual design to account for
practical layout considerations with final design. Assume a vertical spacing of 20 inches (two
10-inch soil lifts) will be used. Therefore, 16 layers of geotextile will be used, with Ta . 930
lb/ft, or greater.

COST ESTIMATE:
Material Costs: Reinforced wall fill 165 ft (27.5 ft) (20 ft) (125 lb/ft3) = 5,670 tons
5,670 tons ($1.75 / ton) = $9,900

FHWA NHI-07-092
Geosynthetic Engineering

9 50

9 MSE Walls
August 2008

Geotextile soil reinforcement (include face area and wrap-tail length) 16 layers (20 + 1.5 + 5 ft) (165 ft) = 7,800 yd2
From the range presented in Appendix H, assume a material cost, on site, of $1.75/yd2
7,800 yd2 ($1.75 / yd2) = $13,650
Engineering and Testing Costs:
A line-and-grade specification will be used. Based upon previous projects, assume cost of
design engineering, soil testing, and site assistance will be approximately $1 per ft2,
because of the height and relatively low total area of wall that will be constructed.
165 ft (27.5 ft) ($1 / ft2) = $4,500
Installation Costs:
Based upon previous bids, assume cost to install will be approximately $5 per ft2
165 ft (27.5 ft) ($5 / ft2) = $22,700
TOTAL ESTIMATED COST:
Materials + Engineering/Site Assistance + Installation =
$23,550 + $4,500 + $22,700 = $50,750
{Check: This is equal to an installed cost of $11.20 / ft2, which is reasonable based the small
size of this project and upon past experience.} Based upon this cost estimate, the geosynthetic
MSE wall option is the most economical for this project. Therefore, it is recommended that
final design proceed using a geosynthetic MSE wall.

9.9

SPECIFICATIONS

9.9-1 Geosynthetic, MBW Unit-Faced Wall


The following example was obtained from New York DOT (2004). It is a special provision
for materials and construction of a geogrid-MBW unit reinforced soil wall. It is noted that
the MBW unit shall conform to NYDOT Standard Specification (2006) 704-07 Segmental
Retaining Wall Blocks. This material specification follows the special provision.

FHWA NHI-07-092
Geosynthetic Engineering

9 51

9 MSE Walls
August 2008

ITEM 554.0517

MECHANICALLY STABILIZED SEGMENTAL BLOCK


RETAINING WALL SYSTEM (EXTENSIBLE REINFORCEMENT)

DESCRIPTION
Construct a Mechanically Stabilized Segmental Block Retaining Wall System (Extensible
Reinforcement), (MSSBRWS) where indicated on the plans.
A MSSBRWS is a mechanically stabilized earth wall system consisting of an un-reinforced concrete
or compacted granular leveling pad, facing and cap units, backfill, underdrains, geotextiles, and an
extensible reinforcement used to improve the mechanical properties of the backfill.
Only MSSBRWS designer-suppliers (designer-supplier) with facing and cap units appearing on the
Department's Approved List of Products for Precast Concrete Retaining Wall Block will be
acceptable for use under this item. The Approved List of Products is available from the Office of the
Director, Materials Bureau.
Upon award of the contract, notify the Deputy Chief Engineer, Technical Services (DCETS) of the
name and address of the chosen designer-supplier. Once designated, the chosen designer-supplier
shall not be changed without written permission from the DCETS. Obtain all necessary materials
(except backfill, unit fill, leveling pad material, underdrains, geotextiles, and cast-in- place concrete)
from the chosen designer-supplier.
Obtain from the designer-supplier and submit to the DCETS for approval, the MSSBRWS design and
installation procedure. Designer-suppliers must submit and have their design reviewed and approved
for use. All MSSBRWS designs must conform to the requirements of Section 5.8 of the Association
of State Highway and Transportation Officials Standard Specifications (AASHTO). All MSSBRWS
designs must be stamped by a Professional Engineer licensed to practice in New York State. The
DCETS requires 20 working days to approve the submission after receipt of all pertinent information.
Begin work only after receiving DCETS approval.
Submit shop drawings and proposed methods for construction to the Engineer for written approval at
least 30 working days before starting work.
Supply on-site technical assistance from a representative of the designated designer-supplier during
the beginning of installation until such time as the Engineer determines that outside consultation is no
longer required.
Provide the Engineer with two copies of the designated designer-supplier Installation Manual two
weeks before beginning construction.
Other definitions that apply within this specification are:
A.

Leveling Pad

An un-reinforeed concrete or compacted granular fill footing or pad


which serves as a flat surface for placing the initial course of facing
units.

B.

Facing Unit

A segmental precast concrete block unit that incorporates an alignment


and connection device and also forms part of the MSSBRWS face area.
A comer unit is a facing unit having two faces.

FHWA NHI-07-092
Geosynthetic Engineering

9 52

9 MSE Walls
August 2008

C.

Alignment and
Connection Device

Any device that is either built into or specially manufactured for the
facing units, such as shear keys, leading/trailing lips, or pins. The
device is used to provide alignment and maintain positive location for a
facing unit and also provide a means for connecting the extensible
reinforcement.

D.

Extensible
Reinforcement

High density polyethylene, polypropylene or high tenacity polyester


geogrid mats of specified lengths which connect to the facing unit and
are formed by a regular network of integrally connected polymer tensile
elements with apertures of sufficiently large size to allow for
mechanical interlock with the backfill.

E.

Unit Fill

Well-graded aggregate fill placed within and/or contiguous to the back


of the facing unit.

F.

Cap Unit

A segmental precast concrete unit placed on and attached to the top of


the finished MSSBRWS.

G.

Backfill

Material placed and compacted in conjunction with extensible


reinforcement and facing units.

H.

Underdrain

A system for removing water from behind the MSSBRWS.

I.

Geotextile

A permeable textile material used to separate dissimilar granular


materials.

MATERIALS
Not all materials listed are necessarily required for each MSSBRWS. Ensure that the proper
materials are supplied for the chosen system design.
A. Leveling Pad
1. For MSSBRWS greater than or equal to 4.6 meters in total height, as measured from the
top of the leveling pad to the top of the cap unit, supply a leveling pad conforming to the
following:
a. Un-reinforced Concrete:
Supply concrete conforming to Section 501 (Class A Concrete).
2. For MSSBRWS less than 4.6 meters in total height, as measured from the top of the
leveling pad to the top of the cap unit, supply a leveling pad conforming to one of the
following:
a. Un-reinforced Concrete:
Supply concrete conforming to Section 501 (Class A Concrete).
b. Granular Fill
Supply select granular fill conforming to 203-2.02C (Select Granular Fill
and Select Structure Fill).

FHWA NHI-07-092
Geosynthetic Engineering

9 53

9 MSE Walls
August 2008

c. Crushed Stone
Supply crushed stone conforming to Section 623 (Screened Gravel,
Crushed Gravel, Crushed Stone, Crushed Slag), Item 623.12, Crushed
Stone (In-Place Measure). Use an approximate 50-50 mix of Size
Designations 1 and 2.
B.

Facing and Cap Units


Supply units fabricated and conforming to 704-07 (Segmental Retaining Wall Blocks). Notify
the Director, Materials Bureau, of the name and address of the units' fabricator no later than 14
days after contract award.

C. Alignment and Connection Devices


Supply devices conforming to the designer-supplier's Installation Manual.
D. Extensible Reinforcement
Supply reinforcement which has been tested and certified to meet the minimum requirements for
the long ten-n design tensile strength, Td, of the latest version of AASHTO.
E.

Unit Fill
Supply unit fill conforming to material and gradation requirements for Type CA-2 Coarse
Aggregate under 501-2.02, B.2. (Coarse Aggregate).

F.

Cast-in-place Concrete
Supply concrete conforming to Section 501 (Class A Concrete).

G. Backfill
Supply backfill material conforming to 203-1.08 (Suitable Material). Backfill material must
come from a single source, unless prior written approval for use of multiple sources is obtained
from the Director, Geotechnical Engineering Bureau.
Stockpile backfill material conforming to the current Geotechnical Control Procedure (GCP)
titled "Procedure for the Control of Granular Materials."
1. Material Test, Control and Acceptance Procedures
The State will perform procedures conforming to the appropriate Departmental publications
in effect on the date of advertisement of bids. These publications are available upon request
to the Regional Director, or the Director, Geotechnical Engineering Bureau.
Acceptance of the backfill will be made in accordance with the procedural directives of the
Geotechnical Engineering Bureau.
2.

Material Properties
a. Gradation
Stockpiled backfill material must meet the gradation requirements listed in Table
17554-2:
TABLE 17554-2
GRADATION
Sieve Size Designation
Percent Passing by Weight
63 mm
100
6.3 mm
30-100
425 m
0-60
75 m
0-15

FHWA NHI-07-092
Geosynthetic Engineering

9 54

9 MSE Walls
August 2008

b. Plasticity Index.
The Plasticity Index must not exceed 5.
c. Durability.
The Magnesium Sulfate Soundness loss must not exceed 30 percent.
H. Separation Geotextile
Supply geotextile material for Separation, Strength Class 2, appearing on the Department's
Approved List of Products for Geosynthetics for Highway Construction, B. Geotextiles (for use
on NYSDOT projects with a Sept. 7, 2000 or later letting date).
I.

Drainage Geotextile
Supply underdrain and geotextile material for drainage as shown on the plans or conforming to
the designer-suppliers Installation Manual:
1. Underdrain Pipe
Supply optional underdrain pipe conforming to Section 605 (Underdrains).
2. Geotextile Drainage
Supply geotextile material for drainage, Strength Class 2, Drainage Class B, appearing on the
Department's Approved List of Products for Geosythetics for Highway Construction, B.
Geotextiles (for use on NYSDOT projects with a Sept. 7, 2000 or later letting date).

J.

Identification Markers
Supply identification markers conforming to the designer-supplier's Installation Manual.

K. Basis of Acceptance
Accept cast-in-place concrete in accordance with the requirements of Section 501 (Portland
Cement Concrete), Class A.
Accept granular and backfill materials by the appropriate Departmental publications.
Accept facing and cap units in accordance with the requirements of 704-07 (Segmental
Retaining Wall Blocks).
Accept other materials by manufacturer's certification. The State reserves the right to sample,
test, and reject certified material in accordance with Departmental written instructions.
CONSTRUCTION DETAILS
A. Excavation, Disposal and MSSBRWS Area Preparation
Excavate, dispose and prepare the area on which the MSSBRWS will rest conforming to Section
203 (Excavation and Embankment), except as modified here:
1. Grade and level, for a width equaling or exceeding the reinforcement length, the area on
which the MSSBRWS will rest. Thoroughly compact this area to the Engineer's satisfaction.
Remove all soils found unsuitable, or incapable of being satisfactorily compacted because of
moisture content, in a manner directed by the Engineer, in conjunction with recommendations
of the Regional Geotechnical Engineer.
FHWA NHI-07-092
Geosynthetic Engineering

9 55

9 MSE Walls
August 2008

2. Remove rock to the limits shown on the plans.


3. Excavate the area for the leveling pad in accordance with the requirements of Section 206,
(Trench, Culvert and Structure Excavation).
B.

Facing and Cap Unit Storage and Inspection


Handle and store facing and cap units with extreme care to prevent damage. The State will
inspect facing and cap units on their arrival at the work site and prior to their installation to
determine any damage that may have occurred during shipment. Facing and cap units will be
considered damaged if they contain any cracks or spalls and/or honey combed areas with any
dimensions greater than 25mm. The State will reject any damaged facing and cap units. Replace
rejected units with facing and cap units acceptable to the Engineer.

C. Facing Unit Erection


1. Provide an un-reinforeed concrete or compacted granular fill leveling pad as shown on the
plans.
a. Place concrete in conformance with 555-3, (Construction Details).
b. Place and compact granular fill in conformance with 203-3.12 (Compaction).
2. Install by placing, positioning, and aligning facing units in conformance with the designersupplier's Installation Manual and within the tolerances in Table 17554-3.
TABLE 17554-3
TOLERANCES
Vertical control
7 mm over a distance of 3 m
Horizontal location control
+13 mm over a distance of 3 m
Rotation from established plan wall batter
13 mm over 3 m in height
3. Correct all misalignments of installed facing units that exceed the tolerances allowed in Table
17554-3 in a manner satisfying the Engineer.
4. Control all operations and procedures to prevent misalignment of the facing units.
Precautionary measures include (but are not limited to) keeping vehicular equipment at least
1 meter behind the back of the facing units. Compaction equipment used within 1 meter of
the back of the facing units must conform to 203-3.12B.6. (Compaction Equipment for
Confined Areas).
D. Unit Fill
1. Place unit fill to the limits indicated on the plans. Before installing the next course of facing
units, compact the unit fill in a manner satisfying the Engineer and brush the tops of the
facing units clean to ensure an even placement area.
2. Protect unit fill from contamination during construction.
E.

Extensible Reinforcement
1. Before placing extensible reinforcement, backfill placed and compacted within a 1 meter
horizontal distance of the back of facing units must be no more than 25 mm above the

FHWA NHI-07-092
Geosynthetic Engineering

9 56

9 MSE Walls
August 2008

required extensible reinforcement elevation. Backfill placed and compacted beyond the 1
meter horizontal distance may be roughly graded to the extensible reinforcement elevation.
2. Place extensible reinforcement normal to facing units unless otherwise indicated on the plans.
Replace all broken, damaged or distorted extensible reinforcement at no additional cost to the
State.
3. Install extensible reinforcement within facing units conforming to the designer- supplier's
Installation Manual. Pull taut and secure the extensible reinforcement before placing the
backfill in a manner satisfying the Engineer.
F.

Backfill
1. Place backfill materials (other than rock) at a moisture content less than or equal to the
Optimum Moisture Content. Remove backfill materials placed at a moisture content
exceeding the Optimum Moisture Content and either rework or replace, as determined by the
Engineer. Determine Optimum Moisture Content in conformance with Soil Test Methods for
compaction that incorporate moisture content determination. Use Soil Test Methods
(excluding STM-6) in effect on the date of advertisement of bids. Cost to rework or replace
backfill materials shall be borne by the Contractor.
2. Place granular backfill material in uniform layers so that the compacted thickness of each
layer does not exceed 250 mm or one block height, whichever is less. Compact each layer to
a minimum of 95 percent of Standard Proctor Maximum Density and in conformance with
203-3.12 (Compaction).
3. Place rock backfill in uniform layers so that the compacted thickness of each layer does not
exceed 250 mm or one block height, whichever is less. Compact each layer in conformance
with 203-3.12 (Compaction). The Engineer will determine by visual inspection that proper
compaction has been attained.

METHOD OF MEASUREMENT
MSSBRWS measurement is computed as the number of square meters of wall face area between the
payment lines shown on the plans.
BASIS OF PAYMENT
A. Mechanically Stabilized Segmental Block Retaining Wall System
Payment in square meters of wall face area includes cost of all labor, equipment and materials
necessary to complete the work, including leveling pad, facing and cap units, backfill,
underdrains and geotextiles.
B.

Excavation and Disposal


Excavation and disposal will be paid for under Item 203.02, Unclassified Excavation and
Disposal or Item 206.01, Structure Excavation.

C.

Water
The unit price bid for the MSSBRWS includes the cost of adding water for backfill compaction,
unless separate items for Furnishing Water Equipment and Applying Water have been included
in the Contract.

FHWA NHI-07-092
Geosynthetic Engineering

9 57

9 MSE Walls
August 2008

NYDOT Standard Specification (2006) 704-07 Segmental Retaining Wall Blocks.


704-07 SEGMENTAL RETAINING WALL BLOCKS
SCOPE. This specification covers the material details and quality requirements for segmental
retaining wall blocks.
MATERIAL REQUIREMENTS. Provide segmental retaining wall block meeting the style and
color requirements in the contract documents. Use materials, meeting the following requirements, in
the manufacture of segmental retaining wall blocks:
Portland Cement
Coarse Aggregate
Mortar Sand
Grout Sand
Concrete Sand
Fly Ash
Ground, Granulated Blast-Furnace Slag
Water

701-01
703-02
703-03
703-04
703-07
711-10
711-12
712-01

Fly ash or ground, granulated blast-furnace slag may be substituted for up to a maximum of 20% by
weight of the total amount of cement plus pozzolan in the mix. Use integral coloring pigments, when
required, meeting the requirements of ASTM C979. Other materials may be used in the manufacture
as approved by the Director, Materials Bureau.
Physical Properties. The minimum acceptable average compressive strength of five-block samples
is 28 MPa, with no individual block sample less than 24 MPa. The maximum acceptable average
freeze/thaw loss of five-block samples, subjected to 42 freeze/thaw cycles in a 3% NaCl solution, is
1.0%, with no 'individual sample exceeding 1.5%.
The formed dimensions of concrete retaining wall block units will not differ more than 5 mm
from the nominal dimensions shown on the approved Materials Detail Drawing. Provide sound
blocks, free from cracks or other defects that would interfere with the proper placing, performance, or
appearance of the blocks.
Materials Details. At the time of application to the Approved List, submit Materials Details
Drawings to the Director, Materials Bureau for approval. Prepare and submit drawings in accordance
with Departmental procedural directives. Submit a unique drawing(s) for each block style under
consideration.
SAMPLING AND TESTING. When samples are requested by the Department, randomly select
them from production-run material. A minimum of 5 samples, prepared by the manufacturer in
accordance with ASTM C140, will be required for compression testing. A minimum of five samples,
prepared by the manufacturer in accordance with ASTM C 1262, will be required for freeze/thaw
testing. Samples will be tested for compressive strength in accordance with ASTM C 140. Samples
will be tested for freeze/thaw durability in accordance with ASTM C 1262.
BASIS OF ACCEPTANCE. Segmental retaining wall blocks will be accepted on the basis of the
manufacturer's name and block style appearing on the Department's Approved List, a material
certification that specifies the product conforms to this specification, and conformance to the
approved materials detail drawing(s).
FHWA NHI-07-092
Geosynthetic Engineering

9 58

9 MSE Walls
August 2008

9.9-2
Modular Block Wall Unit
The following material specification for concrete segmental retaining wall (a.k.a. MBW)
units is from the National Concrete Masonry Association, Design Manual for Segmental
Retaining Walls (NCMA, 1997). Note that NCMA uses the terms SRW and segmental
retaining wall for the FHWA term of MBW unit.
SPECIFICATION FOR SEGMENTAL RETAINING WALL MATERIALS
PART 1: GENERAL
1.01

Description
Work shall consist of furnishing all materials, labor, equipment, and supervision to install a
segmental retaining wall system in accordance with these specifications and in reasonably
close conformity with the lines, grades, design and dimensions shown on the plans or as
established by the Owner or Owners Engineer.

1.02

Related Work
A.
B.

1.03

_________
_________

Section - Site Preparation


Section - Earthwork

Reference Standards
A.

Engineering Design
1.
NCMA Design Manual for Segmental Retaining Walls
2.
NCMA TEK 2-4 - Specifications for Segmental Retaining Wall Units
3.
NCMA SRWU-1 - Determination of Connection Strength between
Geosynthetics and Segmental Concrete Units
4
NCMA SRWU-2 - Determination of Shear Strength between Segmental
Concrete Units

B.

Segmental Retaining Wall Units


1.
ASTM C 140 - Sampling and Testing Concrete Masonry Units
2.
ASTM C 1262 - Evaluating the Freeze-Thaw Durability of Manufactured
Concrete Masonry Units and Related Concrete Units

C.

Geosynthetic Reinforcement
1.
ASTM D 4595 Test Method for Tensile Properties of Geotextiles by the
Wide-Width Strip Method
2.
ASTM D 5262 -Test Method for Evaluating the Unconfined Creep Behavior
of Geosynthetics
3.
GRI GG-1: Single Rib Geogrid Tensile Strength
4.
GRI GG-5: Geogrid Pullout
5.
GRI GT-6: Geotextile Pullout

D.

Soils
1.
2.

FHWA NHI-07-092
Geosynthetic Engineering

ASTM D 698 - Moisture Density Relationship for Soils, Standard Method


ASTM D 422 - Gradation of Soils

9 59

9 MSE Walls
August 2008

3.
4.

1.04

E.

Drainage Pipe
1.
ASTM D 3034 - Specification for Polyvinyl Chloride (PVC) Plastic Pipe
2.
ASTM D 1248 - Specification for Corrugated Plastic Pipe

F.

Where specifications and reference documents conflict, the Owners Engineer shall
make the final determination of applicable document.

Approved Segmental Retaining Wall Systems


A.

Suppliers of segmental retaining wall system material components shall have


demonstrated experience in the supply of similar size and types of segmental
retaining walls on previous projects, and shall be approved by the Owners Engineer.
The supplier must be approved two weeks prior to bid opening. Suppliers currently
approved for this work are:
Segmental Wall Units
1.
2.
3.
1.
2.
3.

1.05

Geosynthetic Reinforcements
1.

Submittals
A.

1.06

ASTM D 424 - Atterberg Limits of Soils


ASTM D G51 - Soil pH

Material Submittals - The Contractor shall submit manufacturers certifications, 30


days prior to the start of work, stating that the SRW units, the geosynthetic
reinforcement, and the drainage aggregate meet the requirements of section 2.0 of
this specification. The Contractor shall provide a list of successful projects with
references showing that the installer for the segmental retaining wall is qualified and
has a record of successful performance.

Delivery, Storage, and Handling


A.

The Contractor shall inspect the materials upon delivery to assure that proper type
and grade material has been received.

B.

The Contractor shall store and handle materials in accordance with manufacturers
recommendations.

C.

The Contractor shall protect the materials from damage. Damaged material shall not
be incorporated into the segmental retaining wall.

PART 2: MATERIAL
2.01 Concrete Segmental Retaining Wall Units

FHWA NHI-07-092
Geosynthetic Engineering

9 60

9 MSE Walls
August 2008

A.

Concrete segmental units shall conform to the requirements of NCMA TEK 2-4 and
have a minimum 28 days compressive strength of 3000 psi and a maximum
absorption of 10 pcf as determined in accordance with ASTM C 140. For areas
subject to detrimental freeze-thaw cycles as determined by the Owner or Owners
Engineer the concrete shall have adequate freeze/thaw protection and meet the
requirements of ASTM C1262.

B.

All units shall be sound and free of cracks or other defects that would interfere with
the proper placing of the unit or significantly impair the strength or permanence of
the construction. Any cracks or chips observed during construction shall fall within
the guidelines outlined in NCMA TEK2-4.

C.
SRW units dimensions shall not differ more than + 1/8 inch except height, which
shall not differ more than, + 1/16 inch, as measured in accordance with ASTM C140.
D.
SRW units shall match the color, surface finish and dimension for height, width,
depth and batter as shown on the plans.
E.
If pins are used by the retaining wall supplier to interconnect SRW units, they shall
consist of a nondegrading polymer or galvanized steel and be made for the express use with
the SRW units supplied.
F.
2.02

Cap adhesive shall meet the requirements of the SRW unit manufacturer.

Geosynthetic Reinforcements

A. Geosynthetic Reinforcements shall consist of high tenacity geogrids or geotextiles


manufactured for soil reinforcement applications. The type, strength and placement
location of the reinforcing geosynthetic shall be as shown on the plans. The design
properties of the reinforcement shall be determined according to the procedures outlines
in this specification and the NCMA Design Manual for Segmental Retaining Walls (1996
Revision.) Detailed test data shall be submitted to the Owners Engineer for approval at
least 30 days prior to construction and shall include tensile strength (ASTM D 4595 or
GRI GG-1), creep (ASTM D 5262) site damage and durability (GRI GG-4) pullout (GRI
GG-5 or GRI GT-6) and connection (NCMA SRWU-1) test data.
Included with the raw test data shall be a report that will show that the proposed geosynthetic
reinforcements have the following minimum properties:

Property
Type 1

Geosynthetic Reinforcement
Type 2

Type 3

Allowable Reinforcement
Tension - Ta (lb/ft)
Coefficient of Interaction Ci
Coefficient of Direct
Sliding - CDs

FHWA NHI-07-092
Geosynthetic Engineering

9 61

9 MSE Walls
August 2008

Calculation of the allowable reinforcement tension shall use the following method:
Allowable Reinforcement Tension:
The allowable reinforcement tension, Ta, at the end of the service life shall consider the timetemperature creep characteristics of the reinforcement, environmental degradation,
construction induced damage and an overall factor of safety.

Ta =

RFD RFID

Tult
RFCR FSUNC

where:
Tult

RFD

RFID

RFCR

FSUNC =

Ultimate (or yield tensile strength) from wide width tensile strength tests
(ASTM D 4595 or GR1 GG-1 for geogrids), based on minimum average roll
value (MARV) for the product.
Durability reduction factor is dependent on the susceptibility of the geosynthetic
to attack by microorganisms, chemicals, thermal oxidation, hydrolysis and
stress cracking, and can vary typically for 1.1 to 2.0.
Installation damage reduction factor can range from 1.05 to 3.0, depending on
backfill gradation and product mass per unit weight.
Creep reduction factor is the ratio of the ultimate strength (Tult) to the creep
limit strength obtained from laboratory creep tests for each product, and can
vary typically from 1.50 to 5.0.
Overall factor of safety or load factor to account for uncertainties in the
geometry of the structure, fill properties, reinforcement properties, and
externally applied loads, and shall be no less than 1.5.

In no case shall the product RFID x RFD x RFCR be less than 2.0.
2.03

Drainage Pipe
A.
B.

2.04

The drainage collection pipe shall be a perforated or slotted, PVC or corrugated


HDPE pipe. The pipe and drainage aggregate may be wrapped with a geotextile that
will function as a filter.
Drainage pipe shall be manufactured in accordance with ASTM D 3034 and/or
ASTM D 1248

Drainage Aggregate
A.

Drainage aggregate shall be a clean crushed stone or granular fill meeting the
following gradation as determined in accordance with ASTM D 422:
Sieve Size
1 inch
3/4 inch
No. 4
No. 40
No. 200

FHWA NHI-07-092
Geosynthetic Engineering

Percent Passing
100
75 - 100
0 - 60
0 - 50
0-5

9 62

9 MSE Walls
August 2008

2.05

Reinforced Backfill

A.

The reinforced backfill shall be free of debris and consist of one of the following
inorganic USCS soil types: GP, GW, SW, SP, SM, meeting the following gradation
as determined in accordance with ASTM D 422.
Sieve Size
4 inch
No. 4
No. 40
No. 200

Percent Passing
100 - 75
100 - 20
0 - 60
0 - 35

The maximum size should be limited to 3/4 inch for reinforced soil SRWs unless
tests have been performed to evaluate potential strength reduction in the geosynthetic
due to installation damage.
The plasticity of the fine fraction of the reinforced soil shall be less than 20.
B.
2.06

The pH of the backfill material shall be between 3 and 9 when tested in accordance
with ASTM G-51.

Geotextile Filter

A. Drainage geotextile shall have the following minimum properties or shall meet the
criteria recommended by the Wall Design Engineer.
AOS
Grab Tensile
Trap Tear
Water Flow Rate
Puncture

ASTM D 4751
ASTM D 4632
ASTM D 4533
ASTM D 4491
ASTM D 4833

PART 3: CONSTRUCTION
3.01

Inspection
A.

The Owner or Owners Engineer is responsible for verifying the materials supplied
by the contractor meets all the requirements of the specification. This includes all
submittals and proper installation of the system.

B.

As requested by the Owners Engineer or Contractor, the segmental retaining wall


system supplier shall provide a qualified and experienced representative on site for
up to 3 days to assist the Contractor regarding proper wall installation. Contractor
site assistant shall be provided at no additional cost to the Owner.

C.

The Contractors field construction supervisor shall have demonstrated experience


and be qualified to direct all work at the site.

FHWA NHI-07-092
Geosynthetic Engineering

9 63

9 MSE Walls
August 2008

3.02

Excavation
A.

3.03

Foundation Preparation
A.

3.04

Following excavation for the leveling pad and the reinforced soil zone foundation
soil shall be examined by the Owners Engineer to assure the actual foundation soil
strength meets or exceeds the assumed design bearing strength. Soils not meeting the
required strength shall be removed and replaced with soil meeting the design criteria,
as directed by the Owners Engineer.

Leveling Pad Preparation


A.

3.05

The Contractor shall excavate to the lines and grades shown on the project grading
plans.
The Contractor shall take precautions to minimize over-excavation.
Excavation support, if required, shall be designed by the Contractor.

A minimum 6 inch thick layer of compacted granular material shall be placed for use
as a leveling pad up to the grades and locations as shown on the construction
drawings. The granular base shall be compacted to provide a firm, level bearing pad
on which to place the first course of concrete segmental retaining wall units.
Compaction should be performed using a lightweight compactor, such as a
mechanical plate compactor to obtain a minimum of 95% of the maximum standard
Proctor density (ASTM D 698).

SRW and Geosynthetic Reinforcement Placement


A.

B.

All materials shall be installed at the proper elevation and orientation as shown in the
wall details on the construction plans or as directed by the Owners Engineer. The
concrete segmental wall units and geosynthetic reinforcement shall be installed in
general accordance with the manufacturers recommendations. The drawings shall
govern in any conflict between the two requirements.
Overlap of the geosynthetic in the design strength direction shall not be permitted.
The design strength direction is that length of geosynthetic reinforcement
perpendicular to the wall face and shall consist of one continuous piece of material.
Adjacent sections of geosynthetic shall be placed in a manner to assure that the
horizontal coverage shown on the plans is provided.

C.

Geosynthetic reinforcement should be installed under tension. A nominal tension


shall be applied to the reinforcement and maintained by staples, stakes, or hand
tensioning until the reinforcement has been covered by at least 6 inches of soil fill.

D.

The overall tolerance relative to the wall design verticality or batter shall not
exceed 1.25 inches maximum over a 10 ft distance; 3 inch maximum.

E.

Broken, chipped, stained or otherwise damaged units shall not be placed in the wall
unless they are repaired and the repair method and results are approved by the
Engineer.

FHWA NHI-07-092
Geosynthetic Engineering

9 64

9 MSE Walls
August 2008

3.06

3.07

3.8

Backfill Placement
A.

The reinforced backfill shall be placed as shown in construction plans in maximum


compacted lift thickness of 10 inches and shall be compacted to a minimum 95% of
standard Proctor density (ASTM D 698) at a moisture content within 2% of optimum.
Backfill shall be placed, spread and compacted in such a manner that eliminates the
development of wrinkles or movement of the geosynthetic reinforcement and the wall
facing units.

B.

Only hand-operated compaction equipment shall be allowed within 3 feet of the front
of the wall face. Compaction within 3 feet of the back face of the facing units shall
be achieved by at least three (3) passes of a lightweight mechanical tamper, plate or
roller. Soil density in this area shall not be less than 90% standard Proctor density.

C.

