You are on page 1of 446

TRAINING COURSE IN GEOTECHNICAL 1111111111111111111111111111111

PB99-153629
AND FOUNDATION ENGINEERING

NHI COURSE NO. 13236 - MODULE 6 PUBLICATION No. FHWA NHI-99-025


APRIL 1999

EARTH RETAINING STRUCTURES

REFERENCE MANUAL

u.s. Department
of Transportation
Federal Highway
Administration

• ~ I:II National Highway Institute

REPRODUCED BY: NlJS.


u.s. Department of Co~merce .
National Technical Information SeNlce
Springfield, Virginia 22161
T echnical Report DocumentatIon
· P'age
1. Report No. 2. Government Accession No. 3. Recipient's Catalog No.
FHWA-NHI-99-025
4. Title and Subtitle 5. Report Date
EARTH RETAINING STRUCTURES April 1999
REFERENCE MANUAL 6. Performing Organization Code

7. Author(s) 8. Performing Organization Report No.

Principal Investigator - George Munfakh


Authors - George A. Munfakh, Naresh C. Samtani, 1111111111111111111111111111111
Raymond J. Castelli, Jaw-Nan (Joe) Wang PB99-153629

9. Performing Organization Name and Address 10. Work Unit No. (TRAIS)
Parsons Brinckerhoff Quade & Douglas, Inc.
One Penn Plaza 11. Contract or Grant No.
New York, NY 10119
DTFH 61-94-C-00104
12. Sponsoring Agency Name and Address 13. Type of Report and Period Covered

National Highway Institute


Federal Highway Administration 14. Sponsoring Agency Code
U. S. Department of Transportation
WashinQ"ton D.C.
15. Supplementary Notes

FHWA Technical Consultants - J.A. DiMaggio, A. Munoz, and P.A. Osborn


FHWA Contracting Officer - J. Mowery III; COTR - L. Jones, National Highway Institute
16. Abstract

The selection, design, construction and performance of earth retaining structures used for support of fills
or excavations are discussed in detail. The walls are grouped according to a classification system that
considers both the nature of the structure and its support mechanism. The factors that affect wall
selection are discussed and a selection process is proposed. Different contracting approaches are
discussed and a particular emphasis is placed on the required bidding documents for each approach.
Example problems are provided for each wall group. Case histories are presented documenting the
selection, design and performance of a variety of earth retaining structures.

17. Key Words 18. Distribution Statement


Retaining walls, earth, groundwater, anchors,
stabilized, gravity, analysis, design,
No restrictions.
construction, lateral pressures, in situ,
contracting, bidding, transportation, highways,
bridQes.
19. Security Classif. (of this report) 20. Security Classif. (of this page) 21. No. of Pages 22. Price

UNCLASSIFIED UNCLASSIFIED 444


Form DOT F 1700.7(8-72) Reproduction of completed page authorized
PREFACE
This module is the sixth in a series of twelve modules that constitute a comprehensive training course in
geotechnical and foundation engineering. Sponsored by the National Highway Institute (NHI) of the
Federal Highway Administration (FHWA), the training course is given at different locations in the U.S.
The intended audience is civil engineers and engineering geologists involved with the design and
construction of transportation facilities.

This module is intended to be a stand-alone document and is geared towards providing the practicing
engineer with a thorough understanding of the various types of retaining walls used for highway
applications. Accordingly, the manual first presents a classification system, discusses earth pressure
theories and the determination of basic soil parameters, and then proceeds to discuss various earth retaining
systems. Discussed also are wall selection, contracting issues, bidding documents, example problems, and
case histories involving selection, design, construction and performance of earth retaining structures. The
organization of the manual is presented below.

Chapter 1 presents a brief historical review of the earth retaining structures and discusses a proposed
classification system.

Chapter 2 describes methods used to determine design earth pressures for different conditions. Discussed
also are other issues, such as surcharge loads and groundwater, that influence the loading diagrams used
in the design.

Chapter 3 presents a discussion of the soil properties that influence lateral pressures. This chapter also
discusses material properties and provides guidance on the choice of suitable infill and backfill materials.

Chapters 4 through 9 discuss the various types of retaining walls identified in Chapter 1. For each type
of wall, the chapters are organized in the following fashion:
available wall types and their geometries
analytical and design procedures
construction techniques
durability and maintenance requirements
example problems

Chapter 10 discusses the factors affecting the selection of an earth retaining system from the design
engineer's viewpoint. A selection process is discussed with a selection matrix proposed to provide a
qualitative evaluation and rating of the most preferred alternatives. General recommendations for selecting
earth retaining structures are then presented. Case studies are provided to demonstrate the selection
process. Chapter 11 provides technical guidance on issues relating to wall system contracting methods
and specifications. Typical contracting approaches are presented. Chapter 12 presents a comprehensive
list of references.

It should be noted that a book on earth retaining structures by G. Munfakh, currently under preparation,
formed the basis of this reference manual. Several chapters of this manual may contain text and graphics
that are adapted from the book.

Finally, this manual is developed to be used as a desktop reference document. After attending the
training session, it is intended that the participant will use it as a manual of practice in everyday
work. Throughout the manual, attention is given to ensure the compatibility of its content with those of
the reference manuals prepared for the other training modules. Special efforts are made to ensure that
the included material is practical in nature and represents the latest developments in the field.
ACKNOWLEDGMENTS

The authors would like to acknowledge the assistance of Jeremy Hung, and a number of professionals from
Parsons Brinckerhoff Quade and Douglas, Inc., in the preparation of this manual. Many thanks to Anu
Yarde for assistance in development of example problems, Wendy Castro and Tony Silva for preparation
of graphics, Ke Fan for producing visual aids, Frank Pepe and David Campo for reviewing Chapter 9 and
Alia Abbas, Wendy Castro, Anthony Rapillo, and Sachin Karnik for overall word processing and
compiling.

NOTICE

The information in this document has been funded wholly or in part by the U.S. Department of
Transportation, Federal Highway Administration (FHWA), under Contract No. DTFH 61-94-C-OOI04 to
Parsons Brinckerhoff Quade & Douglas, Inc. The document has been subjected to peer and administrative
review by FHWA, and it has been approved for publication as a FHWA document.

In this document, certain products have been identified by trade name. Also, photographs of these
products have been included in the document for illustration purposes. Other products which are not
identified in this document may be equally viable to those identified. The mention of any trade name or
photograph of a particular product does not constitute endorsement or recommendation for use by either
the authors or the FHWA.
CONVERSION FACTORS

When When
... --, L
(a) h
inch 25.4 millimeter millimeter 0.039 inch
foot 0.305 meter meter 3.28 foot
yard 0.914 meter meter 1.09 yard
mile 1.61 kilometer kilometer 0.621 mile
(b)
, . Area

square inches 645.2 square millimeters square millimeters 0.0016 square inches
square feet 0.093 square meters square meters 10.764 square feet
acres 0.405 hectares hectares 2.47 acres
square miles 2.59 square kilometers square kilometers 0.386 square miles
(c)
, , Volume
fluid ounces 29.57 milliliters milliliters 0.034 fluid ounces
gallons 3.785 liters liters 0.264 gallons
cubic feet 0.028 cubic meters cubic meters 35.32 cubic feet
cubic yards 0.765 cubic meters cubic meters 1.308 cubic yards
, . Mass
(d)
ounces 28.35 grams grams 0.035 ounces
pounds 0.454 kilograms kilograms 2.205 pounds
short tons (2000 lb) 0.907 me~a~rams (tonne) megagrams (tonne) 1.102 short tons (2000 lb)
(e) Force
pound I 4.448 I Newton I Newton I 0.2248 I pound
(t) Pressure, Stress, Modulus of Elasticit
47.88
6.895
Pascals
kiloPascals
I Pascals
kiloPascals
0.021
0.145

16.019 I kil02rams oer cubic meter Ikil02rams oer cubic meter I 0.0624
(h) Temoerature
Fahrenheit temperatureeF) I 5/geF- 32) I Celsius temperatureeC) I
Celsius temperatureeC) I 915eC) + 32 IFahrenheit temperatureeF)
Notes: 1) The primary metric (SI) units used in civil engineering are meter (m), kilogram (kg), second(s), newton (N) and pascal (Pa = N/m2).
2) In a "soft" conversion, an English measurement is mathematically converted to its exact metric equivalent.
3) In a "hard" conversion, a new rounded metric number is created that is convenient to work with and remember.
MODULE 6
EARTH RETAINING STRUCTURES

TABLE OF CONTENTS

Page

LIST OF FIGURES ix

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. xv

LIST OF PHOTOS xvi

LIST OF NOTATIONS xviii

1.0 Introduction 1-1


1.1 Purpose and Scope of Module . . . . . . . . . . . . . . ..................... 1-1
1.2 History of Earth Retaining Structures 1-2
1.3 Classification of Earth Retaining Structures . . . . . ..................... 1-4
1.3.1 Classification by Load Support Mechanism ..................... 1-4
1.3.2 Classification by Construction Concept 1-4
1.3.3 Classification by System Rigidity 1-6
1.4 Organization of the Manual 1-6
1.5 Primary References . . . . . . . . . . . . . . . . . . . . . ..................... 1-7

2.0 Lateral Earth Pressures 2-1


2. 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-1
2.2 General Considerations 2-1
2.3 At-Rest Earth Pressure 2-3
2.4 Active and Passive Pressures 2-3
2.4.1 Effect of Wall Friction 2-5
2.4.2 Earth Pressures in Stratified Soils 2-11
2.4.3 Empirical Earth Pressures 2-12
2.5 Effect of Groundwater . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-12
2.5.1 Example Problem 2-1 2-13
2.6 Effect of Surface Surcharge Loads 2-14
2.6.1 Example Problem 2-2 2-17
2.7 Earth Pressures Due to Compaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-19
2.7.1 Example Problem 2-3 2-20
2.7.2 Other Compaction Equipment 2-22
2.8 Silo Pressure 2-22
2.9 Seismic Lateral Earth Pressure 2-22
2.10 Other Lateral Pressures 2-22
2.10.1 Frost (Ice) Action 2-22
2.10.2 Swelling Action 2-24
2.10.3 Loads From Jointed Rock Masses. . . . . . . . . . . . . . . . . . . . . . . . . . . 2-24

3.0 Geotechnical Properties 3-1


3. 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-1
3.2 Properties of Natural Soils 3-1
3.2. 1 Shear Strength . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3-1
Total (Undrained) Strength 3-1
Effective (Drained) Strength 3-2
3.2.2 Unit Weight 3-2
3.3 Backfill Properties 3-2
3.3.1 Choice of Backfill Material 3-2
Cohesionless Soils 3-2
Sandy Clays and Clayey Sands 3-3
Silts and Clayey Silts 3-3
Lightweight Aggregates 3-3
Flowable Fills 3-3
3.3.2 Placement and Compaction 3-4
3.4 Wall Friction and Adhesion 3-4
3.5 Other Soil Properties 3-5
3.6 Rock Properties 3-5
3.7 Example Problem 3-1 3-5

4.0 Cast-In-Place (CIP) Gravity and Semi-Gravity Walls 4-1


4.1 Introduction 4-1
4.2 Types of Semi-Gravity CIP Walls 4-3
4.2.1 Cantilever Walls 4-3
4.2.2 Counterfort Walls 4-4
4.2.3 Buttress Walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-4
4.2.4 Other CIP Semi-Gravity Walls 4-6
4.3 Wall Design 4-7
4.3.1 Design Earth Pressures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-7
4.3.2 Analysis of Sliding and Overturning 4-9
4.3.3 Global Stability Analyses 4-9
4.3.4 Bearing Capacity 4-11
4.3.5 Settlement and Tilt 4-11
4.3.6 Drainage 4-12
-4.4 Wall Construction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4-13
4.5 Example Problem 4-1 4-14

5.0 Modular Gravity Walls 5-1


5. 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-1
5.2 Crib Walls 5-3
5.2.1 General 5-3
5.2.2 Wall Design 5-5
5.2.3 Materials 5-7
5.2.4 Wall Construction 5-7
5.3 Concrete Modular Walls 5-8
5.3.1 General 5-8
5.3.2 Wall Design 5-10
5.3.3 Materials.......................................... 5-10
5.3.4 Construction 5-10
5.4 Bin Walls 5-12

ii
5.4.1 General 5-12
5.4.2 Wall Design 5-15
5.4.3 Materials.......................................... 5-15
5.4.4 Wall Construction 5-15
5.5 Gabion Walls 5-18
5.5.1 General 5-18
5.5.2 Wall Design 5-18
5.5.3 Materials 5-21
5.5.4 Wall Construction 5-22
5.6 Example Problem 5-1 5-23

6.0 Mechanically Stabilized Earth Walls 6-1


6. 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-1
6.1.1 Advantages/Limitations 6-5
6.1.2 Cost 6-5
6.2 Reinforcing Elements 6-6
6.2.1 Reinforcement Geometry 6-6
6.2.2 Reinforcement Material 6-6
6.2.3 Reinforcement Extensibility 6-6
6.3 Facing Elements 6-8
6.4 The Concept of Soil Reinforcement 6-16
6.4.1 Basic Soil Reinforcement Concept 6-16
6.4.2 Soil-Reinforcement Interaction 6-16
6.4.3 Pullout Resistance 6-18
Pullout Resistance Factor, F" 6-18
Steel Ribbed Reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . 6-18
Steel Grid Reinforcement 6-19
Geosynthetic Sheet Reinforcement 6-19
Geogrid Reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-19
Scale Effect Correction Factor, IX • • • • • • • • • • • • • . • • • • • • • • • • • • 6-19
6.5 Tensile Strength of Reinforcement 6-19
6.5. 1 Steel Reinforcement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-19
Detennination of Ac • • • • • • . • • • . • . . . . . • • • • • • . . • • • • • • • • • • 6-20
6.5.2 Geosynthetic Reinforcement 6-20
Creep Reduction Factor, RFcR • • . • . • . . . . . . . • . . . . . . . . • • . . . . 6-22
Durability Reduction Factor, RF D • . • • • • . • . • • . • . • . • • • • • • • • • • 6-23
Installation Damage Reduction Factor, RFID . • • • • • • . • • • • • • • • • • • 6-23
Factor of Safety, FS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-23
6.6 Wall Design 6-23
6.6.1 External Stability 6-23
Preliminary Sizing 6-25
Sliding and Limiting Eccentricity (Overturning) 6-25
Bearing Capacity 6-26
Deep Seated (Global) Stability 6-26
Impact of Surface Loads 6-34
6.6.2 Internal Stability 6-34
Step 1: Detennine Critical Failure Surfaces . . . . . . . . . . . . . . . . . . . . 6-34
Step 2: Detennine Maximum Tensile Force In Reinforcement 6-36
Step 3: Detennine Tensile Strength of Reinforcement 6-36

iii
Step 4: Check for Internal Stability Against Reinforcement Breakage 6-36
Step 5: Detennine Pull-out Resistance 6-38
Step 6: Check for Internal Stability Against Pullout 6-38
6.6.3 Required Reinforcement Length 6-38
6.6.4 Connection Design 6-39
6.6.5 Face Design Considerations 6-39
6.6.6 Wall Settlements 6-40
6.6.7 Lateral Wall Displacements 6-41
6.7 Backfill Material 6-41
6.8 Wall Construction 6-43
6.8.1 Construction of MSE Walls Strip Reinforcement and with Precast Facing 6-43
6.8.2 Construction with Geotextile or Mesh Facing 6-47
6.8.3 Backfill Placement and Compaction . . . . . . . . . . . . . . . . . . . . . . . . . 6-47
6.9 Durability and Long Tenn Perfonnance 6-52
6.9.1 Corrosion of Metallic Reinforcements 6-52
6.9.2 Durability of Geosynthetics 6-53
Hydrolysis 6-53
Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-53
Stress Cracking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-53
Ultraviolet Exposure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-53
Biological Degradation ........................ 6-53
6.10 Example Problem 6-1 6-54

7.0 Externally Supported Structural Walls 7-1


7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-1
7.2 Sheet Pile Walls 7-3
7.2.1 General 7-3
7.2.2 Wall Construction 7-6
Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-6
Equipment 7-6
7.3 Soldier Pile and Lagging Walls " 7-8
7.3.1 General............................................ 7-8
7.3. 2 Wall Construction 7-10
General 7-10
Preloading 7-11
Strut Removal and Rebracing 7-11
7.4 Slurry Walls 7-11
7.4.1 General 7-11
7.4.2 Wall Types 7-13
Conventional Reinforced Concrete Wall . . . . . . . . . . . . . . . . . . . . . . 7-13
Soldier-Pile-Tremie-Concrete (SPTC) Wall 7-13
Other Wall Types 7-13
7.4.3 Wall Construction 7-16
Materials 7-16
Equipment 7-17
Procedures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-17
Types of Supports 7-23
7.5 Tangent/Secant Pile Walls 7-23
7.5.1 General 7-23

iv
7.5.2 Wall Construction 7-25
7.6 Wall Design 7-26
7.6.1 Sheet Pile Walls 7-26
Cantilever Wall 7-26
Anchored Walls . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-29
Anchor Systems 7-36
Global Stability Analysis 7-36
7.6.2 Soldier Pile and Lagging Walls 7-41
Earth Pressure Distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-41
Effect of Support Type 7-44
Soldier Pile Spacing 7-45
Methods of Analysis 7-45
Support Spacing and Forces 7-45
Soldier Pile Design 7-45
Lagging Design 7-52
7.6.3 Slurry Walls 7-55
Earth Pressure Distribution 7-55
Methods of Analysis 7-55
7.6.4 Tangent and Secant Pile Walls 7-56
7.7 Anchor Design and Construction 7-56
7.7. 1 Anchor Components . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-56
7.7.2 Suitable Ground Conditions for Anchors 7-58
7.7.3 Determination of Minimum Spacing, Free Length and Inclination 7-59
7.7.4 Selection of Anchor Tendons 7-59
7.7.5 Design of the Bond Length 7-63
Anchor Capacity in Soils 7-63
Anchor Capacity in Rock 7-63
7.7.6 Corrosion Considerations in Design 7-65
Anchor Protection Against Corrosion 7-65
Protection of Free Length 7-65
Protection of Bond Length . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7-66
Protection of the Anchor Head 7-66
7.7.7 Construction........................................ 7-70
Materials 7-71
Construction Sequence 7-71
7.7.8 Testing and Stressing of Anchors 7-73
Preproduction Tests 7-74
Performance Tests 7-75
Proof Tests 7-75
Supplementary Extended Creep Tests 7-76
Acceptance Criteria 7-76
7.8 Wall Movements 7-80
7.9 Example Problems 7-83
7.9.1 Example Problem 7-1 7-83
7.9.2 Example Problem 7-2 7-87
7.9.3 Example Problem 7-3 7-90

8.0 In Situ Reinforced Earth Walls 8-1


8.1 Introduction 8-1

v
8.2 Soil Nail Walls 8-1
8.2.1 General 8-1
8.2.2 Design Issues 8-9
Current Soil Nail Wall Design Methods 8-10
Basic Design Concepts 8-10
Wall Design Considerations 8-13
8.2.3 Recommended Design Procedure 8-15
Step 1 - Set Up Critical Design Cross-Section(s) and Select a Trial Design 8-15
Step 2 - Compute the Allowable Nail Head Load 8-15
Step 3 - Minimum Allowable Nail Head Service Load Check 8-21
Step 4 - Define the Allowable Nail Load Support Diagrams 8-21
Step 5 - Select Trail Nail Spacings and Lengths 8-24
Step 6 - Define the Ultimate Soil Strength . . . . . . . . . . . . . . . . . . . . . 8-24
Step 7 - Calculate the Factor of Safety 8-26
Step 8 - External Stability Check . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-26
Step 9 - Check the Upper Cantilever 8-26
Step 10 - Check the facing Reinforcement Details . . . . . . . . . . . . . . . . 8-27
Step 11 - Serviceability Checks . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-28
Seismic Design 8-29
8.2.4 Simplified Design Charts for Preliminary Design of Cut Slope Walls 8-31
Design Variables 8-31
Design Chart Procedure 8-44
8.2.5 Corrosion Protection 8-45
8.2.6 Wall Drainage 8-46
8.2.7 Construction 8-49
8.2.8 Materials 8-54
8.2.9 Nail Testing 8-55
8.3 Micro-Pile Walls 8-57
8.3.1 General 8-57
8.3.2 Design 8-59
Reticulated Micro-Pile (RMP) Wall 8-59
Type "A" Insert Wall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-62
Pile Design . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8-62
8.3.3 Construction 8-67
Pile Installation .................................. 8-67
CIP Cap; Excavation and Facing 8-69
8.3 .4 Materials.......................................... 8-69
8.3.5 Testing 8-72
8.4 Example Problem 8-1 8-72

9.0 Chemically Stabilized Walls 9-1


9.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9-1
9.2 Jet Grouted Wall 9-1
9.2. 1 Introduction 9-1
9.2.2 Wall Configurations and Applications 9-2
9.2.3 Advantages and Limitations 9-3
9.2.4 Wall Construction 9-4
Jet Grouting Procedure 9-4
Jet Grouting Systems 9-4

vi
Hydrofracture Potential 9-7
Jet Grouting Equipment 9-7
Jet Grouting Materials . . . . . . . . . . ........................ . 9-7
9.2.5 Wall Design 9-10
Introduction 9-10
Factors Affecting Design . . . . . . . . .. ...................... 9-10
Design Procedure . . . . . . . . . . . . . ........................ 9-10
Field Trial 9-11
9.2.6 Quality Control/Quality Assurance 9-11
9.3 Deep Soil Mixing Wall 9-15
9.3. 1 Introduction 9-15
9.3.2 Wall Configuration and Applications 9-15
9.3.3 Advantages and Limitations . . . . . . ........................ 9-15
9.3.4 Wall Construction 9-17
General 9-17
DSM Equipment 9-17
DSM Installation Procedures 9-17
DSM Operational Parameters . . . . . ........................ 9-20
9.3.5 Wall Design 9-21
General 9-21
Factors Affecting Design 9-21
Design Considerations 9-22
9.3.6 Quality Control/Quality Assurance 9-24

10.0 Wall Selection 10-1


10.1 Introduction 10-1
10.2 Factors Affecting Selection 10-1
10.2.1 The Ground 10-1
10.2.2 Groundwater 10-2
10.2.3 Construction Considerations 10-3
10.2.4 Right-of-Way and Space Requirements 10-3
10.2.5 Aesthetics. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ......... . 10-3
10.2.6 Environmental Concerns 10-4
10.2.7 Durability and Maintenance . . . . . . . . . . . . . . . . . . . . ......... . 10-4
10.2.8 Cost 10-5
10.2.9 Tradition ; 10-5
10.2.10 Contracting Practice . . . . . . . . . . . . . . . . . . . . . . . . . ......... . 10-5
10.3 The Selection Process 10-5
10.3.1 Identification of Project Requirements and Site Constraints 10-5
10.3.2 Identification of Appropriate Wall Alternatives 10-8
10.3.3 Qualitative Rating of Alternatives 10-8
10.4 Wall Selection Example 10-8
10.5 Case Studies 10-11
10.5.1 Case 1 - Aviation Corridor Highway, Tucson, Arizona . . . . . . . . . .. 10-11
10.5.2 Case 2 - H-3 Tunnel Access Roads, Hawaii. . . . . . . . . . . . . . . . . .. 10-14
10.5.3 Case 3 - Newark Transit Rehabilitation, New Jersey , 10-16
10.5.4 Case 4 - Mt. McDonald Tunnel Ventilation, Canada , 10-16
10.5.5 Case 5 - Sterling Mountain Tunnel Portal, North Carolina '" , 10-18
10.5.6 Case 6 - New Jersey Turnpike Widening, New Jersey . . . . . . . . . . .. 10-19

vii
10.5.7 Case 7 - Shot Tower Metro Station, Baltimore, Maryland
10-20
10.5.8 Case 8 - Kismayo Port, Somalia 10-22
10.5.9 Case 9 - Canton-Seagirt Facility, Baltimore, Maryland
10-24
10.5.10 Case 10 - Jourdan Road Terminal, New Orleans, Louisiana
10-25
10.5 Summary.............................................. 10-27

11.0 Contracting Methods and Bidding Documents 11-1


11. 1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11-1
11.2 Contracting Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . " 11-1
11.3 Method Contracting Approach 11-2
11.3.1 Introduction 11-2
11.3.2 Bidding Documents 11-3
11.3.3 Special Provisions 11-4
Cast-in-Place Walls (Chapter 4) 11-4
Modular Gravity Walls (Chapter 5) 11-4
Mechanically Stabilized Earth Walls (Chapter 6) 11-4
Externally Supported Structural Walls (Chapter 7) 11-5
In situ Reinforced Earth Walls (Chapter 8) 11-5
Chemically Stabilized Earth Walls (Chapter 9) 11-6
11.4 Performance Contracting Approach . . . . . . . . . . . . . . . . . . . . . . .. 11-7
11.4.1 Introduction 11-7
11.4.2 Bidding Documents 11-7
Geometric and Site Data ........." 11-7
Design Guidelines 11-8
Performance Requirements 11-8
11.4.3 Review and Approval 11-9

12.0 References 12-1

viii
LIST OF FIGURES

HQ... Caption ~
1-1 Schematic of a Retaining Wall and Common Terminology . 1-1
1-2 Variety of Retaining Walls . 1-3
1-3 Classification of Earth Retaining Structures . . . . . . . . . . . . . . . . . . . . . . . . . . 1-5

2-1 Effect of Wall Movement on Wall Pressures . . . . . . . . . . . . . . . . . . . . . . . .. 2-2


2-2 Coefficients KA and ~ for Walls with Sloping Wall and Friction, and Sloping
Backfill . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 2-4
2-3 Computations of Active and Passive Pressures for Homogeneous Soils . . . . . . .. 2-6
2-4 Comparison of Plane and Curved (Log-Spiral) Failure Surfaces. (a) Active Case;
(b) Passive Case 2-8
2-5 Passive Coefficients for Sloping Wall with Wall Friction and Horizontal Backfill . 2-9
2-6 Passive Coefficients for Vertical Wall with Wall Friction and Sloping Backfill ... 2-10
2-7 Pressure Distribution for Stratified Soils . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2-11
2-8 Computation of Lateral Pressures for Static Groundwater Case 2-13
2-9 (a) Problem Geometry and Soil Conditions; (b) Lateral Effective Earth Pressure
Diagram; (c) Lateral Water Pressure Diagram 2-13
2-10 Lateral Pressure Due to Surcharge Loadings . . . . . . . . . . . . . . . . . . . . . . . . . 2-16
2-11 (a) Geometry of the Problem, (b) Variation of Lateral Pressure with Depth Due
to Line Load. . 2-18
2-12 Typical Residual Earth Pressure after Compaction on Backfill Behind an
Unyielding Wall 2-19
2-13 Computation of Earth Pressures due to Compaction with Rollers 2-21
2-14 Estimation of "Silo" Pressures 2-23

3-1 Angle of Wall Friction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 3-4


3-2 Geometry, Properties and Earth Pressures for Example Problem 3-1 3-5

4-1 CIP Gravity Wall. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 4-2


4-2 Cast-In-Place (CIP) Concrete Retaining Walls and Terminology. (a) Cantilever
Wall; (b) Counterfort Wall; and (c) Buttress Wall. . . . . . . . . . . . . . . . . . . . .. 4-2
4-3 Common Proportions of Cantilever Walls. . . . . . . . . . . . . . . . . . . . . . . . . .. 4-3
4-4 Common Proportions of Counterfort Walls. . . . . . . . . . . . . . . . . . . . . . . . .. 4-5
4-5 Other Types of Cast-in-Place (CIP) Walls: (a) Fill Wall with Limited ROW, (b)
Cut Wall with Limited ROW, and (c) U-Wall for Depressed Roadway . . . . . . .. 4-6
4-6 Design Criteria for Cast-in-Place (CIP) Concrete Retaining Walls . . . . . . . . . .. 4-8
4-7 Resistance Against Sliding from Keyed Foundation 4-10
4-8 Typical Modes of Global Instability 4-10
4-9 Typical Movement of Pile Supported Cast-in-Place (CIP) Wall with Soft
Foundation 4-12
4-10 Typical Retaining Wall Drainage Alternatives 4-13
4-11 Geometry and Parameters for Example Problem 4-11 4-14

5-1 Modular Gravity Walls. (a) Metal Bin Wall; (b) Precast Concrete Crib Wall; (c)
Precast Concrete Module Wall; and (d) Gabion Wall 5-1
5-2 Examples of Modular Gravity Wall Applications 5-2

ix
5-3 Crib Walls. (a) Uniform Cross-Section; (b) Stepped Cross-Section; (c) Typical
Details of a Reinforced Concrete Crib Wall 5-4
5-4 Active Earth Pressures for Design of Modular Gravity Walls. (a) Vertical Back
Face; (b) Back Face with Positive Slope; and (c) Back Face with Negative Slope . 5-6
5-5 (a) Typical Section of a Concrete Module Wall; (b) Typical Module; (c) Precast
Parapet . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 5-9
5-6 Typical geometry of Bin Wall. (a) Plan, (b) Elevation and (c) Section A-A ..... 5-13
5-7 (a) Elements of Bin Walls with T-Shaped Vertical Connector for Bin Walls, (b)
Channel Shaped Vertical Connector . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5-14
5-8 Construction of Bin Walls at Curves. (a) Typical Outside Corner, (b) Typical
Inside Corner and (c) Any Wall Configuration 5-19
5-9 Typical Gabion Wall Cross-Sections. (a) Wall with Stepped Front Face; (b) Wall
with Stepped Back Face; (c) Sloped Wall 5-20
5-10 Gabion Baskets. (a) Module Without Diaphragms; (b) Module with Diaphragms. 5-20
5-11 Geometry and Parameters for Example Problem 5-1 5-23

6-1 Principal Components of a Mechanically Stabilized Earth Wall . . . . . . . . . . . .. 6-1


6-2 Examples of MSE Wall Applications , 6-2
6-3 Strip Reinforcements 6-7
6-4 Grid Reinforcements 6-7
6-5 Geotextile Sheet Reinforcements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 6-8
6-6 Various Types of Wall Facings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-11
6-7 MSE Wall Surface Textures 6-12
6-8 Various Types of Modular Blocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6-13
6-9 Mechanisms of Pullout Resistance. (a) By Friction; (b) By Passive Resistance 6-17
6-10 Parameters for Metal Reinforcement Strength Calculations 6-21
6-11 External Stability Analysis Flow Chart 6-24
6-12 Potential External Failure Mechanisms foe a MSE Wall 6-24
6-13 External Analysis: Earth Pressures/Eccentricity. Horizontal Backslope with
Traffic Surcharge 6-27
6-14 External Analysis: Earth Pressures/Eccentricity. Sloping Backfill Case 6-28
6-15 External Stability for Wall with Broken Backslope 6-29
6-16 External Analysis: Calculation of Vertical Stress at the Foundation Level for
Bearing Capacity Calculations - Horizontal Backslope Condition 6-30
6-17 External Analysis: Calculation of Vertical Stress at the Foundation Level for
Bearing Capacity Calculations - Sloping Backslope Condition 6-31
6-18 Distribution of Stress from Concentrated Vertical Load 6-32
6-19 Distribution of Stress from Concentrated Horizontal Loads for External and
Internal Stability Calculations 6-33
6-20 Mechanisms of Internal Failure in MSE Walls. (a) Tension Failure; (b) Pullout
Failure 6-34
6-21 Internal Stability Analysis Flow Chart 6-35
6-22 Location of Potential Failure Surface for Internal Stability Design of MSE Walls.
(a) Inextensible Reinforcements; (b) Extensive Reinforcement 6-37
6-23 Design Values of the Lateral Earth Pressure Coefficient, K, for Various Types of
Soil Reinforcement Systems 6-38
6-24 Construction Procedures: Geotextile Retained Earth Wall with Wrap-Around
Facing 6-48
6-25 Geotextile Retained Earth Wall Details 6-48

x
6-26 (a) Geometry of the Problem, and (b) External Forces to be Considered in
Analysis 6-54

7-1 Primary Types of Externally Supported Structural Walls. (a) Soldier Pile and
Lagging Wall; (b) Soldier Pile and Cast-In-Place Concrete Lagging Wall; (c)
Master Pile Wall; (d) Sheet Pile Wall; (e) Slurry (Diaphragm) Wall; (f) Secant
Pile Wall; (g) Tangent Pile Wall; and (h) Interlocking H-Pile Wall 7-1
7-2 Wall Support Systems. (a) Cantilever Wall; (b) Earth Berm Support; (c) Raker
System; (d) Deadman Anchor; (e) Cross-lot Braced Wall; and (f) Anchored Wall. 7-2
7-3 Steel Sheet Pile Sections Commonly Used for Retaining Walls and Cofferdams.
(a) Z - Section; (b) U - Section; (c) Cold Formed Section 7-5
7-4 Sequence of Construction for a Backfilled Sheet Pile Structure 7-7
7-5 Sequence of Construction for an Excavated Sheet Pile Structure 7-7
7-6 Several Types of Soldier Piles. (a) Wide Flange Section; (b) Double Channel
Section; (c) Pipe Section . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 7-8
7-7 Typical Construction Sequence for a Slurry Wall. (a) Excavation; (b) Insertion of
Steel Tubing (End Stops); (c) Placement of Reinforcement Cage; and (d)
Concrete Placement 7-12
7-8 Conventional Reinforced Concrete Wall 7-14
7-9 Soldier-Pile-Tremie-Concrete (SPTC) Wall 7-14
7-10 Precast-Concrete-Pane1 Wall 7-15
7-11 Post-Tensioned-Concrete Wall 7-15
7-12 Cross-Section of a Guide-Wall. (a) Compact Cohesive Soil; (b) Loose
Cohesionless Soil 7-19
7-13 Reinforced Panel in Cast-In-Place Slurry Wall 7-22
7-14 Various Configurations of Bored-Pile Walls. (a) Tangent Pile Wall; (b)
Staggered Tangent Pile Wall; (c) Secant Pile Wall; (d) Intermittent Pile Wall with
Grouted Openings; and (e) Intermittent Pile Wall with Lagging 7-24
7-15 Construction Sequence. (a) Tangent Pile; and (b) Secant Pile Walls 7-25
7-16 Typical Deformation Conditions and Pressure Distribution for Cantilever Walls.
(a) Yielding Pattern of Cantilever Wall Penetrating a Sand Layer; (b) Net Actual
Earth Pressure Distribution; (c) Simplified Net Earth Pressure Distribution 7-26
7-17 Analysis for Cantilever Wall in Granular Soils 7-27
7-18 Earth Pressure Distributions and Design Procedure for Cantilevered Walls
Embedded in Cohesive Soils - End of Construction Case 7-28
7-19 Simplified Earth Pressure Distributions and Design Procedure for Cantilevered
Walls. (a) for Granular Soils; (b) for Walls Embedded in Cohesive Soil Retaining
Granular Soil; and (c) for Walls Embedded in Cohesive Soil Retaining Cohesive
Soil 7-30
7-20 Variation of Deflection and Bending Moment. (a) Free Earth Support Method;
(b) Fixed Earth Support Method 7-31
7-21 Analysis by Free Earth Support Method. (a) Sheet Piling in Granular Soils 7-32
7-21 Analysis by Free Earth Support Method. (b) Sheet Piling Backfilled With
Granular Soil and Embedded in Cohesive Soil 7-33
7-22 (a) Moment Reduction for Anchored Sheet Pile Wall Analyzed by Free Earth
Support Method for Granular Soils 7-34
7-22 (b) Moment Reduction for Anchored Sheet Pile Wall Analyzed by Free Earth
Support Method for Cohesive Soils 7-35
7-23 Analysis by Equivalent Beam Method 7-37