Tracked construction equipment shall not be operated directly on the geosynthetic


reinforcement. A minimum backfill thickness of 6 inches is required prior to
operation of tracked vehicles over the geosynthetic reinforcement. Turning of
tracked vehicles should be kept to a minimum to prevent displacing the fill and
damaging or moving the geosynthetic reinforcement.

D.

Rubber-tired equipment may pass over the geosynthetic reinforcement, if in


accordance with the manufacturers recommendations, at slow speeds less that 10
mph. Sudden braking and sharp turning should be avoided.

E.

At the end of each days operation, the contractor shall slope the last level of backfill
away from the wall facing to direct runoff of rainwater away from the wall face. In
addition, the contractor shall not allow surface runoff from adjacent areas to enter the
wall construction site.

Drainage Fill Placement


A.

Drainage fill shall be placed to the minimum finished thickness and widths shown on
the construction plans or as modified by the Owners Engineer.

B.

Drainage collection pipes shall be installed to maintain gravity flow of water outside
of the reinforced soil zone. The drainage collection pipe should daylight into a storm
sewer manhole or along a slope at an elevation lower than the lowest point of the
pipe within the aggregate drain.

C.

The main collection drain pipe, just behind the block facing, shall be a minimum of
3 inches in diameter. The secondary collection drain pipes should be sloped a
minimum of two percent to provide gravity flow into the main collection drain pipe.
Drainage laterals shall be spaced at a maximum 50 feet spacing along the wall face.

Cap Block Placement

A. The cap block and/or top SRW unit shall be bonded to the SRW units below using cap
adhesive described in Section 2.01F.

FHWA NHI-07-092
Geosynthetic Engineering

9 65

9 MSE Walls
August 2008

PART 4: MEASUREMENT AND PAYMENT


4.01

Measurement
A.

4.02

The unit of measurement for furnishing the segmental retaining wall system shall be
the vertical square foot of wall surface from the top of the leveling pad to the top of
the wall or wall coping. The quantity to be paid shall include supply and installation
of the segmental retaining wall system. Excavation of unsuitable materials and
replacement with select fill, as directed and approved in writing by the Owner or
Owners Engineer shall be paid for under separate pay items.

Payment
A.

The accepted quantities of segmental retaining wall system will be paid for per
vertical square foot in place as measured from the top of the leveling pad to the top of
wall or coping block. The quantities of the segmental retaining wall system as shown
on plans or as approved by the Owner or Owners Engineer shall be used to
determine the area supplied. Payment will be made under:

Pay Item
Geosynthetic Reinforced SRW

Pay Unit
SQ. FT.

9.9-3 Geosynthetic Wrap Around Wall


The following example was obtained from WashDOT (2008). It is Standard Specification 614 for material and construction for permanent and temporary geosynthetic walls. It is noted
that the geosynthetic shall conform to Standard Specification 9-33 Construction
Geosynthetic. Applicable excerpts from the 9-33 material specification follow the 6-14
Geosynthetic Retaining Walls specification.
6-14

GEOSYNTHETIC RETAINING WALLS

6-14.1
Description
This work consists of constructing geosynthetic retaining walls, including those shown in the
Standard Plans.
6-14.2
Materials
Materials shall meet the requirements of the following sections:
Gravel Borrow For Geosynthetic Retaining Wall
9-03.14(4)
Construction Geosynthetic
9-33
The requirements specified in Section 2-12.2 for geotextile shall also apply to geosynthetic and
geogrid materials used for permanent and temporary geosynthetic retaining walls.
Other materials required shall be as specified in the Special Provisions.

FHWA NHI-07-092
Geosynthetic Engineering

9 66

9 MSE Walls
August 2008

6-14.3
Construction Requirements
Temporary geosynthetic retaining walls are defined as those walls and wall components constructed
and removed or abandoned before the Physical Completion Date of the project or as shown in the
Plans. All other geosynthetic retaining walls shall be considered as permanent.
6-14.3(l) QualityAssurance
The Contractor shall complete the base of the retaining wall excavation to within plus or minus 3inches of the staked elevations unless otherwise directed by the Engineer. The Contractor shall place
the external wall dimensions to within plus or minus 2-inches of that staked on the ground. The
Contractor shall space the reinforcement layers vertically and place the overlaps to within plus or
minus 1-inch of that shown in the Plans.
The completed wall(s) shall meet the following tolerances:
Permanent
Wall
3-inches

Temporary
Wall
5-inches

Deviation from the overall design batter per 10-feet of wall height
shall not exceed:

2-inches

3-inches

Maximum outward bulge of the face between backfill reinforcement


layers shall not exceed:

4-inches

6-inches

Deviation from the design batter and horizontal alignment for the
face when measured along a 10-foot straight edge at the midpoint of
each wall layer shall not exceed:

6-14.3(2) Submittals
A minimum of 14 calendar days prior to beginning construction of each wall the Contractor shall
submit detailed plans for each wall in accordance with Section 6-01.9. As a minimum, the submittals
shall include the following:
1. Detailed wall plans showing the actual lengths proposed for the geosynthetic reinforcing
layers and the locations of each geosynthetic product proposed for use in each of the
geosynthetic reinforcing layers.
2. The Contractor's proposed wall construction method, including proposed forming systems,
types of equipment to be used, proposed erection sequence and details of how the backfill
will be retained during each stage of construction.
3. Manufacturer's Certificate of Compliance, samples of the retaining wall geosynthetic and
sewn seams for the purpose of acceptance as specified.
4. Details of geosynthetic retaining wall corner construction, including details of the positive
connection between the wall sections on both sides of the corner.
5. Details of terminating a top laver of retaining wall geosynthetic and backfill due to a
changing retaining wall profile.
Approval of the Contractor's proposed wall construction details and methods shall not relieve the
Contractor of their responsibility to construct the walls in accordance with the requirements of these
Specifications.
6-14.3(3) Excavation and Foundation Preparation
Excavation shall conform to Section 2-09.3(4), and to the limits and construction stages shown in the
Plans. Foundations soils found to be unsuitable shall be removed and replaced in accordance with
Section 2-09.3(1)C.
FHWA NHI-07-092
Geosynthetic Engineering

9 67

9 MSE Walls
August 2008

The Contractor shall direct all surface runoff from adjacent areas away from the retaining wall
construction site.
6-14.3(4) Erection and Backfill
The Contractor shall begin wall construction at the lowest portion of the excavation and shall place
each layer horizontally as shown in the Plans. The Contractor shall complete each layer entirely
before beginning the next layer.
Geotextile splices shall consist of a sewn seam or a minimum 1-foot 0-inches overlap. Geogrid
splices shall consist of adjacent geogrid strips butted together and fastened using hog rings, or other
methods approved by the Engineer, in such a manner to prevent the splices from separating during
geogrid installation and backfilling. Splices exposed at the wall face shall prevent loss of backfill
material through the face. The splicing material exposed at the wall face shall be as durable and
strong as the material to which the splices are tied. The Contractor shall offset geosynthetic splices in
one layer from those in the other layers such that the splices shall not line up vertically. Splices
parallel to the wall face will not be allowed, as shown in the Plans.
The Contractor shall stretch out the geosynthetic in the direction perpendicular to the wall face to
ensure that no slack or wrinkles exist in the geosynthetic prior to backfilling.
For geogrids, the length of the reinforcement required as shown in the Plans shall be defined as the
distance between the geosynthetic wrapped face and the last geogrid node at the end of the
reinforcement in the wall backfill.
The Contractor shall place fill material on the geosynthetic in lifts such that 6-inches minimum of fill
material is between the vehicle or equipment tires or tracks and the geosynthetic at all times. The
Contractor shall remove all particles within the backfill material greater than 3-inches in size.
Turning of vehicles on the first lift above the geosynthetic will not be permitted. The Contractor shall
not end dump fill material directly on the geosynthetic without the prior approval of the Engineer.

Should the geosynthetic be damaged or the splices disturbed, the backfill around the
damaged or displaced area shall be removed and the damaged strip of geosynthetic replaced
by the Contractor at no expense to the Contracting Agency.
The Contractor shall use a temporary form system to prevent sagging of the geosynthetic facing
elements during construction. A typical example of a temporary form system and sequence of wall
construction required when using this form are detailed in the Plans. Soil piles or the geosynthetic
manufacturer's recommended method, in combination with the forming system shall be used to hold
the geosynthetic in place until the specified cover material is placed.
The Contractor shall place and compact the wall backfill in accordance with the wall construction
sequence detailed in the Plans and Method C of Section 2-03.3(14)C, except as follows:
1. The maximum lift thickness after compaction shall not exceed 10-inches
2. The Contractor shall decrease this fill thickness, if necessary, to obtain the specified density
3. Rollers shall have sufficient capacity to achieve compaction without causing distortion to the
face of the wall in accordance with Section 6-14.3(i).
4. The Contractor shall not use sheepsfoot rollers or rollers with protrusions.
5. The Contractor shall compact the zone within 3-feet of the back of the wall facing panels
without causing damage to or distortion of the wall facing elements (welded wire mats,
backing mats. construction geotextile for wall facing, precast concrete facing panels, and
FHWA NHI-07-092
Geosynthetic Engineering

9 68

9 MSE Walls
August 2008

concrete blocks) by using a plate compactor as approved by the Engineer. No soil density
tests will be taken within this area.
6. For wall systems with geosynthetic reinforcement, the minimum compacted backfill lift
thickness of the first lift above each geosynthetic reinforcement layer shall be 6-inches.
The Contractor shall construct wall comers at the locations shown in the Plans, and in accordance
with the wall corner construction sequence and method submitted by the Contractor and approved by
the Engineer. Wall angle points with an interior angle of less than 150-degrees shall be considered to
be a wall corner. The wall comer shall provide a positive connection between the sections of the wall
on each side of the corner such that the wall backfill material cannot spill out through the comer at
any time during the design life of the wall. The Contractor shall construct the wall corner such that
the wall sections on both sides of the corner attain the full geosynthetic layer embedment lengths
shown in the Plans.
Where required by retaining wall profile grade, the Contractor shall terminate top layers of retaining
wall geosynthetic and backfill in accordance with the method submitted by the Contractor and
approved by the Engineer. The end of each laver at the top of the wall shall be constructed in a
manner that prevents wall backfill material from spilling out the face of the wall throughout the life of
the wall. If the profile of the top of the wall changes at a rate of 1:1 or steeper, this change in top of
wall profile shall be considered to be a corner.
6-14.3(5) Guardrail Placement
The Contractor shall install guardrail posts as shown in the Plans after completing the wall, but before
the permanent facing is installed. The Contractor shall install the posts in a manner that prevents
bulging of the wall face and prevents ripping, tearing, or pulling of the geosynthetic reinforcement.
Holes through the geosynthetic reinforcement shall be the minimum size necessary for the post. The
Contractor shall demonstrate to the Engineer prior to beginning guardrail post installation that the
installation method will not rip, tear, or pull the geosynthetic reinforcement.
6-14.3(6) Permanent Facing
The Contractor shall apply a permanent facing to the surface of all permanent geosynthetic retaining
walls as shown in the Plans. Shotcrete facing, if shown in the Plans, shall conform to Section 6-18.
Concrete fascia panel, if shown in the Plans, shall conform to Section 6-15.3(9).
6-14.3(7) Geosynthetic Retaining Wall Traffic Barrier and Geosynthetic Retaining Wall
Pedestrian Barrier
Geosynthetic wall traffic barrier (single slope and f-shape) and geosynthetic retaining wall
pedestrian barrier shall be constructed in accordance with Sections 6-02.3(11)A and 6-10.3(2), and
the details in the Plans.
6-14.4
Measurement
Permanent geosynthetic retaining wall and temporary geosynthetic retaining wall will be measured by
the square foot of face of completed wall. Corner wrap area and extensions of the geosynthetic wall
beyond the area of wall face shown in the Plans or staked by the Engineer are considered incidental to
the wall constriction and will not be included in the measurement of the square foot of face of
completed geosynthetic retaining wall.
Gravel borrow for geosynthetic retaining wall backfill will be measured as specified in Section 203.4.

FHWA NHI-07-092
Geosynthetic Engineering

9 69

9 MSE Walls
August 2008

Shotcrete facing and concrete fascia panel will be measured by the square foot surface area of the
completed facing or fascia panel, measured to the neat lines or the facing or panel as shown in the
Plans. When a footing is required, the measurement of the fascia panel area will include the footing.
Geosynthetic wall single slope traffic barrier, geosynthetic wall f-shape traffic barrier, and
geosynthetic retaining wall pedestrian barrier will be measured as specified in Section 6-10.4 for castin-place concrete barrier.
Structure excavation Class B, Structure excavation Class B including haul, and shoring or extra
excavation Class B, will be measured in accordance with Section 2-09.4.
6-14.5
Payment
Payment will be made in accordance with Section 1-04.1 for each of the following Bid items when
they are included in the Proposal:
Geosynthetic Retaining Wall, per square foot.
Temporary Geosynthetic Retaining Wall, per square foot.
All costs in connection with constructing the temporary or permanent geosynthetic retaining wall as
specified shall be included in the unit Contract price per square foot for Geosynthetic Retaining
Wall and Temporary Geosynthetic Retaining Wall including compaction of the backfill material
and furnishing and installing the temporary forming system.
Gravel Borrow for Geosynthetic Ret. Wall Incl. Haul, per ton or per cubic yard. All costs in
connection with furnishing and placing backfill material for temporary or permanent geosynthetic
retaining walls as specified shall be included in the unit Contract price per ton or per cubic yard for
Gravel Borrow for Geosynthetic Ret. Wall Incl. Haul.
Concrete Fascia Panel, per square foot.
All costs in connection with constructing the concrete fascia panels as specified shall be included in
the unit Contract price per square foot for Concrete Fascia Panel, including all steel reinforcing
bars, premolded joint filler, polyethylene bond breaker strip, joint sealant, PVC pipe for weep holes,
exterior surface finish, and pigmented sealer (when specified).
Shoterete facing will be paid for in accordance with Section 6-18.5.
Geosynthetic Wall Single Slope Traffic Barrier per linear foot.
Geosynthetic Wall F-Shape Traffic Barrier, per linear foot.
Geosynthetic Retaining Wall Pedestrian Barrier, per linear foot.
The unit Contract price per linear foot for Geosynthetic Wall Single Slope Traffic Barrier,
Geosynthetic Wall F-Shape Traffic Barrier, and Geosynthetic Retaining Wall Pedestrian Barrier
shall be full pay for constructing the barrier on top of the geosynthetic retaining wall.
Structure Excavation Class B, per cubic yard.
Structure Excavation Class B lncl. Haul, per cubic yard.
Shoring Or Extra Excavation Class B, per square foot.

FHWA NHI-07-092
Geosynthetic Engineering

9 70

9 MSE Walls
August 2008

Geosynthetic Reinforcement: Excerpts applicable to wall reinforcement from WashDOT


(2008) Standard Specification 9-33 Construction Geosynthetic follow.
9-33

CONSTRUCTION GEOSYNTHETIC

9-33.1
Geosynthetic Material Requirements
The term geosynthetic shall be considered to be inclusive of geotextiles, geogrids, and prefabricated
drainage mats.
Geotextiles, including geotextiles attached to prefabricated drainage core to form a prefabricated
drainage mat, shall consist only of long chain polymeric fibers or yarns formed into a stable network
such that the fibers or yams retain their position relative to each other during handling, placement,
and design service life. At least 95 percent by weight of the material shall be polyolefins or
polyesters. The material shall be free from defects or tears. The geotextile shall also be free of any
treatment of coating which might adversely alter its hydraulic or physical properties after installation.
Geogrids shall consist of a regular network of integrally connected polymer tensile elements with an
aperture geometry sufficient to permit mechanical interlock with the surrounding backfill. The long
chain polymers in the geogrid tensile elements, not including coatings, shall consist of at least 95
percent by mass of the material of polyolefins or polyesters. The material shall be free of defects,
cuts, and tears.
Prefabricated drainage core shall consist of a three dimensional polymeric material with a structure
that permits flow along the core laterally and which provides support to the geotextiles attached to it.
The geosynthetic shall conform to the properties as indicated in Tables 1 through 8 in Section 9-33.2,
and additional tables as required in the Standard Plans and Special Provisions for each use specified
in the Plans. Specifically the geosynthetic uses included in this section and their associated tables of
properties are as follows:
Geotextile Geosynthetic Application

Applicable Property Tables

Permanent Geosynthetic Retaining Wall


Table 7 and Std. Plans
Temporary Geosynthetic Retaining Wall
Tables 7 and 10
Table 10 will be included in the Special Provisions.
Geogrid and geotextile reinforcement in geosynthetic retaining walls shall conform to the properties
specified in the Standard Plans for permanent walls, and Table 10 for temporary walls.
For geosynthetic retaining walls that use geogrid reinforcement, the geotextile material placed at the
wall face to retain the backfill material as shown in the Plans shall conform to the properties for
Construction Geotextile for Underground Drainage, Moderate Survivability, Class A.
Thread used for sewing geotextiles shall consist of high strength polypropylene, polyester, or
polyamide. Nylon threads will not be allowed. The thread used to sew permanent erosion control
geotextiles, and to sew geotextile seams in exposed faces of temporary or permanent geosynthetic
retaining walls, shall also be resistant to ultraviolet radiation. The thread shall be of contrasting color
to that of the geotextile itself.
FHWA NHI-07-092
Geosynthetic Engineering

9 71

9 MSE Walls
August 2008

9-33.2

Geosynthetic Properties

9-33.2(2) Geosynthetic Properties For Retaining Walls and Reinforced Slopes


Table 7: Minimum properties required for geotextile reinforcement used in geosynthetic reinforced
slopes and retaining walls.
Geotextile Property

ASTM Test Method2

AOS
Water Permittivity
Grab Tensile Strength,
in machine and xmachine direction
Grab Failure Strain, in
machine and xmachine direction
Seam Breaking
Strength
Puncture Resistance
Tear Strength, in
machine and xmachine direction

D 4751
D 4491

Ultraviolet (UV)
Radiation Stability
9-33.4

Geotextile Property Requirements1


Woven
Nonwoven
U.S. No. 20 max.
0.02 sec-1 min.

D 4632

200 lb min.

120 lb min.

D 4632

< 50%

> 50%

D 46323,4

160 lb min.

100 lb min.

D 6241

370 lb min.

220 lb min.

D 4533

63 lb min.

50 lb min.

70% (for polypropylene and polyethylene)


and 50% (for polyester) Strength Retained
min., after 500 hours in a xenon arc device

D 4355

Geosynthetic Material Approval and Acceptance

9-33.4(l) Geosynthetic Material Approval


If the geosynthetic source material has not been previously evaluated, or is not listed in the current
WSDOT Qualified Products List (QPL), a sample of each proposed geosynthetic shall be submitted
to the State Materials Laboratory in Tumwater for evaluation. Geosynthetic material approval will be
based on conformance to the applicable properties from the Tables in Section 9-33.2 or in the
Standard Plans or Special Provisions. After the simple and required information for each
geosynthetic type have arrived at the State Materials Laboratory in Tumwater, a maximum of 14
calendar days will be required for this testing. Source approval shall not be the basis of acceptance of
specific lots of material delivered to the Contractor unless the roll numbers of the lot sampled can be
clearly identified as the rolls tested and approved in the geosynthetic approval process.
For geogrid and geotextile products proposed for use in permanent geosynthetic retaining walls or
reinforced slopes that are not listed in the current QPL, the Contractor shall submit test information
and the calculations used in the determination of Tal, performed in accordance with WSDOT Standard
Practice T925, Standard Practice for Determination of Long-Term Strength for Geosynthetic
Reinforcement, to the State Materials Laboratory in Tumwater for evaluation. The Contracting
Agency will require up to 30 calendar days after receipt of the information to complete the evaluation.
The Contractor shall submit to the Engineer the following information regarding each geosynthetic
material proposed for use:
Manufacturer's name and current address,
Full product name,
FHWA NHI-07-092
Geosynthetic Engineering

9 72

9 MSE Walls
August 2008

Geosynthetic structure, including fiber/yam type,


Geosynthetic polymer type(s) (for temporary and permanent geosynthetic retaining walls),
Proposed geosynthetic use(s), and
Certified test results for minimum average roll values.
9-33.4(2) Vacant
9-33.4(3) Acceptance Samples
When the quantities of geosynthetic materials proposed for use in the following geosynthetic
applications are greater than the following amounts, acceptance shall be by satisfactory test report:
Application
Underground Drainage
Temporary or Permanent Geosynthetic Retaining Walls

Geosynthetic Quantity
600 sq. yd.
All quantities

The samples for acceptance testing shall include the information about each geosynthetic roll to be
used as stated in 9-33.4(4).
Samples will be randomly taken by the Engineer at the job site to confirm that the geosynthetic meets
the property values specified.
Approval will be based on testing of samples from each lot. A lot shall be defined for the purposes
of this Specification as all geosynthetic rolls within the consignment (i.e., all rolls sent to the project
site) that were produced by the same manufacturer during a continuous period of production at the
same manufacturing plant and have the same product name. After the samples have arrived at the
State Materials Laboratory in Tumwater, a maximum of 14 calendar days will be required for this
testing.
If the results of the testing show that a geosynthetic lot, as defined, does not meet the properties
required for the specified use as indicated in Tables 1 through 8 in Section 9-33.2, and additional
tables as specified in the Special Provisions, the roll or rolls which were sampled will be rejected.
Geogrids and geotextiles for temporary geosynthetic retaining walls shall meet the requirements of
Table 7, and Table 10 in the Special Provisions. Geogrids and geotextiles for permanent geosynthetic
retaining wall shall meet the requirements of Table 7, and Table 9 in the Special Provisions, and both
geotextile and geogrid acceptance testing shall meet the required ultimate tensile strength Tult, as
provided in the current QPL for the selected product(s). If the selected product(s) are not listed in the
current QPL, the result of the testing for Tult, shall be greater than or equal to Tult as determined from
the product data submitted and approved by the State Materials Laboratory during source material
approval.
Two additional rolls for each roll tested which failed from the lot previously tested will then be
selected at random by the Engineer for sampling and retesting. If the retesting shows that any of the
additional rolls tested do not meet the required properties, the entire lot will be rejected. If the test
results from all the rolls retested meet the required properties, the entire lot minus the roll(s) that
failed will be accepted. All geosynthetic that has defects, deterioration, or damage, as determined by
the Engineer, will also be rejected. All rejected geosynthetic shall be replaced at no additional
expense to the Contracting Agency.

FHWA NHI-07-092
Geosynthetic Engineering

9 73

9 MSE Walls
August 2008

9-33.4(5) Approval of Seams


If the geotextile seams are to be sewn in the field, the Contractor shall provide a section of sewn scam
that can be sampled by the Engineer before the geotextile is installed. The scam sewn for sampling
shall be sewn using the same equipment and procedures as will be used to sew the production seams.
If production seams will be sewn in both the machine and cross-machine directions, the Contractor
must provide sewn seams for sampling which are oriented in both the machine and cross-machine
directions.
The seam sewn for sampling must be at least 2 yards in length in each geotextile direction. If the
seams are sewn in the factory, the Engineer will obtain samples of the factory scam at random from
any of the rolls to be used. The seam assembly description shall be submitted by the Contractor to the
Engineer and will be included with the scam sample obtained for testing. This description shall
include the seam type, stitch type, sewing thread type(s), and stitch density.

9.10

CONSTRUCTION PROCEDURES

9.10-1
Concrete Faced Wall
The construction of a geosynthetic MSE system with precast facing elements requires the
following steps.
A.

B.

Prepare Subgrade
1.

The foundation should be excavated to the grade as shown on the plans.

2.

The excavated areas should be carefully inspected. Any unsuitable


foundation soils should be compacted or excavated and replaced with
compacted select backfill material.

3.

Foundation soil at the base of the wall excavation should be proofrolled


with a vibratory or rubber-tired roller.

Install Drainage Components


1.

C.

Install base drain, back drain, chimney drain and drainage outlets as
required.

Leveling Pad
1.

A cast-in-place or precast concrete leveling pad should be placed at the


foundation elevation for all MSE structures with concrete (panel and MBW
unit) facing elements. The unreinforced concrete pad is often only 1 foot

FHWA NHI-07-092
Geosynthetic Engineering

9 74

9 MSE Walls
August 2008

(0.3 m) wide and 6 in. (150 mm) thick. The purpose of the pad is to serve
as a guide for facing panel erection and not to act as a structural foundation
support.
D.

Erection of Facing Units


1.

2.

E.

F.

Backfill Placement and Compaction


1.

The backfill material should be placed over a compacted lift thickness as


specified.

2.

The backfill material should be compacted to specified density, usually 95


to 100% of AASHTO T-99 maximum density.

3.

A key to good performance is consistent compaction. Wall fill lift


thickness must be controlled, based upon specification requirements and
vertical distribution of reinforcement elements (and incremental face unit
height).

Reinforcement Placement
1.

G.

The first row of facing panels may be full- or half-height panels, depending
upon the type of facing utilized. The first tier of panels must be shored up
to maintain stability and alignment. For construction with MBW units,
full-sized blocks are used throughout, with no shoring required.
Erection of subsequent rows of facing panels proceed incremental with fill
placement and compaction.

The geosynthetic reinforcements are placed and connected to the facing


units when the fill has been brought up to the level of the connection.
Reinforcement is generally placed perpendicular to the back of the facing
units.

Placement of Fill on Reinforcement


1.

Geosynthetic reinforcement shall be pulled taut and anchored prior to


placing fill.

FHWA NHI-07-092
Geosynthetic Engineering

9 75

9 MSE Walls
August 2008

2.

Fill placement and spreading should prevent or minimize formation of


wrinkles in the geosynthetic. Wrinkles or slack near the facing connection
should be prevented as they can result in differential movement of the wall
face.

3.

A minimum thickness of 6 inches (150 mm) of fill must be maintained


between the tracks of the construction equipment and the reinforcement at
all times.

9.10-2
Geotextile Wrap-Around Wall
Construction procedures for geosynthetic MSE walls are straightforward. Experience of the
U.S. Forest Service, the New York Department of Transportation, and the Colorado
Department of Highways has been very valuable in developing technically feasible and
economical construction procedures. These procedures, detailed in Christopher and Holtz
(1985), are outlined below.
A.

B.

Wall Foundation
1.

The foundation should be excavated to the grade as shown on the plans. It


should be graded level for a width equal to the length of reinforcement plus
1 foot (0.3 m).

2.

The excavated areas should be carefully inspected. Any unsuitable


foundation soils should be compacted or excavated and replaced with
compacted select backfill material.

3.

Foundation soil at the base of the wall excavation should be proofrolled


with vibratory or rubber-tired roller.

4.

Install drainage features (e.g., base drain, back drain, chimney drain and
drainage outlets) as required.

Placement of geosynthetic reinforcement


1.

The geosynthetic should be placed with the principal strength (machine)


direction perpendicular to the face of the wall. It should be secured in
place to prevent movement during fill placement.

FHWA NHI-07-092
Geosynthetic Engineering

9 76

9 MSE Walls
August 2008

C.

D.

2.

It may be more convenient to unroll the geosynthetic with the machine


direction parallel to the wall alignment. If this is done, then the crossmachine design tensile strength must be greater or equal to the design
tensile strength.

3.

A minimum of 6-in. (150 mm) overlap is recommended along edges


parallel to the reinforcement for wrapped-faced walls.

4.

If large foundation settlements are anticipated, which might result in


separation between overlap layers, then field-sewing of adjacent geotextile
sheets is recommended. Geogrids should be mechanically fastened in that
direction.

Backfill placement in reinforced section


1.

The backfill material should be placed over the reinforcement with a


compacted lift thickness of 8 in. (200 mm) or as determined by the
Engineer.

2.

The backfill material should be compacted to at least 90% of the ASTM D


1557 or AASHTO T-99 Standard Proctor maximum density at or below the
optimum moisture.
Alternatively, a relative density compaction
specification could be used. For coarse, gravelly backfills, a method-type
compaction specification is appropriate.

3.

When placing and compacting the backfill material, avoid any folding or
movement of the geosynthetic.

4.

A minimum thickness of 6 in. (150 mm) of fill must be maintained between


the wheels of the construction equipment and the reinforcement at all
times.

Face construction and connections


1.

Place the geosynthetic layers using face forms as shown in Figure 9-13,
unless a precast propped panel facing is to be used.

2.

When using temporary support of forms at the face to allow compaction of


the backfill against the wall face, form holders should be placed at the base

FHWA NHI-07-092
Geosynthetic Engineering

9 77

9 MSE Walls
August 2008

of each layer at about 3 feet (1 m) horizontal intervals. Details of


temporary formwork for geosynthetics are shown in Figure 9-13.

9.11

3.

When using geogrids, it may be necessary to use a geotextile or wire mesh


to retain the backfill material at the wall face (Figure 9-14).

4.

A hand-operated vibratory compactor is recommended when compacting


backfill within 2 feet (0.6 m) of the wall face.

5.

The return-type method shown in Figure 9-4a can be used for facing
support. The geosynthetic is folded at the face over the backfill material,
with a minimum return length of 3 feet (1 m) to ensure adequate pullout
resistance.

6.

Apply facing treatment (shotcrete, precast facing panels, etc.). Figure 9-4
shows several facing alternatives for geosynthetic walls.

INSPECTION

As with all geosynthetic construction, and especially with critical structures such as MSE
walls and abutments, competent and professional construction inspection is absolutely
essential for a successful project. The Engineer should develop procedures to ensure that the
specified material is delivered to the project, that the geosynthetic is not damaged during
construction, and that the specified sequence of construction operations are explicitly
followed. Inspectors should use the checklist in Section 1.7. Other important MSE wall
details that should be monitored and documented include:

Drainage system components and installation


Installation of leveling pad
Vertical spacing of reinforcements
Length of primary and secondary reinforcements as well as wrap-facing return
Construction of the wall face, facing alignment and application of the facing
treatment to minimize geosynthetic exposure to ultraviolet light.

FHWA NHI-07-092
Geosynthetic Engineering

9 78

9 MSE Walls
August 2008

Figure 9-13.

Lift construction sequence for geotextile reinforced soil walls (Steward et


al., 1977).

FHWA NHI-07-092
Geosynthetic Engineering

9 79

9 MSE Walls
August 2008

Figure 9-14. Typical face construction detail for vertical geogrid-reinforced retaining
wall faces.

As-built surveys should be performed on all MSE walls to identify the final vertical and
horizontal position of the face. It is also recommended that permanent survey points be
placed at the top of the wall for future monitoring (e.g., post construction settlement or
unanticipated outward movement), if required. Because geosynthetic MSE retaining walls
are sometimes considered experimental (by some agencies), they often are instrumented.
Sometimes inclinometers and/or multiple-point extensometers are used for observing
potential horizontal movements. On major projects, strain gages are placed on the
geosynthetic to measure internal strains (Allen et al., 1991).