xi
7-24 Sheet Pile Anchorages. (a) Cast-In-Place Deadman; (b, c) Pile Anchorages; (d)
Steel Sheet Pile Wall Used as Anchorage; (e) Soil or Rock Anchor 7-38
7-25 Effect of Anchor Location Relative to the Wall 7-39
7-26 Design Criteria for Deadman Anchor Systems 7-40
7-27 Various Stages of Top-Down Construction. (a) Stage I - Cantilever Condition
(Fixed Earth Support); (b) Stage 2 - Intermediate Condition with One Level of
Support (Fixed Earth Support); (c) Stage 2 - Intermediate Condition with One
Level of Support and Excavation to Second Level of Support (Free Earth
Support); Stage 3 - Multi-level Support (Free Earth Support) 7-42
7-28 Pressure Distribution for Braced Walls 7-43
7-29 Pressure Diagrams for Soldier Pile and Lagging Walls. (a) Cantilevered During
Excavation for Top Row of Supports; (b) Intermediate Excavations for
Subsequent Support Installation; (c) Final Constructed Condition Assuming
Excavation for Utilities in Front of Wall 7-46
7-30 Design Procedure for Walls with Multiple Levels of Supports 7-47
7-31 Stability of Base for Braced Cuts in Cohesionless Soils 7-48
7-32 Stability of Base for Braced Cuts in Cohesive Soils 7-49
7-33 Procedure for Determining Depth of Penetration 7-50
7-34 Analytical Models for Walls with Multiple Levels of Supports by (a) Continuous
Beam Method; (b) Beam on Elastic Foundation Method 7-51
7-35 Arch Action at Soldier Pile and Lagging Walls 7-52
7-36 Schematic of a Typical Ground Anchor 7-57
7-37 Determination of Minimum Spacing and Free Length. (a) Vertical View; and (b)
Plan View 7-60
7-38 Stability Analysis of Anchored Wall Systems 7-61
7-39 Bonded Tendon 7-67
7-40 Simple Corrosion Protected Tendons. (a) Bar Tendon; (b) Strand Tendon 7-68
7-41 Encapsulated Double Corrosion Protection. (a) Bar Tendon; (b) Strand Tendon . 7-69
7-42 Anchor Head Details 7-70
7-43 Main Types of Cement Grout Injection Anchorage 7-71
7-44 Typical Plots of Tendon Movement for a) Performance Test, and b) Proof Test
(PTI, 1997) 7-78
7-45 Evaluation of Externally Supported Wall Movements 7-81
7-45 Evaluation of Externally Supported Wall Movements 7-82
7-46 Geometry of Example Problem 7-1 ., 7-83
7-47 Pressure Diagram 7-84
7-48 Geometry of Example Problem 7-2 7-87
.7-49 Earth Pressure Diagram 7-88
7-50 Geometry for Cantilever Condition 7-91
7-51 Pressure Diagram for Stage 1 7-92
7-52 Geometry for Condition of One Anchor 7-95
7-53 Pressure Diagram for Stage 2 7-95
7-54 Geometry for Final Condition with Two Anchors 7-97
7-55 Earth Pressure Distribution for Stage 3 7-98
7-56 Pressure Diagram for Continuously Supported Span for Stage 3 7-99
7-57 Pressure Diagram for Determination of Pile Penetration 7-100
7-58 Pressure Diagram for Lagging Design 7-101
7-59 Layout of Soil Anchors 7-103

xii
8-1 Soil Nail Wall Application. (a) Temporary Shoring; (b) Roadway Widening
Under Existing Bridge; and (c) Roadway Widening Cut 8-2
8-2 Examples of Micro-Pile Walls. (a) Stabilization of Slopes; (b) Excavation
Support; and (c) Roadway Slope. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 8-3
8-3 Typical Nail Wall Construction Sequence. . . . . . . . . . . . . . . . . . . . . . . . . .. 8-4
8-4 Load Transfer Mechanism in: (a) Ground Anchors; and (b) Soil Nails 8-8
8-5 Comparison of the Construction Sequencing for: (a) a Soil Nail Wall; and (b) a
Mechanically Stabilized Earth Wall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 8-8
8-6 Potential Failure Modes to be Analyzed for Soil Nail Walls. (a) Internal Failure;
(b) External Slope Failure; and (c) Mixed Failure. 8-9
8-7 Soil Nail Design - Basic Concepts and Terminology. (a) Unreinforced Slope;
(b) Reinforced Slope - Single Nail; and (c) Reinforced Slope - Multiple Nails 8-12
8-8 Nail Support Diagram 8-13
8-9 Typical Facing Pressure Distribution 8-16
8-10 Punching Shear of Nail Head Connections. (a) Temporary Bearing Plate
Connection; and (b) Permanent Headed - Stud Connection 8-18
8-11 Nail Length Distribution Assumed for Design 8-25
8-12 Deformation of Soil Nail Walls 8-30
8-13 Definition of "Internal" and "External" Slip Surfaces for Seismic Loading
Conditions 8-31
8-14(a) Preliminary Design Chart 1A. Backslope = 0° (Chart A) 8-32
8-14(b) Preliminary Design Chart lB. Backslope = 0° (Chart B) 8-33
8-14(c) Preliminary Design Chart 1C. Backslope = 0° (Chart C) 8-34
8-15(a) Preliminary Design Chart 2A. Backslope = 10° (Chart A) 8-35
8-15(b) Preliminary Design Chart 2B. Backslope = 10°; Face Batter = 0° (Chart B) . 8-36
8-15(c) Preliminary Design Chart 2C. Backslope = 10°; Face Batter = 10° (Chart C) . 8-37
8-16(a) Preliminary Design Chart 3A. Backslope = 20° (Chart A) 8-38
8-16(b) Preliminary Design Chart 3B. Backslope = 20°; Face Batter = 0° (Chart B) 8-39
8-16(c) Preliminary Design Chart 3C. Backslope = 20°; Face Batter = 10° (Chart C) . 8-40
8-17(a) Preliminary Design Chart 4A. Backslope = 34° (Chart A) 8-41
8-17(b) Preliminary Design Chart 4B. Backslope = 34°; Face Batter = 0° (Chart B) 8-42
8-17(c) Preliminary Design Chart 4C. Backslope = 34°; Face Batter = 10° (Chart C) .8-43
8-18 (a) & (b) Alternative Soil Nail Corrosion Protection Details 8-47
8-18 (c) & (d) Alternative Soil Nail Corrosion Protection Details 8-48
8-19 Typical Test Nail Detail 8-57
8-20 Design Model for RMP Walls - Gravity Wall Concept: a) Reticulated Pile
Structure, b) Calculation of Earth Pressure, c) "Gravity Wall" Model, and d)
Transformed Section 8-60
8-21 Preliminary Design Chart for Ultimate Horizontal Resistance of Piles 8-63
8-22 Ultimate Stress Transfer from Soil to Pile Versus Shear Strength of Soil 8-64
8-23 (a) Influence of Grouting Pressure on Ultimate Load Holding Capacity; and (b)
Effect of Post-Grouting on Skin Friction 8-68
8-24 Typical Construction Steps for a Micro-Pile Wall 8-68
8-25 Cross-Section of a Completed Micro-Pile Wall 8-71
8-26 Different Types of Reinforcement for Micro-Piles. (a) Reinforcing Bar; «b)
Pipe; (c) Strand; and (d) Rebar Cage 8-72
8-27 Sample Problem - Trial Design Cross-Section 8-74
8-28 Preliminary Design Charts for Backslope 20° and Face Batter 10°. (a)
Determination of To; and (b) Determination of LlH 8-75
8-29 Facing Details . . . . . . . . . . . . . . . 8-78

xiii
8-30 Construction Stability 8-84
8-31 Final Design Section 8-87

9-1 Grout Columns Layouts 9-3


9-2 Jet Grouted Wall Applications 9-3
9-3 Range of Soil Types Treatable by Chemical and Jet Grouting . . . . . . . . . . . . .. 9-4
9-4 Jet Grouting Procedure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 9-5
9-5 Schematic of Jet Grouting Systems: (a) Single Fluid System, (b) Double Fluid
System, and (c) Triple Fluid System. . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 9-5
9-6 Jet Grout Set Up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 9-8
9-7 Details of Jet Grouting Monitors: (a) Single Fluid System, and (b) Triple Fluid
System 9-8
9-8 Plan View of Typical DSM Wall Layouts: (a) Cut-off Wall, (b) and (c)
Excavation-Support Wall, and (d) Lattice Pattern for Liquefaction Control. 9-16
9-9 Various Wall Applications 9-16
9-10 DSM Wall Equipment. (a) Mixing Shaft for General Use; (b) Mixing Shaft for
Soil with Boulders; and (c) Mixing Shaft for Cohesive Soil. 9-18
9-11 Installation Procedure 9-20
9-12 DSM Wall Design. (a) Analysis for Punch-Through Shear; (b) Analysis for
Compressive Action of Arching Effects; (c) Empirical Guidelines for Avoiding
Bending Failure 9-23

10-1 Wall Selection Flow Chart . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-9


10-2 Single Wall Alternatives Considered for Aviation Corridor Highway 10-12
10-3 Double Wall Alternatives Considered for Aviation Corridor Highway 10-13
10-4 Retaining Structure Alternatives for North Halawa Access Road 10-15
10-5 Haiku Access Road Typical Wall Cross-Section 10-17
10-6 Cross-Section of Geotextile Wall for Newark Transit 10-17
10-7 Element Wall for Mt. McDonald Ventilation Building . . . . . . . . . . . . . . . . . . 10-19
10-8 Soil Nail Retaining Structure for New Jersey Turnpike Widening 10-20
10-9 Shot Tower Station Retaining Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-21
10-10 Cross-Section of Bulkhead for Kismayo Port . . . . . . . . . . . . . . . . . . . . . . . . 10-23
10-11 Master Pile Retaining Structure Proposed for Canton Seagirt Facility . . . . . . . . 10-25
10-12 Cellular Structure Reinforced with Stone Columns Used for the Canton-Seagirt
Facility .... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10-26
10-13 Composite Wharf Structure for Jourdan Road Terminal 10-26

xiv
LIST OF TABLES

~
2-1 Ultimate Friction Factors and Adhesion for Dissimilar Materials 2-7
2-2 Computation of Lateral Pressures due to Line Load 2-17

5-1 Tolerance of Modular Gravity Walls to Different Settlement 5-7


5-2 Typical Height-Thickness Relationship for Bin Walls 5-12

6-1 Summary of Reinforcement and Face Panel Details for Selected MSE Wall
Systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . 6-4
6-2 Creep Reduction Factors, RFcR . . . • . • • . • . . . • . . . . . . . . • . . . . . . • . . • . 6-22
6-3 Minimum Embedment Requirements for MSE Walls 6-25
6-4 Relationship Between Joint Width and Limiting Differential Settlement for MSE
Walls 6-40
6-5 Backfill Requirements for MSE Walls ............................, 6-42

7-1 Strength Properties for Typical Grades of Timber 7-53


7-2 Recommended Thickness of Wood Lagging 7-54
7-3 Critical Values and Test Standards to Evaluate Soil Aggressiveness 7-59
7-4 Typical Steel Properties and Dimensions for Tendons 7-62
7-5 Relationship Between Tendon Size and Drilled Hole Diameter 7-62
7-6 Presumptive Ultimate Values of Load Transfer for Preliminary Design of
Anchors in Soil 7-64
7-7 Presumptive Ultimate Values of Load Transfer for Preliminary Design of
Anchors in Rock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 7-64

8-1 Soil Nail Design Methods 8-11


8-2 Load Combinations in AASHTO Specifications 8-15
8-3 Facing Pressure Factors Recommended Values for Design 8-17
8-4 Nail Head Strength Factors 8-20
8-5 Strength Factors and Factors of Safety . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 8-22
8-6 Ultimate Bond Stress - Cohesionless Soils . . . . . . . . . . . . . . . . . . . . . . . . .. . 8-23
8-7 Ultimate Bond Stress - Cohesionless Soils . . . . . . . . . . . . . . . . . . . . . . . . .. . 8-23
8-8 Ultimate Bond Stress - Rock 8-23
8-9 Ground Aggressiveness Indicators . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 8-46
8-10 Types of Soil Nail Tests 8-56
8-11 Values of K and I 8-66
8-12 Shotcrete Construction Facing 8-80
8-13 Pennanent Facing 8-80
8-14 Nail Length Distribution 8-81

9-1 Summary of System Variables and Their Impact on Basic Design Elements 9-12
9-2 Typical Range of Jet Grouting Parameters and Jet-Grouted Soil Properties 9-13

10-1 System Selection Chart for Fill Walls 10-6


10-2 System Selection Chart for Cut Walls 10-7
10-3 Selection Matrix 10-10
10-4 Summary of Selection Factors 10-28

xv
LIST OF PHOTOS

& Ii1k ~
2-1 (a) Retaining Wall with Uniform Surcharge Load; (b) Retaining Wall with Line
Loads and Point Loads 2-15

4-1 CIP Cantilever Retaining Wall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . 4-4


4-2 CIP Counterfort Wall Construction in a Cut Application 4-5
4-3 CIP U-Wall for Depressed Roadway 4-7
4-4 CIP Abutment with Integral Wing Walls 4-10

5-1 Setting Precast Elements for an Open Faced Crib Wall 5-8
5-2 Construction of Concrete Module Wall 5-11
5-3 Bin Wall with (a) Corrugated Steel Facing, and (b) Precast Concrete Facing 5-16
5-4 Construction of a Bin Wall (a) Setting Preassembled Panels, and (b) Filling the
Completed . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 5-17
5-5 Gabion Wall Construction 5-24

6-1 to 6-5 Examples of MSE Walls 6-3


6-6, 6-7 Metallic Reinforcements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . . 6-9
6-8,6-9 Geosynthetic Reinforcements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 6-10
6-10,6-11 Modular Block Facing 6-14
6-12 Concrete Pad Construction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 6-44
6-13 Erection of Facing Elements. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 6-44
6-14 Placement of Reinforcements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 6-44
6-15 Placement of Backfill. 6-45
6-16 Spreading of Backfill. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 6-45
6-17 Backfill Compaction. 6-45
6-18 Placement of Backfill Over Geotextile Sheets 6-49
6-19 Backfill Compaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 6-49
6-20 Placement of Wire Mesh Cover Over Wrapped Geotextile Facing 6-49
6-21 Gunite Facing and Drainage Pipes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. . 6-49
6-22,6-23 Lightweight Compaction Adjacent to Wall Facing 6-51

7-1 Sheet Pile Wall for Earth Support Behind a Cast-in-Place Wall 7-4
7-2 Sheet Pile Cofferdam for Footing Construction on Land 7-4
7-3 Sheet Pile Cofferdam for Construction of Foundations in Water 7-4
7-4 Anchored Sheet Pile Bulkhead for Waterfront Structure 7-4
7-5 Soldier Pile and Lagging Wall for Temporary Excavation Support 7-9
7-6 Permanent Soldier Pile and Lagging Walls for (a) Roadway Embankment,
and (b) Roadway Cut. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .' . . 7-9
7-7 Soldier Pile and Logging Wall: (a) Drilling for Pile Installation, and (b)
Excavation Between Soldier Piles in Cohesive Soil for Installation of Lagging. .. 7-10
7-8 Slurry Trench Excavation Equipment: (a) Hydraulically Operated Clamshell
Bucket, and (b) Hydromill 7-18
7-9 Slurry Wall Reinforcement Cage (a) On Fabrication Bed, Showing Styrofoam
Knock-Out Panel, and (b) During Lifting for Installation into Slurry Filled
Trench 7-21
7-10 Slurry Filled Trench with Tremie Pipes Just Prior to Concrete Placement 7-21

xvi
7-11 Tangent Pile Wall a) With Structural Steel Section as Core Reinforcement, and
b) Face of Completed Wall 7-24
7-12 Drilling Equipment for Installation of Ground Anchors: (a) Hollow Stem
Auger, and (b) Down-the-Hole Hammer 7-72
7-13 Installation of Anchor Tendon 7-73
7-14 Testing Arrangement for Strand Tendon 7-74

8-1 Soil Nail Wall Applications: (a) Temporary Shoring; (b) Roadway Widening
Under Existing Bridge, (c) Roadway Cut; and (d) Slope Stabilization 8-6
8-2 Soil Nail Construction: (a) Excavation; (b) Drilling with Hollow System Auger;
and (c) Shotcrete Construction Facing 8-51
8-3 Soil Nail Wall Final Facing: (a) Cast-in-Place (CIP) Facing; (b) Shotcrete
Facing; and (c) Precast Concrete Facing. 8-53
8-4 Exposed Piles of Reticulated Micro-Pile (RMP) Wall 8-58
8-5 Construction of a Reticulated Micro-Pile (RMP) Wall: (a) Drilling for Pile
Installation; (b) Placement of CIP Concrete Cap; and (c) Excavation and
Shotcrete Facing. . 8-70

9-1 Grout Mixing and Injection Plant 9-9


9-2 Drilling and Grouting Rig 9-9
9-3 Forming a Jet-Grouted Column 9-9
9-4 Jet Grouting Monitor . . . . . . . . . . . . . . . . . . . . . . . . .. . . 9-9
9-5 Monitor for a Triple Fluid Grouting System 9-10
9-6 Drilling and Mixing Unit 9-19
9-7 Mixing Plant 9-19

xvii
LIST OF NOTATIONS

CHAPTER 2

Ph Lateral or horizontal pressure


Pa Active earth pressure
Pp Passive earth pressure
Po At-rest earth pressure
K Coefficient of lateral pressure
Ka Coefficient of active earth pressure
Kp Coefficient of passive earth pressure
Ko Coefficient of at-rest earth pressure
0 Total Stress
0' Effective Stress
°vo Vertical pressure
o'vo Effective vertical stress
H Wall height
Y/H Wall rotation
u Pore pressure
Yw Unit weight of water
y' Total Unit weight
Ph' Lateral earth pressure
4> Angle of internal friction
4>' Angle of effective (drained) friction
Q Dimensionless coefficient

"
y
Angle of wall friction
Unit weight of soil
z Depth from top of wall
Zo Depth of tensile zone
z Location of resultant from base of wall
Pw Hydrostatic Force
c Cohesion
Angle of friction between soil and wall
"
~ Backfill slope angle
e Angle of back face of the wall with the vertical
t Lift thickness
w Roller thickness

xviii
q Line load intensity

CHAPTER 3

't Shear strength


c Cohesion
<p Angle of internal friction
'tu Undrained shear strength
cu Undrained cohesive strength
<Pu Undrained friction angle
a Total stress
't
d
Drained shear strength
cl Drained cohesive strength
<pI Drained friction angle
u Pore pressure
(, Angle of wall friction
c. Wall adhesion
y Unit weight
K Coefficient of active earth pressure

p. Lateral pressure at base of wall
p Active earth force

H Height

CHAPTER 4

H Height
(, Angle of wall friction
~ Backslope angle
e Eccentricity
B Width of the base of the footing
qMAX Maximum contact pressure on foundation
qMIN Minimum contact pressure on foundation
W Weight of wall and soil above footing, for cantilever and counterfront walls
c Cohesion of foundation soil
c. Adhesion of concrete on soil
Pp Passive resistance
FS Factor of Safety

xix
p. Resultant of active pressure
Pb Horizontal component of active pressure
Pv Vertical component of active pressure
F Total Horizontal Resistance
K. Coefficient of active earth pressure
<I> Internal friction angle
0b Friction angle between concrete base and foundation soil
MR Resisting moment about point A
M0 Overturning moment about point A
MpH PHb
Mw MR
M pv Pvt
FS s Factor of safety against sliding
FS o Factor of safety against overturning
d Distance of Resultant from point A
quit Ultimate bearing capacity
FS bt Factor of safety against bearing failure

CHAPTERS

o Angle of wall friction


<1>' Friction angle of the compacted backfill soil
Yg Unit weight of the wall
or Porousity of the rock fill
Gs Specific gravity of the rock
yw Unit weight of water
K. Lateral earth pressure
FS Factor of safety
FS s Factor of safety against sliding
FS 0 Factor of Safety against overturning
f.L Friction between rockfill and foundation materials
d Distance of Resultant from point A
e Eccentricity
qmax Maximum pressure
qmill Minimum pressure

xx
CHAPTER 6

Pr Pullout resistance
F* Pullout resistance factor
IX Scale effect correction factor

avo
I Effective vertical stress at the soil-reinforcement interfaces
L. Embedment or adherence length in the resisting zone behind the failure surface
C Reinforcement effective unit perimeter
p Soil-reinforcement interaction friction angle
Fq Embedment (or surcharge) bearing capacity factor
IXp Bearing factor for passive resistance
Cu Uniformity coefficient of the backfill
<I> Friction angle
SI Distance between transverse bars
(X Scale correction factor
Ta Allowable tensile force per unit width of the reinforcement
b Gross width of the strip, or unit width of the sheet, grid, or bar mat
Fy Yield stress of steel
A. Design cross section area of the steel, defined as the original cross section area minus corrosion
losses anticipated to occur during the design life of the wall
t. Thickness of the reinforcement at the end of the design life
to Nominal thickness at construction
t5 Sacrificial thickness of metal expected to be lost by uniform corrosion during the service life of
the structure
Nb Number of longitudinal bars per unit width b
D* Diameter of bar or wire corrected for corrosion loss
b Unit width of reinforcement (if reinforcement is continuous count number of bars for reinforcement
width of 1 unit)
FS Factor of safety
R. Reinforcement coverage ratio
T max Maximum load applied to reinforcement
T ULT Ultimate (or yield tensile) strength from wide strip tensile strength tests based on minimum average
roll value for the product
RF Dimensionless product of all applicable reduction factors
RF CR Creep reduction factor
RF 0 Durability reduction factor
RF 10 Installation damage reduction factor

xxi
FS Overall factor of safety
H Design height
FS o Factor of safety against overturning
FS s Factor of safety against sliding
e Eccentricity
q Traffic live load
R Resultant of venical forces
4>r Friction angle of reinforced eanh
Yr Unit weight of the reinforced eanh
Kr Coefficient of Lateral pressure
VI Weight of the reinforced zone
DI Effective width of applied load at any depth
bf Width of applied load
Py Load per linear meter of strip footing
PI
y
Load on isolated rectangular footing or point load
Zz Depth where effective width intersects back of wall face
F1 Lateral force due to eanh pressure
Fz Lateral force due to traffic surcharge
PHI Lateral force due to superstructure or other concentrated lateral loads
P HZ Lateral force due to superstructure or other concentrated loads
e Eccentricity of load on footing
K Lateral eanh pressure coefficient
CJ Iy Venical soil stress
z Depth to the reinforcement layer under consideration
q Uniform surcharge load
FS po Safety factor against pullout
Le Length of embedment in the resisting zone
CJIy Overburden pressure
Zp Depth of soil at reinforcement layer at beginning of resistance zone, for pullout calculations
L1 Total length of reinforcement
Yf Unit weight of the retained backfill
PI Internal inenial force due to the weight of the backfill within the active zone
Lei Length of the reinforcement in the resistant zone of the i'th layer
T md Load per unit wall width applied to each reinforcement due to dynamic forces
T all Allowable strength
OR Relative displacement
0max Maximum displacement

xxii
PI Plasticity index
MR Resisting moment
M0 Overturning moment
AI Tributary area
N Number of reinforcing strips
Sb Two times panel length
PR Resistance
z Depth

CHAPTER 7

<1>' Effective friction angle


c' Effective Cohesion
N Standard penetration resistance
D Depth of penetration
Z Distance
Coefficient of passive earth pressure
Coefficient of active earth pressure
Vertical pressure behind exposed base at wall dredging level due to backfill
Unconfined compressive strength of clay
Unit weight of soil
Active pressure
Passive pressure
Active earth pressure coefficient
Point of action
Embedment
Magnitude of active pressure at the exposed wall base due to surcharge load
Passive earth pressure coefficent
Horizontal anchor load per unit horizontal length of wall
Force from earth pressure above point a + other horizontal forces (except T)
Equivalent unit weight
Active pressure coefficient for the soil below the exposed base of wall
Design or reduced moment
Theoretical maximum moment
Section modulus
Moment of inertia
Flexibility of wall

xxiii
E Modulus of elasticity
aan Allowable stress in pile material
SD Stability number
C
u
Undrained cohesion below the exposed base of the wall
ca Wall adhesion
Y.H Weight of backfill and surcharge load above the exposed base of the wall
K I Coefficient of active earth pressure for the soil below the exposed base of the wall
a
y Distance from the expose base of wall to the point of zero pressure
R Horizontal reaction at point 0, obtained by assuming the piling is simply supported at point 0 and
at the anchor level
pI Passive earth pressure in the soil below the exposed base of the wall
p
pI Active earth pressure in the soil below the exposed base of the wall
a
KI Coefficent of passive earth pressure
p
Po Resultant
Ap Ultamate anchor resistance
Apt One-half of the ultimate anchor resistance
PI Trapezoid force
Pa Rankine force
Pb Unbalanced force
N y2 Bearing capacity factor
YI Moist unit weight
Y2 Submerged unit weight
Nt Bearing capacity factor
c Unit strength of clay
H Excavation depth
Y Unit weight of soil
q Surcharge
N tD Bearing capacity factor
c1 Unit strength of soft stratum
V Shear at tip of sheeting
Rd Horizontal component of load at bottom brace or anchor level
fb Allowable flexural stress
W Weight of soil mass within the failure surface
S4> Frictional component of soil resistance
Sc Component of soil resistance due to cohesive soil strength
PA Active earth force between point A and point C
T Tieback force

xxiv
Ta Allowable anchor load capacity
TI Test load capacity
AL Alignment load
P Design load
<p Angle of internal friction
(, Friction angle between sheetpiling and sand
~ Backslope angle
e Angle of wall with vertical
R Reduction factor
Ysal Saturated unit weight
Ka Active earth pressure coefficient
Kp Passive earth pressure coefficient
Pd Earth pressure on lagged depth
Ps Earth pressure below lagged depth
p Iq Lateral surcharge pressure on lagged depth
P sq Lateral surcharge below lagging up to pivot point
P Total load on lagged depth
TH Load in each tieback
L Length

CHAPTERS

N SPT blow count


W Service load (F)
C Ultimate soil cohesion (F/L:!)
<p Ultimate soil friction angle (0)
F Global factor of safety applied to soil shear strength
T FN Nominal nail head strength (F)
TF Allowable nail tendon load (F)
T NN Nominal nail tendon strength (F)
TN Allowable nail tendon load (F)
Qu Ultimate pullout resistance (F/L)
Qd Allowable pullout resistance (F/L)
T Allowable nail load (F)
S Resisting shear force (F)
N Nominal force (F)
Q Nail-ground pullout resistance (allowable)

xxv
D Dead load
L Live load
E Earth pressure
B Buoyancy
T FN Critical nominal nail head strength
my,NEG Vertical nominal unit moment resistance at the nail head
my,pos Vertical nominal unit moment resistance at the midspan locations
SH Horizontal nail spacings
Sv Vertical nail spacings
CF The pressure factor for facing flexure
T FN Critical nominal head strength
Cs Shear pressure factor
D GC Diameter of soil nail
Ac Area of punching shear cone base at back of facing
A GC Cross-sectional area of soil nail borehole
VN Nominal internal punching shear strength of the facing
DI Effective punching cone diameter
C
he Effective cone depth
T FN Punching shear strength
S HS Stud spacing
(XF Nail head strength factor
FF Empirical factor
KA Active earth pressure coefficient
y Soil unit weight
H Vertical height of soil nail wall
tF Nail head service load
PATotal active load
AB Nominal cross-sectional area of the bar
Fy Yield strength of steel
(X Strength factor
(XN Tendon yield strength
(XQ Pullout resistance strength factor
L Maximum nail length
H Wall height
QD Dimensionless pullout resistance
LlH Nail length-to-wall-height ratio
Apt Earthquake peak ground acceleration

xxvi
Horizontal acceleration
Horizontal top of wall deflection
Vertical top of wall deflection
Horizontal distance behind top of wall which may be influenced by construction induced wall
deformation
T} Face batter (degrees)
L Nail length
Damping coefficient
Backslope angle
Face or batter angle
Factored friction angle
Dimension less cohesion
Dimension less nail tensile capacity
Eccentricity
FS Factor of safety
R Horizontal component of the soil shear resisting forces along the critical slip surface
Additional shear resistance due to micro-piles
Horizontal component of the driving force along the critical slip surface (p. 8-59)
Depth
Unit skin friction
Vertical effective stress at depth z
Coefficient of earth pressure at the wall of the pile
Angle of wall friction
Unstrained shear strength of soils
Ultimate pile capacity
D Nominal diameter (cm)
L Pile length (cm)
K Empirical coefficient representing average pile-soil interaction (krn/cm2)
I Empirical factor related to the pile diameter
UCS Uniaxial compressive strength
Earth pressure coefficient
Parameter reflecting the effect of increased pile diameter due to grouting pressure
In situ density
Internal friction angle
Cohesion
v One-way unit shear force at the top row of nails
M Allowable one-way moment resistance

xxvii
V NS One-way unit shear strength of the facing

CHAPTER 9

DSM Deep soil mixing


SCC Soil-cement-column
't r Shear strength
qu Unconfined compressive strength
Eso Modulus of elasticity measured at 50 percent strength
qu(28l Unconfined compressing strength measured on 28th day
Qu(7l Unconfined compressing strength measured on 7th day
ex Coefficient of elasticity of the soil-cement
D Column diameter
Q Resistance to punching shear

xxviii·
CHAPTER 1.0
INTRODUCTION

1.1 PURPOSE AND SCOPE OF MODULE

Earth retaining structures (or retaining walls) are used to hold back earth and maintain a difference in the
elevation of the ground surface as shown in Figure 1-1. The retaining wall is designed to withstand the
forces exerted by the retained ground (commonly referred to as the "backfill"), and to transmit these forces
safely to a foundation and/or to restraining elements located beyond the failure surface.
Q2
Q( ~
Surcharge
/ loads I
I
c 1jI, f ~ Failure
'i-l~~'fIlf' Backfill I
surface
.r I
/"::
/ ::~ Restraining
Pl=Later;1 clement
Front pressure (e.g. Deadman)
batter ,
/
/

Backface

Figure 1-1: Schematic of a Retaining Wall and Common Terminology.

In general, the cost of constructing a retaining wall is usually high compared with the cost of forming a
new slope. Therefore, the need for a retaining wall should be assessed carefully during preliminary design
and an effort should be made to keep the retained height as low as possible.

In highway construction, retaining walls are used along cuts or fills where space is inadequate for
construction of cut slopes or embankment slopes. Bridge abutments and foundation walls, which must
support earth fills, are also designed as retaining walls. Typical earth retaining structure applications for
highway construction include:

• new or widened highways in developed areas


• new or widened highways at mountain or steep slopes
• grade separation
• bridge abutments, wing walls and approach embankments
• culvert walls
• tunnel portals and approaches
• flood walls, bulkheads and waterfront structures
• cofferdams for construction of bridge foundations
• stabilization of new or existing slopes and protection against rockfalls
• groundwater cut-off barriers for excavations or depressed roadways

1- 1
Figure 1-2 presents some of the wide variety of retaining walls. A great number of wall systems have been
developed in the past two decades by specialty contractors who have been promoting either a special
product or a specialized method of construction, or both. Due to the rapid development of these diversified
systems and their many benefits, the design engineer is now faced with the difficult task of having to select
the best possible system; design the structure; and ensure its proper construction. The purpose of this
module is to provide the practicing engineer with a thorough understanding of various retaining walls and
their application in highway facilities.

The scope of this module is limited to retaining walls in highway practice. Although many of the walls and
their design principles can be applied to retaining walls in underground facilities, such as basements and
cut-and-cover tunnels, this manual does not include specific guidance for such cases.

1.2 mSTORY OF EARTH RETAINING STRUCTURES

Historical reviews of the developments in retaining walls have been presented by Kerisel (1992), Gould
(1990), and O'Rourke and Jones (1990). Following is a brief review of these historical developments.

Examples of wall construction have been traced back to approximately 3000 B.C. (Kerisel, 1992). These
earlier walls were primarily gravity structures that relied on their weights to resist the earth pressures.
They were constructed of stone masonry (with or without mortar) or various types of cribs or bins with
different filling materials.

The development of "modern" retaining walls gained impetus in the late 19405, 1950s and 1960s due to
the innovations in construction technology. New construction methods were developed at that time, largely
through adaptations and improvements in specialized excavation and drilling equipment. In that period,
North America was introduced to new forms of cast-in-place below-grade walls, tiebacks in soil and rock
(both temporary and permanent), and the very important concept of earth reinforcement (Vidal, 1966)
which led to a wide variety of earth retaining structures. Most of the procedures now being utilized for
construction of earth retaining structures were well developed by 1970 or had appeared in full scale
experiments.

The most important breakthrough in the design of earth retaining structures which occured in this era was
the recognition that the earth pressure acting on a wall is a function of the type of wall and the extent of
its movement. Classical earth pressure theories, which were developed by Coulomb (1776) and Rankine
(1857), were formalized for use by Caquot and Kerisel (1948) and others. Sophisticated analyses of soil-
structure interaction and wall/soil movements began in the 19605 using finite difference and finite element
analytical procedures. The simultaneous advancement of geotechnical instrumentation equipment and
monitoring procedures made the "observational method" of design (Peck, 1969) popular and cost effective.

Since 1970 there has been a dramatic growth in the methods and products for retaining soil. O'Rourke and
Jones (1990) describe two trends in particular which have emerged since 1970. First, there has been an
increasing use of reinforcing elements, either by incremental burial to create reinforced soils (MSE walls),
or by systematic in situ installation to reinforce natural soils (soil nailing). Mechanically stabilized earth
and soil nailing have changed the ways we construct fill or cut walls, respectively, by providing
economically attractive alternatives to traditional construction methods. Second, there has been an
increasing use of polymeric products to reinforce the soil and control drainage. Rapid developments in
polymer manufacturing have supplied a wide array of geosynthetic materials. The use of these products
in construction has encouraged a multitude of different earth retention schemes which seem to be limited
only by the imagination of the engineer.

1-2
Potential failure wedge

Cantilever
Wall ~ Gravity
Elements

Cantilever Gravity Element

In situ wall In situ wall


"'-...------.."""""" "'-...------r"""""'"
Strut
Raker Anchors
Tiebacks

Braced Tied-back (Anchored)

(a) Externally Stabilized Systems

Facing
panels
~

Reinforced Soil Soil Nailing

(b) Internally Stabilized Systems

Figure 1-2: Variety of Retaining Walls. (Adapted from O'Rourke and Jones, 1990)

1-3
The rapid development of these new trends, and the increased awareness of the impact of construction on
the envirorunent, have led to the emergence of the concept of "earth walls". In this concept, the soil
supports itself or is incorporated into the structure and assumes a major structural or load carrying function.
By adopting this concept, the structural members requirements of the system are reduced, or eliminated
altogether. Examples of recently developed earth walls include the soil-reinforcement systems discussed
above, as well as systems involving chemical treatment of the in situ soil such as jet grouting or deep soil
mixing.