9.12

IMPLEMENTATION

There are two primary issues regarding implementation of geosynthetic MSE wall
technology within a transportation agency. They are: determining who will complete and be
responsible for final design; and specifying determination of the allowable and design tensile
strengths of geosynthetic reinforcement.
FHWA NHI-07-092
Geosynthetic Engineering

9 80

9 MSE Walls
August 2008

9.12-1
Design Responsibility
MSE walls may be contracted using two different approaches. MSE walls can be contracted
on the basis of:
C in-house (agency) design with geosynthetic reinforcement, facing, drainage, and
construction execution specified; or
C system or end-result design approach using approved systems with lines and grades
noted on the drawings.
Both options are acceptable, but of course, the in-house design approach is the preferred
engineering approach. The in-house option will enable agency engineers to examine more
facing and reinforcement options during design. This option requires engineering staff
trained in MSE technology. Design responsibility is, however, well defined. This trained
staff would be valuable during construction, when questions and/or design modification
requests arise.
A variation of in-house design is the use of standardized designs. An example of this
approach is discussed in 9.12-2.
The end-result approach, with sound specifications and prequalification of suppliers and
materials, may offer some benefits. Design of the MSE structure is completed by staff
experienced with the specific system, though not necessarily experienced with local soil and
construction conditions. The prequalified material components of geosynthetic and facing
units have been successfully and routinely used together. While the system specification
approach lessens the engineering requirements for an agency, and transfers some of a
project's design cost to construction, design responsibility must be clearly established: In this
case, the designer does not work for the owner!

9.12-2 Standardized Designs (after Elias et al., 2001)


MSEW structures are customarily designed on a project-specific basis. Most agencies use an
end result (i.e., line-and-grade) contracting approach, with the contractor selected MSEW
vendor providing the detailed design after contract bid and award. This approach works well
for segmental and full-height panel faced walls, and can be used for MBW unit faced walls.
However, standard designs can be developed and implemented by an agency for MSEW
structures, somewhat similar to standard concrete cantilever wall designs used by many
agencies.
Use of standard designs for MSEW structures could offer the following advantages over an
end result approach:
C Agency is more responsible for design details and integrating wall design with other
components.
FHWA NHI-07-092
Geosynthetic Engineering

9 81

9 MSE Walls
August 2008

C
C
C
C
C

Pre evaluation and approval of materials and material combinations, as opposed to


evaluating contractor submittal post bid.
Economy of agency design versus vendor design/stamping of small walls.
Agency makes design decisions versus vendors making design decisions.
More equitable bid environment as agency is responsible for design details, and
vendors are not making varying assumptions.
Filters out substandard work, systems and designs with associated approved product
lists.

The Minnesota Department of Transportation (MN/DOT), with support of the FHWA (via
Demo 82 project) recently developed and implemented standardized MSEW designs (Berg,
2000) for MBW unit faced and geosynthetic reinforced MSEW structures. The use of the
standard designs are limited by material1, geometric, subsurface and economic constraints.
Structures outside of these constraints should be designed on a project-specific basis. The
general approach used in developing these standards could be followed by other agencies to
develop their own, agency-specific standard designs.
Standardized designs require generic designs and generic materials. Generic designs require
definition of wall geometry and surcharge loads, soil reinforcement strength, structure height
limit, and MBW unit properties of width and batter. As an example, the MN/DOT standard
designs address three geometric and surcharge loadings, and can be used for walls up to 12
feet1 (3.7 m) in height.
Definition of generic material properties for the standard designs requires the development of
approved product lists for MBW units, soil reinforcement, and MBW unit-soil reinforcement
combinations. The combinations require a separate approved product list as the connection
strength is specific to each unique combination of MBW unit and reinforcement, and often
controls the reinforcement design strength. An additional requirement for MBW units is an
approved manufacturing quality control plan on file with the agency. This requirement is a
result of the stringent durability (to freeze-thaw and deicing salt conditions) specifications for
the units and the long duration testing used to demonstrate durability.
An example design cross section and reinforcement layout table from the MN/DOT standard
designs is presented in Figure 9-15. Note that the MN/DOT standard designs are not directly
applicable to, nor should they be used by, other agencies.

Concern over long-term durability of MBW units in freeze-thaw and highway deicing salt
environment has resulted in a 12-foot height limit of MBW unit walls on MN/DOT projects. Original
designs were for walls up to 23 feet (7 m) in height.

FHWA NHI-07-092
Geosynthetic Engineering

9 82

9 MSE Walls
August 2008

a. Typical cross section.


Figure 9-15. Example of standard MSE wall design (Mn/DOT, 2006).

FHWA NHI-07-092
Geosynthetic Engineering

9 83

9 MSE Walls
August 2008

b. Cross section notes.

c. Table of reinforcement layout options.


Figure 9-15. Example of standard MSE wall design (Mn/DOT, 2006). (cont.)
FHWA NHI-07-092
Geosynthetic Engineering

9 84

9 MSE Walls
August 2008

9.12-3 Geosynthetic Design Strength


Another issue difficult for many agencies to address is the evaluation and specification of the
allowable and design tensile strengths of geosynthetic reinforcement. The procedure for
evaluation, as summarized within this chapter and in previous chapters on reinforced slopes,
is detailed in Appendix H.
This recommended procedure is based upon the assumption that materials will be
prequalified and listed on an approved products list as specifications. The recommended
requirements for supplier submissions and for agency review, and recommended delineation
of responsibilities within a typical agency, are presented in the FHWA guidelines (Berg,
1993). This procedure has been cumbersome for agencies that do not use approved-products
lists or that review and approve products based upon specific project submittals.
Because of these implementation problems, an alternative starting point procedure for
determining long-term allowable design strength of geosynthetic soil reinforcing elements
has been prepared and is presented in Appendix H. This new, proposed procedure is meant
to complement, and not supersede, the recommended procedure of detailed testing and
evaluation of geosynthetic reinforcement materials.

9.13

SUMMARY OF LOAD RESISTANCE FACTOR DESIGN

9.13-1 Introduction
FHWA and the AASHTO subcommittee on Bridges and Substructures established an
October 1, 2007 deadline for full implementation of load resistance factor design (LRFD).
For many years, engineers have designed walls for highway and other applications using
allowable stress design (ASD) methods. In an ASD design, all uncertainty in applied loads
and material resistance are combined in a factor of safety or allowable stress. In LRFD,
uncertainty in load and material resistance are accounted for separately. The uncertainty in
load is represented by a load factor and the uncertainty in material resistance is represented
by a resistance factor. The load factors are generally greater than one, and differ in
magnitude for various load types. The resistance factor value is usually less than one.
Although the implementation of LRFD requires a change in design procedures for engineers
accustomed to ASD, many advantages do exist. LRFD accounts for variability in both
resistance and load and can provide more consistent levels of safety in the superstructure and
substructure as both are designed using the same loads. Section 11 of the LRFD AASHTO

FHWA NHI-07-092
Geosynthetic Engineering

9 85

9 MSE Walls
August 2008

Bridge and Substructures Specification provides a detailed summary of the LRFD design of
earth retaining structures including mechanically stabilized earth (MSE) walls.
NHI Course 132036 titled, Earth Retaining Structures and reference manual (Tanyu et al.,
2007) contains comprehensive information on LRFD for wall systems. In this section,
LRFD-based analyses for MSE walls are summarized and similarities and differences
between existing ASD design recommendations and LRFD design recommendations are
described. This section is not intended to provide the information necessary to perform the
complete design or design review of an MSE wall system using the LRFD method.

9.13-2 Background
LRFD accounts for the variability in both resistance and load by utilizing separate resistance
and load factors as shown in the following equation.

Q
i

Rn

[9-22]

where, i = Load factor;


Qi = Load (e.g. dead load, live load, resultant lateral earth pressure load, etc.);
= Resistance factor; and
Rn = Nominal resistance or strength of the element under consideration.
Load and resistance factors have been calibrated in an effort to obtain a more uniform level
of safety for different limit states and types of materials and to provide reasonably consistent
designs to those based on ASD. The most rigorous method for developing and adjusting
resistance factors requires availability of statistical data.
LRFD is a limit state design procedure. A limit state is a condition beyond which a
structural component ceases to satisfy the provisions for which it was designed. Limit states
include strength limits, service limits, and extreme event limits. Strength limit states describe
a condition that results in the total or partial collapse of the structure. Examples of strength
limit states in earth retaining structures include foundation bearing failure, substructure
sliding, and failure of a wall element in flexure. Service limit states affect the function of the
structure under regular service conditions. For earth retaining structures, service limit states
can be reached through excessive total and differential settlement, excessive lateral wall
movement, or overall stability. Extreme event limit states represent a low probability design
condition including vehicle impact or seismic events.

FHWA NHI-07-092
Geosynthetic Engineering

9 86

9 MSE Walls
August 2008

As described by Equation 9-21, factored loads or combinations of factored loads are


compared to the available factored resistance of the earth retaining system. Section 3 of
AASHTO (2007) provides a comprehensive list of all load combinations to be evaluated for a
bridge substructure or earth retaining system. While all load groups should be considered, in
general, the following load groups are typically used in the LRFD design of earth retaining
systems: Strength Ia, Strength Ib, and Service I. Strength Ia is the critical load combination
to evaluate sliding and eccentricity of retaining structures and Strength Ib is the critical load
combination to evaluate bearing resistance. The uncertainty in the loads applied to the
retaining structure are addressed by applying a load factor, (. The load factors for each of
the three loading groups are given in Table 9-4, where DC is the dead load of structural
components and attachments, EV is the vertical pressure from dead load of earth fill, EH is
the horizontal earth pressure load, ES is the earth surcharge load, and LS is the live load
surcharge (transient load).

Table 9-4. Common LRFD Load Groups for Walls.


Group

DC

EV

EH
(Active)

EH
(At Rest)

ES

LS

Strength Ia

0.90

1.00

1.50

1.35

1.50

1.75

Strength Ib

1.25

1.35

1.50

1.35

1.50

1.75

Service I

1.00

1.00

1.00

1.00

1.00

1.00

9.13-3 MSE Wall Design


With few exceptions, the procedure for design of MSE walls using LRFD is identical to that
followed using ASD. External stability considerations include bearing resistance, sliding,
evaluation of eccentricity and overall stability. Internal stability considerations include
resistance to pullout and tensile rupture, resistance to connection rupture at the wall face, and
structural resistance of the wall facing. Settlement and lateral wall deformations must also be
checked.

9.13-4 External Stability


In ASD, the surcharge over the reinforced zone is only accounted for in certain situations.
Figure 9-16 illustrates the typical application of live load surcharge over the retained fill for
an MSE wall in LRFD design. When a live load surcharge is considered in design, the
factored surcharge force is generally included over the backfill immediately above the wall
only for evaluation of foundation bearing resistance and tensile resistance of the
FHWA NHI-07-092
Geosynthetic Engineering

9 87

9 MSE Walls
August 2008

reinforcement strips. The live load surcharge is not included over the backfill for evaluation
of eccentricity, sliding or other failure mechanisms for which the surcharge load would
increase resistance to failure.

Figure 9-16. Typical application of live load surcharge for MSE walls.

An equivalent footing having a width equal to the length of the reinforcing elements at the
foundation level can be assumed for the evaluation of bearing resistance. Due to the
flexibility of the MSE wall and the inability of the flexible reinforcement to transmit
moment, a uniform base pressure distribution is assumed over this equivalent footing width.
The required ASD factor of safety for bearing resistance ranges from 2.0 to 2.5 (AASHTO,
2002). Depending on the method of calculation, the LRFD resistance factor for bearing
resistance ranges from 0.35 to 0.60 (see Table 10.5.5-1 in AASHTO, 2007).

9.13-5 Internal Stability


For LRFD, the maximum load in a reinforcement within the wall cross-section; Tmax, is
evaluated at each reinforcement level. The maximum load is compared to the pullout
resistance and to the tensile resistance in the reinforcement. Geosynthetic reinforcement
tensile rupture resistance factors are 0.90 for static loading, and 1.20 for combined static
/earthquake loading.
In AASHTO Equation 11.10.6.3.2-1, the value for Tmax is based on calculated horizontal loads
with a load factor of 1.35 for the vertical earth pressure, EV. It is noted that increases in
horizontal loads resulting from surcharges, i.e. H, are also factored by the load factor of
1.35. A pullout resistance factor of 0.9 (see Table 11.5.6-1 in AASHTO, 2007) is used for
static loading. All other parameters in AASHTO Equation 11.10.6.3.2-1 are formulated
exactly as for ASD.
FHWA NHI-07-092
Geosynthetic Engineering

9 88

9 MSE Walls
August 2008

9.14 REFERENCES
AASHTO (2007). AASHTO LRFD Bridge Design Specifications, American Association of
Highway and Transportation Officials, 4th Edition, Washington, DC.
AASHTO (2002). Standard Specifications for Highway Bridges, American Association of
Highway and Transportation Officials, 17th Edition, Washington, DC.
AASHTO (1996). Standard Specifications for Highway Bridges, Sixteenth Edition, with
1997 Interims, American Association of State Transportation and Highway Officials,
Washington, D.C.
AASHTO (1990). Design Guidelines for Use of Extensible Reinforcements (Geosynthetic)
for Mechanically Stabilized Earth Walls in Permanent Applications, Task Force 27
Report In-Situ Soil Improvement Techniques, American Association of State
Transportation and Highway Officials, Washington, D.C.
ASTM Test Methods see Appendix E.
Allen, T.M., Christopher, B.R. and Holtz, R.D. (1992). Performance of a 12.6 m High
Geotextile Wall in Seattle, Washington, Geosynthetics Reinforced Soil Retaining
Walls, J.T.H. Wu Editor, A.A. Balkema, Rotterdam, pp. 81-100.
Allen, T.M. and Holtz, R.D. (1991). Design of Retaining Walls Reinforced with
Geosynthetics, Geotechnical Special Publication No. 27, Proceedings of ASCE
Geotechnical Engineering Congress 1991, American Society of Civil Engineers, New
York, Vol. II, pp. 970-987.
Allen, T.M. (1991). Determination of Long-Term Tensile Strength of Geosynthetics: A
State-of-the-Art Review, Proceedings of Geosynthetics '91 Conference, Atlanta, GA,
Vol. 1, pp. 351-380.
Bathurst, R. J. and Simac, M.R. (1994). Geosynthetic Reinforced Segmental Retaining Wall
Structures in North America, Preprint of Special Lecture and Keynote Lectures of
Fifth International Conference on Geotextiles, Geomembranes and Related Products,
Singapore, September, pp. 31-54.
Bell, J.R., Barrett, R.K. and Ruckman, A.C. (1983). Geotextile Earth-Reinforced Retaining
Wall Tests: Glenwood Canyon, Colorado, Transportation Research Record 916, pp.
59-69.
Bell, J.R., Stiley, A.N. and Vandre, B. (1975). Fabric Retained Earth Walls, Proceedings of
the 13th Annual Engineering Geology and Soils Engineering Symposium, Moscow,
ID, pp. 271-287.

FHWA NHI-07-092
Geosynthetic Engineering

9 89

9 MSE Walls
August 2008

Berg, R.R. (1993). Guidelines for Design, Specification, & Contracting of Geosynthetic
Mechanically Stabilized Earth Slopes on Firm Foundations, FHWA-SA-93-025,
Federal Highway Administration, Washington, D.C., 87p.
Berg, R.R. (1991) The Technique of Building Highway Retaining Walls, Geotechnical
Fabrics Report, Industrial Fabrics Association International, St. Paul, MN, July, pp.
38-43.
Bonaparte, R. and Berg, R.R. (1987). Long-Term Allowable Tension for Geosynthetic
Reinforcement, Proceedings of Geosynthetics '87 Conference, Volume 1, New
Orleans, LA, pp. 181-192.
Carroll, R.G. and Richardson, G.N. (1986). Geosynthetic Reinforced Retaining Walls,
Proceedings of The Third International Conference on Geotextiles, Austria, Vienna,
Vol. II, pp. 389-394.

Christopher, B.R., Gill, S.A., Giroud, J.P., Juran, I. Scholsser, F., Mitchell, J.K. and
Dunnicliff, J. (1989). Reinforced Soil Structures, Volume I. Design and
Construction Guidelines, FHWA-RD-89-043, Federal Highway Administration,
Washington, D.C., 287 p.
Christopher, B.R. and Holtz, R.D. (1989). Geotextile Design and Construction Guidelines,
FHWA-HI-90-001, Federal Highway Administration, Washington, D.C., 297 p.
Christopher, B.R. and Holtz, R.D. (1985). Geotextile Engineering Manual, FHWA-TS86/203, Federal Highway Administration, Washington, D.C., 1044 p.
Claybourn, A.F. and Wu, J.T.H. (1993). Geosynthetic-Reinforced Soil Wall Design,
Geotextiles and Geomembranes, Vol. 12, No. 8, pp. 707-724.
Collin, J.G., Berg, R.R. and Meyers, M. (2002). Segmental Retaining Wall Drainage
Manual, National Concrete Masonry Association, Herdon, VA, 96 p.

Elias, V., Christopher, B.R. and Berg, R.R. (2001). Mechanically Stabilized Earth
Walls and Reinforced Soil Slopes, Design & Construction Guidelines, FHWANHI-00-043, Federal Highway Administration, Washington, D.C., 418 p.
Elias, V. (2000). Corrosion/Degradation of Soil Reinforcements for Mechanically
Stabilized Earth Walls and Reinforced Soil Slopes, FHWA-NHI-00-044, FHWARD--89-043, Federal Highway Administration, Washington, D.C., 94 p.
Elias, V. and Christopher, B.R. (1997). Mechanically Stabilized Earth Walls and Reinforced
Soil Slopes, Design & Construction Guidelines, FHWA-SA-96-071, Federal
Highway Administration, Washington, D.C., 371 p.

GRI Test Standards - see Appendix E.


FHWA NHI-07-092
Geosynthetic Engineering

9 90

9 MSE Walls
August 2008

GRI (1990). Test Method GG4a, Determination of Long-Term Design Strength of Stiff
Geogrids, Geosynthetic Research Institute, Drexel University, Philadelphia, PA.
GRI (1991). Test Method GG4b, Determination of Long-Term Design Strength of Flexible
Geogrids, Geosynthetic Research Institute, Drexel University, Philadelphia, PA.
GRI (1992). Test Method GT7, Determination of Long-Term Design Strength of Geotextiles,
Geosynthetic Research Institute, Drexel University, Philadelphia, PA.
Holtz, R.D., Allen, T.M. and Christopher, B.R. (1991). Displacement of a 12.6 m High
Geotextile-Reinforced Wall, Proceedings of the Tenth European Conference on Soil
Mechanics and Foundation Engineering, Florence, Italy, Vol. 2, 1991, pp. 725-728.
HITEC (2003). Evaluation of Anchor Wall Systems Landmark Reinforced Soil Wall
System with T.C. Mirafis Miragrid and Miratex Geogrid Soil Reinforcement,
CERF Report #40677, American Society of Civil Engineers, Reston, VA.
Jaky, J. (1948). Earth Pressure in Silos, Proceedings of the Second International Conference
on Soil Mechanics and Foundation Engineering, Rotterdam, Vol. I, pp.103-107.
Koerner, J., Soong, T-Y, and Koerner, R. M. (1998). Earth Retaining Wall Costs in the USA,
GRI Report #20, Geosynthetics Institute, Folsom, PA.
Lee, K.L., Adams, B.D., and Vagneron, J.M.J. (1973). Reinforced Earth Retaining Walls,
Journal of the Soil Mechanics and Foundations Division, American Society of Civil
Engineers, New York, Vol. 99, No. SM10, 1973, pp. 745-764.
Minnesota Department of Transportation (2006). Retaining Wall Standards, Office of
Technical Support, Minnesota Department of Transportation, www.dot.state.mn.us.
Mitchell, J.K. and Villet, W.C.B. (1987). Reinforcement of Earth Slopes and Embankments,
NCHRP Report No. 290, Transportation Research Board, Washington, D.C.
Munfakh G., Arman, A., Collin, J.G., Hung, J. and Brouillette, R.P. (2001). Shallow
Foundations Reference Manual, FHWA-NHI-01-023, Federal Highway
Administration, Washington, D.C., 222 p.
NCMA (1997). Design Manual for Segmental Retaining Walls, 2nd Edition, J.G. Collin Ed.,
National Concrete Masonry Association, Herdon, VA, 289 p.
New York Department of Transportation (2006). 704-07 Segmental Retaining Wall Blocks,
Standard Specifications, New York Department of Transportation, www.nysdot.gov.

Special Provision Item 554.05


New York Department of Transportation (2004).
Mechanically Stabilized Segmental Block Retaining Wall System (Extensible
Reinforcement), New York Department of Transportation, www.nysdot.gov.
FHWA NHI-07-092
Geosynthetic Engineering

9 91

9 MSE Walls
August 2008

Perloff, W.H. and Baron, W. (1976). Soil Mechanics: Principles and Applications, Ronald
Press, 504 p.
Samtani, N.C. and Nowatzki, E.A. (2006). Soils and Foundations Reference Manual
Volume II, FHWA-NHI-06-089, Federal Highway Administration, Washington,
D.C., 594 p.
Simac, M.R., Bathurst, R.J., Berg, R.R. and Lothspiech, S.E. (1993). Design Manual for
Segmental Retaining Walls, National Concrete Masonry Association, Herdon, VA,
336 p.
Steward, J., Williamson, R. and Mohney, J. (1977). Guidelines for Use of Fabrics in
Construction and Maintenance of Low-Volume Roads, USDA, Forest Service,
Portland, OR. Also reprinted as Report No. FHWA-TS-78-205.
Tanyu, B.F., Sabatini, P.J. and Berg, R.R. (2007). Earth Retaining Structures, U.S.
Department of Transportation, Federal Highway Administration, Washington, D.C.,
FHWA-NHI-07-071, 668 p.
Terzaghi, K. and Peck, R.B. (1967). Soil Mechanics in Engineering Practice, John Wiley &
Sons, New York, 729 p.
U.S. Department of the Navy (1986). Design Manual 7.01 - Soil Mechanics, Department of
the Navy, Naval Facilities Engineering Command, Alexandria, VA. (can be
downloaded from http://www.geotechlinks.com).
Washington State Department of Transportation (2008). 6-14 Geosynthetic Retaining Walls
and 9-33 Construction Geosynthetic, Standard Specifications, Washington
Department of Transportation, www.wsdot.wa.gov.
Wrigley, N.E. (1987). Durability and Long-Term Performance of 'Tensar' Polymer Grids for
Soil Reinforcement, Materials Science and Technology, Vol. 3, pp. 161-170.
Wu, T.H. (1975). Retaining Walls, Chapter 12 in Foundation Engineering Handbook,
Winterkorn and Fang, Editors, Von Nostrand Reinhold, pp. 402-417.

FHWA NHI-07-092
Geosynthetic Engineering

9 92

9 MSE Walls
August 2008

10.0 GEOMEMBRANES AND OTHER GEOSYNTHETIC BARRIERS

10.1 BACKGROUND
Barriers are used in earthwork construction to control movement of water, other liquids and
sometimes vapors. Barriers are used to waterproof structures, to prevent moisture changes
beneath roadways, to contain water and wastes, and to support other applications in
transportation works. The function of these barriers is to either prevent damage to highway
pavements and structures or to contain water or waste materials. Barriers must be engineered
to perform their intended function for the particular application and project being designed.
Traditional barriers, or liners, are field-constructed of soil or aggregate-based materials.
Thick compacted clay layers, cast-in-place concrete, and asphalt concrete are used to
construct liners. Another conventional liner material is geomembranes, which have been
used in transportation applications for more than fifty years. The U.S. Bureau of
Reclamation has been using geomembranes in water conveyance canals since the 1950s
(Staff, 1984). Other types of geosynthetic barriers have also been used in transportation
applications. These include thin-film geotextile composites, geosynthetic clay liners, and
field-impregnated geotextiles.
While soil or aggregate-based liners are well-suited to some applications, geomembrane and
other geosynthetic barriers are more appropriate for other projects. Suitability may be
defined during design and with due consideration to material availability, long-term
performance, and cost. For example, rigid concrete and asphalt liners or semi-stiff
compacted clay liners are not well-suited to sites where barriers are subject to foundation
settlements; conversely a geomembrane which has adequate flexibility would be suitable.

10.2 GEOSYNTHETIC BARRIER MATERIALS


Geosynthetic barrier materials can be classified as geomembranes, thin-film geotextile
composites, geosynthetic clay liners, or field-impregnated geotextiles. Materials within each
classification are reviewed herein, starting with a general material definition. The
components, manufacturing processes, resultant product characteristics, and typical
dimensions are presented. Available test standards used to quantify property values are listed
in Appendix E.

FHWA NHI-07-092
Geosynthetics Engineering

10-1

Geomembranes and Barriers


August 2008

10.2-1 Geomembranes
The term geomembrane is defined as an essentially impermeable geosynthetic composed of
one or more synthetic sheets (ASTM D 4439). However, within this document the term
geomembrane will be used to specifically describe materials which are manufactured of
continuous polymeric sheets. Commonly available geomembranes are manufactured of the
polymers listed in Table 10-1.
Table 10-1
Common Types of Geomembranes
Abbreviation

Available without
scrim
reinforcement

Available with
scrim
reinforcement

Chlorinated Polyethylene

CPE

Chlorosulfonated Polyethylene

CSPE

POLYMER TYPE

Ethylene Interpolymer Alloy


High-Density Polyethylene
Polypropylene
Polyvinyl Chloride
Linear Low-Density Polyethylene

EIA

HDPE

PP

PVC

LLDPE

The HDPE and LLDPE geomembranes are supplied in roll form. Widths of approximately
15 to 35 feet (4.6 to 10.5 m) are available. Roll lengths of 650 to 1,000 feet (200 to 300 m)
are typical, although custom roll lengths are available. These materials are generally
available in sheet thicknesses of 40, 60, 80 and 100 mil (1.0, 1.5, 2.0, and 2.5 mm). These
PE geomembranes are usually of singular (i.e., not composite) manufacture. However, thick
coextruded PE composites are available with light-colored heat reflective surfaces, electrical
conductive surfaces (for leakage testing), or a LLDPE layer sandwiched between an upper
and lower layer of HDPE.
Geomembranes manufactured of CPE, CSPE, PVC, and PP are supplied in large panels,
accordion-folded onto pallets. The panels are fabricated by factory-seaming of rolls,
typically 4.6 to 8.2 feet (1.4 to 2.5 m) in width. Panels as large as 20,000 ft2 (1,800 m2) are
available. These geomembranes may be of singular manufacture or composites with a fabric
scrim incorporated to modify the mechanical properties (see Table 10-2).

FHWA NHI-07-092
Geosynthetics Engineering

10-2

Geomembranes and Barriers


August 2008

Geomembranes are relatively impermeable materials - i.e., all materials are permeable, but
the permeability of geomembranes (on the order of 10-14 m/s) is significantly lower than that
of compacted clays. Hence, geomembranes are sometimes referred to as being impermeable,
relative to soil. Theoretically, multiple layers of geomembranes and drains can be utilized to
construct impermeable structures (Giroud, 1984).
Leakage, and not permeability, is the primary concern when designing geomembrane
containment structures. Leakage can occur through poor field seams, poor factory seams,
pinholes from manufacture, and puncture holes from handling, placement, or in-service
loads. Leakage of geomembrane liner systems is minimized by attention to design,
specification, testing, quality control (QC), quality assurance (QA) of manufacture, and QC
and QA of construction.
Geomembrane materials are better-defined than other geosynthetic barriers, due to their
widespread use in environmental applications and their historical use in other applications.
Manufacturing QC/QA standards, index property test methods, performance test methods,
design requirements, design detailing, and construction QC/QA are well-established (Daniel
and Koerner, 1993).
10.2-2 Thin-Film Geotextile Composites
The moisture barrier commonly used in roadway reconstruction is a thin-film geotextile
composite. These composites are used to prevent or minimize moisture changes in pavement
subgrades, as discussed in Section 10.3. Two styles of composites, as illustrated in Figure
10-1, are available.
One commercially available composite consists of a very lightweight PP nonwoven
geotextile sandwiched between two layers of PP film (Figure 10-1a). This product has a
mass per unit area of 4 oz/yd2 (140 g/m2) and is available in roll widths of 7.5 feet (2.3 m)
and roll lengths of 300 feet (91 m).
Another commercially available composite consists of a polyethylene (PE) film sandwiched
between two layers of nonwoven PP geotextiles (Figure 10-1b). This product has a mass per
unit area of 9 oz/yd2 (300 g/m2) and is available in roll widths of 12 feet (3.65 m) and roll
lengths of 300 feet (91 m). Similar products are available with PVC, CSPE, and PP
geomembranes as the core.

FHWA NHI-07-092
Geosynthetics Engineering

10-3

Geomembranes and Barriers


August 2008

Figure 10-1. Thin-film geotextile composites: (a) PP film / PP geotextile / PP film; and (b)
PP geotextile - PE film - PP geotextile.
10.2-3 Geosynthetic Clay Liners
Geosynthetic clay liners (GCLs) are another type of composite barrier materials. A dry
bentonite clay soil is supported between two geotextiles or on a geomembrane carrier, as
illustrated in Figure 10-2. Geotextiles used above and below the dry clay may or may not be
connected with threads or fibers, to increase the in-plane shear strength of a hydrated GCL.
Approximately 1 lb/ft2 (5 kg/m2) of dry sodium bentonite is used in the manufacture of
GCLs. The bentonite is at a moisture content of 6 to 20% in its dry condition. This dry
bentonite hydrates and swells upon wetting, creating a very-low permeability barrier. The
fully hydrated bentonite typically will have a permeability in the range of 1 to 5 x 10-11
m/sec.
GCLs are supplied in roll form. Widths of approximately 14.5 to 17.5 feet (4.4 to 5.3 m) are
available. Roll lengths of 100 to 200 feet (30 to 60 m) are typical, although custom roll
lengths may be used for large projects.