1.3 CLASSIFICATION OF EARTH RETAINING STRUCTURES

The wide variety of earth retaining systems may be classified by:

• Load support mechanism, i.e., externally or internally stabilized walls


• Construction concept, i.e., fill or cut walls
• System rigidity, i.e., rigid or flexible walls

1.3.1 Classification by Load Support Mechanism

For the purpose of this manual, the walls have been organized according to two principal categories of
externally and internally stabilized systems (O'Rourke and Jones, 1990). An externally stabilized system
uses an external structural wall, against which stabilizing forces are mobilized. An internally stabilized
system involves reinforcements installed within and extending beyond the potential failure mass. Hybrid
systems combine elements of both internally and externally supported walls. Figure 1-3 presents the
organization of walls used in this manual.

Virtually all traditional types of walls may be regarded as externally stabilized systems. Gravity walls, in
the forms of cantilever structures or gravity elements (e.g., bins, cribs and gabions), support the soil
through weight and stiffness to resist the sliding, overturning, and shear. Bracing systems, such as cross-
lot struts and rakers, provide temporary support for in situ structural and chemically stabilized walls.
Ground anchors provide support through the pullout capacity of anchors established in stable soil outside
of the zone of potential failure.

It is in the area of internally stabilized systems that a fundamentally new concept has been introduced.
Shear transfer to mobilize the tensile capacity of closely spaced reinforcing elements has freed retaining
structures of the need for a structural wall, and has substituted instead a composite system of reinforcing
elements and soil as the primary structural entity. A facing is required on an internally stabilized system,
but its role is to prevent raveling and deterioration, rather than to provide primary structural support.

1.3.2 Classification by Construction Concept

Earth retaining structures can also be classified according to the construction method, i.e., fill construction
or cut construction. Fill wall construction refers to a wall system in which the wall is constructed from
the base of the wall to the top (i.e., "bottom-up" construction). Cut wall construction refers to a wall
system in which the wall is constructed from the top of the wall to the base (i.e., "top-down" construction)
concurrent with excavation operations. The classification of each wall system per its construction concept
is also presented in Figure 1-3.

It is important to recognize that the "cut" and "fill" designations refer to how the wall is constructed, not
necessarily the nature of the earthwork (i.e., cut or fill) associated with the project. For example, a

1-4
EARTH RETAINING STRUCTURES

I I i

Externally Stabilized Internally Stabilized

I I I iii
In situ Walls Gravity Walls In situ Reinforced

'---..
Mechanically Stabilized

, , I i
(Fill Walls)
(Chapter 6)
(Cut Walls)
(Chapter 8)

Structural Chemical Gravity Semi-Gravity Modular Gravity Metallic and Polymeric Soil Nailing
(Cut Walls) (Cut Walls) (Fill Walls) (Fill Walls) (Fill Walls) reinforcing strips, Micro Pile
(Chapter 7) (Chapter 9) (Chapter 4) (Chapter 4) (Chapter 5) grids and sheets
Anchored Earth
....
I Sheet-pile* Jet Grout Mass Concrete Cantilever Crib Reinforced Soil Slopes
VI
Soldier pile-Iagging* Deep Soil Mix Stone Masonry Counterfort Bin
Cast-in-situ Buttress Gabion
- Slurry Walls Concrete Module
Bored pile
- Contiguous
Tangent pile
Secant pile Hybrid Walls
- Non contiguous
Tailed Segmental
Low Density Fills

Braced Anchored

* can also be used in fill conditions


Figure 1-3: Classification of Earth Retaining Structures (Modified from O'Rourke and Jones, 1990)
prefabricated modular gravity wall, which may be used to retain earth for a major highway cut, is
considered a fill wall as it is constructed "bottom-up" after the excavation for the cut has reached its final
grade.

1.3.3 Classification by System Rigidity

The rigidity (or flexibility) of a wall system is fundamental to the understanding of the development of earth
pressures (discussed in Chapter 2). In simple terms, a wall is considered to be rigid if it moves as a unit
(rigid body rotation and/or translation) and does not experience bending deformations. Most gravity walls
can be considered rigid walls.

Flexible walls are those walls which undergo bending deformations in addition to rigid body motion. Such
deformations result in a redistribution of lateral pressures from the more flexible to the stiffer portions of
the system. Virtually all wall systems, except the gravity walls, may be considered to be flexible.

1.4 ORGANIZATION OF THE MANUAL

This manual is intended to be a stand-alone document and is geared towards providing the practicing
engineer with a thorough understanding of the various types of retaining walls used for highway
applications. Accordingly, the manual starts with a discussion of earth pressure theories and the
determination of basic soil parameters, and then proceeds to discuss each of the earth retaining systems
identified in Section 1.3. The organization of the manual is presented below.

Chapter 2 describes methods used to determine design earth pressures for different conditions and discusses
other issues, such as surcharge loads and groundwater, that influence the design earth pressures.

Chapter 3 presents a discussion of the soil properties that influence lateral earth pressures. This chapter
also discusses backfill materials and provides guidance on the choice of suitable backfill material.

Chapters 4 through 9 discuss the various types of retaining walls as identified in Section 1.3. For each
type, the chapters are organized in the following fashion:

• background, uses, advantages and limitations


• analytical and design procedures
• construction materials and techniques
• maintenance and durability issues
• example problems

Chapter 10 discusses the factors affecting the selection of an earth retaining system from the design
engineer's viewpoint. A selection process is discussed and a selection matrix is proposed to provide a
qualitative evaluation and rating of the most preferred alternatives. General recommendations for selecting
earth retaining structures are then presented. Case studies are provided documenting the selection, design
and construction of a variety of earth retaining systems, with special emphasis placed on wall selection.

Chapter 11 provides technical guidance for issues relating to contracting methods and bidding documents.
Typical contracting approaches and key plans and specifications items are presented.

1-6
I.S PRIMARY REFERENCES

A detailed list of references is provided in Chapter 12. However, several AASHTO and FHWA references
were utilized as primary references in the preparation of this manual. Following is a listing of these
primary references for this manual.

AASHTO. (1992). Standard Specifications for Highway Bridges, 15th Edition, American Association
of State Highway and Transportation Officials, Washington, D.C.

Byrne, R.I., Cotton, D., Porterfield, J., Wolschlag, C., and Ueblacker, G. (1996). "Manual for design
and construction monitoring of soil nail walls." Federal Highway Administration, FHWA-SA-96-
069.

Cheney, R. S. (1988). "Permanent ground anchors." FHWA-DP-68-1R, Federal Highway


Administration, Washington, D. C.

Christopher, B. R., Gill, S. A., Giroud, J. P., Juran, I., Mitchell, J. K., Schlosser, F., and Dunnicliff,
J. (1990). "Design and construction guidelines for reinforced soil structure - Volume 1." FHWA-
RD-89-043, Federal Highway Administration, Washington, D. C.

Clouterre. (1991). "Recommendations clouterre - soil nailing recommendations." French National


Research Project, FHWA-SA-93-026, Federal Highway Administration, Washington, D.C. (in
English).

Dash, U. (1987). "Pin-pile wall evaluation." FHWA-PA-86-047-84-36, Federal Highway Administration,


Washington, D.C.

Elias, V. (1996). "Corrosion/degradation of soil reinforcement for mechanically stabilized earth walls and
reinforced soil slopes." FHWA-SA-96-072, Federal Highway Administration, Washington, D.C.

Elias, V., and Christopher, B. R. (1996). "Mechanically stabilized earth walls and reinforced soil slopes
design and construction guidelines." FHWA-SA-96-071, Federal Highway Administration,
Washington, D. C.

Elias, V., and Juran, I. (1991). "Soil nailing for stabilization of highway slopes and excavations."
FHWA-RD-89-198, Federal Highway Administration, Washington, D. C.

Elias, V., Welsh, J., Warren, J., and Lukas, R. (1996). "Ground improvement technology manual."
Demonstration Project No. 82 - Ground Improvement, FHWA-DP-3, Federal Highway
Administration, Washington, D.C.

Federal Highway Administration. (1988). "Earth retaining structure - review and acceptance procedures."
Geotechnical Guideline No.2, FHWA Geotechnical Engineering Notebook, Washington, D. C.

Goldberg, D. T., Jaworski, W. E., and Gordon, M. D. (1976). "Lateral support systems and
underpinning." VoU Design and construction, April, FHWA-RD-75-128, Federal Highway
Administration, Washington, D. C.

1-7
Holtz, R. D., Christopher, B. R., and Berg, R. R. (1995). Geosynthetic design and construction guidelines,
May, FHWA- Hl-95-038, Federal Highway Administration, McLean, Va.

Porterfield, J. A., Cotton, D. M. and Byrne, R. J. (1994). "Soil nailing field inspectors manual." FHWA-
SA-93-068, Federal Highway Administration, Washington, D. C.

Sabatini, PJ., Elias, V., Schmertmann, G. R., and Bonaparte, R. (1996). "Earth retaining systems."
Geotechnical Engineering Circular No.2, FHWA-SA-96-038, Federal Highway Administration,
Washington, D. C.

Seismic Design ofHighway Bridges. (1986). (NHI Course No. 13048), FHWA/93/040, Federal Highway
Administration, Washington, D. C.

Weatherby, D. E. (1982). "Tiebacks." FHWA/RD-82/047, Federal Highway Administration, Washington,


D. C.

1- 8
CHAPTER 2.0
LATERAL EARTH PRESSURES

2.1 INTRODUCTION

This chapter describes the basic principles of lateral earth pressure used in the design of earth retaining
structures. Issues such as ground displacement, groundwater, compaction and surcharge loads, which
influence the value of the lateral earth pressure, are also discussed. The use of lateral earth pressures for
analysis and design of various types of earth retaining structures is discussed in the following chapters of
this manual.

2.2 GENERAL CONSIDERATIONS

The lateral, or horizontal pressure, Ph' acting on a retaining structure at any point is directly proportional
to the vertical pressure, avo, at that point as follows:

Ph = Ko vo (2-1)

where K is the coefficient of lateral pressure which is a function of the retained material, the type of
retaining structure and the lateral strain in the soil. When computed in terms of the effective overburden
0:
stress, 0 , the lateral earth pressure is as follows I :

(2-2)

The lateral strain in the soil will alter its lateral stress condition. Depending on the magnitude and
direction of the strain involved, the final lateral stress is bounded by two limiting (failure) conditions. The
limiting stresses occur at the active failure state when the wall moves away from the soil mass, and at the
passive failure state when the wall moves into the soil mass. The existing undisturbed state, corresponding
to zero lateral strain, is referred to as the at-rest state.

The lateral stresses corresponding to active, passive and at-rest states are referred to as the active. Pat
passive, Pp' and at-rest, Po' lateral earth pressures, respectively; the corresponding coefficients of lateral
pressure are denoted as Ka , ~ and Ko ' respectively, and are expressed in terms of effective stresses.

The effect of lateral strain in mobilizing the various states of stress in the soil is illustrated in Figure 2-1.
The magnitudes of wall rotation (YIH) required to achieve the limiting states of stress in various soil types
are presented in the figure. Figure 2-1 can also be used to estimate the state of stress for walls with
uniform horizontal translation equal to Y. As illustrated in this figure, significantly larger lateral
displacements are required to mobilize the passive resistance than those required to develop active
pressures. When the estimated wall movement is less than the value required to fully mobilize active or
passive pressure, the earth pressure coefficient can be adjusted proportionally based on the graphical
relationship presented in Figure 2-1.

The concept of effective stress is presented in Chapter 7 of Course Module I - Subsurface Investigations. In simplest
terms, the principle of effective stress states that the total stress (a) on any plane within a soil mass is equal to the
sum of the effective stress (a') and the pore pressure (u). In general, soil deposits below the water table will be
considered saturated and the ambient pore pressure at any depth may be computed by multiplying the unit weight of
water (y w ) by the height of water above that depth. The total stress at that depth may be calculated by multiplying
the total unit weight (y) of the soil by the depth. The effective stress is obtained as a' =a - u .

2- 1
10.0
8.0
6.0
~
5.0
vr
en
4.0

...~
en
3.0

';i 2.0
u
.€
~
...
0
1.0
...c
';i
0.8
0
N 0.6 Active State Passive State
'C
0 0.5
..c:
0.4
...
.2
~ 0.3
~
0.2

0.1
0.005 0.004 0.003 0.002 0.00 1 0 0.0 I0 0.020 0.030 0.040 0.050

Wall rotation, -.!..


H

Magnitude of Wall Rotation to Reach Failure

Soil type and Rotation, Y/H


condition Active Passive
Dense cohesionless 0.001 0.02
Loose cohesionless 0.004 0.06
Stiff cohesive 0.010 0.02
Soft cohesive 0.020 0.04

Figure 2-1: Effect of Wall Movement on Wall Pressures. (Canadian Geotechnical Manual, 1992)

2-2
2.3 AT-REST EARTH PRESSURE

The at-rest earth pressure represents the lateral effective stress that exists in a natural soil in its undisturbed
state. For normally consolidated soils, the coefficient of at-rest earth pressure, Ko ' can be approximated
by the equation (laky, 1944):

Ko = I-sin <1>/ (2-3)

where <1>/ is the effective (drained) friction angle of the soil at large strains (see Course Module 1 -
Subsurface Investigations). Ko is known to increase with the overconsolidation ratio, OCR, of the soil.
In overconsolidated soils, Ko can be determined as follows (Schmidt, 1966):

(2-4)

where Q is a dimensionless coefficient, which for most soils, can be taken as sin <1>/ (Mayne and Kulhawy,
1982).

Placement and compaction of backfill adjacent to rigid retaining structures may increase the lateral earth
pressure to at-rest conditions, or even higher. Therefore, walls which are rigid, or restrained against lateral
displacement, such as culvert box structures or V-wall structures, should be designed for at rest earth
pressures. Walls which may experience limited lateral displacement, such as pile-supported concrete
cantilever walls, are commonly designed for lateral earth pressures between the at-rest and active
conditions.

2.4 ACTIVE AND PASSIVE PRESSURES

In general, active and passive pressures are developed as a result of soil displacement within a failure zone
developed behind the wall. The failure zone is typically bounded by the back face of the wall and a failure
surface through the soil mass along which the soil has attained limiting equilibrium (Figure 2-2). In
addition to the effect of lateral movements as shown in Figure 2-1, the magnitude of the active and passive
earth pressures are also influenced by the characteristics of the wall, such as its geometry, the structural
stiffness and frictional properties of its back face; and the characteristics of the retained soil, such as its
geometry, properties, stratification and groundwater conditions.

Active and passive earth pressure coefficients based on a plane wedge theory, which considers the effect
of wall friction, sloping backfill and sloping wall face, were first proposed by Coulomb (1776) and are
shown in Figure 2-2. The pressures calculated using these coefficients are commonly known as the
Coulomb earth pressures. Since Coulomb's method is based on limit equilibrium of a wedge of soil, only
the magnitude and direction of the earth pressure is found. Pressure distributions and the location of the
resultant are assumed to be triangular.

For simple cases involving vertical walls retaining homogeneous soil with a level ground surface, without
friction between the soil and the wall face, and without the presence of groundwater, the formulas for
e
computing the earth pressure coefficients can be simplified considerably by substituting =0, ~ =0 and
0=0 in the Coulomb equations, as shown in Figure 2-2. For such simplified cases, I<a and I<.p can be
expressed as follows:

(2-5)

2-3
ACTIVE CASE
__ ~fJ:r

~..:-:'-:7."''''''Pf/A "/'. •

_--~:;:""-"""'7" .
....: "'~Failure Surface
z '.' .;' .
.". / : ' . y ,4>, c

.I .'. Active Pressure


H ; .... Diagram

H/3

2
K = cos (cI> - 9) ?

a cos 2 9cos(9 + 0)[1 + sin(cI> + o)sin(cI> - r3) ] ~


cos(9 + 0)cos(9 - r3)

PASSIVE CASE
.............. v . ./
;"c. .' •

~ .. ." .. '.":/ .
' '

.... . . ' . ." ./. .. '


z . . .../ ~
y ,4>, c / .' ... , Failure
. ' . ' / ... ' Surface
.. / '.
H : .. ': -;.. .' 'Passive Pressure
: ." ....: .. /. . . Diagram, pp=yzK p
. ./
-yoI-~

H/3

Pp
2
K = cos (9+cI» ?

p cos 2 9cos(9 _ 0)[1- sin(cI> + o)sin(cI> + r3) ] ~


cos(9 - 0)cos(9 - r3)

~ values are satisfactory for 0:$:<1>/3 but are unconservative for 0> <1>/3 and therefore should not be
used. Use Figure 2-5 when ~ = O. Use Figure 2-6 when 6 = O. Use above equation when both
6..0, ~ ..O.

Figure 2-2: Coefficients KA and ~ for Walls with Sloping Wall and Friction, and Sloping Backfill.
(Coulomb 1776; NAVFAC, 1982)

2-4
Kp = 1 + sin <t> = (2-6)
1 - sin <t>

These simplified equations were also derived independently by Rankine (1857). Hence, the earth pressures
computed using these equations are commonly known as the Rankine earth pressures.

Figure 2-3 illustrates how the Rankine earth pressures are computed in a frictional soil, a cohesive soil,
and a soil with combined frictional and cohesive (c - 4» characteristics.

Simple fonnulas are used for computing the earth pressures in granular soils (Figure 2-3a). In a combined
c - <I> material, a cohesive component contributes to these pressures (Figure 2-3b) and the pressures should
be computed in tenns of effective stresses. For purely cohesive soils, 4> = 0 is substituted in the latter
equations resulting in the equations shown in Figure 2-3c.