Figure 10-2. Geosynthetic clay liners: (a) geotextile/bentonite clay/geotextile; (b) stitched
bonded geotextile GCL; (c) needle punched geotextile GCL; and (d)
bentonite clay/PE geomembrane (after Koerner, 1998). 1 mm = 39 mil
FHWA NHI-07-092
Geosynthetics Engineering

10-4

Geomembranes and Barriers


August 2008

10.2-4 Field-Impregnated Geotextiles


Impregnated geotextiles are also used as moisture and liquid barriers. The coating treatment
is applied in the field, after the geotextile is deployed and anchored. A nonwoven geotextile
is used with a variety of coatings, including asphalt, rubber-bitumen, emulsified asphalt, or
polymeric formulations. The coating may be proprietary. The geotextile type and mass per
unit area will be a function of the coating treatment, although use of lightweight nonwoven
geotextiles, in the range of 6 to 12 oz/yd2 (200 to 400 g/m2), is common. Heavier-weight,
nonwoven geotextiles may be used to provide gas venting, if gas potential exists on a site.
The barrier is formed as sprayed-on liquid solidifies into a seam-free membrane. Although
sprayed-on membranes are seam-free, bubbles and pinholes may form during installation and
can cause performance problems. Proper preparation of the geotextile (i.e., clean and dry) to
be sprayed is important. These types of barriers have been used in canals, small reservoirs,
and ponds for water control. Water storage applications have used air-blown asphalt
coatings. (Matrecon, 1988)
Engineers also use field-impregnated geotextiles to provide moisture control in friable
roadway soils. Pavement application of a barrier is called membrane encapsulated soil
layers (MESLs).

10.3 APPLICATIONS
Geomembranes and other geosynthetic barriers are used in wide variety of applications for
transportation construction and maintenance. Geomembrane transportation applications, as
summarized by Koerner and Hwu (1989), are summarized below. The applications are noted
as representing regular use or limited use of geomembranes and other synthetic barriers in
highway works.

Control of vertical infiltration of moisture into a subgrade of expansive soil. This


minimizes the change in soil water content and subsequent volume changes.
Placement of the geosynthetic barrier is illustrated in Figure 10-3. Thin-film
geotextile composites or geomembranes are often used in this application.
Control of horizontal infiltration of moisture into a subgrade of expansive soil. This
minimizes change in soil water content and volume changes. Placement of the barrier
is illustrated in Figure 10-4. Depth of moisture barrier is approximately 18 to 24 in.
(450 to 600 mm) beneath estimated swell depth, or a typical total depth of 3 to 5 feet
(1.0 to 1.5 m). Thin-film geotextile composites or geomembranes are usually used in
this application, although field-impregnated geotextiles are also used.

FHWA NHI-07-092
Geosynthetics Engineering

10-5

Geomembranes and Barriers


August 2008

Maintenance of water content of frost-sensitive soils with a horizontally placed


barrier. This application is illustrated in Figure 10-5, as the MESL previously
described. Thin-film geotextile composites or geomembranes are usually used in this
application.

Waterproofing of tunnels, as illustrated in Figure 10-6. Geomembranes (in


conjunction with heavyweight nonwoven geotextiles) and GCLs are used in this
application.

Transport of water in canals lined with a geomembrane, as illustrated in Figure 10-7,


or a GCL.

Geomembranes are used for secondary containment of underground fuel storage


tanks, as illustrated in Figure 10-8. GCLs and geomembranes are also used for
secondary containment of above ground fuel storage tanks.

Rest area waste water treatment lagoons may be lined with geomembranes.

Sealing of berms for wetland mitigation.

Storm water retention and detention ponds also may be lined with geomembranes,
GCLs, or coated geotextiles.

Geomembranes are used beneath structures as methane and radon gas barriers.

Geomembranes are used for containment of waste, caustic soils (e.g., pyritic soils)
and construction debris.

Deicing salt and aviation deicing fluid runoff may be contained in geomembranelined facilities.

Geomembranes, GCLs, and coated geotextiles may be used to waterproof walls and
bridge abutments. Geomembranes may be used to prevent infiltration of corrosive
deicing salt runoff into metallic MSE walls, as illustrated in Figure 10-9.

Railroads use geosynthetic barriers to waterproof subgrades, to prevent upward


groundwater movement in cuts, and to contain diesel spills in refueling areas.

FHWA NHI-07-092
Geosynthetics Engineering

10-6

Geomembranes and Barriers


August 2008

Figure 10-3. Control of expansive soils (from Koerner and Hwu, 1989).

Figure 10-4. Control of horizontal infiltration of base.

FHWA NHI-07-092
Geosynthetics Engineering

10-7

Geomembranes and Barriers


August 2008

Figure 10-5. Maintenance of optimum water content (from Koerner and Hwu, 1989).

Figure 10-6. Waterproofing of tunnels (from Koerner and Hwu, 1989).

FHWA NHI-07-092
Geosynthetics Engineering

10-8

Geomembranes and Barriers


August 2008

Figure 10-7. Water conveyance canals.

Figure 10-8. Secondary containment of underground fuel tanks (from Koerner and Hwu,
1989).
FHWA NHI-07-092
Geosynthetics Engineering

10-9

Geomembranes and Barriers


August 2008

Figure 10-9. Waterproofing of walls (from Koerner and Hwu, 1989).

10.4 DESIGN CONSIDERATIONS


All geosynthetic barriers are continuous materials which are relatively impermeable as
manufactured. However, to fulfill its barrier function the geosynthetic system must remain
leak-proof and relatively impermeable when installed and throughout its design life. The
following steps are part of the design process for geosynthetic barrier systems:
FHWA NHI-07-092
Geosynthetics Engineering

10-10

Geomembranes and Barriers


August 2008

define performance requirements;


design for in-service conditions;
durability design for project-specific conditions;
design for installation, under anticipated project conditions;
peer review (optional); and
economic analysis.

10.4-1 Performance Requirements


The required function and performance of a geosynthetic barrier must be defined
prior to design and material selection. The purpose of the barrier significantly affects
the design and installation requirements. Some of the questions that should be asked
regarding performance include the following.
Is the barrier functioning as a primary or a secondary liner?
Is a barrier less-permeable than adjacent soil required, or is an impermeable barrier
system (liner(s) and drain(s)) required?
What are the consequences of leakage?
Can an acceptable leakage rate be defined?
What is the anticipated life of the system, i.e., is it temporary or permanent?
10.4-2 In-Service Conditions
Applications of a barrier vary, and the in-service exposure and stresses that the barrier must
withstand likewise vary. Some of the questions that should be addressed regarding in-service
conditions include the following.
Will the barrier be placed within soil, or remain exposed to environmental elements
throughout the design life?
What environmental conditions (e.g., temperature variations, sunlight exposure, etc.)
will the geosynthetic barrier be exposed to throughout its design life?
Will the barrier be subject to deformation-controlled (e.g., due to post-construction
movements caused by settlement of underlying soil) stresses ?
Will the geosynthetic barrier be subject to downdrag forces (e.g., on a side slope of a
surface impoundment) or to load-controlled stresses?
Will the barrier system be exposed to varying stress levels due to fluctuating water
loads?
Will the barrier system result in a low-friction interface that must be analyzed?
Will the barrier trap gases and/or liquids generated beneath the liner, and require
venting?
Are performance requirements such that abrasion or puncture protection is required?

FHWA NHI-07-092
Geosynthetics Engineering

10-11

Geomembranes and Barriers


August 2008

10.4-3 Durability
Many geomembranes and other geosynthetic barriers in transportation applications contain
nonpolluted water. As such, geosynthetic chemical durability is not normally a concern.
However, chemical resistance is a concern when liquids such as fuel or other contaminants
must be contained. The chemical resistance of candidate geosynthetic barriers, as well as of
their components (if applicable), must be specifically evaluated when other than nonpolluted
water is to be contained. The EPA 9090 test is available for such an assessment. Available
geosynthetic barriers have a wide range of chemical resistance to various elements and
compounds.
Resistance to ultraviolet light must be assessed for those applications where the barrier
remains exposed over its design life. Oxidation or hydrolytic degradation potential may also
be assessed. Biological degradation potential should also be checked. Degradation due to
vegetation growth, burrowing animals, or microorganisms may be a concern. Biological
degradation of materials beneath a liner can result in gas formation that must be vented
around or through the liner.
10.4-4 Installation Conditions
Installation conditions are a design consideration for all geosynthetics in all applications.
Installation of geomembranes and other geosynthetic barriers is a primary design
consideration. Location and installation time of year can affect barrier material selection.
Environmental factors such as temperature, temperature variation, humidity, rainfall, and
wind must be considered. Some geosynthetics are more sensitive to temperature than others,
and moisture and wind affect field-seaming ability. Barriers constructed of fieldimpregnated coated geotextiles must be placed during carefully defined weather conditions.
Placement, handling, and soil covering operations can also affect geosynthetic design and
selection. The panel weight and size must be compatible with project requirements and
constraints. The timing of soil placement over the liner may dictate ultraviolet light
resistance requirements. And the geomembrane or other geosynthetic barrier must be
capable of withstanding the rigors of installation.
The subgrade material, subgrade preparation, panel deployment method, overlying soil fill
type, and placement and compaction of overlying fill soil all affect the geosynthetic barrier's
survivability. Recommended properties of geomembrane barriers (Koerner, 1998) are
presented in Table 10-2.

FHWA NHI-07-092
Geosynthetics Engineering

10-12

Geomembranes and Barriers


August 2008

Table 10-2
Recommended Minimum Properties for General
Geomembrane Installation Survivability
(from Koerner, 1998)
Required degree of survivability
Property and test method

Low1

Medium2

High3

Very high4

Thickness, mm - ASTM D 1593

0.63

0.75

0.88

1.00

Tensile (25 mm strip), kN/m ASTM D 882

7.0

9.0

11

13

Tear (Die C), N - ASTM D 1004

33

45

67

90

Puncture, N ASTM D 4833

110

140

170

200

Impact, J
ASTM D 3998, modified

11

15

19

NOTES:
1. Low refers to careful hand-placement on very uniform, well-graded subgrade with light loads of a static
nature - typical of vapor barriers beneath building floor slabs.
2. Medium refers to hand- or machine-placement on machine-graded subgrade with medium loads - typical of
canal liners.
3. High refers to hand- or machine-placement on machine-graded subgrade of poor texture with high loads typical of landfill liners and covers.
4. Very high refers to hand- or machine-placement on machine-graded subgrade of very poor texture with high
loads - typical of reservoir covers and liners for heap leach pads.

Geotextiles and other geosynthetics are often used with geomembranes to enhance the
barrier's puncture resistance during installation and in-service. Geotextiles, and other
geosynthetics, act as cushions and further prevent puncture of the geomembrane. The
cushion can be placed below the geomembrane to resist rocks, roots, etc., in the subgrade,
and/or above the geomembrane to resist puncture from subsequently placed fill or waste.
Selection of the most-effective geotextile will depend upon several factors (Richardson and
Koerner, 1990), including:
mass per unit area;
geotextile type;
fiber type;
thickness under load;
polymeric type; and
geomembrane type and thickness.

FHWA NHI-07-092
Geosynthetics Engineering

10-13

Geomembranes and Barriers


August 2008

The level of puncture protection provided by the geotextile is directly related to the mass per
unit area.
10.4-5 Peer Review
A peer, or design quality assurance, review is recommended for landfill barrier systems
(Rowe and Giroud, 1994). A peer review of a geosynthetic barrier structure for a
transportation application may likewise be warranted, depending upon the critical nature of
the structure, experience of design team, and project location and function. The goal of such
a project review is to enhance the quality of the constructed project.
A peer review is recommended (Berg, 1993) for:
projects where performance is crucial to public safety and/or the environment;
projects that are controversial or highly visible;
proposed designs that incorporate new materials or construction techniques;
projects requiring state-of-the-art expertise;
designs that lack redundancy in primary components;
designs that have a poor performance record;
projects with accelerated design and/or construction schedules; and
projects with overlapping design and construction schedules.
10.4-6 Economic Considerations
Cost should be considered in design after function, performance, and installation design
criteria are addressed. Material and in-place costs will obviously vary with the type of
geosynthetic barrier and the quantity of barrier specified. In-place cost of geosynthetic
barriers can vary from approximately $2.50/yd2 to $16.00/yd2. Cost of conventional
compacted clay liners can vary between approximately $5.00/yd2 to $30.00/yd2 in-place.
10.5 INSTALLATION
A well-designed geosynthetic must be installed correctly to perform its function as a barrier.
Handling and installation specifications for geomembrane and other geosynthetic barriers
should, as a minimum, conform to the manufacturer's recommendations. Special project
requirements should also be noted in the construction specifications and plans.
Geosynthetic barrier handling and storage requirements at the construction site should be
specifically designated. Layout of the geosynthetic normally should be predetermined and
documented on a roll or panel layout plan. The installer or geosynthetic supplier is normally
required, by specification, to provide the layout plan.
FHWA NHI-07-092
Geosynthetics Engineering

10-14

Geomembranes and Barriers


August 2008

Three areas of construction which are critical to a successful installation are:


subgrade preparation;
field seaming; and
sealing around penetrations and adjacent structures.
The subgrade must provide support to the geosynthetic barrier and minimal point loadings.
The subgrade must be well-compacted and devoid of large stones, sharp stones, grade stakes,
etc., that could puncture the geosynthetic barrier. In general, no objects greater than -in.
(12 mm) should be protruding above the prepared subgrade (Daniel and Koerner, 1993).
Geotextiles are often used as cushions for geomembranes to increase puncture resistance, as
previously discussed. Geotextiles and geocomposite drains are also used beneath
geosynthetic barriers to vent underlying gas (e.g., from decomposing organic deposits) or
relieve excess hydrostatic pressure.
The method of seaming is dependent upon the chosen geosynthetic material and the project
design. Overlaps, of a designated length, are typically used for thin-film geotextile
composites and geosynthetic clay liners. Geomembranes are seamed with thermal methods
or solvents. Temperature, time, and pressure must be specified and maintained within
tolerances for thermal seaming. With solvent seams, solvent application is important,
because too much solvent can weaken the geomembrane and too little solvent can result in a
weak or leaky seam. Pressure, or heat, is used in conjunction with solvents. Seaming
procedures for a variety of geomembranes is detailed in several waste containment manuals
(Daniel and Koerner, 1993; Landreth and Carson, 1991; Eastern Research Group, Inc., 1991;
Matrecon, 1988).
Construction details around penetrations and adjacent structures depend upon the chosen
geosynthetic material and the project design. As such, they must be individually designed
and detailed. Geosynthetic manufacturers and several waste containment manuals (Daniel
and Koerner, 1993; Matrecon, 1988; Richardson and Koerner, 1988) can provide design
guidance.

10.6 INSPECTION
Quality assurance (QA) and quality control (QC) are recognized as critical factors in the
construction of geomembrane-lined waste containment facilities. QA and QC may or may
not be as important for highway-related barrier works. The extent of QA and QC for
highway barrier works should be project, and barrier product, specific.

FHWA NHI-07-092
Geosynthetics Engineering

10-15

Geomembranes and Barriers


August 2008

10.6-1 Manufacture
Manufacturing quality control (MQC), normally performed by the geosynthetic
manufacturer, is necessary to ensure minimum (or maximum) specified values in the
manufactured product (Daniel and Koerner, 1993). Additionally, manufacturing quality
assurance (MQA) programs are used to provide assurance that the geosynthetics were
manufactured as specified. Quality of raw materials and of finished geosynthetic products
are monitored in an MQA program. The MQA program may be conducted by the
manufacturer, in a department other than manufacturing, or by an outside organization.
Details on MQA and MQC for geomembrane and geosynthetic clay liners, along with other
geosynthetic components, are presented in an EPA Technical Guidance Document (Daniel
and Koerner, 1993).
10.6-2 Field
Construction quality assurance (CQA) and construction quality control (CQC) programs
should be used for most geosynthetic barrier structure construction. CQC is normally
performed by the geosynthetic installer to ensure compliance with the plans and
specifications. CQA is performed by an outside organization to provide assurance to the
owner and regulatory authority (as applicable) that the structure is being constructed in
accordance with plans and specifications. Typically, for waste containment facilities, the
CQA-performing organization is not the installer or designer, i.e., it is a third-party
organization. CQA may be performed by the transportation agency for highway works.
Details on CQA and CQC for geomembrane and geosynthetic clay liners and for traditional
compacted clay barriers are presented in an EPA Technical Guidance Document (Daniel
and Koerner, 1993).

10.7 SPECIFICATION
Geosynthetic barrier specifications should contain the following components:
statement on purpose of barrier;
material specification for the barrier and all associated geosynthetic components,
including component property requirements, product requirements, manufacturing
quality control requirements, and manufacturing quality assurance;
shipping, handling, and storage requirements;
installation requirements;
requirements for sealing to and around penetrations and appurtenances;
seaming requirements, including pass/fail criteria;
anchoring requirements; and
FHWA NHI-07-092
Geosynthetics Engineering

10-16

Geomembranes and Barriers


August 2008

statement on construction quality control and construction quality assurance.


Again, detailed information on geomembrane and GCL specifications is presented in waste
management manuals (Daniel and Koerner, 1993; Matrecon, 1988).
As discussed under Installation, Section 10.5, subgrade preparation is crucial to a successful
installation. Therefore, it is imperative that the accompanying subgrade preparation
specification be written specifically for the geosynthetic barrier to be installed.
The
subgrade must be inspected and approved prior to placement of the geosynthetic barrier.

10.8

REFERENCES

Berg, R.R. (1993). Project Peer Review in Geosynthetic Engineering, Geotechnical Fabrics
Report, Industrial Fabrics Association International, St. Paul, MN, April, pp. 34-35.
Daniel, D.E. and Koerner, R.M. (1993). Quality Assurance and Quality Control for Waste
Containment Facilities, U.S. EPA Technical Guidance Document, EPA/600/R-93/182,
Cincinnati, OH, 305 p.
Frobel, R.K. (1989). Transportation Tunnel Waterproofing Using Geosynthetics, Transportation Record 1248, Transportation Research Board, Washington, D.C., pp. 45-52.
Giroud, J.P. (1984). Impermeability: The Myth and a Rational Approach, Proceedings of
the International Conference on Geomembranes, Vol. 1, Denver, June, pp. 157-162.
Koerner, R.M. (1998). Designing With Geosynthetics, 4th Edition, Prentice-Hall Inc.,
Englewood Cliffs, NJ, 761p.
Koerner, R.M. and Hwu, B-L (1989). Geomembrane Use in Transportation Systems,
Transportation Research Record 1248, Transportation Research Board, Washington,
D.C., pp. 34-44.
Landreth, R.E. and Carson, D.A. (1991). Inspection Techniques for the Fabrication of
Geomembrane Field Seams, U.S. EPA Technical Guidance Document, EPA/530/SW91/051, Cincinnati, OH, 174 p.
Matrecon, Inc. (1998). Lining of Waste Containment and Other Impoundment Facilities,
U.S. EPA Technical Guidance Document, EPA/600/2-88/052, Cincinnati, OH, 1026 p.
Richardson, G.N. and Koerner, R.M., Editors (1990). A Design Primer: Geotextiles and
Related Materials, Industrial Fabrics Association International, St. Paul, MN, 166 p.

FHWA NHI-07-092
Geosynthetics Engineering

10-17

Geomembranes and Barriers


August 2008

Richardson, G.N. and Koerner, R.M. (1988). Geosynthetic Design Guidance for Hazardous
Waste Landfill Cells and Surface Impoundments, EPA Contract No. 68-03-3338, U.S.
Environmental Protection Agency, Cincinnati, OH.
Rowe, R.K. and Giroud, J.P. (1994). Quality Assurance of Barrier Systems for Landfills,
IGS News, International Geosynthetics Society, Vol. 10, No. 1, March, pp. 6-8.
Staff, C.E. (1984). The Foundation and Growth of the Geomembrane Industry in the United
States, Proceedings of International Conference on Geomembranes, Denver, June, pp. 58.

FHWA NHI-07-092
Geosynthetics Engineering

10-18

Geomembranes and Barriers


August 2008

APPENDICES
A.
B.
C.
D.
E.

Geosynthetic Literature
Geosynthetic Terms
Notation and Acronyms
AASHTO M288 Specification
Geosynthetic Test Standards
E-1. ASTM
E-2. GRI
F. Representative List of Geosynthetic Manufacturers and Suppliers
G. General Properties and Costs of Geotextiles and Geogrids
H. Geosynthetic Reinforcement Structural Design Properties

FHWA NHI-07-092
Geosynthetic Engineering

Appendices
August 2008

[ BLANK ]

FHWA NHI-07-092
Geosynthetic Engineering

Appendices
August 2008

Appendix A
GEOSYNTHETIC LITERATURE

Geosynthetic literature is referenced at the end of each chapter. Key references and website
locations where to locate additional references are summarized in this appendix. Only
current references are noted. The geosynthetic applications and/or functions that the
reference covers are noted in brackets, following the reference.

AASHTO www.transportation.org
AASHTO (2007). LRFD Bridge Design Specifications, Fourth Edition, American
Association of State Transportation and Highway Officials, Washington, D.C.
[MSE retaining walls]
AASHTO (2006). Geosynthetic Reinforcement of the Aggregate Base Course of Flexible
Pavement Structures PP 46-0, Standard Specifications for Transportation Materials
and Methods of Sampling and Testing, 26th Edition, and Provisional Standards,
American Association of State Transportation and Highway Officials, Washington,
D.C.
[Geosynthetic reinforcement of flexible pavements]
AASHTO (2006). Geotextile Specification for Highway Applications - M 288-06, Standard
Specifications for Transportation Materials and Methods of Sampling and Testing,
26th Edition, and Provisional Standards, American Association of State
Transportation and Highway Officials, Washington, D.C.
[Geotextiles in subsurface drainage; separation; stabilization; erosion control;
temporary silt fence; and pavement overlay applications]
AASHTO (2002). Standard Specifications for Highway Bridges, Seventeenth Edition,
American Association of State Transportation and Highway Officials, Washington,
D.C.
[MSE retaining walls]

ASTM www.astm/org
ASTM (2006). Volume 4.13 Geosynthetics, Section Four: Construction, ASTM Standards,
ASTM International, West Conshohocken, Pennsylvania.
[All applications and functions]
FHWA NHI-07-092
Geosynthetic Engineering

A-1

Appendix A
August 2008

FHWA www.fhwa.dot.gov/engineering/geotech/
(2007). Geosynthetic Design and Construction Guidelines, FHWA NHI-07-092, NHI course
No. 132013 reference manual; by Holtz, R.D., Christopher, B.R. and Berg, R.R., U.S.
Department of Transportation, Federal Highway Administration, Washington, D.C.
[All applications and functions]
(2006). Ground Improvement Methods, FHWA NHI-06-019 Volume I and FHWA NHI-06020 Volume II; NHI course No. 132034 reference manual; by Elias, V., Welsh, J.,
Warren, J., Lukas, R., Collin, J.G. and Berg, R.R.; U.S. Department of Transportation,
Federal Highway Administration, Washington, D.C., 884 p.
[load transfer platforms, geofoam light-weight fill, MSE wall and reinforced slope
applications]
(2001). Mechanically Stabilized Earth Walls and Reinforced Soil Slopes, Design &
Construction Guidelines, FHWA NHI-00-043; NHI courses No. 132042 and 132043
course reference; by Elias, V., Christopher, B.R. and Berg, R.R.; U.S. Department of
Transportation, Federal Highway Administration, Washington, D.C., 418 p.
[MSE wall and reinforced slope applications]
(2001). Corrosion/Degradation of Soil Reinforcements for Mechanically Stabilized Earth
Walls and Reinforced Soil Slope, FHWA NHI-00-044, by Elias, V., U.S. Department of
Transportation, Federal Highway Administration, Washington, D.C., March, 94 p.
[MSE wall and reinforced slope applications]
Magazines
Geosynthetics (formerly GFR), published by Industrial Fabrics Association International,
Roseville, MN; bi-monthly publication. www.ifai.com
[All applications and functions]
Manufacturers
A number of manufactures have brochures describing typical applications, and listing the
general and index properties of their products. Some have also produced detailed design
manuals, often written by reputable consulting engineers and professors. In some
instances, they have technical assistance personnel on staff. This information and
assistance is usually free upon request and should be considered along with the other
information given above. See Appendix F for a list of manufacturers and website
addresses.
[All applications and functions]

FHWA NHI-07-092
Geosynthetic Engineering

A-2

Appendix A
August 2008

Proceedings of Conferences and Symposiums


Specialty conferences on geosynthetics have been held since 1977. International conference
have been organized under the International Geosynthetics Society (IGS). North
American conferences have been organized by the North American Geosynthetics
Society (NAGS) and the International Fabrics Association International (IFAI).
Information on ordering proceedings from these conferences can be located at:
www.geosyntheticssociety.org/proceedings
www. bookstore.ifai.com
www.nagsigs.org
[All applications and functions]
The International Erosion Control Association (IECA) organizes conferences. Information
on ordering proceedings from these conferences can be located at:
www.ieca.org/Bookstore
[Erosion control applications]
The GeoInstitute (G-I), of the American Society of Civil Engineers (ASCE) organizes
specialty geotechnical conferences. Geosynthetic applications are included in some of
these conferences. Information on ordering proceeding from the G-I can be located at:
www.geoinstitute.org/publications
[All applications and functions]
ASTM International (ASTM) periodically organizes symposiums on geosynthetics.
Proceedings from these symposium are published as Special Technical Publications
(STPs). Information on STPs and ordering information can be located at:
www.astm.org
[All applications and functions]

Professional Societies
Publications (in addition to conference proceedings) such as magazines, recommended
practices, etc. are available at professional societies such as IGS, NAGS, IECA, ASTM and
ASCE. Websites of these societies are:
www.geosyntheticssociety.org
www.nagsigs.org
www.ieca.org
www.astm.org
www.geoinstitute.org
www.asce.org

FHWA NHI-07-092
Geosynthetic Engineering

A-3

Appendix A
August 2008

Technical Journals
Geotextiles & Geomembranes (Elsevier Applied Science Publishers) publishes technical
papers technical papers, technical notes, discussion and book reviews on all topics
relating to geosynthetics, research, behavior, performance analysis, testing, design,
construction methods, case histories and field experience.
Geotextiles and
Geomembranes is dedicated to the mission of the IGS which is to promote the scientific
and engineering development of geotextiles, geomembranes, related products and
associated technologies.
www.geosyntheticssociety.org
[All applications and functions]
Geosynthetics International was a peer-reviewed technical journal that published technical
papers, technical notes, discussions, and book reviews on all topics relevant to
geosynthetic materials (including natural fiber products), research, behavior, performance
analysis, testing, design, construction methods, case histories, and field experience. Past
issues of Geosynthetics International is electronically available and free for all individual
IGS members.
www.geosyntheticssociety.org
[All applications and functions]
Technical journals on geotechnical engineering occasionally include papers on geosynthetic
topics. Journals and websites include: Journal of Geotechnical and Geoenvironmental
Engineering (ASCE, G-I); Geotechnical Testing Journal (ASTM); Canadian
Geotechnical Journal (CGS); and Transportation Research Record (TRB).
www.geoinstitute.org/publications
www.astm.org
www.cgs.ca
www.trb.org/bookstore
[All applications and functions]
Trade Groups
Geosynthetic Materials Association (GMA) serves as the central resource for information
regarding geosynthetics. Publications include GMA Handbook of Geosynthetics; and
GMA-sponsored White Paper "Geosynthetic Reinforcement of the Base and Subbase
Courses of Pavement Structures;" both are downloadable from the GMA website.
www.gmanow.com
[All applications and functions]

FHWA NHI-07-092
Geosynthetic Engineering

A-4

Appendix A
August 2008

National Concrete Masonry Association (NCMA) publishes design guides and software for
geosynthetic reinforced MSE walls. Description of products and ordering information is
located on the NCMA website.
www.ncma.org
[Non-highway MSE walls]
Textbooks
There are several textbooks on geosynthetics and many more that include geosynthetic topics
in some chapters or sections. The most notable textbook is by Dr. Robert Koerner,
Bowman Professor of Civil Engineering of Drexel University and Founder and Director
of the Geosynthetic Institute. This textbook is:
Koerner, R.M. (2005) Designing with Geosynthetics, 5th Edition, Prentice-Hall, Inc.
[All applications and functions]

FHWA NHI-07-092
Geosynthetic Engineering

A-5

Appendix A
August 2008

[ BLANK ]

FHWA NHI-07-092
Geosynthetic Engineering

A-6

Appendix A
August 2008

Appendix B
GEOSYNTHETIC TERMS

apparent opening size (AOS, O95) - a property which indicates the approximate largest
particle that would effectively pass through a geotextile
blinding - condition whereby soil particles block the surface openings of a geotextile,
thereby reducing hydraulic conductivity
California Bearing Ratio (CBR) - the ratio of (1) the force per unit area required to
penetrate a soil mass with a 3-square-inch circular piston (approximately 2-inch diameter) at
the rate of 0.05 inches/minute to (2) the force per unit area required for corresponding
penetration of a standard material
clogging - condition where soil particles move into and are retained in the openings of a
geotextile, thereby reducing hydraulic conductivity
cross-machine direction - the direction in the plane of the geosynthetic perpendicular to the
direction of manufacture
filtration - the process of retaining soils while allowing the passage of water (fluid)
geocell - a three-dimensional comb-like structure, to be filled with soil, aggregate or concrete
geocomposite - a geosynthetic material manufactured of two or more materials
geogrid - a geosynthetic formed by a regular network of tensile elements and apertures,
typically used for reinforcement applications
geomembrane - an essentially impermeable geosynthetic, typically used to control fluid
migration
geonet - a geosynthetic consisting of integrally connected parallel sets of ribs overlying
similar sets of ribs, for planar drainage of liquids or gases
geosynthetic - a planar product manufactured from polymeric material used with soil,
aggregate, or other geotechnical engineering materials

FHWA NHI-07-092
Geosynthetic Engineering

B-1

Appendix B
August 2008

geotextile - a permeable geosynthetic comprised solely of textiles


index test - a test procedure which may contain a known bias but which may be used to
establish an order for a set of specimens with respect to the property of interest
machine direction - the direction in the plane of the geosynthetic parallel to the direction of
manufacture
minimum average roll value (MARV) a quality control tool used by geosynthetic
manufacturers to establish and publish minimum (or maximum) property values.
permeability - the rate of flow of a liquid under a differential pressure through a material
permittivity - the volumetric flow rate of water per unit cross sectional area per unit head
under laminar flow conditions, in the normal direction through a geotextile

Bibliography
ASTM (2006). D 4439 Standard Terminology for Geosynthetics, ASTM International, West
Conshohocken, PA, 5 p.
Frobel, R.K. (1987). Geosynthetics Terminology - An Interdisciplinary Treatise, Industrial Fabrics
Association International, St. Paul, 126 p.