In cohesive soils, from the ground surface to a critical depth (z) where yz is less than 2c, the active earth
pressure behind the wall becomes negative (tension). In the design of the wall, however, the pressure
acting on the wall is assumed equal to zero from the ground surface to that critical depth. The depth Zo
of the tensile zone where tension cracks may develop is estimated as shown in Figures 2-3b and c.

2.4.1 Effect of Wall Friction

For walls with a rough face, the magnitude of the earth pressure is affected by the wall friction. Unless
a wall is settling, friction on its back acts upward on the active wedge (angle 0 is positive, see Figure 2-2),
reducing the active pressure. Generally, wall friction acts downward against the passive wedge (angle 0
is negative), resisting its upward movement and increasing passive pressure. The effect of wall friction on
the active pressure is small and ordinarily is disregarded except in the case of a settling wall where it can
be very significant. The effect of wall friction on the passive pressure is large but relative movement
between the wall and the adjacent soil is necessary for mobilization of wall friction. Table 2-1 presents
guidelines for estimating wall friction/adhesion in the absence of specific test data.

The wall friction also affects the shape of the failure surface. If the angle of wall friction (0) is low, the
failure surface is almost plane. For high values of 0, the failure surface is curved and can be approximated
by a log-spiral curve. The deviation of the curved surface from a plane surface is minor for the active case
and significant for the passive case as shown in Figure 2-4. For the passive case, a large angle of wall
friction causes downward tangentional shear forces to act on the passive wedge of soil adjacent to the wall,
increasing its resistance to upward movement. This, in tum, causes a curved failure surface to occur in
the soil, as shown in Figure 2-4b. The soil fails on this curved surface of least resistance and not on the
Coulomb plane, which would require greater lateral driving force. Hence, passive pressures computed
on the basis of the plane wedge theory are always higher than those calculated on the basis of a log-spiral
failure surface and may be on the unsafe side.

Based on the above discussion, it is recommended that the log-spiral theory be used for detennination of
the passive earth pressure coefficients. Charts for two common wall configurations (sloping wall with level
backfill and vertical wall with sloping backfill) based on the log-spiral theory (Caquot and Kerisel, 1948;
NAVFAC, 1982) are presented in Figures 2-5 and 2-6. For walls which have sloping backface and
sloping backfill, the passive pressure coefficient can be calculated as indicated in Figure 2-2.

For the active case, the resultant load predicted using coefficients based on the plane wedge theory is
within 10 percent of that obtained with the more exact log-spiral theory. Hence, for the active case,
Coulomb's theory (Figure 2-2) can be used.

2-5
Granular Soil Combined cohesive and Cohesive Soil (<I> = 0)
frictional soil (c-<I»

Active Pressures

n1 ~M=m.m ~:::~:-' \ t
\. If
.,
/ 'Y.c.~
I~ /~
~-
surface <// 4ailure
H I Pa / ~~ Pa /
surface

1\ Resultant

~
j,I /\45-~2 I \

\
Pa /

(c)
p. =Yz -2c
p. = YH 2/2 - 2 c H + 2c 2/y
Zo =2c/y

The active earth pressure within the tensile zone is assumed to be zero. Unless positive drainage
measures are provided, water infiltration into the tension crack may result in hydrostatic pressure on the
retaining structure.
Passive Pressures

-l' fMovement

~F~i1urc
~2c

U
.../)//
surface /
Pp /

H
./
X "' P
P

~
(l//\45+~2 """

(d) (e) (f)

Pp =K p yz P p =yz +2c
Pp =K p yH 2 /2 P p =yH 2 /2 +2cH

Figure 2-3: Computations of Active and Passive Pressures for Homogeneous Soils.

2-6
TABLE 2-1
ULTIMATE FRICTION FACTORS AND ADHESION FOR DISSIMILAR MATERIALS
(NAVFAC, 1982)

Interface Materials Friction Friction


Factor, tan 0 angle, 0
degrees
Mass concrete on the following foundation materials:
Clean sound rock 0.70 35
Clean gravel, gravel sand mixtures, coarse sand 0.55 to 0.60 29 to 31
Clean fine to medium sand, silty medium to coarse sand, silty or clayey gravel 0.45 to 0.55 24 to 29
Clean fine sand, silty or clayey fine to medium sand 0.35 to 0.45 19 to 24
Fine sandy silt, nonplastic silt 0.30 to 0.35 17 to 19
Very stiff and hard residual or preconsolidated clay 0.40 to 0.50 22 to 26
Medium stiff and stiff clay and silty clay (Masonry on foundation 0.30 to 0.35 17 to 19
materials has same friction factor)
Steel sheet piles against the following soils:
Clean gravel, gravel-sand mixtures, well-graded rock fill with spalls 0.40 22
Clean sand, silty sand-gravel mixtures, single size hard rock fill 0.30 17
Silty sand, gravel or sand mixed with silt or clay 0.25 14
Fine sandy silt, nonplastic silt 0.20 11
Formed concrete or concrete sheet piling against the following soils:
Clean gravel, gravel-sand mixture, well-graded rock fill with spalls 0.40 to 0.50 22 to 26
Clean sand, silty sand-gravel mixture, single size hard rock fill 0.30 to 0.40 17 to 22
Silty sand, gravel or sand mixed with silt or clay 0.30 17
Fine sandy silt, nonplastic silt 0.25 14
Various structural materials:
Masonry on masonry, igneous and metamorphic rocks:
Dressed soft rock on dressed soft rock 0.70 35
Dressed hard rock on dressed soft rock 0.65 33
Dressed hard rock on dressed hard rock 0.55 29
Masonry on wood (cross grain) 0.50 26
Steel on steel at sheet pile interlocks 0.30 17
Interface Materials (Cohesion) Adhesion c. (kPa)
Very soft cohesive soil (0 - 12 kPa) 0- 12
Soft cohesive soil (12 - 24 kPa) 12 - 24
Medium stiff cohesive soil (24 - 48 kPa) 24 - 36
Stiff cohesive soil (48 - 96 kPa) 36 - 45
Very stiff cohesive soil (96 - 192 kPa) 45 - 62

2-7
Straight Line

...-.-Log Spiral

(a)

Straight Line

,,
,,
,,

/"'
Log Spiral
............
....... --------

(b)

Figure 2-4: Comparison of Plane and Curved (Log-Spiral) Failure Surfaces (a) Active Case, (b)
Passive Case.

2-8
REDUCTION FACTOR (R) OF Kp
45° -4>/2 45~°-4>/2
FOR VARIOUS RATIOS OF - 0/4>
~. ,,/~/, <":'/ A~/'" ~ /
~ o/<fl -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0
10 .978 .962 .946 .929 .912 .898 .881 .864 11 +a . Failure Pp=Kr:t!f

P~
........ Surface>v'" 2
,/
15 .961 .934 .907 .881 .854 .830 .803 .775 ........
H ,/ Pn=Ppcoso
20 .939 .901 .862 .824 .787 .752 .716 .678 / ........
P,=Ppsino
\
25
30
.912 .860 .808 .759 .711 .666 .620 .574
.878 .811 .746 .686 .627 .574 .520 .467
/3
P.
Pp= KpyH
- ' - Logarithm ic
Spiral
35 .836 .752 .674 .603 .536 .475 .417 .362 Example: 4>=30°; 6=-10; 6/4>=-0.6
40 .783 .682 .592 .512 .439 .375 .316 .262 Kp=R(K p for 6/4>= -1)= (0.811)(8.2)=6.65
45 .718 .600 .500 .414 .339 .276 .221 .174
Note: Curves shown are for 6/4> =-1

14
I I I
'I II 11 7 1
13
12
,*j .4 I II 1.4
'I
<1Ji:~1
J .4
11
10 r); ~ 7q I II I
9
J ]I11~ fI J 'I

If Ifi/~ If 11
8

If
J I J; 'I II <1J
II
W) I
~

J I I ~ I ~I
I If If 1/ <1J
/j
/
J J j 1/ 1/
I If II If J ) /j~
/1 IIV j" JII i/ j
/) rJ / )' j ' V . ~. ,!;j
rr
Ii 'J I ) I "V ~ ~
~ ,.~ ~ 1I 1/ ~
II JI
~ 1I
1I
~ V ~ 1I II I.JII'1I
i
r7 [7 1.-1 V LI
V l/ ,,-
~ ~V

V ~ l/"
~, ~V
I." V
.~
~
~ ~
V ~ ~
",.
~
1
l/ -- .... ~
.8
o 10 20 30 40 45
Angle of Internal Friction, <1>, Degrees

Figure 2-5: Passive Coefficients for Sloping Wall with Wall Friction and Horizontal Backfill (Caquot
and Kerisel, 1948; NAVFAC, 1982)

2-9
REDUCTION FACTOR (R) OF ~ "" ~.... /
FOR VARIOUS RATIOS OF - 6/4> ~
_ / Failure
~o/<l> -0.7 -0.6 -0.5 -0.4 -0.3 -0.2 -0.1 0.0
.- ._ ) Surface

-'-
10 .978 .962 .946 .929 .912 .898 .881 .864
15 .961 .934 .907 .881 .854 .830 .803 .775 H .- / 90°-4>
20 .939 .901 .862 .824 .787 .752 .716 .678 ./ /
25 .912 .860 .808 .759 .711 .666 .620 .574 .....--Logarithm ic
30 .878 .811 .746 .686 .627 .574 .520 .467 .... Spiral
35 .836 .752 .674 .603 .536 .475 .417 .362 Pp= KpyH
40 .783 .682 .592 .512 .439 .375 .316 .262
45 .718 .600 .500 .414 .339 .276 .221 .174 Pp=K pyH 2J2; p.=Ppcoso; P.=Ppsino
Example: 4> = 25 ° ;~/4> = -0.2;0 14> = -0.3
Note: Curves shown are for 6/4> =-1 Kp= R(K p for 6/4> = -1) = (0.711 )(3 .62) = 2.58

., J I~ 11/41 = +.6
90 ~~=+I c;..
t-l 1 ,I Il/~ = +.4
80 ~~=+.8 '..14 'rI 'I
70
60
'IJ J J
50 J1/ If if Il/~ = +.2

If) II ) J
40
I II ., II
Il/~ = 0
Ifi / / ,. J IJ
0. 30
~
~ 20
) If~
~ If V LI
,.~
V
)
., )
1l/41 = -.2
:s
en
en
CI)
'"'
VIVV1I!/" II
J ~ ~~
l:l-.
CI) V 1I V
>
'Vi LI ~ ~ 11/41 = -.4
en 10 I'
~ 9 ~ V
~ ~ I I I~ 1-'
l:l-.
e- 8 '~
~ " ~
-"

-O
c::
CI)
7
6 "
~ 1/ If'
~ ~ V'~If' ~ l/
~
~
~ 1/
1/
~

l/
11/41 = -.6
'0 5
E /. ~ ~ ~ 1.....- 'f'" I;' ~
~

U
CI)
0
4

3
~.~

~ , i./ ~
'/
.".
~
i..' """
I""""
~
~ I""""

~41 = -.8
':.oIl~ " ~ ~""
2 ~ ~ ~ " ". io""'" " -~
~
""""
~:::: ~ ~ """""

t::
--- ---
~
10""'" -- --- ~~ ~41 = -.9

1.0
0.9
0.8
.~
"
0.7
0.6
o 10 20 30 40 45
Angle of Internal Friction, cp, Degrees

Figure 2-6: Passive Coefficients for Vertical Wall with Wall Friction and Sloping Backfill. (Caquot
and Kerisel, 1948; NAVFAC, 1982)

2 - 10
2.4.2 Earth Pressures in Stratified Soils

For stratified (or nonhomogeneous) soils, the earth pressures are assumed to be distributed as shown in
Figure 2-7. Unless the computed earth pressures vary widely with depth, the total applied lateral force
determined from the computed pressure diagram may be redistributed to a corresponding simplified
equivalent triangular pressure diagram as indicated in Figure 2-7.

For complex cases with layered soils, irregular backfill, irregular surcharges, wall friction, sloping
groundwater level, etc., pressures can be determined by graphical solutions. Among the many graphical
solutions are Culmann's (ca. 1866) and the trial wedge method (ca. 1877). These procedures can be found
in Bowles (1988) or NAVFAC (1982).

Pressure, Ph

LH.. . - ---__~
H G
z

Theoretical Earth Simplified (Design) Earth


Pressure Diagram Pressure Diagram

The lateral force is equal to the area of the pressure diagram. Thus,

Pal = Area ABC Pa2 = Area CBDE Pa3 = Area EFGH

Resultant (total) active force per unit length, Pa = Pal + Pa2 + Pa3
P L + P a2 L 2 + P a3 L 3
Location of resultant from base of wall, Z = _a_I- 1 - - - - - - -
Pa

• Use buoyant unit weight for soils below water table.

• Add water pressure as appropriate (see Figure 2-8) to obtain total lateral pressure.

• The simplified distribution may not be justified for all soil conditions. Use judgement to
determine validity of such simplified distributions.

Figure 2-7 Pressure Distribution for Stratified Soils.

2 - 11
2.4.3 Empirical Earth Pressures

The earth pressures discussed in the above sections are strictly applicable to rigid wall systems; i.e., walls
which move as a unit (rigid body rotation and/or translation) and do not experience bending deformations
(Figure 2-1). Most gravity walls can be considered rigid walls.

If a wall system undergoes bending deformations in addition to rigid body motion then such a wall system
is considered flexible. Virtually all wall systems, except the gravity walls, may be considered to be
flexible. The bending deformations result in a redistribution of the lateral pressures from the more flexible
to the stiffer portions of the system. Thus, in these walls the final distribution and magnitude of the lateral
earth pressure may be considerably different than those used for rigid walls. For example, the soldier-pile
and lagging walls with mutiple levels of support are usually designed using empirical earth pressure
distributions based on observed data. These empirical earth pressure distributions may vary from
rectangular to trapezoidal shapes and may have varying magnitudes depending on the soil type (see Figure
7-28 in Chapter 7).

Other factors which may influence the development of earth pressures are the type of construction (e.g.,
"bottom-up" or "top-down"), the wall support mechanism (e.g., tie-backs, struts, rakers, soil nails,
reinforcing elements, single or multiple levels of support, etc.), the geometry of the retained soil (e.g., silo
pressure), the superimposed or surcharge loads (e.g, strip, line, concentrated or equipment loads) and the
type of analysis (static or seismic). In addition, for cases of soil reinforced by inclusions such as MSE
walls or soil-nailed walls, different types of earth pressure distributions are used to evaluate the internal
and external stability of the wall system. The empirical earth pressure distributions are generally related
to the basic earth pressure coefficients ~,~ and Ko which are a function of the shear strength of the soil.

Specific empirical earth pressures used for various wall types are presented and discussed in detail in the
following chapters. The rest of this chapter is devoted to lateral pressures induced by groundwater,
surcharge loads, compaction equipment and silo conditions. Other lateral pressures from frost and swelling
action and jointed rock masses are briefly discussed at the end of the chapter.

2.5 EFFECT OF GROUNDWATER

In retaining wall design, it is general practice to provide drainage paths (commonly known as "weep
holes") through the earth retaining structure, or use other methods to drain groundwater that may otherwise
collect behind the structure, and thereby avoid the development of water pressure on the structure.
Occasionally, however, it may not be feasible nor desirable to drain the water from behind the structure
to safeguard against potential settlement of adjacent structures, or prevent contaminated groundwater from
entering the excavation. In such instances, the earth retaining structure must be designed for both lateral
earth pressure and water pressure.

Computation of lateral pressures for the case of a uniform backfill and static groundwater is illustrated in
Figure 2-8. In this case, the water pressure represents a hydrostatic condition since there is no seepage
or flow of water through the soil. The lateral earth pressure below the water level is based on the effective
. I stress, avo'
vertlca I

The lateral pressure due to water is added to the lateral earth pressure to obtain the total lateral pressure
on the wall. For water, the lateral pressure at any depth Zw is equal to the vertical pressure ywZw at that
point (K= 1 for water.). The lateral pressure computations should consider the greatest unbalanced water
head anticipated to act on the wall, since this generally results in the largest total lateral load.

2 - 12
z
H

• ...It---P a
= 1/2 K r
a
Z2

(No water Pw
level)

Figure 2-8: Computation of Lateral Pressures for Static Groundwater Case.

For cases where seepage may occur through or beneath the earth retaining structure, the resulting seepage
gradients will result in a reduction in the lateral water pressure. For such cases, flow net procedures can
be used to compute the lateral pressure distribution due to water (NAVFAC, 1982).

2.5.1 Example Problem 2-1

For the wall configuration shown in Figure 2-9 a, construct the lateral pressure diagram. Assume the face
of the wall to be smooth.

4> = 30° 3
Y = 18 kN/m "fable _
to d W atet
_-.- ..-- ---
G un
...1--.--
.-,.-

4> = 30°
Ysat = 19 kN 1m 3
+

2
27.21 kN 1m 2 (c) 39.24 kN/m
(a) (b)

Figure 2-9: (a) Problem Geometry and Soil Conditions, (b) Lateral Effective Earth Pressure Diagram,
(c) Lateral Water Pressure Diagram.

2 - 13
Solution:

From the equation for K. in Figure 2-2, K. = 0.374, 4> = 30°, P= 10°, e= 0, (, = o.
The pressures at various depths are tabulated below. The lateral pressure diagrams due to earth and water
are shown in Figure 2-9 band c, respectively.

Effective Lateral Earth Pressures


Z, m O'YO' kN/m 2 Lateral pressure, P., kN/m2

0 0 K. O'YO = 0 0

2 (18)(2) = 36 K. o'YO = 0.374(36) = 13.46

6 36 + (19-9.81)4 = 72.76 K. o'vn = 0.374(72.76) = 27.21

Hydrostatic Pressure, p,.


Z, m Zw' m 0w=z,.Yw' kN/m2 Lateral pressure, Pw' kN/m2

0 0 0 0

2 0 0 0

6 4 4 (9.81) = 39.24 K..,o", = 1.0(39.24) = 39.24

2.6 EFFECT OF SURFACE SURCHARGE LOADS

Surcharge loads on the backfill surface near an earth retaining structure cause increased lateral earth
pressures on the structure. Typical surcharge loadings may result from railroads, highways, sign/light
structures, electric/telecommunications towers, buildings, construction equipment, material stockpiles, etc.

The loading cases of particular interest in the determination of lateral earth pressures are:

• point loads,
• line loads parallel to the wall,
• strip loads parallel to the wall, and
• uniform surcharge.

Photo 2-1 shows examples of retaining walls with surcharge loads. Figure 2-10 presents procedures for
computing lateral pressures due to these surcharge loadings.

For a uniform surcharge load over an extensive area, the lateral pressure can be computed by treating the
surcharge as an equivalent height of backfill. Lateral pressures due to surcharge loads applied on a limited
area (point, line and strips loads), however, are generally computed using semi-empirical methods of
analysis developed from elastic theory and experiments on unyielding walls (Figure 2-10). The assumption
of an unyielding wall results in higher (approximately double) lateral pressures as compared to
corresponding values of lateral stress in an elastic half space. Thus, the assumption of an unyielding rigid
wall is conservative, and its applicability should be evaluated for each specific wall.

2 - 14
(a)

(b)

Photo 2-1: (a) Retaining Wall with Uniform Surcharge Load; (b) Retaining Wall with Line Loads (see
Railway Tracks) and Point Loads (see Catenary Structure).

2 - 15
Point Load Line Load
x ="'inH Qp _
For msO.4:
ForiiisO.4:
H2) 028ii2
Ph [Q; (016+ii2)3" H)_ 0200
=
c:l
III
Ph(Q; - (Ol6+ii2)2
Form> 0.4: H N t-~;;'.-Pb Ph =0. 55QI

H2]_ 1.77fti2ft2 For iii > 0.4:


Ph[Q; - (ffi2 +ii2)3 H)_ 128m2n
Ph [Q;- (M2 + ft2)2
Ph' = Ph cos2 (1.19) O.64Q
Resultant Ph = (ffi2 + l)
Pressures from Line Load Q,
(Bous~inesq Equation Modified by Experiment)

Pressures from Point Lold Q,


0.0
(Bouuinesq Equation Modified By Experiment)
0.0 r--r---,-.....,.-r---r---r---r---.,-.,..--, 0.1

0.1 0.2
0.2 0.3

-
:
N
II
0.3
0.4
0.5
-
:
N
II
IC
0.4
0.5
0.6
I
I
--:;-'--i--
IC 0.6 z I
!
i 0.7
.60H
0.7
0.8 .60H
0.8 .56H
0.9 0.9 ,4SH
1.0 I..-&-.&L:<.....a...............-'-..I....I...............~.........................~ 1.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0 0.0 0.2 0.4 0.6 0.8 1.0

Strip Load (pressure) Uniform Surcharge

'f?

,"~~ .... ~

\ ;'a (due 10 q)= qK

Ph = ~U3-sin~cos2a] K=K••Ko.K p
As Appropl1ale
ex and ~ in radians

Pressure from Strip Load Pressures from Uniform Surcharge


Solutions for point. line and strip loading are semi-empirical and based on an assumption of unyielding walls.

Figure 2-10: Lateral Pressure Due to Surcharge Loadings. (After USS Steel Sheet-Pile Manual, 1975)

2-16
2.6.1 Example Problem 2-2

Construct the lateral pressure diagram due to a line load of 10 kN/m located 5 m behind the top of a
10 m high unyielding wall shown in Figure 2-11a.

Solution:

From Figure 2-10

iii = 5/10 = 0.5 > 0.4

For iii > 0.4, the lateral pressure is given by:

P = 1 28 Q
_I J [ m- n
2- ]
h •
( H (iii 2 + ii 2r

For iii = 0.5, QI = 10 kN/m and H = 10 m, the lateral pressure is given by:

Ph = 1.28 [ 0.25n ]
(0.25 + nf 2

The computations for lateral pressures at various depths using the above formula are shown in Table 2-2.
Figure 2-13 presents the variation of the lateral pressure with depth for this example.

TABLE 2-2
COMPUTATION OF LATERAL PRESSURES DUE TO LINE LOAD
n = z/H ] Depth below top of wall (m) I Ph' (kPa)
o 0 0.00
0.1 1 0.47
0.2 2 0.76
0.3 3 0.83
0.4 4 0.76
0.5 5 0.64
0.6 6 0.52
0.7 7 0.41
0.8 8 0.32
0.9 9 0.26
1.0 10 0.21

2 - 17
Q1 = 10 kN/m

Sm

(a)

--- --- -.......


0 --..
-1 .....
............
-2
r---.... """'-
e "'" ~
0.>
Co)
-3
1/.
~
...
:s -4 ~
CIl
"0
c:
V
:s -5 ./
...
0 ~
Q/)

~
0 -6 /
]
oS
Q"
-7 /
0
4.l

-8 /
-9
j "
-10 V
o 0.2 0.4 0.6 0.8
Lateral Pressure, kPa

(b)

Figure 2-11: (a) Geometry of the Problem, (b) Variation of Lateral Pressure with Depth Due to Line
Load.

2 - 18
2.7 EARTH PRESSURES DUE TO COMPACTION

In walls supporting fill, compaction of the backfill can generate relatively high horizontal pressures on the
upper portion of the wall. It is customary, therefore, to reduce the compaction criteria for the zone
immediately behind the wall. However, if a relatively high level of compaction is needed in that zone to
reduce future ground deformations, the horizontal stresses generated by that compaction should be taken
into account in the wall design. The value and distribution of the compaction stresses are dependent on
the compaction equipment, procedures and applied energy, as well as, on the type and rigidity of the wall.
The following discussion illustrates a semi-empirical procedure developed for estimating the earth pressures
due to compaction immediately behind a rigid wall.

When compaction equipment moves over the backfill adjacent to a wall, it induces added earth pressures
on the waIl. The added pressure due to the equipment load can be estimated using procedures described
in Section 2.6. When the compaction equipment moves away, a portion of the added earth pressure
continues to act on the wall due to the inelastic behavior of the soil. These residual horizontal pressures
in a compacted soil mass can be considerably larger than the horizontal pressures in an uncompacted mass.

A typical distribution of the residual horizontal pressures due to the compaction process is shown in Figure
2-12. It can be seen that the residual earth pressure increases sharply with depth in the upper 1.5 m, but
the rate of increase is reduced at greater depths. At depths below 5.5 m, in this particular example, there
is no residual earth pressure due to compaction. Below about 5.5 m, the earth pressure is equal to the
normal earth pressure at rest.
Lateral Pressure After Compaction (kPa)
15 20 25 30 35 40 45 50 55 60 65 70 75

I
oJ:.
4
Q.
<>
Cl
5

Roller Length = 2.13 m


6 Distance from Wall = I SO mm
Lift Thickness = I SO mm
~= 35°
7
ROLLER COMPACTION
Static + Dynamic Load
q=
8 Roller Width

Figure 2-12: Typical Residual Earth Pressure after Compaction of Backfill Behind an Unyielding Wall.
(Clough and Duncan, 1991)

2 - 19
Figure 2-13 presents a chart and a table that can be used to estimate the magnitudes of these residual earth
pressures due to compaction by rollers. The linear pressure variations shown by the dash-dot lines
correspond to various values of earth pressures at-rest, as reflected by the value Koy. These dash-dot lines
represent the stresses in the soil prior to compaction. The horizontal earth pressures after compaction are
shown by solid lines for various values of line load intensity, q, which is computed as shown in the lower
left hand corner of the chart. For cohesive soils, the earth pressures due to compaction within
approximately 0.5 m depth are higher than those for cohesionless soils; this is identified by the dotted lines
in the upper portion of the chart. The final pressure diagram consists of the curved solid line from the
ground surface to the depth where it intersects the appropriate dash-dot line, followed by the dash-dot line
which represents the at-rest earth pressure, as shown in Figure 2-12. It is important to note that the
pressures estimated from the chart at any depth include the at-rest horizontal earth pressures.

For conditions other than those on which the chart is based, adjustments are made using the multiplication
factors given in the table shown in the lower part of Figure 2-13.

2.7.1 Example Problem 2-3

It is proposed to compact a backfill with a friction angle 4> of 30°, in lift thicknesses of 300 nun using a
vibratory roller. For the vibratory roller, the total (static plus dynamic) force exerted by the roller drum
is 128 leN, and the drum width is 1.68 m. Estimate the horizontal earth pressure due to compaction at a
depth of 1.2 m when the distance between the roller and the wall is 300 nun.

Solution:

Step 1: Compute line load intensity, q


q= 128/1.68=76.2 kN/m.

Step 2: Find Ph at a depth of 1.2 m for the above value of q

From Figure 2-13, at a depth of 1.2 m, for for q =76.2 leN/m we find Ph=27.5 kPa
Step 3: From the table in Figure 2-13, fmd the adjustment factors:

For t=3oo nun and x=300 nun, multiplier=0.84;


For w= 1.68 m, multiplier=0.98; and
For 4>=30°, multiplier =0.83.

Step 4: Compute the lateral pressures

The pressure is calculated by multiplying the value of Ph by the adjustment factors as follows:

Ph = (27.5)(0.84)(0.98)(0.83) = 18.8 kPa

Step 5: Construct the pressure diagram with depth

By repeating Steps 1 through 4 at various depths, a complete chart for a given case can be
developed, as shown in Figure 2-12.

2 - 20
Lateral Pressure After Compaction (kPa)
00 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75

6
RoUcrL~-2.13m
Dist:mce &om WaD -ISO 111m
7 Lift lbiclalcss - 1SO mm
+- 3S·
ROLLER. COMPAcnON
_ Sialic + Dynamic Load
8 q- Roller Wid!h

ADJUSTMENT (MULTIPLICATION) FACTORS


Multiplier Factors for z =
Variables 0.6m 1.2 m 2.4 m 4.8m
x = 0 1.00 1.00 1.00 1.00
t=150 x = 60mm 1.00 1.00 1.00 1.00
mm lifts x = 150mm 1.00 1.00 1.00 1.00
Lift thickness (t) and x = 300mm 0.87 0.88 0.89 0.90
distance from wall (x)
x = 0 0.94 0.95 0.95 0.96
(Adjustments for these two
t=3oo x = 60mm 0.94 0.95 0.95 0.96
factors are combined)
mm lifts x = 150mm 0.94 0.95 0.95 0.96
x = 300mm 0.83 0.84 0.86 0.88
Roller width (w) w = 0.38 m 0.80 0.80 0.80 0.88
w = 1.07 m 0.96 0.94 0.94 0.97
w = 2.13 m 1.00 1.00 1.00 1.00
w = 3.05 m 1.00 1.01 1.02 1.04
Friction angle (<<1» «1> = 25° 0.59 0.70 0.81 0.96
cI> = 30° 0.75 0.83 0.89 0.98
cI> = 35° 1.00 1.00 1.00 1.00
d> = 40° 1.23 1.16 1.10 1.03

Figure 2-13: Computation of Earth Pressures due to Compaction with Rollers. (Duncan, et aI., 1993)

2 - 21
It should be noted that the post compaction earth pressures estimated using Figure 2-13 apply to conditions
where the wall is stiff and unyielding. These pressures would provide a conservative estimate of pressures
on flexible walls or massive walls whose foundations conditions allow them to shift laterally or tilt away
from the backfill during compaction. Such movements would reduce the earth pressures. The reduction
would be expected to be less near the surface, where the compaction-induced loads tend to "follow" the
wall as it is deflected or yielded (Clough and Duncan, 1991).

2.7.2 Other Compaction Equipment

If other compaction equipment such as vibratory plates or rammer plates are used instead of rollers, the
reader is referred to the work by Duncan, et al. (1991, 1993) which presents charts and tables similar to
those presented above, but for the cases of vibratory and rammer plates.

2.8 SILO PRESSURE

Occasionally, retaining walls are built in front of a stable rock face or an existing wall, and granular fill
is placed between the new wall and the existing structure or the rock face (Figure 2-14). Due to the
proximity of the two structures, the fill is partially supported by friction between the soil and the confining
"walls" similar to the way grain or other granular materials act in a silo. Consequently, the vertical stress
in the fin and the horizontal stress acting on the wan are reduced. The lateral pressures in such wall
configurations are commonly referred to as "silo" pressures. The design of these walls using the earth
pressure coefficients presented in Figures 2-2 and 2-3 may. therefore, be overconservative particularly for
cases where the ratio of wall height to width of the fill zone is large. Figure 2-14 provides a procedure for
estimating the lateral "silo" pressures on retaining wans.

2.9 SEISMIC LATERAL EARTH PRESSURE

During an earthquake, the soil behind a retaining wan exerts a horizontal dynamic earth thrust that is
greater than the static force. The method most frequently used for the calculation of this dynamic earth
thrust is that developed by Okabe (1926) and Mononobe (1929). This pseudo-static approach and other
methods are discussed in detail in Module 9 - Geotechnical Earthquake Engineering.

2.10 OTHER LATERAL PRESSURES

In addition to the lateral pressures described in the above sections, earth retaining structures may be
subjected to lateral pressures due to frost action, swelling of soils or jointed rock masses. These are briefly
discussed below:

2.10.1 Frost (Ice) Action

Lateral pressures can be caused by the volume expansion of ice in fine grained soils (fine sand, silt and
clay). These lateral pressures are difficult to predict and may achieve high values. Since structures usuany
are not designed to withstand frost generated stresses, provisions should be made so that frost related
stresses will not develop behind the structure or be kept to a minimum. The use of one or more of the
following measures may be necessary:

• Isolate the backfin from underground sources of water either by providing a permeable drainage
system or an impervious barrier.

• Use pervious backfill and provide weep holes in the structure.

2 - 22
Use following equation to estimate Psh (Spangler
and Handy, 1984)

x
Psh yx
2 tano
[I -exp (-2K ~x tano)]
where,
Psh = horizontal silo pressure
z x = distance between the walls
z = depth at which Psh is calculated
K = coefficient of lateral earth pressure
H y = unit weight of the fill
0= angle of friction between the wall and
the fill

\
\
z Silo.'
I

Wall cross-section Typical "Silo" Pressure Distribution

• In the absence of specific test data, estimate the wall friction angle, 0, from Table 2-1.
• Use K = Ko = I-sin 4>' for walls with no-movement criterion where 4>1 is the backfill friction angle.
In this case, small variations in placement procedures such as localized compaction effects and
slight variations in density may induce significant variations from the estimated "silo"pressure.
A conservative approach could be to use a smaller ~ value in calculating K so as to obtain an upper
envelope to the expected pressure values (Frydman and Keissar, 1987).
• Use K=K.a values in the following table (after Frydman and Keissar, 1987) for walls expected to
mobilize active state. Under active conditions the estimated "silo" pressures are less sensitive to
small variations in placement procedures. Progressive failure, which occurs within the soil mass
adjacent to the wall during its movement, may result in a decrease in <1>', and this decreased value
should be used in estimating the pressure acting on the wall.

Values of Ksa
Wall friction Backfill Friction Angle, <1>' (degrees)
angle, 0 (deg) 28 30 32 34 36 38 40
0 0.3610 0.3333 0.3073 0.2827 0.2596 0.2379 0.2174
5 0.3632 0.3351 0.3086 0.2838 0.2605 0.2386 0.2180
10 0.3701 0.3406 0.3130 0.2873 0.2633 0.2408 0.2198
15 0.3838 0.3512 0.3213 0.2938 0.2684 0.2449 0.2230
20 0.4093 0.3701 0.3357 0.3049 0.2770 0.2515 0.2282
25 0.4664 0.4077 0.3621 0.3241 0.2913 0.2623 0.2364
30 - - 0.4265 0.3635 0.3179 0.2812 0.2501

Figure 2-14: Estimation of "Silo" Pressures.

2 - 23
• Provide an impervious soil layer near the ground surface, and grade the ground behind the wall
to drain surface water away from the wall.

2.10.2 Swelling Action

Expansive clays can cause very high lateral pressures on the back of a retaining structure. Clay backfills,
therefore, should be avoided whenever possible. In the cases where expansive clays are present behind
the wall, swelling pressures may be evaluated based on laboratory tests and the wall may be designed to
withstand these swelling pressures. Alternatively, one of the following measures can be taken:

• A granular filter material can be provided between the clay backfill and the back of the wall. This
material will drain the groundwater away from the expansive soil and, at the same time, act as a
buffer zone between the expansive soil and the structure.

• The expansive soil can be treated with lime to significantly reduce, or eliminate, its swelling
potential.

2.10.3 Loads from Jointed Rock Masses

Rock is not a continuum but a regulated discontinuum and most rock slope failures occur along
discontinuites or zones of weakness (Bell 1992). For structures retaining rock, it is important to know
the possible failure modes of the specific rock masses. The actual rupture or sliding plane behind a
retaining wall depends on the spatial orientation, frequency and distribution of the discontinuities, and the
involved shear and interlock resistance to shear among them (Jumikis 1983).

Depending on the type and structure of the rock, the lateral loads may be different. For example, if the
rock is extensively weathered, highly disintegrated, densely jointed, loosely fragmented or decomposed,
it behaves more or less like "soil"; in such cases the theories discussed in Sections 2.2 to 2.4 are suitable.
For other cases, the rupture or sliding plane behind a wall must be determined in order to calculate the
lateral pressure. The reader is referred to Brandl (1992) for detailed procedures for determining lateral
loads from jointed rock masses.