FHWA NHI-07-092
Geosynthetic Engineering

B-2

Appendix B
August 2008

Appendix C
NOTATION AND ACRONYMS
C-1 NOTATION
a
A
AOS
b
B
c
c'
ca
cv
Cc
Cr
Cu
CBR
d
D

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

e
F*
FS
FScr
FSID
FSCD
FSBD
FSJNT
g
GL
GH
GVW
H
i
k
K
KA
L

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

radius of tire contact area


area
apparent opening size
a dimension; horizontal length of embankment slope
a coefficient; width of geosynthetic or embankment
undrained shear strength ("cohesion") in terms of total stresses
effective stress strength parameters
soil-geosynthetic adhesion
coefficient of consolidation
compressive index
recompression index
uniformity coefficient, D60/D10
California Bearing Ratio
depth
grain size (subscript indicates percent smaller than); depth of embankment;
thickness of soft layers
eccentricity
the pullout resistance (or friction-bearing interaction) factor
factor of safety
partial factor for creep deformation, ratio of Tult to creep limiting strength,
partial factor of safety for installation damage
partial factor of safety for chemical degradation
partial factor of safety for biological degradation
partial factor of safety for joints, seams, and connections
acceleration due to gravity
lower strength geogrid
higher strength geogrid
gross vehicle weight
head difference (gradient ratio test); embankment, slope or wall height
hydraulic gradient
coefficient of permeability
stress ratio; force coefficient
active earth coefficient of the retained backfill
length; length of reinforcement; length of failure arc

FHWA NHI-07-092
Geosynthetic Engineering

C-1

Appendix C
August 2008

Le
M
n
N
NC
O
Pa
Pb
PI
Pq
PQ
q
qa
qult
QL
R
Rv
S
SF
t
T
Ta
Td
Tult
Tl
v
Vq
w
W
x
y
"
$
(
)
,
1
:

=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=
=

embedment length to resist pullout


moment
porosity
number of layers
bearing capacity factor for cohesive soils
opening size; subscript indicates percent smaller than
active earth pressure
resultant active earth pressure due to the retained backfill
plasticity index
resultant active earth pressure due to the uniform surcharge
resultant of live load
flow rate; surcharge load
allowable bearing capacity
ultimate bearing capacity
live load
radius of critical failure circle
resisting force (Meyerhof's approach)
vertical spacing between horizontal geogrid layer
safety factor
thickness of geogrid
tensile strength of the geosynthetic
allowable tensile strength of the geosynthetic
design tensile strength of the geosynthetic (usually at a given strain)
ultimate tensile strength of a geosynthetic
creep limit tensile strength of a geosynthetic
vertical
vertical force due to surcharge
water content
vertical force due to the weight of the fill
a dimension or coordinate
a dimension or coordinate
peak horizontal acceleration for seismic loading
slope of soil surface; angle of reinforcement force
unit weight
change in some parameter or quantity
strain
inclination of the wall face
friction coefficient along the sliding plane, which depends on the location
plane, i.e., tan Nr or tan Nf

FHWA NHI-07-092
Geosynthetic Engineering

C-2

Appendix C
August 2008

:*
R
2
Fh
Fo
Fp '
Fv
N
N'
J

=
=
=
=
=
=
=
=
=
=

the pullout resistance of shearing friction between soil and geogrid


permittivity
transmissivity; an angle; angle of failure plane
horizontal stress
overburden stress
preconsolidation stress
vertical stress
angle of internal friction
effective angle of internal friction
shear resistance

C-2 ACRONYMS
AASHTO
CPE
CSPE
CQA
CQC
EIA
EPA
FHWA
HDPE
GCL
MBW
MSE
MQA
MQC
MESL
PP
PVC
QA
QC
RSS
SRW
VRSS
VLDPE
USFS

American Association of State Highway and Transportation Officials


chlorinated polyethylene
chlorosulfonated polyethylene
construction quality assurance
construction quality control
ethylene interpolymer alloy
U.S. Environmental Protection Agency
U.S. Department of Transportation, Federal Highway Administration
high-density polyethylene
geosynthetic clay liner
modular (masonry) block wall unit
mechanically stabilized earth
manufacturing quality assurance
manufacturing quality control
membrane encapsulated soil layer
polypropylene
polyvinyl chloride
quality assurance
quality control
reinforced soil slope
segmental retaining wall (unit) - see MCU
vegetated reinforced soil slope
very low-density polyethylene
U.S. Department of Agriculture, Forest Service

FHWA NHI-07-092
Geosynthetic Engineering

C-3

Appendix C
August 2008

[ BLANK ]

FHWA NHI-07-092
Geosynthetic Engineering

C-4

Appendix C
August 2008

Appendix D
AASHTO M288 SPECIFICATION

(Taken from Standard Specifications for Transportation Materials and Methods of


Sampling and Testing. Copyright 2006. By the American Association of State
Highway and Transportation Officials, www.transportation.org. Reproduced with
permission.)

FHWA NHI-07-092
Geosynthetic Engineering

D-1

Appendix D
August 2008

Standard Specification for

Geotextile Specification for Highway Applications


AASHTO DESIGNATION: M 288-06

1.

SCOPE

1.1.

This is a materials specification covering geotextile fabrics for use in subsurface


drainage; separation; stabilization; erosion control; temporary silt fence; and paving
fabrics. This is a material purchasing specification and design review of use is
recommended.

1.2.

This specification sets forth a set of physical, mechanical and endurance properties
that must be met, or exceeded, by the geotexile being manufactured.

1.3.

In the context of quality systems and management, this specification represents a


manufacturing quality control (MQC) document. However, its general use is
essentially as a recommended design document.

1.4.

This specification is intended to assure both good quality and performance of


geotextiles used as listed in Section 1.1, but is possibly not adequate for the complete
specification in a specific situation. Additional tests, or more restrictive values for
the tests indicated, may be necessary under conditions of a particular application.

1.5.

This specification is based on geotextile survivability from installation stresses.


Designers should be aware that the classes in this specification reflect this basic
premise. Refer to Appendix XI of this specification for geotextile construction
guidelines.

2.

REFERENCED DOCUMENTS

2.1.

AASHTO Standards
T 88, Particle Size Analysis of Soils
T 90, Determining the Plastic Limit and Plasticity Index of Soils
T 99, The Moisture-Density Relations of Soils Using a 2.5 kg (5.5 lb)
Rammer and a 305 mm (12 in.) Drop

FHWA NHI-07-092
Geosynthetic Engineering

D-2

Appendix D
August 2008

2.2.

ASTM Standards:1
D 123, Standard Terminology Relating to Textiles
D 276, Test Method for Identification of Fibers in Textiles
D 4354, Practice for Sampling of Geosynthetics for Testing
D 4355, Test Method for Deterioration of Geotextiles from Exposure to
Ultraviolet Light and Water (Xenon-Arc Type Apparatus)
D 4439, Terminology for Geosynthetics
D 4491, Test Methods for Water Permeability of Geotextiles by Permittivity
D 4533, Test Method for Trapezoid Tearing Strength of Geotextiles
D 4632, Test Method for Grab Breaking Load and Elongation of Geotextiles
D 4751, Test Method for Determining Apparent Opening Size of a Geotextile
D 4759, Practice for Determining the Specification Conformance of
Geosynthetics
D 4873, Guide for Identification, Storage, and Handling of Geotextiles
D 5141, Test Method to Determine Filtering Efficiency and Flow Rate for Silt
Fence Application of a Geotextile Using Site Specific Soils
D 5261, Test Method for Measuring Mass per Unit Area of Geotextiles
D 6140, Test Method for Determining the Asphalt Retention of Paving Fabrics
D 6241, Test Method for Static Puncture Strength of Geotextiles and Geotextile
Related Products Using a 50-mm Probe

3.

DEFINITIONS

3.1.

Formulation The mixture of a unique combination of ingredients identified by type,


properties and quantity. For nonwoven geotextiles, a formulation is defined as the
exact percentages and types of resin(s), additives and/or carbon black.

3.2.

Manufacturing Quality Control (MQC) A planned system of inspections that is used


to directly monitor and control the manufacture of a material which is factory
originated. MQC is normally performed by the manufacturer of geosynthetic
materials and is necessary to ensure minimum (or maximum) specified values in the
manufactured product. MQC refers to measures taken by the manufacturer to
determine compliance with the requirements for materials and workmanship as stated
in certification documents and contract specifications [ref. EPA/600/R-93/182].

3.3.

Minimum Average Roll Value (MARV) For geosynthetics, a manufacturing quality


control tool used to allow manufacturers to establish published values such that the
user/purchaser will have a 97.7 percent confidence that the property in question will
1

Available from ASTM.

FHWA NHI-07-092
Geosynthetic Engineering

D-3

Appendix D
August 2008

meet published values. For normally distributed data, "MARV" is calculated as the
typical value minus two standard deviations from documented quality control test
results for a defined population from one specific test method associated with one
specific property.
3.4.

Minimum Value The lowest sample value from documented manufacturing quality
control test results for a defined population from one test method associated with one
specific property.

3.5.

Maximum Value The highest sample value from documented manufacturing quality
control test results for a defined population from one test method associated with one
specific property.

4.

PHYSICAL REQUIREMENTS

4.1.1

Fibers used in the manufacture of geotextiles, and the threads used in joining
geotextiles by sewing, shall consist of long-chain synthetic polymers, composed of at
least 95 percent by weight of polyolefins or polyesters. They shall be formed into a
stable network such that the filaments or yarns retain their dimensional stability
relative to each other, including selvages.

4.2

Geotextiles used for subsurface drainage, separation, stabilization, and permanent


erosion control applications shall conform to the physical requirements of Section 8.
Geotextiles used for temporary silt fence shall conform to the physical requirements
of Section 9 and geotextiles used as paving fabrics shall conform to the physical
requirements of Section 10.

4.3

All property values, with the exception of apparent opening size (AOS), in these
specifications represent minimum average roll values (MARV) in the weakest
principle direction (i.e., average test results of any roll in a lot sampled for
conformance or quality assurance testing shall meet or exceed the minimum values
provided herein). Values for AOS represent maximum average roll values.

5.

CERTIFICATION

5.1.

The contractor shall provide to the engineer a certificate stating the name of the
manufacturer, product name, style number, chemical composition of the filaments or
yarns, and other pertinent information to fully describe the geotextile.

FHWA NHI-07-092
Geosynthetic Engineering

D-4

Appendix D
August 2008

5.2.

The Manufacturer is responsible for establishing and maintaining a quality control


program to assure compliance with the requirements of the specification.
Documentation describing the quality control program shall be made available upon
request.

5.3.

The manufacturers certificate shall state that the furnished geotextile meets MARV
requirements of the specification as evaluated under the manufacturers quality
control program. A person having legal authority to bind the manufacturer shall attest
to the certificate.

5.4.

Either mislabeling or misrepresentation of materials shall be reason to reject those


geotextile products.

6.

SAMPLING, TESTING, AND ACCEPTANCE

6.1.

Geotextiles shall be subject to sampling and testing to verify conformance with this
specification. Sampling shall be in accordance with the most current ASTM D 4354
using the section titled, Procedure for Sampling for Purchasers Specification
Conformation Testing. In the absence of purchasers testing, verification may be
based on manufacturers certifications as a result of a testing by the manufacturer of
quality assurance samples obtained using the procedure for Sampling or
Manufacturers Quality Assurance (MQA) Testing. A lot size shall be considered to
be the shipment quantity of the given product or a truckload of the given product,
whichever is smaller.

6.2.

Testing shall be performed in accordance with the methods referenced in this


specification for the indicated application. The number of specimens to test per
sample is specified by each test method. Geotextile product acceptance shall be
based on ASTM D 4759. Product acceptance is determined by comparing the
average test results of all specimens within a given sample to the specification
MARV. Refer to ASTM D 4759 for more details regarding geotextile acceptance
procedures.

7.

SHIPMENT AND STORAGE

7.1.

Geotextiles labeling, shipment, and storage shall follow ASTM D 4873. Product
labels shall clearly show the manufacturer or supplier name, style name, and roll

FHWA NHI-07-092
Geosynthetic Engineering

D-5

Appendix D
August 2008

number. Each shipping document shall include a notation certifying that the material
is in accordance with the manufacturers certificate.
7.2.

Each geotextile roll shall be wrapped with a material that will protect the geotextile,
including the ends of the roll, from damage due to shipment, water, sunlight, and
contaminants. The protective wrapping shall be maintained during periods of
shipment and storage.

7.3

During storage, geotextile rolls shall be elevated off the ground and adequately
covered to protect them from the following: site construction damage, precipitation,
extended ultraviolet radiation including sunlight, chemicals that are strong acids or
strong bases, flames including welding sparks, temperatures in excess of 71C
(160F), and any other environmental condition that may damage the physical
property values of the geotextile.

8.

GEOTEXTILE PROPERTY REQUIREMENTS FOR


SUBSURFACE DRAINAGE, SEPARATION, STABILIZATION,
AND PERMANENT EROSION CONTROL

8.1

General Requirements:

8.1.1

Table 1 provides strength properties for three geotextile classes. The geotextile shall
conform to the properties of Table 1 based on the geotextile class required in Table 2,
3, 4, 5, or 6 for the indicated application.

8.1.2

All numeric values in Table 1 represent MARV in the weaker principal direction.
The geotextile properties required for each class are dependent upon geotextile
elongation. When sewn seams are required, the seam strength, as measured in
accordance with ASTM D 4632, shall be equal to or greater than 90 percent of the
specified grab strength.

FHWA NHI-07-092
Geosynthetic Engineering

D-6

Appendix D
August 2008

Table 1 Geotextile Strength Property Requirements


Geotextile Class
Class 1
Test
Methods

a,b

Class 2

Class 3

Elongation

Elongation

Elongation

Elongation

Elongation

Elongation

Units

< 50%

> 50%

< 50%

> 50%

< 50%

> 50%

Grab strength

ASTM
D 4632

1400

900

1100

700

800

500

Sewn seam
d
strength

ASTM
D 4632

1260

810

990

630

720

450

Tear strength

ASTM
D 4533

500

350

400

250

300

180

Puncture
strength

ASTM
D 6241

2750

1925

2200

1375

1650

990

Permittivity

ASTM
D 4491

sec

Apparent
opening size

ASTM
D 4751

mm

Ultraviolet
stability
(retained
strength)

ASTM
D 4355

b
c
d
e

-1

Minimum property values for permittivity, AOS, and UV stability are


based on geotextile application. Refer to Table 2 subsurface drainage,
Table 3 and Table 4 for separation, Table 5 for stabilization, and Table 6
for permanent erosion control.

Required geotextile class is designated in Tables 2, 3, 4, 5, or 6 for the indicated application. The severity of
installation conditions for the application generally dictate the required geotextile class. Class 1 is specified
for more severe or harsh installation conditions where there is a greater potential for geotextile damage, and
Classes 2 and 3 are specified for less severe conditions.
All numeric values represent MARV in the weaker principal direction. (See Section 8.1.2)
As measured in accordance with ASTM D 4632.
When sewn seams are required. Refer to Appendix for overlap seam requirements.
The required MARV tear strength for woven monofilament geotextiles is 250 N.

8.2

Subsurface Drainage Requirements:

8.2.1

Description This specification is applicable to placing a geotextile against a soil to


allow for long-term passage of water into a subsurface drain system retaining the in
situ soil. The primary function of the geotextile in subsurface drainage applications is
filtration. Geotextile filtration properties are a function of the in situ soil gradation,
plasticity, and hydraulic conditions.

FHWA NHI-07-092
Geosynthetic Engineering

D-7

Appendix D
August 2008

8.2.2

Geotextile Requirements The geotextile shall meet the requirements of Table 2.


Woven slit film geotextiles (i.e., geotextiles made from yarns of a flat, tape-like
character) will not be allowed. All numeric values in Table 2, except AOS represent
MARV in the weaker principal direction. Values of AOS represent maximum
average roll values.

8.2.3

The property values in Table 2 represent default values that provide sufficient
geotextile survivability under most construction conditions. Note b of Table 2
provides for a reduction in the minimum property requirements when sufficient
survivability information is available. The engineer may also specify properties
different from that listed in Table 2 based on engineering design experience.

8.3

Separation Requirements:

8.3.1

Description This specification is applicable to the use of a geotextile to prevent


mixing of a subgrade soil and an aggregate cover material (subbase, base, select
embankment, etc.). This specifications may also apply to situations other than
beneath pavements where separation of two dissimilar materials is required, but
where water seepage through the geotextile is not a critical function.

8.3.2

The separation application is appropriate for pavement structures constructed over


soils with a California Bearing Ratio equal to or greater than 3 (CBR > 3) (shear
strength greater than approximately 90 kPa). It is appropriate for unsaturated
subgrade soils. The primary function of a geotextile in this application is separation.

8.3.3

Geotextile Requirements The geotextile shall meet the requirements of Table 3. All
numeric values in Table 3 except AOS represent MARV in the weakest principal
direction. Values for AOS represent maximum average roll values.

8.3.4

The property values in Table 3 represent default values that provide for sufficient
geotextile survivability under most construction conditions. The engineer may also
specify properties different from those listed in Table 3 based on engineering design
and experience.

FHWA NHI-07-092
Geosynthetic Engineering

D-8

Appendix D
August 2008

Table 2 Subsurface Drainage Geotextile Requirements


Requirements
Percent In Situ Soil Passing 0.075 mma
Test Methods

Units

< 15

15 to 50

> 50

Class 2 from Table 1b

Geotextile class
Permittivityc,d

ASTM D 4491

sec-1

0.5

0.2

0.1

Apparent opening
sizec,d

ASTM D 4751

mm

0.43 max
avg roll value

0.25 max
avg roll value

0.22e max avg


roll value

Ultraviolet stability
(retained strength)

ASTM D 4355

a
b

50% after 500 hrs of exposure

Based on grain size analysis of in situ soil in accordance with T 88.


Default geotextile selection. The engineer may specify a Class 3 geotextile from Table 1 for trench drain
applications based on one or more of the following:
1) The engineer has found Class 3 geotextiles to have sufficient survivability based on field experience.
2) The engineer has found Class 3 geotextiles to have sufficient survivability based on laboratory testing
and visual inspection of a geotextile sample removed from a field test section constructed under
anticipated field conditions.
3) Subsurface drain depth is less than 2 m; drain aggregate diameter is less than 30 mm; and compaction
requirement is less than 95 percent of T 99.
These default filtration property values are based on the predominant particle size of in situ soil. In addition to
the default permittivity value, the Engineer may require geotextile permeability and/or performance testing
based on engineering design for drainage systems in problematic soil environments.
Site specific geotextile design should be performed especially if one or more of the following problematic soil
environments are encountered: unstable or highly erodible soils such as non-cohesive silts; gap graded soils;
alternating sand/silt laminated soils; dispersive clays; and/or rock flour.
For cohesive soils with a plasticity index greater than seven, geotextile maximum average roll value for
apparent opening size is 0.30 mm.

Table 3 Separation Geotextile Property Requirements


Test Methods

Units

Geotextile Class

Requirements
See Table 4

Permittivity

ASTM D 4491

sec-1

0.02a

Apparent opening size

ASTM D 4751

mm

0.60 max avg roll value

Ultraviolet stability
(retained strength)

ASTM D 4355

50% after 500 hrs of exposure

Default value. Permittivity of the geotextile should be greater than that of the soil (g > s). The engineer
may also require the permeability of the geotextile to be greater than that of the soil (kg > ks).

FHWA NHI-07-092
Geosynthetic Engineering

D-9

Appendix D
August 2008

Table 4 Required Degree of Survivability as a Function of Subgrade Conditions,


Construction Equipment, and Lift Thickness (Class 1, 2, and 3 properties are given in Table
1; Class 1+ properties are higher than Class 1, but not defined at this time and if used must be
specified by the purchaser)a

Subgrade has been cleared of all obstacles except grass,


weeds, leaves, and fine wood debris. Surface is
smooth and level so that any shallow depressions and
humps do not exceed 450 mm (18 in.) in depth or
height. All larger depressions are filled. Alternatively,
a smooth working table may be placed.
Subgrade has been cleared of obstacles larger than
small to moderate-sized tree limbs and rocks. Tree
trunks and stumps should be removed or covered with
a partial working table. Depressions and humps should
not exceed 450 mm (18 in.) in depth or height. Larger
depressions should be filled.

Low-ground
pressure
equipment
< 25 kPa
(3.6 psi)

Medium groundpressure
equipment
> 25 to < 50 kPa
(>3.6 to < 7.3 psi)

High groundpressure
equipment
> 50 kPa
(> 7.3 psi)

Low
(Class 3)

Moderate
(Class 2)

High
(Class 1)

Moderate
(Class 2)

High
(Class 1)

Very high
(Class 1+)

Minimal site preparation is required. Trees may be


felled, delimbed, and left in place. Stumps should be
cut to project not more than + 150 mm (+ 6 in.) above
subgrade. Geotextile may be draped directly over the
High
Very high
Not
tree trunks, stumps, large depressions and humps,
(Class 1)
(Class 1+)
recommended
holes, stream channels, and large boulders. Items
should be removed only if placing the geotextile and
cover material over them will distort the finished road
surface.
a
Recommendations are for 150 to 300 mm (6 to 12 in.) initial lift thickness. For other initial lift
thicknesses:
300 to 450 (12 to 18 in.): reduce survivability requirement by one level;
450 to 600 mm (18 to 24 in.): reduce survivability requirement two levels;
> 600 mm ((24 in.): reduce survivability requirement three levels
For special construction techniques such as prerutting, increase the geotextile survivability requirement one
level. Placement of excessive initial cover material thickness may cause bearing failure of the soft subgrade.

Stabilization Requirements:
8.3.5

Description This specification is applicable to the use of a geotextile in wet,


saturated conditions to provide the coincident functions of separation and filtration.
In some installations, the geotextile can also provide the function of reinforcement.
Stabilization is applicable to pavement structures constructed over soils with a
California Bearing Ratio between one and three (1<CBR<3) (shear strength between
approximately 30 kPa and 90 kPa).

FHWA NHI-07-092
Geosynthetic Engineering

D - 10

Appendix D
August 2008

8.3.6

The stabilization application is appropriate for subgrade soils that are saturated due to
a high groundwater table or due to prolonged periods of wet weather. This
specification is not appropriate for embankment reinforcement where stress
conditions may cause global subgrade foundation or stability failure. Reinforcement
of the pavement section is a site specific design issue.

8.3.7

Geotextile Requirements The geotextile shall meet the requirements of Table 5. All
numeric values in Table 5 except AOS represent MARV in the weakest principal
direction. Values for AOS represent maximum average roll values.

8.3.8

The property values in Table 5 represent default values that provide for sufficient
geotextile survivability under most construction conditions. Note 2 of Table 5
provides for a reduction in the minimum property requirements when sufficient
survivability information is available. The engineer may also specify properties
different from that listed in Table 5 based on engineering design and experience.

Table 5 Stabilization Geotextile Property Requirements


Test Methods

Units

Requirements
Class 1 from Table 1a

Geotextile class
Permittivity

ASTM D 4491

sec-1

0.05b

Apparent opening size

ASTM D 4751

mm

0.43 max avg roll value

Ultraviolet stability
(retained strength)

ASTM D 4355

50% after 500 hrs of exposure

Default geotextile selection. The engineer may specify a Class 2 or 3 geotextile from Table 1 based on one or
more of the following:
1. The engineer has found the class of geotextile to have sufficient survivability based on field experience.
2. The engineer has found the class of geotextile to have sufficient survivability based on laboratory
testing and visual inspection of a geotextile sample removed from a field test section constructed under
anticipated field conditions.
Default value. Permittivity of the geotextile should be greater than that of the soil (g > s). The engineer
may also require the permeability of the geotextile to be greater than that of the soil (kg > ks).

FHWA NHI-07-092
Geosynthetic Engineering

D - 11

Appendix D
August 2008

8.4

Permanent Erosion Control:

8.4.1

Description This specification is applicable to the use of a geotextile between


energy absorbing armor systems and in the in situ soil to prevent soil loss resulting in
excessive scour and to prevent hydraulic uplift pressures causing instability of the
permanent erosion control system. This specification does not apply to other types of
geosynthetic soil erosion control materials such as turf reinforcement mats.

8.4.2

The primary function the geotextile serves in permanent erosion control applications
is filtration. Geotextile filtration properties are a function of hydraulic conditions,
and in situ soil gradation, density, and plasticity.

8.4.3

Geotextile Requirements The geotextile shall meet the requirements of Table 6.


Woven slit film geotextiles (i.e., geotextiles made from yarns of flat, tape-like
character) will not be allowed. All numeric values in Table 6 except AOS represent
MARV in the weaker principal direction. Values for AOS represent maximum
average roll values.

8.4.4

The property values in Table 6 represent default values which provide for sufficient
geotextile survivability under conditions similar to or less severe than those described
under Note b of Table 6. Note c of Table 6 provides for a reduction in the minimum
property requirements when sufficient survivability information is available or when
the potential for construction damage is reduced. The engineer may also specify
properties different from that listed in Table 6 based on engineering design and
experience.

9.

TEMPORARY SILT FENCE REQUIREMENTS

9.1

Description This specification is applicable to the use of a geotextile as a vertical,


permeable interceptor designed to remove suspended soil from overland water flow.
The function of a temporary silt fence is to filter and allow settlement of soil particles
from sediment-laden water. The purpose is to prevent the eroded soil from being
transported off the construction site by water runoff.

9.2

Geotextile Requirements The geotextile used for temporary silt fence may or may
not be supported between posts with wire or polymeric mesh. The temporary silt
fence geotextile shall meet the requirements of Table 7. All numeric values in Table
7 except AOS represent MARV. Values for AOS represent maximum average roll
values.

FHWA NHI-07-092
Geosynthetic Engineering

D - 12

Appendix D
August 2008

9.3 Field monitoring shall be performed to verify that the armor system placement does not
damage the geotextile. The minimum height above ground for all silt fence shall be 750 mm.
Minimum embedment depth shall be 150 mm. Refer to Appendix X1 for more detailed
installation requirements.

Table 6 Permanent Erosion Control Geotextile Requirements


Requirements,
Percent In Situ Soil Passing 0.075 mma
Test Methods

Units

< 15

15 to 50

Geotextile class:
Woven monofilament geotextiles
All other geotextiles

> 50

Class 2 from Table 1b


Class 1 from Table 1b,c

Permittivitya,d

ASTM D 4491

sec-1

0.7

0.2

0.1

Apparent opening
sizec,d,e

ASTM D 4751

mm

0.43 max
avg roll value

0.25 max
avg roll value

0.22e max avg


roll value

Ultraviolet stability
(retained strength)

ASTM D 4355

a
b

50% after 500 hrs of exposure

Based on grain size analysis of in situ soil in accordance with AASHTO T 88.
As a general guideline, the default geotextile selection is appropriate for conditions of equal or less severity
than either of the following:
1. Armor layer stone weights do not exceed 100 kg, stone drop height is less than 1 m, and no aggregate
bedding is required.
2. Armor layer stone weighs more than 100 kg, stone drop height is less than 1 m, and the geotextile is
protected by a 150-mm thick aggregate bedding layer designed to be compatible with the armor layer.
More severe applications require an assessment of geotextile survivability based on a field trial section
and may require a geotextile with higher strength properties.
The engineer may specify a Class 2 geotextile from Table 1 based on one or more of the following:
1. The engineer has found Class 2 geotextiles to have sufficient survivability based on field experience.
2. The engineer has found Class 2 geotextiles to have sufficient survivability based on laboratory testing
and visual inspection of a geotextile sample removed from a field test section constructed under
anticipated field conditions.
3. Armor layer stone weighs less than 100 kg, stone drop height is less than 1 m, and the geotextile is
protected by a 150-mm thick aggregate bedding layer designed to be compatible with the armor layer.
4. Armor layer stone weights do not exceed 100 kg, and stone is placed with a zero drop height.
These default filtration property values are based on the predominant particle size of in situ soil. In addition to
the default permittivity value, the engineer may require geotextile permeability and/or performance testing
based on engineering design for drainage systems in problematic soil environments.
See the following:
1. Site specific geotextile design should be performed especially if one or more of the following
problematic soil environments are encountered: unstable or highly erodible soils such as non-cohesive
silts; gap graded soils; alternating sand/silt laminated soils; dispersive clays; and/or rock flour.
2. For cohesive soils with a plasticity index greater than seven, geotextile maximum average roll values
for apparent opening size is 0.30 mm.

FHWA NHI-07-092
Geosynthetic Engineering

D - 13

Appendix D
August 2008

Table 7 Temporary Silt Fence Property Requirements


Requirements
Unsupported Silt Fence
Test Methods

Units

Maximum post spacing

Supported
Silt Fencea

Geotextile
Elongation
> 50%b

Geotextile
Elongation
< 50%b

1.2 m

1.2 m

2m

400
400

550
450

550
450

Grab strength
Machine direction
X-Machine direction

ASTM D 4632

Permittivityc

ASTM D 4491

sec-1

0.05

0.05

0.05

Apparent opening size

ASTM D 4751

mm

0.60 max
avg roll value

0.60 max
avg roll value

0.60 max avg


roll value

Ultraviolet stability
(retained strength)

ASTM D 4355

70% after 500


hrs of
exposure

70% after 500 hrs of exposure

a
b
c

Silt fence support shall consist of 14-gage steel wire with a mesh spacing of 150 mm by 150 mm
or prefabricated polymeric mesh of equivalent strength.
As measured in accordance with ASTM D 4632.
These default filtration property values are based on empirical evidence with a variety of
sediments. For environmentally sensitive areas, a review of previous experience and/or site or
regionally specific geotextile tests should be performed by the agency to confirm suitability of
these requirements.

10.

PAVING FABRIC REQUIREMENTS

10.1

Description This specification is applicable to the use of a paving fabric, saturated


with asphalt cement, between pavement layers. The function of the paving fabric is
to act as a waterproofing and stress relieving membrane within the pavement
structure. This specification is not intended to describe fabric membrane systems
specifically designed for pavement joints and localized (spot) repairs.

10.2

Paving Fabric Requirements The paving fabric shall meet the requirements of
Table 8. All numeric values in Table 8 represent MARV in the weaker principal
direction.

FHWA NHI-07-092
Geosynthetic Engineering

D - 14

Appendix D
August 2008

Table 8 Paving Fabric Property Requirementsa

a
b

Test Methods

Units

Requirements

Grab Strength

ASTM D 4632

450

Ultimate Elongation

ASTM D 4632

> 50

Mass Per Unit Area

ASTM D 5261

gm/m2

140

Asphalt Retention1

ASTM D 6140

l/m2

bc

Melting Point

ASTM D 276

150

All numeric values represent MARV in the weaker principal direction. (Refer to Section 10.2.)
Asphalt required to saturate paving fabric only. Asphalt retention must be provided in manufacturer
certification. (Refer to Section 5.) Value does not indicate the asphalt application rate required for
construction. Refer to Appendix for discussion of asphalt application rate.
Product asphalt retention property must meet the MARV value provided by the manufacturer
certification. (Refer to Section 5.)

______________

FHWA NHI-07-092
Geosynthetic Engineering

D - 15

Appendix D
August 2008

APPENDIX
(Nonmandatory Information)
X1.

CONSTRUCTION/INSTALLATION GUIDELINES

X1.1.

GENERAL

X1.1.1.

This Appendix is intended for use in conjunction with M 288 for Geotextiles.
The specification details materials properties for geotextiles used in drainage,
erosion control, separation/stabilization, silt fences, and pavement overlay
application. The material properties are only one factor in a successful
installation involving geotextiles. Proper construction and installation
techniques are essential in order to ensure that the intended function of the
geotextile is fulfilled.

X1.1.2.

Geotextile Identification, Packaging, and Storage:

X1.1.2.1.

Refer to ASTM D 4873.

X.1.1.3

Geotextile Exposure Following Placement:

X1.1.3.1.

Atmospheric exposure of geotextiles to the elements following lay down shall


be a maximum of 14 days to minimize damage potential.

X1.1.4.

Seaming:

X1.1.4.1.

If a sewn seam is to be used for the seaming of the geotextile, the thread used
shall consist of high strength polypropylene, or polyester. Nylon thread shall
not be used. For erosion control applications, the thread shall also be resistant
to ultraviolet radiation. The thread shall be of contrasting color to that of the
geotextile itself.

X1.1.4.2.

For seams that are sewn in the field, the contractor shall provide at least a twometer length of sewn seam for sampling by the engineer before the geotextile
is installed. For seams that are sewn in the factory, the engineer shall obtain
samples of the factory seams at random from any roll of geotextile that is used
on the project.

FHWA NHI-07-092
Geosynthetic Engineering

D - 16

Appendix D
August 2008

X1.1.4.2.1.

For seams that are field sewn, the seams sewn for sampling shall be sewn
using the same equipment and procedures as will be used for the production
seams. If seams are sewn in both the machine and cross machine direction,
samples of seams from both directions shall be provided.

X1.1.4.2.2.

The contractor shall submit the seam assembly description along with the
sample of the seam. The description shall include the seam type, stitch type,
sewing thread, and stitch density.

X1.2.

DRAINAGE GEOTEXTILES2 (See Sections 8.1 & 8.2)

X1.2.1.

Construction:

X1.2.1.1.

Trench excavation shall be done in accordance with details of the project


plans. In all instances excavation shall be done in such a way so as to prevent
large voids from occurring in the sides and bottom of the trench. The graded
surface shall be smooth and free of debris.

X1.2.1.2.

In the placement of the geotextile for drainage applications, the geotextile


shall be placed loosely with no wrinkles or folds, and with no void spaces
between the geotextile and the ground surface. Successive sheets of
geotextiles shall be overlapped a minimum of 300 mm, with the upstream
sheet overlapping the downstream sheet.

X1.2.1.2.1.

In trenches equal to or greater than 300 mm in width, after placing the


drainage aggregate the geotextile shall be folded over the top of the backfill
material in a manner to produce a minimum overlap of 300 mm. In trenches
less than 300 mm but greater than 100 mm wide, the overlap shall be equal to
the width of the trench. Where the trench is less than 100 mm the geotextile
overlap shall be sewn or otherwise bonded. All seams shall be subject to the
approval of the engineer.

X1.2.1.2.2.

Should the geotextile be damaged during installation or drainage aggregate


placement, a geotextile patch shall be placed over the damaged area extending
beyond the damaged area a distance of 300 mm, or the specified seam
overlap, whichever is greater.

Geotextiles used as sheet drains are not included in the discussions in this section.

FHWA NHI-07-092
Geosynthetic Engineering

D - 17

Appendix D
August 2008

X1.2.1.3.

Placement of drainage aggregate should proceed immediately following


placement of the geotextile. The geotextile should be covered with a
minimum of 300 mm of loosely placed aggregate prior to compaction. If a
perforated collector pipe is to be installed in the trench, a bedding layer of
drainage aggregate should be placed below the pipe, with the remainder of the
aggregate placed to the minimum required construction depth.

X1.2.1.3.1.

The aggregate should be compacted with vibratory equipment to a minimum


of 95 percent Standard AASHTO density unless the trench is required for
structural support. If higher compactive effort is required, a Class 1 geotextile
as per Table 1 of the M 288 is needed.

X1.2.1.4.

Figures X1.1 through X1.3 illustrate various geotextile drainage application


details.

Figure X1.1 Geotextile Drain Requirements for Permeable Bases

Figure X1.2 Geotextile Wrapped Longitudinal Edge Drain


FHWA NHI-07-092
Geosynthetic Engineering

D - 18

Appendix D
August 2008

Figure X1.3 Geotextile Wrapped Pavement Under Drain

X1.3

SEPARATION/STABILIZATION GEOTEXTILES (See Sections 8.1,


8.3, and 8.4)

X1.3.1.

Construction:

X1.3.1.1.

The installation site shall be prepared by clearing, grubbing, and excavation or


filling the area to the design grade. This includes removal of topsoil and
vegetation.
NOTE X1 Soft spots and unsuitable areas will be identified during site
preparation or subsequent proof rolling. These areas shall be excavated and
backfilled with select material and compacted using normal procedures.

X1.3.1.2.

The geotextile shall be laid smooth without wrinkles or folds on the prepared
subgrade in the direction of construction traffic. Adjacent geotextile rolls
shall be overlapped, sewn or joined as required in the plans. Overlaps shall be
in the direction as shown on the plans. See Table X1.1 for overlap
requirements.

FHWA NHI-07-092
Geosynthetic Engineering

D - 19

Appendix D
August 2008

TABLE X1.1 Overlay Requirements


Soil CBR

Minimum Overlap

Greater than 3

300 - 450 mm

1-3

0.6 - 1 m

0.5 - 1

1 m or sewn

Less than 0.5

Sewn

All roll ends

1 m or sewn

X1.3.1.2.1.

On curves the geotextile may be folded or cut to conform to the curves. The
fold or overlap shall be in the direction of construction and held in place by
pins, staples, or piles of fill or rock.

X1.3.1.2.2.

Prior to covering, the geotextile shall be inspected to ensure that the geotextile
has not been damaged (i.e., holes, tears, rips) during installation. The
inspection shall be done by the engineer or the engineers designated
representative. It is recommended that the designated representative be a
certified inspector.
Damaged geotextiles, as identified by the engineer, shall be repaired
immediately. Cover the damaged area with a geotextile patch that extends an
amount equal to the required overlap beyond the damaged area.

X1.3.1.3.

The subbase shall be placed by end dumping onto the geotextile from the edge
of the geotextile, or over previously placed subbase aggregate. Construction
vehicles shall not be allowed directly on the geotextile. The subbase shall be
placed such that at least the minimum specified lift thickness shall be between
the geotextile and equipment tires or tracks at all times. Turning of vehicles
shall not be permitted on the first lift above the geotextile.
NOTE X2 On subgrades having a CBR value of less than one, the subbase
aggregate should be spread in its full thickness as soon as possible after
dumping to minimize the potential of localized subgrade failure due to
overloading of the subgrade.

X1.3.1.3.1.

Any ruts occurring during construction shall be filled with additional subbase
material, and compacted to the specified density.

FHWA NHI-07-092
Geosynthetic Engineering

D - 20

Appendix D
August 2008

X1.3.1.3.2.

If placement of the backfill material causes damage to the geotextile, the


damaged area shall be repaired as previously described in Section X3.1.2.1.
The placement procedures shall be then be modified to eliminate further
damage from taking place (i.e., increase initial lift thickness, decrease
equipment loads, etc.).
NOTE X3 In stabilization applications, the use of vibratory compaction
equipment is not recommended with the initial lift of subbase material, as it
may cause damage to the geotextile.

X1.4.

EROSION CONTROL GEOTEXTILES (See Section 8.5.)

X1.4.1

Construction:

X1.4.1.1

The geotextile shall be placed in intimate contact with the soils without
wrinkles or folds and anchored on a smooth graded surface approved by the
engineer. The geotextile shall be placed in such a manner that placement of
the overlying materials will not excessively stretch so as to tear the geotextile.
Anchoring of the terminal ends of the geotextile shall be accomplished
through the use of key trenches or aprons at the crest and toe of slope. Refer
to Figures X1.4 through X1.7 for construction details.
NOTE X4 In certain applications to expedite construction, 450-mm
anchoring pins placed on 600- to 1800-mm centers, depending on the slope of
the covered area, have been used successfully.

X1.4.1.1.1.

The geotextile shall be placed with the machine direction parallel to the
direction of water flow which is normally parallel to the slope for erosion
control runoff and wave action (Figure X1.4), and parallel to the stream or
channel in the case of streambank and channel protection (Figure X1.6).
Adjacent geotextile sheets shall be joined by either sewing or overlapping.
Overlapped seams of roll ends shall be a minimum of 300 mm except where
placed under water. In such instances the overlap shall be a minimum of 1 m.
Overlaps of adjacent rolls shall be a minimum of 300 mm in all instances.
NOTE X5 When overlapping, successive sheets of the geotextile shall be
overlapped upstream over downstream, and/or upslope over downslope. In
cases where wave action or multidirectional flow is anticipated, all seams
perpendicular to the direction of flow shall be sewn.

FHWA NHI-07-092
Geosynthetic Engineering

D - 21

Appendix D
August 2008

Figure X1.4 Method of Placing Geotextile for Protection of Cut and Fill Slopes

Figure X1.5 Cross Section of Slopes with Riprap

Figure X1.6 Geotextile Placement Scheme for Streambank Protection


FHWA NHI-07-092
Geosynthetic Engineering

D - 22

Appendix D
August 2008

Figure X.17Key Detail at Top of Slope for Geotextiles Used for Permanent Erosion Control

X1.4.1.1.2.

Care shall be taken during installation so as to avoid damage occurring to the


geotextile as a result of the installation process. Should the geotextile be
damaged during installation, a geotextile patch shall be placed over the
damaged area extending 1 m beyond the perimeter of the damage.

X1.4.1.2.

The armor system placement shall begin at the toe and proceed up the slope.
Placement shall take place so as to avoid stretching and subsequent tearing of
the geotextile. Riprap and heavy stone filling shall not be dropped from a
height of more than 300 mm. Stone with a mass of more than 100 kg shall not
be allowed to roll down the slope.

X1.4.1.2.1.

Slope protection and smaller sizes of stone filling shall not be dropped from a
height exceeding 1 m, or a demonstration provided showing that the
placement procedures will not damage the geotextile. In underwater
applications, the geotextile and backfill material shall be placed the same day.
All void spaces in the armor stone shall be backfilled with small stone to
ensure full coverage.

X1.4.1.2.2.

Following placement of the armor stone, grading of the slope shall not be
permitted if the grading results in movement of the stone directly above the
geotextile.

FHWA NHI-07-092
Geosynthetic Engineering

D - 23

Appendix D
August 2008

X1.4.1.3.

Field monitoring shall be performed to verify that the armor system placement
does not damage the geotextile.

X1.4.1.3.1.

Any geotextile damaged during backfill placement shall be replaced as


directed by the engineer at the contractors expense.

X1.5.

SILT FENCE GEOTEXTILES (See Section 9.)

X1.5.1.

Related Material Requirements:

X1.5.1.1.

Wood, steel, or synthetic support posts having a minimum length of 1 m plus


the burial depth may be used. They shall be of sufficient strength to resist
damage during installation and to the support applied loads due to material
build up behind the silt fence.
NOTE X6 It has been found that hardwood post having dimensions of at
least 30 mm x 30 mm, No. 2 Southern Pine at least 65 mm x 65 mm, or steel
posts of U, T, L, or C shape, weighing 600 g per 300 mm have performed
satisfactorily.

X1.5.1.2.

Wire or polymer support fence shall be at least 750 mm high and strong
enough to support applied loads. Polymer support fences shall meet the same
ultraviolet degradation requirements as the geotextile.
NOTE X7 Wire support fences having at least six horizontal wires and
being at least 14-gauge wire have performed satisfactorily. Vertical wires
should be a maximum of 150 mm apart.

X1.5.2.

Construction:

X1.5.2.1.

The geotextile at the bottom of the fence shall be buried in a J configuration


to a minimum depth of 150 mm in a trench so that no flow can pass under the
silt fence. The trench shall be backfilled and the soil compacted over the
geotextile.

X1.5.2.1.1.

The geotextile shall be spliced together with a sewn seam only at a support
post, or two sections of fence may be overlapped instead.

FHWA NHI-07-092
Geosynthetic Engineering

D - 24

Appendix D
August 2008

X1.5.2.1.2.

The contractor must demonstrate to the satisfaction of the engineer that the
geotextile can withstand the anticipated sediment loading.

X1.5.2.1.3.

See Figure X1.8 for details.

Figure X1.8 Typical Silt Fence Detail

X1.5.2.2.

The posts shall be placed at spacing as shown on the project plans. Posts
should be driven or placed a minimum of 500 mm into the ground. Depth
shall be increased to 600 mm if fence is placed on a slope of 3:1 or greater.
NOTE X8 Where 500-mm depth is impossible to attain, the posts should be
adequately secured to prevent overturning of the fence due to sediment
loading.

X1.5.2.3.

The support fence shall be fastened securely to the upslope side of the fence
post. The support fence shall extend from the ground surface to the top of the
geotextile.

FHWA NHI-07-092
Geosynthetic Engineering

D - 25

Appendix D
August 2008

X1.5.2.4.

When self-supported fence is used, the geotextile shall be securely fastened to


fence posts.

X1.5.2.5.

Silt fences should be continuous and transverse to the flow. The silt fence
should follow the contours of the site as closely as possible. The fence shall
also be placed such that the water cannot runoff around the end of the fence.

X1.5.2.5.1.

The silt fence should be limited to handle an area equivalent to 90 square


meters per 3 m of fence. Caution should be used where the site slope is
greater than 1:1, and water flow rates exceed 3 L per second per 3 m of fence.

X1.5.3.

Maintenance:

X1.5.3.1.

The contractor shall inspect all temporary silt fences immediately after each
rainfall, and at least daily during prolonged rainfall. The Contractor shall
immediately correct any deficiencies.

X1.5.3.1.1.

The contractor shall also make a daily review of the location of silt fences in
areas where construction activities have altered the natural contour and
drainage runoff to ensure that the silt fences are properly located for
effectiveness. Where deficiencies exist as determined by the engineer,
additional silt fence shall be installed as directed by the engineer.

X1.5.3.1.2.

Damaged or otherwise ineffective silt fences shall be repaired or replaced


promptly.

X1.5.3.2.

Sediment deposits shall either be removed when the deposit reaches half the
height of the fence, or a second silt fence shall be installed as directed by the
engineer.

X1.5.3.3.

The silt fence shall remain in place until the engineer directs it be removed.
Upon removal, the contractor shall remove and dispose of any excess
sediment accumulations, dress the area to give it a pleasing appearance, and
cover with vegetation all bare areas in accordance with contract requirements.

X1.5.3.3.1.

Removed silt fence may be used at other locations provided the geotextile and
other material requirements continue to be met to the satisfaction of the
engineer.

FHWA NHI-07-092
Geosynthetic Engineering

D - 26

Appendix D
August 2008

A6.

PAVING FABRICS (See Section 10.)

X1.6.

Materials:

X1.6.1.1.

The sealant material used to impregnate and seal the paving fabric, as well as
bond it to both the base pavement and overlay, shall be a paving grade asphalt
recommended by the paving fabric manufacturer, and approved by the
engineer.

X1.6.1.1.1.

Uncut asphalt cements are the preferred sealants; however, cationic and
anionic emulsions may be used provided the precautions outlined in Section
X6.3.3 are followed. Cutbacks and emulsions that contain solvents shall not
be used.

X1.6.1.1.2.

The grade of asphalt cement specified for hot-mix design in each geographic
location is generally the most acceptable material.

X1.6.1.2.

Washed concrete sand may be spread over an asphalt saturated paving fabric
to facilitate movement of equipment during construction or to prevent tearing
or delamination of the paving fabric. Hot-mix broadcast in front of
construction vehicle tires may also be used to serve this purpose. If sand
applied, excess quantities shall be removed from the paving fabric prior to
placing the surface course.

X1.6.1.2.1.

Sand is not usually required. However, ambient temperatures are occasionally


sufficiently high to cause bleed-through of the asphalt sealant resulting in
undesirable paving fabric adhesion to construction vehicle tires.

X1.6.2.

Equipment:

X1.2.1.

The asphalt distributor shall be capable of spraying the asphalt sealant at the
prescribed uniform application rate. No streaking, skipping, or dripping will
be permitted. The distributor shall also be equipped with a hand spray having
a single nozzle and positive shut-off valve.

X1.6.2.2.

Mechanical or manual lay down equipment shall be capable of laying the


paving fabric smoothly.

FHWA NHI-07-092
Geosynthetic Engineering

D - 27

Appendix D
August 2008

X1.6.2.3.

The following miscellaneous equipment shall be provided: stiff bristle


brooms or squeegees to smooth the paving fabric; scissors or blades to cut the
paving fabric; brushes for applying asphalt sealant to paving fabric overlaps.

X1.6.2.4.

Pneumatic rolling equipment to smooth the paving fabric into the sealant and
sanding equipment may be required for certain jobs. Rolling is especially
required on jobs where thin lifts or chip seals are being placed. Rolling helps
ensure paving fabric bond to the adjoining pavement layers in the absence of
heat and weight associated with thicker lifts of asphaltic pavement.

X1.6.3.

Construction:

X1.6.3.1.

Neither the asphalt sealant nor the paving fabric shall be placed when weather
conditions, in the opinion of the engineer, are not suitable. Air and pavement
temperatures shall be sufficient to allow the asphalt sealant to hold the paving
fabric in place. For asphalt cements, air temperature shall be 10C and rising.
For asphalt emulsions, air temperature shall be 15C and rising.

X1.6.3.2.

The surface on which the paving fabric is to be placed shall be reasonably free
of dirt, water, vegetation, or other debris. Cracks exceeding 3 mm in width
shall be filled with suitable crack filler. Potholes shall be properly repaired as
directed by the engineer. Fillers shall be allowed to cure prior to paving fabric
placement.

X1.6.3.3.

The specified rate of asphalt sealant application must be sufficient to satisfy


the asphalt retention properties of the paving fabric, and bond the paving
fabric and overlay to the old pavement.
NOTE X9 When emulsions are used, the application rate must be increased
to offset water content of the emulsion.

X1.6.3.3.1.

Application of the sealant shall be by distributor spray bar, with hand spraying
kept to a minimum. Temperature of the asphalt sealant shall be sufficiently
high to permit uniform spray pattern. For asphalt cements the minimum
temperature shall be 145C. To avoid damage to the paving fabric, however,
the distributor tank temperature shall not exceed 160C.

FHWA NHI-07-092
Geosynthetic Engineering

D - 28

Appendix D
August 2008

X1.6.3.3.2.

Spray patterns for asphalt emulsion are improved by heating. Temperatures in


the 55C to 70C range are desirable. A temperature of 70C shall not be
exceeded since higher temperatures may break the emulsion.

X1.6.3.3.3.

The target width of asphalt sealant application shall be the paving fabric width
plus 150 mm. The asphalt sealant shall not be applied any farther in advance
of paving fabric placement than the distance the contractor can maintain free
of traffic.

X1.6.3.3.4.

Asphalt spills shall be cleaned from the road surface to avoid flushing and
paving fabric movement.

X1.6.3.3.5.

When asphalt emulsions are used, the emulsion shall be cured prior to placing
the paving fabric and final wearing surface. This means essentially no
moisture remaining.

X1.6.3.4.

The paving fabric shall be placed onto the asphalt sealant with minimum
wrinkling prior to the time the asphalt has cooled and lost tackiness. As
directed by the engineer, wrinkles or folds in excess of 25 mm shall be slit and
laid flat.

X1.6.3.4.1.

Brooming and/or pneumatic rolling will be required to maximize paving


fabric contact with the pavement surface.

X1.6.3.4.2.

Overlap of paving fabric joints shall be sufficient to ensure full closure of the
joint, but should not exceed 150 mm. Transverse joints shall be lapped in the
direction of paving to prevent edge pickup by the paver. A second application
of asphalt sealant to the paving fabric overlaps will be required if in the
judgement of the engineer additional asphalt sealant is needed to ensure
proper bonding of the double paving fabric layer.

X1.6.3.4.3.

Removal and replacement of paving fabric that is damaged will be the


responsibility of the contractor.
NOTE X10 The problems associated with wrinkles are related to thickness of the
asphalt lift being placed over the paving fabric. When wrinkles are large enough to
be folded over, there usually is not enough asphalt available from the tack coat to
satisfy the requirement of multiple layers of paving fabric. Therefore, wrinkles
should be slit and laid flat. Sufficient asphalt sealant should be sprayed on the top of
the paving fabric to satisfy the requirement of the lapped paving fabric.

FHWA NHI-07-092
Geosynthetic Engineering

D - 29

Appendix D
August 2008

NOTE X11 In overlapping adjacent rolls of paving fabric it is desirable to


keep the lapped dimension as small as possible and still provide a positive
overlap. If the lapped dimension becomes too large, the problem of
inadequate tack to satisfy the two lifts of paving fabric and the old pavement
may occur. If this problem does occur, then additional asphaltic sealant
should be added to the lapped areas. In the application of the additional
sealant, care should be taken not to apply too much since excess will cause
flushing.
X1.6.3.4.4.

Trafficking the paving fabric will be permitted for emergency and


construction vehicles only.

X1.6.3.5.

Placement of the hot-mix overlay should closely follow paving fabric


laydown. The temperature of the mix shall not exceed 160C. In the event
asphalt bleeds through the paving fabric causing construction problems before
the overlay is placed, the affected areas shall be blotted by spreading sand. To
avoid movement of, or damage to, the seal-coat saturated paving fabric,
turning of the paver and other vehicles shall be gradual and kept to a
minimum.

X1.6.3.6.

Prior to placing a seal coat (or thin overlay such as an open-graded friction
2

course), lightly sand the paving fabric at a spread rate of 0.65 to 1 kg per m ,
and pneumatically roll the paving fabric tightly into the sealant.

ADVISORY
It is recommended that for safety considerations, trafficking of the paving
fabric should not be allowed. However, if the contracting agency elects to
allow trafficking, the following verbiage is recommended:
If approved by the engineer, the seal-coat saturated paving fabric may be
opened to traffic for 24 to 48 hours prior to installing the surface course.
Warning signs shall be placed which advise the motorist that the surface may
be slippery when wet. The signs shall also post the appropriate safe speed.
Excess sand shall be broomed from the surface prior to placing the overlay.
If, in the judgement of the engineer, the fabric surface appears dry and lacks
tackiness following exposure to traffic, a light tack coat shall be applied prior
to the overlay.
FHWA NHI-07-092
Geosynthetic Engineering

D - 30

Appendix D
August 2008

Appendix E
GEOSYNTHETIC TEST STANDARDS

E-1 American Society for Testing and Materials

GENERAL
D 4439 Standard Terminology for Geosynthetics
D 4873 Identification, Storage, and Handling of Geotextiles
MECHANICAL PROPERTIES
Specification for:
D 4533 Trapezoid Tearing Strength of Geotextiles
D 4595 Tensile Properties of Geotextiles by the Wide Strip Method
D 4632 Breaking Load and Elongation of Geotextiles (Grab Method)
Practices for:
D 4354 Sampling of Geosynthetics for Testing
D 4759 Specification Conformance of Geosynthetics, Determining
D 5818 Obtaining Samples of Geosynthetics from a Test Section for Assessment of
Installation Damage
Test Method for:
D 4833 Index Puncture Resistance of Geotextiles, Geomembranes, and Related Products
D 4884 Seam Strength of Sewn or Thermally Bonded Seams of Geotextiles
D 5261 Measuring Mass Per Unit Area of Geotextiles
D 5321 Determining the Coefficient of Soil and Geosynthetic or Geosynthetic and
Geosynthetic Friction by the Direct Shear Method
D 6241 Static Puncture Strength of Geotextiles and Geotextile-Related Products Using a
50-mm Probe
D 6364 Determining Short-Term Compression Behavior of Geosynthetics
D 6637 Determining Tensile Properties of Geogrids by the Single or Multi-Rib Tensile
Method
D 6638 Determining Connection Strength Between Geosynthetic Reinforcement and
Segmental Concrete Units (Modular Concrete Blocks)
D 6706 Measuring Geosynthetic Pullout Resistance in Soil
D 6916 Determining the Shear Strength Between Segmental Concrete Units (Modular
Concrete Blocks)
FHWA NHI-07-092
Geosynthetic Engineering

E-1

Appendix E
August 2008

D 7005 Determining the Bond Strength (Ply Adhesion) of Geocomposites


D 7179
Determining Geonet Breaking Force
PERMEABILITY AND FILTRATION
Specification for:
D 6707
Circular-Knit Geotextile for Use in Subsurface Drainage Applications
D 7001
Geocomposites for Pavement Edge Drains and Other High-Flow Applications
Practices for:
D 6088
Practice for Installation of Geocomposite Pavement Drains
Test Method for:
D 4491 Water Permeability of Geotextiles by the Permittivity Method
D 4716 Constant Head Hydraulic Transmissivity (In Plane Flow) of Geotextiles and
Geotextile Related Products
D 4751 Apparent Opening Size of a Geotextile, Determining
D 5101 Measuring the Soil-Geotextile System Clogging Potential (By the Gradient Ratio)
D 5141 Filtering Efficiency and Flow Rate of a Geotextile for Silt Fence Application
Using Site-Specific Soil
D 5199 Measuring Nominal Thickness of Geotextiles and Geomembranes
D 5321 Coefficient of Soil and Geosynthetic or Geosynthetic and Geosynthetic Friction
by the Direct Shear Method, Determining the
D 5493 Permittivity of Geotextiles Under Load
D 5567 Hydraulic Conductivity Ratio (HCR) Testing of Soil/Geotextile Systems
D 6244 Vertical Compression of Geocomposite Pavement Panel Drains
D 6574 Determining the (In-Plane) Hydraulic Transmissivity of a Geosynthetic by Radial
Flow
D 6767 Pore Size Characteristics of Geotextiles by Capillary Flow Test
D 6918 Testing Vertical Strip Drains in the Crimped Condition
Guide for:
D 6917 Selection of Test Methods for Prefabricated Vertical Drains (PVD)
ENDURANCE PROPERTIES
Specification for:
D 4355 Deterioration of Geotextiles from Exposure to Ultraviolet Light and Water
(Xenon-Arc Type Apparatus)
D 4594 Effects of Temperature on Stability of Geotextiles
D 5322 Immersion Procedures for Evaluating the Chemical Resistance of Geosynthetics
to Liquids
FHWA NHI-07-092
Geosynthetic Engineering

E-2

Appendix E
August 2008

Test Methods for:


D 1987 Biological Clogging of Geotextile or Soil/Geotextile Filters
D 4886 Abrasion Resistance of Geotextiles (Sand/Sliding Block Method)
D 5262 Unconfined Tension Creep Behavior of Geosynthetics, Evaluating the
D 5397 Stress Crack Resistance of Polyolefin Geomembranes Using Notched Constant
Tensile Load Test, Evaluation of
D 5596 Microscopic Evaluation of the Dispersion of Carbon Black in Polyolefin
Geosynthetics
D 5885 Oxidative Induction Time of Polyolefin Geosynthetics by High Pressure
Differential Scanning Calorimetry
D 6992 Accelerated Tensile Creep and Creep-Rupture of Geosynthetic Materials Based
on Time-Temperature Superposition Using the Stepped Isothermal Method
Practice for:
D 5496 In Situ Immersion Testing of Geosynthetics
D 5721 Air-Oven Aging of Polyolefin Geomembranes
D 5747 Tests to Evaluate the Chemical Resistance of Geomembranes to Liquids
D 5970 Deterioration of Geotextiles from Outdoor Exposure
D 6213 Tests to Evaluate the Chemical Resistance of Geogrids to Liquids
D 6388 Tests to Evaluate the Chemical Resistance of Geonets to Liquids
D 6389 Tests to Evaluate the Chemical Resistance of Geotextiles to Liquids
Guide for:
D 5819 Selecting Test Methods for Experimental Evaluation of Geosynthetic Durability

GEOMEMBRANES
Specification for:
D 2643 Prefabricated Asphalt Reservoir, Pond, Canal, and Ditch Liner (Exposed Type)
D 3020 Polyethylene and Ethylene Copolymer Plastic Sheeting for Pond, Canal, and
Reservoir Lining
D 3083 Flexible Poly(Vinyl Chloride) Plastic Sheeting for Pond, Canal, and Reservoir
Lining
D 4885 Determining Performance Strength of Geomembranes by the Wide Strip Tensile
Method
D7177 Air Channel Evaluation of Polyvinyl Chloride (PVC) Dual Track Seamed
Geomembranes

Practices for:
FHWA NHI-07-092
Geosynthetic Engineering

E-3

Appendix E
August 2008

D 4437 Integrity of Field Seams Used in Joining Flexible Polymeric Sheet


Geomembranes, Determining
D 4545 Integrity of Factory Seams Used in Joining Manufactured Flexible Sheet
Geomembranes, Determining
D 5323 2% Secant Modulus for Polyethylene Geomembranes, Determination of
D 5641 Geomembrane Seam Evaluation by Vacuum Chamber
D 5820 Pressurized Air Channel Evaluation of Dual Seamed Geomembranes
D6365 Nondestructive Testing of Geomembrane Seams using the Spark Test
D7006 Ultrasonic Testing of Geomembranes
D7007 Electrical Methods for Locating Leaks in Geomembranes Covered with Water or
Earth Materials
D7240 Leak Location using Geomembranes with an Insulating Layer in Intimate Contact
with a Conductive Layer via Electrical Capacitance Technique (Conductive
Geomembrane Spark Test)
Test Method for:
D 5494 Pyramid Puncture Resistance of Unprotected and Protected Geomembranes,
Determination of
D 5514 Large Scale Hydrostatic Puncture Testing of Geosynthetics
D 5617 Multi-Axial Tension Test for Geosynthetics
D 5884 Determining the Tearing Strength of Internally Reinforced Geomembranes
D5994 Measuring Core Thickness of Textured Geomembrane
D6214 Determining the Integrity of Field Seams Used in Joining Geomembranes by
Chemical Fusion Methods
D6392 Determining the Integrity of Nonreinforced Geomembrane Seams Produced
Using Thermo-Fusion Methods
D6636 Determination of Ply Adhesion Strength of Reinforced Geomembranes
D7004 Grab Tensile Properties of Reinforced Geomembranes
D7056 Determining the Tensile Shear Strength of Pre-Fabricated Bituminous
Geomembrane Seams
Guide for:
D 5886 Selection of Test Methods to Determine the Rate of Fluid Permeation Through
Geomembranes for Specific Applications
D6434 Selection of Test Methods for Flexible Polypropylene (fPP) Geomembranes
D6455 Selection of Test Methods for Prefabricated Bituminous Geomembranes (PBGM)
D6497 Mechanical Attachment of Geomembrane to Penetrations or Structures
D7106 Selection of Test Methods for Ethylene Propylene Diene Terpolymer (EPDM)
Geomembranes

FHWA NHI-07-092
Geosynthetic Engineering

E-4

Appendix E
August 2008

GEOSYNTHETIC CLAY LINERS


Test Method for:
D 5887 Measurement of Index Flux Through Saturated Geosynthetic Clay Liner
Specimens Using Flexible Wall Permeameter
D 5890 Swell Index of Clay Mineral Component of Geosynthetic Clay Liners
D 5891 Fluid Loss of Clay Component of Geosynthetic Clay Liners
D 5993 Measuring Mass Per Unit of Geosynthetic Clay Liners
D 6243 Determining the Internal and Interface Shear Resistance of Geosynthetic Clay
Liner by the Direct Shear Method
D 6496 Test Method for Determining Average Bonding Peel Strength Between Top and
Bottom Layers of Needle-Punched Geosynthetic Clay Liners
D 6766 Evaluation of Hydraulic Properties of Geosynthetic Clay Liners Permeated
with Potentially Incompatible Liquids
D 6768 Tensile Strength of Geosynthetic Clay Liners
D 5993 Measuring Mass Per Unit of Geosynthetic Clay Liners
D 6248 Determining the Internal and Interface Shear Resistance of Geosynthetic Clay
Liner by the Direct Shear Method
D 6496 Determining Average Bonding Peel Strength Between Top and Bottom Layers of
Needle-Punched Geosynthetic Clay Liners
D 6766 Evaluation of Hydraulic Properties of Geosynthetic Clay Liners Permeated with
Potentially Incompatible Liquids
D 6768 Tensile Strength of Geosynthetic Clay Liners
Practices for:
D 5888 Storage and Handling of Geosynthetic Clay Liners
D 5889 Quality Control of Geosynthetic Clay Liners
Guide for:
D6072 Obtaining Samples of Geosynthetic Clay Liners
D6102 Installation of Geosynthetic Clay Liners
D6141 Screening Clay Portion of Geosynthetic Clay Liner (GCL) for Chemical
Compatibility to Liquids
D6495 Acceptance Testing Requirements for Geosynthetic Clay Liners
D6072 Obtaining Samples of Geosynthetic Clay Liners
D6102 Installation of Geosynthetic Clay Liners
D6141 Screening Clay Portion of Geosynthetic Clay Liner (GCL) for Chemical
Compatibility to Liquids
D6495 Acceptance Testing Requirements for Geosynthetic Clay Liners

FHWA NHI-07-092
Geosynthetic Engineering

E-5

Appendix E
August 2008

TURF REINFORCEMENT MATS


Test Method for:
D 6454 Determining the Short-Term Compression Behavior of Turf Reinforcement Mats
D 6524 Measuring the Resiliency of Turf Reinforcement Mats
D 6525 Measuring Nominal Thickness of Permanent Rolled Erosion Control Products
D 6566 Measuring Mass per Unit Area of Turf Reinforcement Mats
D 6567 Measuring the Light Penetration of a Turf Reinforcement Mat
D 6575 Determining Stiffness of Geosynthetics Used as Turf Reinforcement Mats
D 6818 Ultimate Tensile Properties of Turf Reinforcement Mats

MISCELLANEOUS
Specification for:
D 6817 Rigid Cellular Polystyrene Geofoam
D 6826 Sprayed Slurries, Foams and Indigenous Materials Used As Alternative Daily
Cover for Municipal Solid Waste Landfills
D 7008 Geosynthetic Alternate Daily Covers
Test Method:
D 6140 Determine Asphalt Retention of Paving Fabrics Used in Asphalt Paving for FullWidth Applications
Guide for:
D 6523 Evaluation and Selection of Alternative Daily Covers (ADCs) for Sanitary
Landfills
D 7180 Use of Expanded Polystyrene (EPS) Geofoam in Geotechnical Projects

E-2

GT1
GT2
GT3
GT4
GT5

Geosynthetic Research Institute


GRI TEST METHODS & STANDARDS
Geotextile (GT) Related
Clogging Potential via Long-Term Flow Rate
Biological Clogging of Geotextile or Soil/Geotextile Filters (discontinued, see
ASTM D 1987)
Deterioration of Geotextiles from Outdoor Exposure (discontinued, see ASTM D
5970)
Geotextile Permittivity-Under-Load (discontinued, see ASTM D 5493)
Tension Creep Testing of Geotextiles (discontinued, see ASTM D 5262)

FHWA NHI-07-092
Geosynthetic Engineering

E-6

Appendix E
August 2008

GT6
GT7
GT8
GT 9
GT10

Geotextile Pullout (discontinued, see ASTM D 6706)


Determination of Long-Term Design Strength of Geotextiles
Fine Fraction Filtration Using Geotextile Filters
Grip Types for Use in Wide Width Testing of Geotextiles and Geogrids
Test Methods, Properties and Frequencies for High Strength Geotextile Tubes Used
as Coasted and Riverine Structures
GT11 Installation of Geotextile Tubes Used as Coastal and Riverine Structures
GT12a Test Methods and Properties for Nonwoven Geotextiles Used as Protection (or
Cushioning) Materials ASTM Version
GT12b Test Methods and Properties for Nonwoven Geotextiles Used as Protection (or
Cushioning) Materials ISO Version
GT13 Test Methods and Properties for Geotextiles Used as Separation Between Subgrade
Soil and Aggregate
GT14 Hanging Bag Test for Field Assessment of Fabrics Used for Geotextile Tubes and
Containers

GG9

Geogrid (GG) Related


Geogrid Rib Tensile Strength
Geogrid Junction Strength
Tension Creep Testing of Stiff Geogrids (discontinued, see ASTM D 5262)
Tension Creep Testing of Flexible Geogrids (discontinued, see ASTM D 5262)
Determination of the Long-Term Design Strength of Stiff Geogrids
Determination of the Long-Term Design Strength of Flexible Geogrids
Test Method for Geogrid Pullout (discontinued, see ASTM D 6706)
Grip Types for Use in Wide Width Testing of Geotextiles and Geogrids
Carboxyl End Group Content of PET Yarns
Determination of the Number Average Molecular Weight of PET yarns based on a
Relative Viscosity Value
Torsional Behavior of Bidrectional Geogrids When Subjected to In-Plane Rotation

GN1

Geonet (GN) Related


Compression Behavior of Geonets (discontinued, see ASTM D 6364)

GM1
GM2
GM3

Geomembrane (GM) Related


Seam Evaluation by Ultrasonic Shadow Method
Embedment Depth for Anchorage Mobilization
Large Scale Hydrostatic Puncture Test

GG1
GG2
GG3a
GG3b
GG4a
GG4b
GG5
GG6
GG7
GG8

FHWA NHI-07-092
Geosynthetic Engineering

E-7

Appendix E
August 2008

GM4
GM5a
GM5b
GM5c
GM6
GM7
GM 8
GM9
GM10
GM11
GM12
GM13
GM14
GM15
GM16
GM17
GM17

GM 19
GM 20
GM 21

GCL1

Three Dimensional Geomembrane Tension Test (discontinued, see ASTM D 5617)


Notched Constant Tensile Load (NCTL) Test for Polyolefin Resins or
Geomembranes (discontinued, see ASTM D 5397)
Single Point NCTL Test for Polyolefin Resin or Geomembranes (discontinued, see
ASTM D 5397 Appendix)
Seam Constant Tensile Load (SCTL) Test for Polyolefin Geomembrane Seams
Pressurized Air Channel Test for Dual Seamed Geomembranes
Accelerated Curing of Geomembrane Test Strip Seams Made by Chemical Fusion
Methods
Measurement of the Core Thickness of Textured Geomembranes
Cold Weather Seaming of Geomembranes
Specification for Stress Crack Resistance of HDPE Geomembrane Sheet
Accelerated Weathering of Geomembranes Using a Fluorescent UVA Device
Asperity Measurement of Textured Geomembranes Using a Depth Gage
Test Properties, Testing Frequency and Recommended Warrant for High Density
Polyethylene (HDPE) Smooth and Textured Geomembranes
Selecting Variable Intervals for Taking Geomembrane Destructive Seam Samples
Using the Method of Attributes
Determination of Ply Adhesion of Reinforced Geomembranes (discontinued, see
ASTM D 6636)
Observation of Surface Cracking of Geomembranes
Test Properties, Testing Frequency and Recommended Warrant for Linear Low
Density Polyethylene (LLDPE) Smooth and Textured Geomembranes
Test Properties, Testing Frequency and Recommended Warrant for Flexible
Polyproplylene (fPP and fPP-R) Nonreinforced and Reinforced Geomembranes
(Temporarily suspended: 3 May 2004)
Seam Strength and Related Properties of Thermally Bonded Polyolefin
Geomembranes
Selecting Variable Intervals for Taking Geomembrane Destructive Seam Samples
Using Control Charts
Test Methods, Properties, Frequency and Recommended Warrant for Ethylene
Propylene Diene Terpolymer (EDPM) Nonreinforced and Scrim Reinforced
Geomembranes

Geosynthetic Clay Liner (GCL) Related


Swell Measurement of the Clay Component of GCL's (discontinued, see ASTM D
5890)

FHWA NHI-07-092
Geosynthetic Engineering

E-8

Appendix E
August 2008

GCL2
GCL3

GC1
GC2
GC3
GC4
GC5
GC6
GC 7
GC 8

GS1
GS2
GS3
GS4
GS5
GS6
GS7
GS8
GS9
GS 10

Permeability of Geosynthetic Clay Liners (GCLs) (discontinued, see ASTM D 5887


and ASTM D 6766)
Test Methods, Required Properties, and Testing Frequencies of Geosynthetic Clay
Liners (GCLs)
Geocomposite (GC) Related
Soil-Filter Core Combined Flow Test
Strip (Wick) Drain Flow Rate Under Load
Strip Drain Kinking Efficiency
Compression Behavior of Prefabricated Edge Drains and Sheet Drains
Erosion Control Systems to Protect Against Soil Detachment by Rainfall Impact and
Overload Flow Transport
Erosion Control Systems for High Velocity Flows in Channels
Determination of Adhesion and Bond Strength of Geocomposites (discontinued, see
ASTM D 7005)
Determination of the Allowable Flow Rate of a Drainage Geocomposite

Geosynthetic (GS) Related (i.e., Multipurpose)


CBR Puncture Strength (discontinued, see ASTM D 6241)
Rupture Strength by Pendulum Impact for Geotextiles-GeomembranesGeocomposites
In-Situ Monitoring of the Mechanical Performance of Geosynthetics
Time Dependent (Creep) Deformation Under Normal Pressure
Impregnation of Geosynthetics Under Load
Interface Friction Determination by Direct Shear Testing (discontinued, see ASTM
D 5321)
Determining the Index Friction Properties of Geosynthetics
Determining the Connection Strength of Mechanically Anchored Geosynthetics
Oxidative Induction Time of Polyethylene Geosynthetics by High Pressure
Differential Scanning Calorimetry (discontinued, see ASTM D 5885)
Accelerated Tensile Creep and Creep Rupture of Geosynthetic Materials Based on
Time Temperature Superposition Using the Stepped Isothermal Method
(discontinued, see ASTM D 6992)

FHWA NHI-07-092
Geosynthetic Engineering

E-9

Appendix E
August 2008

[ BLANK]

FHWA NHI-07-092
Geosynthetic Engineering

E - 10

Appendix E
August 2008

Appendix F
REPRESENTATIVE1 LIST OF GEOSYNTHETIC
MANUFACTURERS AND SUPPLIERS
Agru/America Inc.

Carthage Mills

Address: 500 Garrison Rd


Georgetown, SC 29440-9688

Contact: Alan L Ossege


Address: 4243 Hunt Rd
Cincinnati, OH 45242-6645

Phone: 843 546 0600


Toll Free: 800 373 2478

Phone: 513 794 1600

Fax: (843) 546-0516

Toll Free: 800 543 4430

Email: salesmkg@agruamerica.com

Fax: (513) 794-3434

Website: www.agruamerica.com

Email: info@carthagemills.com
Website: www.carthagemills.com

BBA Fiberweb/Biobarrier
Address: 70 Old Hickory Blvd
Old Hickory, TN 37138-3159

Colbond Inc.
Address: PO Box 1057
1301 Sand Hill Rd
Enka, NC 28728-1057

Phone: 615 847 7132


Toll Free: 800 284 2780 x7132

Phone: 828 665 5000

Fax: (615) 847-7068

Toll Free: 800 365 7391


Fax: (828) 665-5009

Belton Industries Inc.

Email: info@colbond-usa.com

Contact: Bob Moran

Website: www.colbond-usa.com

Address: 5600 Oakbrook Pkwy Ste 150


Norcross, GA 30093-1843
Phone: (770) 248-1927

Contech Construction Products

Toll Free: +1 800 225 4099

Address: 9025 Centre Pointe Dr


West Chester, OH 45069-4984

Fax: +1 770 248 1655, 800 851

Phone: (513) 645-7165


Toll Free: 800 338 1122

Bradley Industrial Textiles, Inc.

Fax: (513) 645-7993

Address: 101 John Simms Parkway


Valparaiso, FL 32580
Phone: 850 678 6111

Cooley Group

Fax: (850) 729-1052

Address: 50 Esten Ave


Pawtucket, RI 02860-4840

Email: geotextile@aol.com

Phone: 401 724 9000

Website: www.bradley-geotextile.com

Toll Free: 800 333 3048


Fax: (401) 728-1910
1.

List is from the Geosynthetics Materials


Association membership lists.

FHWA NHI-07-092
Geosynthetic Engineering

Email: info@cooleygroup.com
Website: www.cooleygroup.com

F - 11

Appendix F
August 2008

Firestone Building Products Co.

In-Line Plastics LC
Address: 8615 Golden Spike Ln
Houston, TX 77086-3238

Address: 310 E 96th St


Indianapolis, IN 46240-3702
Phone: +1 317 575 7232

Phone: 281 272 1660

Toll Free: +1 800 428 4442

Toll Free: 800 364 7688


Fax: (281) 272-1673

Fax: +1 317 575 7002

Website: www.in-lineplastics.com
Geo Products LLC d/b/a
Envirogrid

LINQ Industrial Fabrics Inc.

Address: 8615 Golden Spike Ln


Houston, TX 77086-3238

Address: 2550 W 5th North St


Summerville, SC 29483-9665

Phone: (281) 820-5493

Phone: 843 873 5800

Toll Free: 800 434 4743

Toll Free: 800 445 4675

Fax: (281) 820-5499

Fax: (843) 875-8276

Website: www.geoproducts.org

Website: www.linq.com

Geotech Solutions, Inc.

Luckenhaus Technical Textiles


Inc.

Address: 94-155 C
Leowaena Street
Waipahu, HI 96797

Address: 175 Mehler Ln


Martinsville, VA 24112-2037

Phone: +1 808 677 1580

Phone: 276 632 1605

Fax: +1 808 671 5919

Toll Free: 866 513 6275


Fax: (276) 638-4704

GSE Lining Technology Inc.


Address: 19103 Gundle Rd
Houston, TX 77073-3515

Maccaferri Inc.
Contact: Massimo Ciarla

Phone: 281 443 8564

Address: 10303 Governor Lane Blvd


Williamsport, MD 21795-3115

Toll Free: 800 435 2008


Fax: (281) 230-8650

Phone: 301 223 6910


Toll Free: 800 638 7744

Huesker Inc.

Fax: (301) 223-6134

Contact: Thomas G Collins

Email: hdqtrs@maccaferri-usa.com

Address: PO Box 411529


Charlotte, NC 28241-1529

Website: www.maccaferri-usa.com

Phone: 704 588 5500


Toll Free: 800 942 9418
Fax: (704) 588-5988
Website: www.hueskerinc.com

FHWA NHI-07-092
Geosynthetic Engineering

F - 12

Appendix F
August 2008

Strata Systems Inc.


Div. John Boyle & Co. Inc.

Poly-Flex Inc./Poly-Flex
Construction

Address: 380 Dahlonega St Ste 200


Cumming, GA 30040-8218

Address: 2000 W Marshall Dr


Grand Prairie, TX 75051-2709
Phone: 972 337 7280

Phone: 770 888 6688

Toll Free: 888 765 9359

Toll Free: 800 680 7750


Email: strata@geogrid.com

Fax: (972) 337-7233

Website: www.geogrid.com

Website: www.poly-flex.com
Propex Fabrics Inc.

Tenax Corp.

Address: 6025 Lee Highway, Ste. 425


P.O. Box 22788
Chattanooga, TN 37422

Address: 4800 E Monument St


Baltimore, MD 21205-3031
Phone: 410 522 7000

Phone: (423) 892-8080

Toll Free: 800 356 8495

Toll Free: 800 621 1273

Fax: (410) 522-7015

Fax: (423) 899-7619

Website: www.tenax.net

Website: www.propexinc.com
TenCate Geosynthetics
Saint-Gobain Technical Fabrics

Address: 365 S Holland Dr


Pendergrass, GA 30567-4625

Address: 1795 Baseline Rd


Grand Island, NY 14072-2010

Phone: (706) 693-2237

Phone: (716) 775-3900

Toll Free: 800 685 9990

Toll Free: 888 549 7667

Fax: (706) 693-4400

Fax: (716) 775-3901

Website: www.mirafi.com

Skaps Nonwoven Div.

Tensar Earth Technologies Inc.

Address: 335 Athena Dr


Athens, GA 30601-1616

Address: 5883 Glenridge Dr NE Ste 200


Atlanta, GA 30328-5571

Phone: (706) 354-3700

Phone: (404) 260 1290

Fax: (706) 354-3737

Toll Free: 888 836 7271


Fax: (404) 260 1290

Stevens-Geomembranes

Email: info@tensarcorp.com

Address: 9 Sullivan Rd
Holyoke, MA 01040-2841

Website: www.tensarcorp.com

Phone: 413 533 8100


Toll Free: 800 621 2281
Fax: (413) 552-1198

FHWA NHI-07-092
Geosynthetic Engineering

F - 13

Appendix F
August 2008

Layfield Geosynthetics &


Industrial Fabrics Ltd.

WEBTEC LLC
Address: 6509-C Northpark Blvd
Charlotte, NC 28216-2368

Address: 11603 180th St SW


Edmonton, T5S 2H6 Canada

Phone: 704 398 0954

Phone: (780) 453-6731

Toll Free: 800 438 0027

Toll Free: 800 840 2884

Fax: (704) 394-7946

Fax: (780) 455-5218

Email: info@webtecgeos.com

Website: www.geomembranes.com

Website: www.webtecgeos.com

Nilex, Inc.

Cosella-Doerken Products Inc.

Address: 9304 39th Ave


Edmonton, T8A 5Z6 Canada

Address: 4655 Delta Way


Beamsville, L0R 1B4 Canada

Phone: (780) 463-9535

Phone: (905) 563 3255

Toll Free: 800 667 4811

Toll Free: 888 433 5824

Fax: (780) 463-1773

Fax: (905) 563 5582

Website: www.nilex.com

Website: www.cosella-dorken.com

Solmax Intl. Inc.


Address: 2801 Marie-Victorin Blvd
Varennes, J3X 1P7 Canada
Phone: (450) 929-1234
Toll Free: 800 571 3904
Fax: (450) 929-2547

FHWA NHI-07-092
Geosynthetic Engineering

F - 14

Appendix F
August 2008

2,400 14,400
1,700 24,000*

7.0 - 22
4.0 21

1.5 7.0
2.0 7.0
7.0 25

4.0 7.0
7.0 - 21
7.0 21

Nonwoven
Continuous Filament-Melt Bonded
Needlepunched (lightweight)
Needlepunched (heavyweight)

Geogrid
Polypropylene
High-Density Polyethylene
Polyester
10 - 20
10 - 20
5 - 15

30 - 100
40 - 150
40 - 150

15 - 40
10 - 30

20 - 40
20 - 40

Strain at Ultimate4
Tensile Strength
(%)

6,200 16,000
3,800 4,800
24,000 178,000

1,200 6,200
140 1,700
600 3,800

12,000 48,000
1,200 72,000*

4,800 18,000
3,500 18,000

Secant4 Modulus
at 10% Strain
(lb/ft)

n/a
n/a
n/a

40 - 400
40 - 250
160 - 520

160 - 1400
160 - 2000*

160 - 520
70 - 360

Grab5
Strength
(lb)

n/a
n/a
n/a

18 - 100
45 - 125
100 - 250

160 - 250
45 - 315

70 - 160
18 - 135

Puncture6
Strength
(lb)

FHWA NHI-07-092
Geosynthetic Engineering

G-1

Appendix G
August 2008

1. Data was obtained from numerous sources, in some cases estimated, and represents an average range. There may be products outside this range. No
relation should be inferred between maximum and minimum limits for different tests.
2. Both directions
3. Method 1.1.84, Appendix B, FHWA Geotextile Engineering Manual
4. Wide Width Method, ASTM D-4595
5. ASTM D-4632
6. ASTM D-4833
7. ASTM D-3786
8. ASTM D-4533
9. ASTM D-4491
* Limited by test machine

550 2,400
550 6,200
2,400 9,600

270 2,400
270 1,200
550 2,400

1,100 4,800
820 3,100

Ultimate
Tensile4
Strength (kN/m)

3.5 7.0
1.5 5.0

Weight 3
(oz/yd2)

Woven
Monofilament-Polypropylene
Silt-Film
Fibrillated Tape and Multifilament
Polypropylene
Multifilament-Polyester

Geotextile Type

n/a
n/a
n/a

80 - 500
150 - 400
300 - 1000

600 - 1500*
500 - 1500*

400 - 700
200 - 700

Burst7
Strength
(psi)

Table G1 - General Range of Strength and Permeability


Properties1, 2 for Representative Types of Geotextiles and Geogrids

Appendix G
GENERAL PROPERTIES AND COSTS OF GEOTEXTILES AND GEOGRIDS

n/a
n/a
n/a

28 - 200
28 - 160
70 - 200

100 - 400
80 - 520

45 - 100
45 - 360

Tear8
Strength
(lb)

> 10
> 10
> 10

10-4 - 10-2
10-3 - 10-2
10-4 - 10-2

10-4 - 10-3
10-4 - 10-3

10-4 - 10-2
10-4 - 10-3

Equivalent9 Darcy
Permeability
(m/sec)

Table G2
General Description of Geotextiles
Description
Property

LOW

MODERATE

HIGH

Tensile Strength (lb/ft)

< 1,000

1,000 3,500

> 3,500

, @ ult.

< 20%

20% - 50%

> 50%

Burst (psi)

< 200

200 - 500

> 500

Permeability, k (m/s)

< 0.0001

0.0001 - 0.001

> 0.001

Cost ($/yd2)

< $1

$1 - $2

> $2

Approximate Cost

1,2

Table G3
Range of Geotextiles and Geogrids
Material Cost1,2
( $/yd2 )

Geosynthetic
Filtration Geotextiles - high survivability

1.00 - 2.00

Erosion Control Mats

2.50 - 9.50

Temporary Erosion Control Blankets

0.75 - 2.50

Roadway Geotextile Separators - high survivability

1.00 - 1.75

Subgrade Stabilization Geotextiles

1.50 - 3.00

Asphalt Overlay Geotextiles

0.60 - 1.25

Geotextile Embankment Reinforcement3

2.50 - 12.00

Biaxial Geogrid for Subgrade Stabilization

2.50 - 4.50

Geogrid/Geotextile Wall and Slope Reinforcement4,5


- per 1,000 lb/ft long-term allowable strength

1.50 - 3.50

NOTES:
1. Typical costs for materials delivered on-site, for use in Engineer's estimate. Costs are exclusive of installation
and Contractor's markup.
2. Installation cost of geosynthetics typically cost $0.30 to $0.90/yd2, except for very soft ground and underwater
placement.
3. Assumes design strength is based upon a 5% to 10% strain criteria with an ASTM D 4595 test.
4. Assumes allowable design strength is based upon a complete evaluation of partial safety factors, as detailed in
Appendix H.
5. Material costs of $14.00 to $20.00 should be anticipated if the alternative procedure for determination of longterm design strength (Appendix H, section H.5) is used.

FHWA NHI-07-092
Geosynthetic Engineering

G-2

Appendix G
August 2008

Appendix H
GEOSYNTHETIC REINFORCEMENT
STRUCTURAL DESIGN PROPERTIES

Table of Contents
H.1BACKGROUND
H.2TENSILE STRENGTHS
H.2-1 Long-Term Tensile Strength
H.2-2 Allowable Strength
H.3REDUCTION FACTORS
H.3-1 Ultimate Strength
H.3-2 Creep
H.3-3 Installation Damage
H.3-4 Durability
H.3-5 Joints, Seams, and Connections
H.4IMPLEMENTATION
H.5ALTERNATIVE LONG-TERM STRENGTH DETERMINATION
H.6SOIL-REINFORCEMENT INTERACTION
H.6-1 Pullout Resistance
H.6-2 Direct Shear Resistance
H.7
REFERENCES
H.1

BACKGROUND

The tensile properties of geosynthetics are affected by environmental factors such as creep,
installation damage, aging, temperature, and confining stress. Therefore, the long-term
tensile strength, Tal, should be determined by thorough consideration of creep rupture,
allowable elongation, and all possible strength degradation mechanisms.
An inherent advantage of geosynthetics is their longevity in fairly aggressive soil conditions.
The anticipated half-life of some geosynthetics in normal soil environments is in excess of
1,000 years. However, as with steel reinforcements, strength characteristics must be adjusted
to account for potential degradation in the specific environmental conditions, even in
relatively neutral soils.

FHWA NHI-07-092
Geosynthetic Engineering

H-1

Appendix H
August 2008

Polymeric reinforcement, although not susceptible to corrosion, may degrade due to


physicochemical activity in the soil such as hydrolysis, oxidation, and environmental stress
cracking depending on polymer type. In addition, these materials are susceptible to
installation damage and the effects of (potential) high temperatures at the facing and
connection of retaining wall structures. Temperatures can be as high as 120F compared with
the normal range of in-ground temperature of 54F in cold and temperate climates to 85F in
arid desert climates.
Degradation most commonly occurs from mechanical damage, long-term time dependent
degradation caused by stress (creep), deterioration from exposure to ultraviolet light, and
chemical or biological interaction with the surrounding environment. Because of varying
polymer types, quality, additives and product geometry, each geosynthetic is different in its
resistance to aging and attack by different chemical and biological agents. Therefore, each
product must be investigated individually.
This appendix is broken into several categories, to aid the user. First, the partial reduction
factor methodology for determining long-term strength is presented. Next, application of
long-term strength to allowable (i.e., design) strength is reviewed for wall and slope
applications. Then, a discussion on use of the partial reduction factor method and evaluation
of supplier submittals by agencies is presented. Implementation obstacles are noted within
this discussion. An alternative procedure, less cumbersome to implement, for determination
of a long-term strength is then presented. This simplified procedure is for complementary
use with the more-detailed methodology, and is presented as a means to encourage use of
geosynthetics in reinforced soil structures.

H.2

TENSILE STRENGTHS

H.2-1 Long-Term Tensile Strength


Long-term tensile strength (Tal) of the geosynthetic shall be determined using a partial factor
approach (Bonaparte and Berg, 1987). Reduction factors are used to account for installation
damage, chemical and biological conditions and to control potential creep rupture and
deformation of the polymer. The total reduction factor is based upon the mathematical
product of these factors. The long-term tensile strength, Tal, thus can be obtained from:

Tal =

FHWA NHI-07-092
Geosynthetic Engineering

Tult
RF

H-2

[H-1]

Appendix H
August 2008

with RF equal to the product of all applicable reduction factors:

RF = RFCR RFID RFD

[H-2]

where:
Tal
= long-term geosynthetic tensile strength,(lb/ft {kN/m});
Tult
= ultimate geosynthetic tensile strength, based upon MARV, (lb/ft {kN/m});
RFCR = creep reduction factor, ratio of Tult to creep-limiting strength,
(dimensionless);
RFID = installation damage reduction factor, (dimensionless); and
RFD = durability reduction factor for chemical and biological degradation,
(dimensionless).
H.2-2 Allowable Strength
Additionally, the following factors should be considered. The long-term strength determined
by dividing the ultimate strength by RF does not include an overall factor of safety to account
for variation from design assumptions (e.g., heavier loads than assumed, construction
placement, fill consistency, etc.). Therefore, an allowable strength needs to be defined and
quantified. If seams or joints in the reinforcement are allowed, they likely will reduce the
strength of the material, and an applicable reduction factor should be included in computation
of an allowable strength. Thus, the allowable strength of a geosynthetic for RSS applications
can be defined as:

Ta =

Tal
FS RFJNT

[H-3]

where:
Ta
FS

RFJNT

= allowable geosynthetic tensile strength,(kN/m);


= overall factor of safety to account for uncertainties in the geometry of the
structure, fill properties, reinforcement properties, and externally applied
loads; and
= partial reduction factor for joints (seams and connections),
(dimensionless).

For MSE wall structures, a minimum factor of safety, FS, of 1.5 is recommended. Of course,
the FS value should be dependent upon the specifics of each project.
For RSS structures, the FS value will be dependent upon the analysis tools utilized by the
designer. For most design charts, a FS is incorporated into the soil shear strength and a FS =
FHWA NHI-07-092
Geosynthetic Engineering

H-3

Appendix H
August 2008

1 should be used in Eq. H-3 to determine the allowable geosynthetic strength. With
computerized analyses, the FS value is dependent upon how the specific program accounts
for the reinforcement tension is computing a stability factor of safety.
The FHWAs ReSSA program, and many others, assume the reinforcement force as
contributing to the resisting moment in the computation of a stability factor of safety; simply
stated:
FS STABILITY =

M R + TS R
MD

[H-4]

With this assumption, the allowable strength is equal to the long-term strength (if no
reduction is required for seams or joints). The factor of safety on the reinforcement is equal
to the stability factor of safety. Therefore, a value of FS = 1 is used in Eq. H-3.
Some computer programs use an assumption that the reinforcement force is a negative
driving component, thus the stability FS is computed as:
FS STABILITY =

MR
M D TS R

[H-5]

With this assumption, the stability factor of safety is not applied to TS. Therefore, the
allowable strength should be computed as the long-term strength divided by the safety factor
(e.g., target stability factor of safety). A FS value of 1.3 to 1.5 is typically used in Eq. H-3, in
this case.

H.3

REDUCTION FACTORS

H.3-1 Ultimate Strength


Ultimate strength values shall be based upon minimum average roll values (MARV), as
defined in ASTM D 4759. Ultimate strength for agency quality assurance (QA) purposes
may be determined according to ASTM D 4595 Test Method for Tensile Properties of
Geotextiles by the Wide-Width Strip Method or ASTM D 6637 Determining Tensile
Properties of Geogrids by the Single or Multi-Rib Tensile Method. The test procedure used
to determine ultimate strength, however, must be the same as that used to define RFCR.

FHWA NHI-07-092
Geosynthetic Engineering

H-4

Appendix H
August 2008

H.3-2 Creep
The creep reduction factor, RFCR, is the ratio of the ultimate strength, TULT, to the creep
limited strength obtained from laboratory creep tests for each product. Typical ranges of
reduction factors as a function of polymer type, are summarized in Table H-1.
Table H-1.
Typical Ranges Of Creep Reduction Factors
Polymer Type

Creep Reduction Factor

Polyester
Polypropylene
Polyethylene

2.5 to 2.0
5 to 4.0
5 to 2.5

The long-term tension-strain-time polymeric reinforcement behavior shall be determined


from results of controlled laboratory creep tests conducted for a minimum duration of 10,000
hours for a range of load levels on samples of the finished product. At present, creep tests
are conducted in-isolation rather than confined in soil, even though in-isolation creep tests
tend to over predict creep strains and under predict the true creep strength when used in a
structure. Creep testing should be conducted in accordance with ASTM D 5262, Test
Method for Evaluating the Unconfined Tension Creep Behavior of Geosynthetics. For
confined creep testing (NOTE: a standardized test procedure is not available), projectspecific or representative backfill material and typical placement and confinement conditions
should be in the testing.
The creep reduction factor is determined by comparing the long-term creep strength, Tl, to
the average ultimate tensile strength of the samples tested for creep. The samples tested for
ultimate strength should be taken from the same lot, and preferably the same roll, of material
which is used for the creep testing. For ultimate limit state design, the strength reduction
factor to prevent long-term creep rupture is determined as follows:

RFCR =

Tultlot
Tl

[H-6]

where,
Tultlot = the average lot specific ultimate tensile strength for the lot of material used for
creep testing; and
Tl
= the long-term creep rupture strength at the design life and design
temperature.

FHWA NHI-07-092
Geosynthetic Engineering

H-5

Appendix H
August 2008

Creep test data at a given temperature may be directly extrapolated over time up to one order
of magnitude, in accordance with standard polymeric practices. Considering that typical
design lives for permanent MSE wall and RSS structures are 75 years or more, extrapolation
of creep data is required. Therefore, accelerated testing (e.g., testing under elevated
temperatures or possibly other corroborating evidence) is required to extrapolate 10,000-hour
creep test data to a minimum 75-year design life. Accelerated testing is used to extrapolate
to a 75-year design life and to ensure that the rupture failure mechanism, does not change
(e.g., transition from ductile to brittle rupture). A step-by-step procedure for extrapolating
and interpolating stress rupture data in presented in Appendix B of FHWA NHI-00-043
(Elias et al., 2001).
Additional judgement should be exercised when selecting a creep limited value for
geosynthetics whose creep response is established with confined (in-soil) creep testing. Field
performance of structures designed on the basis of confined creep test data is limited.
However, it is well-established and generally recognized that unconfined creep test results
are consistently conservative.
The requirement for a 10,000-hour minimum creep test period for geogrids and geotextiles
may be waived for a new product if it can be demonstrated that it is sufficiently similar to a
proven 10,000 hour creep tested product of a similar nature. Product similarity must consider
base resin, resin additives, product manufacturing process, product geometry, and creep
response. Creep testing of all products, regardless of similarity, shall be conducted for a least
1,000 to 2,000 hours. See FHWA NHI-00-043 (Elias et al., 2001) for details on use of creep
data from similar products.
H.3-3 Installation Damage
The installation damage factor reduces the long-term strength to account for the effect of
installation damage on geosynthetic reinforcement. RFID can range from 1.05 to 3.0, and is a
function of the backfill gradation, compaction techniques, product structure, and product
mass per unit area. A minimum RFID value of 1.10 is recommended. Installation damage
factors shall be determined from the results of full-scale construction damage tests. Damaged
samples should be obtained following ASTM D 5818, Practice for Obtaining Samples of
Geosynthetics from a Test Section for Assessment of Installation Damage. This practice
requires that the geosynthetic material is subjected to a backfilling and compaction cycle,
consistent with field practice. The ratio of the initial reinforcement strength to the strength of
the retrieved samples defines the RFID.

FHWA NHI-07-092
Geosynthetic Engineering

H-6

Appendix H
August 2008

For reinforcement applications a minimum weight of 8 oz/yd2 for geotextiles is


recommended to minimize installation damage. This roughly corresponds to a Class 1
geotextile as specified in AASHTO M288.
See Table H-2 for typical ranges of installation damage factors, by geosynthetic type.
Table H-2. Installation Damage Reduction Factors
Reduction Factor, RFID
No.

Geosynthetic

Type 1 Backfill Max.


Size 4 in. (102 mm)
D50 about 1- in.(30 mm)

Type 2 Backfill Max.


Size in. (20 mm)
D50 about #25 (0.7 mm)

HDPE uniaxial geogrid

1.20 - 1.45

1.10 - 1.20

PP biaxial geogrid

1.20 - 1.45

1.10 - 1.20

PVC coated PET geogrid

1.30 - 1.85

1.10 - 1.30

Acrylic coated PET geogrid

1.30 - 2.05

1.20 - 1.40

Woven geotextiles (PP & PET)

1.40 - 2.20

1.10 - 1.40

Non woven geotextiles (PP & PET)

1.40 - 2.50

1.10 - 1.40

Slit film woven PP geotextiles

1.60 - 3.00

1.10 - 2.00

H.3-4 Durability
The durability reduction factor, RFD, is dependent on the susceptibility of the geosynthetic to
attack by microorganisms, chemicals, thermal oxidation, hydrolysis and stress cracking, and
can vary typically from 1.1 to 2.0. Because of varying polymer types, quality, additives and
product geometry, each geosynthetic is different in its resistance to aging and attack by
different chemical and biological agents. Therefore, each product must be investigated
individually.
Typically, polyester products (PET) are susceptible to aging strength reductions due to
hydrolysis (water availability) and high temperatures. Hydrolysis and fiber dissolution are
accelerated in alkaline regimes, below or near piezometric water levels or in areas of
substantial rainfall where surface water percolation or capillary action ensures water
availability over most of the year.

FHWA NHI-07-092
Geosynthetic Engineering

H-7

Appendix H
August 2008

Polyolefin products (PP and HDPE) are susceptible to aging strength losses due to oxidation
(contact with oxygen) and or high temperatures. The level of oxygen in reinforced fills is a
function of soil porosity, ground water location and other factors not yet fully understood.
However, it is considerably less than oxygen levels in the atmosphere (21 percent).
Therefore, oxidation of geosynthetics in the ground should proceed at a slower rate than
those used above ground. Oxidation is accelerated by the presence of transition metals (Fe,
Cu, Mn, Co, Cr) in the backfill as found in acid sulphate soils, slag fills, other industrial
wastes or mine tailings containing transition metals. It should be noted that the resistance of
polyolefin geosynthetics to oxidation is primarily a function of the proprietary antioxidant
package added to the base resin, which differs for each product brand, even when formulated
with the same base resin.
The artificial relative resistance of polymers to these identified regimes is shown in Table H4. Polymeric reinforcements may, therefore, be chosen consistent with the preliminary data
shown in Table H-4. Limits, based upon current research, on the pH limits of the reinforced
fill, by polymer type, are listed in Table H-5. Specific aging reduction factor values for PET
and for polyolefin geosynthetics are discussed below.
Based on the developed data, PET geosynthetics are recommended for use in environments
characterized by 3 < pH < 9, only. The following (Table H-3) reduction factors for PET
aging (RFD) are presently indicated for a 100-year design life in the absence of product
specific testing.
Table H-3. Aging Reduction Factors, PET
Reduction Factor, RFID
No.

Geosynthetic
5 < pH < 8

3 < pH < 5
8 < pH < 9

Geotextiles
Mn < 20,000, 40 < CEG < 50

1.6

2.0

Coated geogrids
Mn > 25,000, CEG < 30

1.15

1.3

Use of materials outside the indicated pH or molecular property range requires specific product testing.

A rough measure of antioxidant effectiveness for PP products formulated without significant


carbon black is resistance to UV degradation measured in accordance with ASTM D 4355.
A retained strength of 90% at 500 hours or more generally indicates an effective antioxidant
blend and potentially a reduction factor as low as 1.1 at 20EC and 100 years. For HDPE
FHWA NHI-07-092
Geosynthetic Engineering

H-8

Appendix H
August 2008

geogrids presently available (Tensar UX series), current research data indicates a Reduction
Factor of 1.1 for use at 20EC and 100 years.
Most geosynthetic reinforcement is buried, and therefore ultraviolet (UV) stability is only of
concern during construction and when the geosynthetic is used to wrap the wall or slope face.
If used in exposed locations, the geosynthetic should be protected with coatings or facing
units to prevent deterioration. Vegetative covers can also be considered in the case of open
weave geotextiles or geogrids. Thick geosynthetics with ultraviolet stabilizers can be left
exposed for several years or more without protection; however, long-term maintenance
should be anticipated because of both UV deterioration and possible vandalism.
Table H-4. Anticipated Resistance of Polymers to Specific Environments
Polymer

Soil Environment
Acid Sulphate Soils
Organic Soils
Saline Soils pH < 9
Calcareous Soils
Modified Soils/Lime, Cement
Sodic Soils, pH > 9
Soils with Transition Metals
NE
?

=
=

PET

PE

PP

NE
NE
NE
?
?
?
NE

?
NE
NE
NE
NE
NE
?

?
NE
NE
NE
NE
NE
?

No Effect
Questionable Use, Exposure Tests Required

Table H-5. Recommended pH Limits for Reinforced Fill Soils


Base Polymer

Test Method

Limits

Polyester (PET)
Polyolefin (PP & HDPE)

AASHTO T289
AASHTO T289

>3 and < 9


>3

The protocols for testing to obtain this reduction factor are under development. In general, it
consists for oven aging polyolefins (PP and HDPE) samples to accelerate oxidation and
measure their strength reduction, as a function of time, temperature and oxygen
concentration. This high temperature data must then be extrapolated to a temperature
consistent with field conditions. For polyesters (PET) the aging is conducted in an aqueous
FHWA NHI-07-092
Geosynthetic Engineering

H-9

Appendix H
August 2008

media at varying pHs and relatively high temperature to accelerate hydrolysis, with data
extrapolated to a temperature consistent with field conditions.
H.3-5 Joints, Seams, and Connections
The effect of the joint strength must be factored into design strength when separate lengths of
geosynthetics are connected together or overlapped in the direction of primary force
development. The value of FSJNT should be taken as the ratio of the unjointed specimen
strength to the joined specimen strength. Testing should be conducted in accordance with
ASTM D 4595 for mechanically connected joints and GRI:GG5 or GRI:GT6 for overlap
joints. Sustained tension tests of 1,000-hour minimum duration should also be conducted on
mechanically connected joints, according to GRI:GG4 and GRI:GT7. A load level equal to
the allowable strength, Ta, is suggested for long-term testing. Limits on number and location
of joints and seams in a slope or wall structure should be addressed in the project
specifications.
As discussed in Chapter 9, MSE Retaining Walls and Abutments, the connection
geosynthetic strength to the wall facing element may limit strength and, therefore, control the
allowable design tensile strength. Connection strength must be addressed in wall designs.
Connection strength requirements, testing, and test data interpretation are addressed in
FHWA NHI-00-043 (Elias et al., 2001).

H.4

IMPLEMENTATION

The determination of the reduction factors for creep (including extrapolation) and
chemical/biological durability require extensive testing and is product-specific. Testing
standards for determination of these partial factors are not fully developed; thus, test
procedures and/or result interpretation can vary. Therefore, evaluation of supplier submittals
is not an easy process for many agencies.
Detailed lists of items to be supplied by potential geosynthetic reinforcement material
suppliers and system suppliers, are presented in the Guidelines for Design, Specification, &
Contracting of Geosynthetic Mechanically Stabilized Earth Slopes on Firm Foundations,
U.S. Department of Transportation, Federal Highway Administration, FHWA-SA-93-025,
(Berg, 1993). Guidelines for assigning evaluation responsibilities to the various agency
organizations are also provided in these guidelines. However, these guidelines are focused
upon scenarios where agencies evaluate materials and use an approved products or vendor
list system.

FHWA NHI-07-092
Geosynthetic Engineering

H - 10

Appendix H
August 2008

In 1994, the Civil Engineering Research Foundation (CERF) established the Highway
Innovative Technology Evaluation Center (HITEC) in cooperation with the Federal Highway
Administration (FHWA), the American Association of State Highway and Transportation
Officials (AASTO), and the Transportation Research Board (TRB).
HITECs purpose is to accelerate the introduction of technological advances in products,
systems, services, materials, and equipment to the highway and bridge markets. The
evaluation of new and more cost-effective retaining wall systems is performed through
HITECs nationally-focused, earth retaining system (ERS) group evaluation program. This
program collaborates directly with AASHTOs T-15 Technical Committee on Retaining
Walls. The benefits of HITEC for suppliers and users/owners include the following:
For ERS Suppliers
timely and less costly verification of recommended design and construction
procedures and the overall performance characteristics of a wall system;
provides independent, high quality evaluations;
fosters greater private-public sector collaboration; and
establishes a consistent level of quality within the industry.
For ERS Users
minimizes the time and effort required of individual agencies to review and approve
wall systems;
provides a benchmark of current evaluation/review criteria to help ensure selection of
safe, effective, and economical wall systems;
expands the retaining wall knowledge base among the user community; and
assists the development of national codes and standards of practice.
A HITEC evaluation is performed for wall systems that incorporate new materials or use old
ones in new ways, rely on new design or construction methods, or perform differently than
conventional structures. The requirements and full scope of the ERS evaluation program are
described in Guidelines for Evaluating Earth Retaining Systems (HITEC, 1998) (which is
available from http://www.cerf.org/hitec/news/reports.htm).
A copy of the HITEC
application form can be found in HITEC (1998). Although a HITEC evaluation is not free, it
is believed that this evaluation is far less expensive than agency-by-agency review.
HITEC evaluation is organized and reviewed with regard to the following categories:
Materials;
Design;
FHWA NHI-07-092
Geosynthetic Engineering

H - 11

Appendix H
August 2008

Construction;
Performance; and
Quality Assurance.

Since established, some of the walls evaluated by HITEC include: (1) MSE wall with
geosynthetic grids, steel grids, and steel grids with anchor plates; (2) precast wingwall with
deadman anchors; (3) anchored wall with concrete wales; and (4) soil mixed gravity wall.
As of 2004, 15 states in the U.S. require and 12 states recommend HITEC evaluation for
innovative wall systems. States that require HITEC evaluations typically use HITEC reports
as their primary basis to list a wall system in their specifications. Based on the evaluation
results, these DOTs determine whether or not they would consider that wall type for their
projects. It should be kept in mind that HITEC reports do not approve, recommend, or
endorse the technology, but rather they present comprehensive findings and a summary of the
performance data that state and local agencies can use to make informed decisions about
product selections.
Implementation of geosynthetic RSS and MSE wall technologies has been significantly
hampered by the review required to assess geosynthetic long-term allowable strength. When
an approved products or vendor list is not used, many agencies find the procedure too
laborious and time-consuming for post-bid evaluation of materials. Other agencies are not
able to implement a complete review process, amongst their various organizational units, for
a variety of reasons, or do not feel that they have the in-house expertise to complete an
evaluation.
Impeded implementation of well-documented, cost-effective geosynthetic reinforced slope
and wall technologies is costly. Funds saved by implementing these technologies can be
well-spent on other transportation projects.
It is recognized that some agencies need an easier and/or quicker procedure for implementing
long-term allowable strength quantification and geosynthetic product acceptance. Therefore,
an alternative procedure for determining long-term strength is presented in Section H.5. This
alternative procedure is to be used in conjunction with, and to compliment, the detailed
procedure presented in Section H.2.

FHWA NHI-07-092
Geosynthetic Engineering

H - 12

Appendix H
August 2008

H.5

ALTERNATIVE LONG-TERM STRENGTH DETERMINATION

An alternative for computing a long-term allowable strength (Section H.2) is presented in this
section. This alternative procedure is for complementary use with the more detailed
procedure.
The goal of providing an alternative method is to foster widespread use of geosynthetic
reinforced slopes and walls in transportation facilities. Specific objectives include:
providing an easy-to-use method for determining design strengths;
providing agencies with a method to generically specify geosynthetic reinforcement with
a defined default allowable strength (through a defined default long-term strength and a
design safety factor), that suppliers will be required to use unless a detailed evaluation of
long-term strength on their specific products has been completed by the agency;
providing long-term strengths that are conservative and economical;
providing long-term strengths that are sufficiently punitive that thorough testing and
evaluation of geosynthetic materials by manufacturers and suppliers is still promoted;
providing an interim method until a sufficient number of materials and/or systems are
approved by an agency and an approved list specifications is developed;
providing a method for use with conservative soil environment parameters; and
maintaining the current state-of-practice.
These objectives and goals can be achieved using a single default reduction factor to account
for creep, installation damage, durability, and connections. A reasonable default reduction
factor, RF, which meets the stated goal and objectives, is presented below.
For preliminary design of permanent structures and for applications defined by the user as
not having severe consequences should poor performance or failure occur, the long-term
tensile strength, Tal, may be evaluated without product specific data, as:

Tal =

Tult
7

[H-7]

with the default RF equal to 7. This reduction factor RF=7, should be limited to projects
where the project environment meets the following requirements:
granular soils (sands, gravels) used in the reinforced fill;
4.5 < pH < 9;
site temperature < 85F;
biologically inactive environments; and
FHWA NHI-07-092
Geosynthetic Engineering

H - 13

Appendix H
August 2008

maximum backfill particle size of -inch.


Other qualifiers on application of this default reduction factor is limiting use to projects
where:
maximum retaining walls height is 33 ft (10 m);
face element (for walls) shall be a nonaggressive environment for the geosynthetic;
maximum reinforced slope height is 50 ft (15 m);
geotextile reinforcement meets AASHTO M 288 specification strength requirements
for High Survivability Level; and
the manufacturer certifies that the supplied geosynthetic is intended for and fit to use
as long-term soil reinforcement.
Site temperature is defined as the temperature which is halfway between the average yearly
air temperature and normal daily air temperature for the hottest month at the site.
The total reduction factor of 7 has been established by multiplying lower bound partial
reduction factors obtained from currently available test data, for products which meet the
minimum requirements in Table H-6. See Berg et al. (1998) for a discussion on the range of
total reduction values used over the past twenty-five years.
It should be noted that the total Reduction Factor may be reduced significantly with
appropriate test data. It is not uncommon for products with creep, installation damage
and aging data, to develop total reduction factors in the range of 4 to 6.

Use of this alternative allowable strength procedure for structurally-faced, MSE retaining
walls does not eliminate the requirement of connection strength testing. Testing shall be
conducted to define the ultimate, short-term connection strengths, Tultc (see Chapter 9).
For temporary applications not having severe consequences should poor performance or
failure occur, a default value for RF of not less than 3 could be used. For more critical
projects, a higher default reduction factor could be used.
Again, use of a default reduction factor, RF, is for complementary use with the moredetailed procedure. Blanket, long-term use of a default reduction factor will penalize
many current suppliers and limit economic benefit of geosynthetic reinforced
structures. Those manufacturers and suppliers that have or will conduct the extensive
testing to document partial safety factors should be allowed to use those factors, after
agency review and approval. Exclusive use of a blanket default value will also severely
impede further evolution of this technology.

FHWA NHI-07-092
Geosynthetic Engineering

H - 14

Appendix H
August 2008

Table H-6. Minimum Requirements for Use of Default Reduction Factor


Geosynthetic
Type

Property

Test Method

Criteria to Allow Use of


Default RF

Polypropylene

UV Oxidation
Resistance

ASTM D 4355

Min. 70% strength


retained after 500 hours in
weatherometer.

Polyethylene

UV Oxidation
Resistance

ASTM D 4355

Min. 70% strength


retained after 500 hours in
weatherometer.

Polyester

Hydrolysis
Resistance

Intrinsic Viscosity Method (ASTM


D 4603) with Correlation or
Determine Directly Using Gel
Permeation Chromatography

Min. Number (Mn)


Molecular Weight of
25,000

Polyester

Hydrolysis
Resistance

ASTM D 2455

Max. Carboxyl End Group


Number of 30

All Polymers

Survivability
(mass per unit
area)

All Polymers

% Postconsumer
recycled
material by
weight

H.6

ASTM D 5261

Min. 270 g/m2


(8 oz/yd2)

Certification of material used

Maximum 0%

SOIL-REINFORCEMENT INTERACTION

Two types of soil-reinforcement interaction coefficients or interface shear strengths must be


determined for design: pullout coefficient, and direct shear coefficient. Pullout coefficients
are used in stability analyses to compute mobilized tensile force at the front and tail of each
reinforcement layer in slopes and at the tail in walls. Direct shear coefficients are used to
check factors of safety against outward sliding of the entire reinforced mass.
H.6-1 Pullout Resistance
Design of reinforced slopes requires evaluation of the long-term pullout performance with
respect to three basic criteria:
(i) pullout capacity; i.e., the pullout resistance of each reinforcement should be adequate
to resist the design working tensile force in the reinforcement with a specified
factor of safety (FSPO), where FSPO is typically set at a minimum of 1.5; and
(ii) allowable displacement; i.e., the relative soil-to-reinforcement displacement required
FHWA NHI-07-092
Geosynthetic Engineering

H - 15

Appendix H
August 2008

to mobilize the design tensile force should be smaller than the allowable
displacement; and
(iii)long-term displacement; i.e., the pullout load should be smaller than the critical creep
load.(Christopher et al., 1989)
The pullout resistance of the reinforcement is mobilized through one or a combination of two
basic soil-reinforcement interaction mechanisms. The two mechanisms by which load may
be transferred between soil and geosynthetic are: i) interface friction; and ii) passive soil
resistance. Geotextile pullout resistance is developed with an interface friction mechanism.
Geogrid pullout resistance may be developed by both interface friction and passive soil
resistance against transverse elements. The load transfer mechanisms mobilized by a specific
geogrid depends primarily upon its structural geometry (i.e., composite reinforcement versus
linear or planar elements, thickness of in-plane or out-of-plane transverse elements, and
aperture dimension to grain-size ratio). The soil-to-reinforcement relative movement
required to mobilize the design tensile force depends mainly upon the load transfer
mechanism, the extensibility of the reinforcement material, and the soil type. (Christopher et
al., 1989)
The long-term pullout performance (i.e., displacement under constant design load) is
predominantly controlled by the soil's creep characteristics and the reinforcement material.
Soil reinforcement systems will generally not be used with cohesive soils susceptible to
creep. Therefore, creep is primarily an issue of the reinforcement type. The basic aspects of
pullout performance in terms of the major load transfer mechanism, relative soil-toreinforcement displacement required to fully mobilize the pullout resistance, and creep
potential of the reinforcement in granular (and low-cohesive) soils for generic extensible
reinforcement types are presented in Table H-7. (Christopher et al., 1989)
Pullout resistance of geosynthetic reinforcement is defined by the lower value of:
i) the ultimate tensile load required to generate outward sliding of the reinforcement
through the soil mass; or
ii) the tensile load which produces a 0.6-inch (15 mm) displacement as measured at the
end of the embedded sample.

FHWA NHI-07-092
Geosynthetic Engineering

H - 16

Appendix H
August 2008

Table H-7. Basic Aspects of Reinforcement Pullout Performance in


Granular and Low Cohesive Soils
(after Christopher et al., 1989)

Generic
Reinforcement
Type

geogrids

geotextiles

Major Load
Transfer
Mechanism

Displacement to
Pullout

Long-Term
Performance

frictional
+ passive H.D.

dependent on
reinforcement
extensibility
(1 to 2 in.)
{25 to 50 mm}

dependent on
reinforcement
structure and
polymer creep

frictional
(interlocking) L.D.

dependent on
reinforcement
extensibility
(1 to 4 in.)
{25 to 100 mm}

dependent on
reinforcement
structure and
polymer creep

NOTE: L.D. - low-dilatency effect; H.D. - high-dilatency effect

Several approaches and design equations have been developed and are currently being used
to estimate pullout resistance by considering frictional resistance, passive resistance, or a
combination of both. The design equations use different interaction parameters, therefore it
is, difficult to compare the pullout performance of different reinforcements for a specific
application (Christopher et al., 1989).
A normalized definition is recommended as presented in FHWA Reinforced Soil Structures,
Volume I - Design and Construction Guidelines (Christopher et al., 1989). The ultimate
pullout resistance, Pr, of the reinforcement per unit width of reinforcement is given by:
Pr = F* C " C Fv C Le C C
where:
Le C C = the total surface area per unit width of the reinforcement in the
resistance zone behind the failure surface, (ft2 {m2});
Le
= the embedment or adherence length in the resisting zone behind the
failure surface, (ft {m});
C
= the reinforcement effective unit perimeter; e.g., C=2 for geogrids
and
FHWA NHI-07-092
Geosynthetic Engineering

H - 17

Appendix H
August 2008

F*
"
Fv

geotextiles, (dimensionless);
= the pullout resistance (or friction-bearing-interaction) factor,
(dimensionless);
= a scale effect correction factor, (dimensionless); and
= the effective vertical stress at the soil-reinforcement interfaces,
(lb/ft2 {kN/m2}).

The pullout resistance factor, F*, can be most accurately obtained from pullout tests
performed on the specific, or representative, backfill to be used on the project. Refer to
ASTM D 6706 Measuring Geosynthetic Pullout Resistance in Soil for pullout test
procedures. Note that this test method produces pullout interaction coefficients that are
classified as either short-term or long-term. Design of MSE walls and slopes for
permanent applications requires use of long-term interaction coefficients.
For standard backfill materials, with the exception of uniform sands (i.e., coefficient of
uniformity, Cu < 4), it is acceptable to use conservative default values for F* and as shown in
Table H-8.
Table H-8. Default Values for F* and Pullout Factors

Reinforcement Type

Default F*

Default

Geogrid

0.67 tan

0.8

Geotextile

0.67 tan

0.6

Passive resistance is applicable only to geogrids, and not to geotextiles. The passive
resistance portion of the above equation assumes that long-term passive resistance will occur
across transverse geogrid ribs. This requires sufficient long-term junction strength between
the transverse and longitudinal ribs to assure stress transfer. Long-term stress transfer is
assured if the geogrid is creep-tested with the through-the-junction method per GRI:GG3a.
Long-term pullout interaction coefficients should be quantified for geogrids with either:
i)
quick, effective stress pullout tests and through-the-junction creep-testing of the
geogrid per GRI:GG3a test method;
ii)
quick, effective stress pullout tests of the geogrid with severed transverse ribs;
iii)
quick, effective stress pullout tests of the entire geogrid structure if summation of
shear strengths of the joints occurring in a 300 mm length of grid sample is equal
to or greater than the ultimate strength of the grid element to which they are
attached; or
FHWA NHI-07-092
Geosynthetic Engineering

H - 18

Appendix H
August 2008

iv)

v)

long-term effective stress pullout tests of the entire geogrid structure.


Long-term pullout interaction coefficients should be quantified for geotextiles
with:
quick, effective stress pullout tests. (Berg, 1993)

Test Method a) Controlled Strain Rate Method For Short-Term Testing per GRI:GG5 and
GRI:GT6 is recommended for testing under conditions i), ii), iii), and v) above. Test method
d) Constant Stress (Creep) Method For Long-Term Testing per GRI:GG5 is recommended
for condition iv) above. Joint shear strength shall be measured in accordance with GRI:GG2
and ultimate strength shall be measured with either GRI:GG1 or ASTM D 4595 for condition
iii) above.
Long-term testing may also be required if cohesive soils are utilized, to define long-term
effective stress (drained) pullout resistance. Procedures and results for long-term testing in
cohesive soils have been presented by Christopher and Berg (1990). Their method is a
combination of GRI:GG5 and GRI:GT6 methods c) Incremental Stress Method for ShortTerm Testing, and d) Constant Stress (Creep) Method For Long-Term Testing.
H.6-2 Direct Shear Resistance
Soil-geosynthetic direct shear resistance should be determined in accordance with ASTM D
5321, Test Method for Determining the Coefficient of Soil and Geosynthetic or Geosynthetic
and Geosynthetic Friction by the Direct Shear Method.

H.7

REFERENCES

Berg, R.R. Allen, T.M. and Bell, J.R. (1998). Design Procedures for Geosynthetic
Reinforced Soil Walls A Historical Perspective, Proceedings of Sixth International
Conference on Geotextiles, Geomembranes and Related Products, Atlanta, March, pp. 491496.
Berg, R.R. (1993). Guidelines for Design, Specification, & Contracting of Geosynthetic
Mechanically Stabilized Earth Slopes on Firm Foundations, Report No. FHWA-SA-93-025,
Federal Highway Administration, Washington, D.C., 87 p.
Christopher, B.R. and Berg, R.R. (1990). Pullout Evaluation of Geosynthetics in Cohesive
Soils, Proceedings of the 4th International Conference on Geotextiles, Geomembranes, and
Related Products, The Hague, Netherlands, pp. 731-736.
Bonaparte, R. and Berg, R.R. (1987). Long-Term Allowable Tension for Geosynthetic
Reinforcement, Proceedings of Geosynthetics '87 Conference, Volume 1, New Orleans, LA,
pp. 181-192.
FHWA NHI-07-092
Geosynthetic Engineering

H - 19

Appendix H
August 2008

Christopher, B.R., Gill, S.A., Giroud, J.P., Juran, I. Scholsser, F., Mitchell, J.K. and
Dunnicliff, J. (1989). Reinforced Soil Structures, Volume I. Design and Construction
Guidelines, Federal Highway Administration, Washington, D.C., Report No. FHWA-RD-89-043, 287 p.
Elias, V., Christopher, B.R. and Berg, R.R. (2001). Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes, Design & Construction Guidelines, U.S. Department of
Transportation, Federal Highway Administration, Washington, D.C., FHWA-NHI-00-043,
418 p.
Elias, V. (1997). Corrosion/Degradation of Soil Reinforcements for Mechanically
Stabilized Earth Walls and Reinforced Soil Slope, FHWA-SA-96-072, Federal Highway
Administration, U.S. Department of Transportation, 105 p.
Elias, V. and Christopher, B.R. (1997). Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes, Design & Construction Guidelines, FHWA-SA-96-071, U.S.
Department of Transportation, Federal Highway Administration, Washington, D.C.,
371 p.

FHWA NHI-07-092
Geosynthetic Engineering

H - 20

Appendix H
August 2008

You might also like