2 - 24
CHAPTER 3.0
GEOTECHNICAL PROPERTIES

3.1 INTRODUCTION

Prior to the design of a retaining wall, information must be obtained regarding the subsurface conditions
and the available backfill material (retained soil). For all cut walls, the properties of the undisturbed
natural soil deposits behind the proposed walls must also be investigated. The topic of subsurface
investigations is treated in Module 1 - Subsurface Investigations. This chapter gives guidance on the
evaluation of the soil and properties and the selection of geotechnical parameters to be used in the
determination of lateral earth pressures for analysis and design of earth retaining structures.

3.2 PROPERTIES OF NATURAL SOILS

The key soil properties required for the calculation of the lateral earth pressure (Chapter 2) are the shear
strength and unit weight of the retained soil. Typical properties for various in situ soils and backfill
materials are provided in Module 1 - Subsurface Investigations.

3.2.1 Shear Strength

The shear strength, "t, of the retained soil may be expressed as:

"t = c + a tan <I> (3-1)

in which c denotes the cohesion and <I> the angle of internal friction. As discussed in Module 1, several
laboratory testing methods are available for determining these shear strength parameters. Selection of the
appropriate test method, and the specific shear strength values to be used in the earth pressure calculations,
must consider a variety of factors including: the type of soil being tested, the drainage or permeability
characteristics of the soil, the type and rate of wall construction, the period of interest (Le., immediately
after construction, or long term), and the anticipated in situ stress conditions.

Total (Undrained) Strength

Generally, total (undrained) shear strength parameters are appropriate for saturated cohesive soils since
these soils have low permeability and would require a long period of time to dissipate excess pore water
pressure. For soils of intermediate permeability, such as silts and clayey sands, the undrained shear
strength provides a crude estimate of strength immediately after construction. The undrained shear
strength, "tu' is expressed in terms of the undrained cohesive strength, cu ' undrained friction angle, <l>u' and
the total stress, a, as follows:

oru =c u + a tan ~u
A- (3-2)

For most saturated cohesive soils, tested under quick undrained conditions, the undrained friction angle,
<l>u' is zero. The undrained shear strength is therefore equal to the undrained cohesive strength, cu ' at a
specific moisture content. This value of Cu is often normalized with respect to the vertical overburden
stress, a'v, prior to the application of the load, and is correlated with the plasticity index (PI) or the over
consolidation ratio (OCR) of the soil. The reader is referred to Chapter 9, Section 9.5.1 of Module 1 -
Subsurface Investigations for a more detailed coverage of the undrained shear strength of cohesive soils

3-1
wherein several correlations of Cu / 0' v with PI and OCR are also presented.

Effective (Drained) Strength

The effective or drained shear strength, t d , can be expressed in tenns of the drained cohesion, c', the
drained friction angle, 4>', and the effective stress, a', as follows:

t d = C, + a / tan 'fI
A-.' = c I + (a - u) tan 'fI
""
(3-3)

where u is the pore pressure in the cohesive soil. A long time after construction, the pore pressure, u, is
equal to the hydrostatic pressure, and can be estimated with reasonable accuracy. Hence, the effective
stress parameters (c' and 4>') are most commonly used to calculate the long-tenn lateral earth pressure in
cohesive soils.

For cohesionless soils (such as clean sand and gravel) the excess hydrostatic pore pressure may be taken
as zero. Hence the drained friction angle may be used for all periods during and following construction.

The reader is referred to Chapter 9, Section 9.5.2 of Module 1 - Subsurface Investigations, for a more
detailed coverage of the effective (drained) shear strength of cohesive soils wherein guidance is provided
for selection of the effective stress parameters (c' and 4>') .

3.2.2 Unit Weight

The unit weight of the soil, y, is used to estimate the vertical stress in the soil which, in tum, is used to
calculate the lateral earth pressure applied to the wall. The unit weight, y, of cohesive soils may be
determined from laboratory testing of undisturbed soil samples. Where the soil is granular in nature or
is lightly cohesive, disturbance in sampling renders laboratory detennination of unit weight unreliable.
In such cases, conservatively assumed values based on soil classification and correlation with field testing
data (e.g., SPT N-values) are usually used in the design. However, where accurate values of unit weight
are needed, special sampling techniques are available. Typical values of unit weights for various soils are
provided in Chapter 9 of Module 1 - Subsurface Investigations.

3.3 BACKFILL PROPERTIES

In general, clean, free-draining granular materials should be used for backfill. The shear strength of such
soils is relatively independent of the soil's moisture content. For economical reasons, however, locally-
available cohesive soils are sometimes used for backfill. The shear strength and consequently the lateral
earth pressure of these soils may fluc~ate widely with changes in the water content, such as in the cases
of flooding or heavy precipitation. In cold climates, frost action in poorly drained fine-grained soils may
increase the lateral earth pressure to many times the values commonly used in design (Penner, 1970).

3.3.1 Choice of Backfill Material

Cohesionless Soils

Soils classified as GW, GP, SW or SP in accordance with the Unified Soil Classification System (USeS)
are excellent backfill materials. The important characteristics of these soils are their high frictional
resistance and high penneability. With adequate drainage measures (such as weep holes, collector drains,
etc.) the build-up of pore water pressure in the backfill would be prevented.

3-2
Sandy Clays and Clayey Sands

Soils classified as SC, SM, GC or GM, according to the USCS, may be suitable as wall backfill if kept
dry, but are subject to frost action when wet. Such soils have low permeability and do not drain rapidly.
Hence, their water content may increase substantially as a result of precipitation or flooding. Also, these
soils cannot be properly compacted when wet. Where adequate drainage measures are provided, the
theoretical earth pressures discussed in Chapter 2 are valid for design purposes.

Silts and Clayey Silts

Soils classified as CL, MH, ML or OL, according to the USCS, are often subject to excessive frost action
and swelling when used as wall backfill, and the resulting wall movement is likely to be excessive.
Accordingly, the use of these materials should be avoided. If they must be used, however, frost protection
should be provided and an earth-pressure coefficient of 1.0 should be used in the design (Canadian
Geotechnical Society, 1992).

Lightweight Materials

Lightweight materials such as expanded shale may be used in place of a select soil backfill to control the
settlement of walls constructed over soft compressible soils. The dry loose unit weight of expanded shale,
for example, typically varies from 7 to 10 kN/m3 , while its compacted (Standard Proctor) dry unit weight
ranges from 11 to 13 kN/m3 . The internal friction angle for these materials may be 40° or higher, which
corresponds to lower active earth pressure coefficients. The lower earth pressure coefficient coupled with
lower unit weight results in considerably lower earth pressures and corresponding reduction in bending
moments in the wall structure and a reduction of the wall section. An example problem comparing the
earth pressures from lightweight fill with those from a regular fill is presented at the end of this chapter.

In addition to the above advantages, the gradation of lightweight aggregates can be easily controlled to
produce a free-draining backfill, thereby avoiding development of water pressures. Before using these
materials in construction, however, laboratory testing should be performed to define their shear strength,
unit weight, corrosivity and durability (soundness and resistance to abrasion). The lightweight aggregates
must have adequate hardness and durability to resist degradation during placement and compaction, and
to resist long-term deterioration in the underground environment. A more detailed coverage of lightweight
materials is included in Module 4 "Ground Improvement. "

The applicability of lightweight material may not be appropriate for all cases. For example, due to their
light weight such materials may not be suitable for mechanically stabilized earth walls which rely on the
overburden weight to generate friction along the reinforcing elements within the backfill material.

Flowable Fill

A variation of the lighweight materials is the flowable fill. This backfill material is composed of cement
in which air voids are distributed in the form of small, homogeneous, non-interconnected foam cells. High
flowability and pumpability permits complete backfilling. In its hardened state, this type of backfill is
lightweight and stable. However, since it is in a fluid state initially, formwork is usually required.

3-3
3.3.2 Placement and Compaction

The backfill should be placed in layers of not more than 300 mm thick, and should be properly compacted
by rollers and/or tampers. The backfill immediately behind the structure is typically compacted with light
tamping equipment to avoid development of excessive lateral earth pressures or displacement of the wall
face. Depending on the type of wall and its use, a dry unit weight not less than 95 percent of the maximum
dry density achieved in a Standard Proctor test (AASHTO T99) is commonly specified. To obtain proper
compaction, the moisture content of the furnished backfill material should be controlled, typically within
±2 percent of the optimum moisture content determined from laboratory compaction tests. Fills which
are formed by uncontrolled dumping of the material are generally unsatisfactory, and typically lead to long
term settlement problems. During placement, moisture-density tests must be performed to verify that the
in-place backfill material meets the specified compaction requirements.

3.4 WALL FRICTION AND ADHESION

The angle of wall friction refers to the angle between the direction of the resultant force and the normal
to the wall face (Figure 3-1). While the value of the angle of wall friction, O. is often expressed as a
function of the angle of soil friction, <1>, it is not solely a property of the soil, but is also a function of the
wall facing material type and roughness.

Pa
o = Angle of
Wall Friction

Figure 3-1: Angle of Wall Friction.

The effect of wall friction on the magnitude of lateral earth pressures was discussed in Chapter 2. In
general, the wall friction depends not only on the wall characteristics and soil properties, but also on the
magnitude and relative direction of wall movement. In addition, since the relative soil-wall movement
is not constant, the wall friction may not be constant along the wall. Furthermore, the maximum wall
friction may not occur simultaneously with maximum shearing resistance along the failure surface within
the soil mass behind the wall. As a simplification in design, however, a constant value of 0 is assumed.
In the absence of specific data, Table 2-1 in Chapter 2 can be used to estimate 0 for various wall and soil
conditions.

Wall adhesion, ca' develops from cohesion in the soil. In the absence of specific data, Table 2-1 in Chapter
2 can be used to estimate the wall adhesion for a range of cohesive soils. However, in estimating the
adhesion force acting on the wall, an adjustment may be necessary to account for a potential tension crack
that may form in the upper region of the wall. The depth of the tension crack is typically a function of the
plasticity, shear strength and moisture condition of the retained soil, and can be estimated as shown in
Figure 2-3 in Chapter 2.

In designing walls which have a geotextile drainage layer on the back face or granular drainage material

3-4
between the wall and the backfill, care should be taken to properly evaluate the wall friction.
Manufacturer's literature may give guidance on the interface friction values of geotextile products. For
walls retaining soft cohesive soils or granular soils that will be subjected to significant vibration (e.g., walls
near railway tracks or machine foundations), 0 or ca should be assumed to be zero in the design.

For some wall types, such as a L-shaped cantilever retaining wall, the "interface" is within the retained
soils wherein the earth pressures are computed along a vertical soil plane passing through the heel of the
base slab (Figure 1-2). In such cases, there is soil-to-soil contact and the resultant may be oriented at the
angle of mobilized friction. The angle of mobilized friction depends on the factor of safety used for the
angle of internal friction. For these cases, it is generally conservatively assumed that the earth pressure
is parallel to the slope of the backfill.

3.5 OTHER SOIL PROPERTIES

Certain wall types, which derive their stability from metallic or geosynthetic inclusions in backfill or in
situ material, require additional criteria regarding soil properties. Such criteria may include electrical
resistivity, pH, sulfate content, chloride content, etc. These criteria are discussed in Chapter 6.

3.6 ROCK PROPERTIES

For structures retaining rock, it is important to know the possible failure modes of the specific rock masses.
As discussed in Chapter 2, the actual rupture or sliding plane behind a retaining wall depends on the spatial
orientation, frequency and distribution of the discontinuities, and the involved shear and interlock
resistance to shear among them (Jumikis 1983). All of these rock properties are necessary to properly
determine the lateral pressure on a retaining structure. For further detailed discussion of rock properties
the reader is referred to Module 5 - Rock Slopes, and Bell (1992).

3.7 EXAMPLE PROBLEM 3-1

Consider a 5 m high wall shown in Figure 3-2a. Compare the active earth pressure resulting from a
regular fill with that from a lightweight fill. The fill properties are shown in the figure.

For Regular Fill


y = 19 kN/m 3
e <1>= 30°
11'\
II
::c For Lightweight Fill
y = 12 kN/m 3 78.38 kN 1~-+-.....;33 kN
<I> = 40° ~=====~

31.35 kPa 13.2 kPa

(a) (b) Regular fill (c) Lightweight fill

Figure 3-2: Geometry, Properties and Earth Pressures for Example Problem 3-1

3-5
Solution:

Step 1. Compute coefficient of active earth pressure, 1(,., using Equation 2-5.

Ka = tan2 (45 - 4>/2) = 0.33 for regular fill


= 0.22 for lightweight fill

Step 2. Compute the lateral pressure at base of wall, Pa

Pa = Ka YZ = 31.35 kPa for regular fill


= 13.20 kPa for lightweight fill
Step 3. Compute the active earth force, Pa

Pa = In Ka yH 2 = 78.38 kN for regular fill = PaR


= 33.00 kN for lightweight fill = PaL

Step 4. Compute percent reduction in total lateral force


PaR - PaL
Percent reduction in total lateral force: x 100% = 58%
PaR

3-6
CHAPTER 4.0
CAST-IN-PLACE (CIP) GRAVITY AND SEMI-GRAVITY WALLS

4.1 INTRODUCTION

This chapter discusses the design and construction of Cast-In-Place (CIP) concrete gravity walls (also
sometimes identified as "rigid gravity walls". or "mass gravity walls") and CIP concrete semi-gravity walls
(also identified as "conventional CIP walls", or "concrete cantilever walls").

CIP gravity walls have been used as low-height retaining walls for both roadway cut and fill applications.
CIP gravity walls are generally trapezoidal in shape (Figure 4-1) and are generally constructed of
unreinforced or minimally reinforced mass concrete. This type of wall is also occasionally constructed of
stone masonry. CIP gravity walls are rigid type walls that rely entirely on their self-weight to resist
overturning and sliding, and are generally proportioned to avoid any tensile stresses within the structure.
Considering the volume of material required for this type of wall, and the related construction cost, CIP
gravity walls are usually limited to a maximum height of about 3 m. The base width of these walls
generally range from 0.5 to 0.7 times the wall height.

CIP semi-gravity walls are commonly used for earth retaining structures and bridge abutments in fill
situations. They can also be used in cut situations, but for such applications a temporary support system
is typically required. In addition to its own weight, this type of wall uses bending action to resist lateral
earth pressures. The CIP semi-gravity walls include the cantilever, counterfort and buttress walls illustrated
in Figure 4-2. These, and other types of CIP semi-gravity walls, are discussed in Section 4.2.

At sites underlain by competent soils, the base of the CIP wall can be designed as a spread footing bearing
directly on the foundation soils. At locations where foundation bearing capacity or settlement is a concern,
the CIP wall can be provided with a deep foundation, using the wall base as the pile cap. Procedures for
analysis and design of spread footings and deep foundations are presented in Module 7 - Shallow
Foundations and Module 8 - Deep Foundations, respectively.

Since there are substantial labor and material costs associated with CIP walls they may not be economical
in comparison with alternative types of walls, particularly for wall heights more than about 2 m for CIP
gravity walls or 5 m for CIP semi-gravity walls. They also may not be economical for walls in cut
applications. In evaluating the use of CIP walls, therefore, consideration should also be given to other wall
alternatives as discussed in the following chapters of this manual.

The primary advantages of CIP walls are:

• Control of all construction materials

• Well-established design procedures

• Compatibility with deep foundations

• Minimal wall movements with deep foundations

• Resistance to degradation (corrosion)

4- 1
/
/
Min. /
Batter /
H
IH:48V /
/
/
/
_------- --'/
......
I. O.SH to O.7H .1
Figure 4-1: CIP Gravity Wall

Back face

Batter
Key between successive
concrete pours for high
walls

Heel

Base slab, footing or pile cap


(a)

Front
Face

Base slab, footing


or pile cap

(b) (c)
Figure 4-2: Cast-In-Place (CIP) Concrete Retaining Walls and Terminology. (a) Cantilever Wall
(Bowles, 1988); (b) Counterfort Wall; and (c) Buttress Wall. (Teng, 1962)

4-2
However, the use of CIP walls has major disadvantages, including:

• Construction is labor intensive

• Relatively long construction time is required

• A bottom-up sequence of construction is followed:


Temporary excavation support may be necessary
Increased quantity of excavation
Increased quantity of backfill
Wider work area required; this may be a significant problem at urban sites and on steep
hillsides

• Low tolerance to differential settlement

• Higher cost

4.2 TYPES OF SEMI-GRAVITY CIP WALLS

4.2.1 Cantilever Walls

Cantilever walls are constructed in the form of an inverted T wherein the projecting members act as
cantilever elements. Typical proportions of cantilever walls are shown in Figure 4-3. This type of wall
is generally suitable for heights up to about 9 m; higher walls may require an excessively dense
arrangement of reinforcing bars at the base of the stem wall. Photo 4-1 shows an example of a CIP
cantilever retaining wall.

1
Min. Batter
(IH:48V)

H
Note: The footing level should
be below depth of seasonal
volume change and frost line
0.7m min.
.H. to H.
12 10

.1
Figure 4-3: Common Proportions of Cantilever Walls. (Adapted from Teng, 1962)

4-3
Photo 4-1 : CIP Cantilever Retaining Wall

4.2.2 Counterfort Walls

Counterfort walls can be used for structures higher than about 9 m. This type of wall is a variation of the
cantilever wall wherein both the base slab and wall face span horizontally between vertical brackets known
as counterforts. The counterforts increase wall stability allowing the stem thickness to be reduced
substantially without excessive outward deflection.

The proportions of counterfort walls, Figure 4-4, vary to a greater extent than those for cantilever walls
because the thicknesses of the face and base slab depend primarily on the spacing of the counterforts. For
walls with a height of about 9 m, the counterforts may be spaced as far as two-thirds of the height of the
wall. As wall heights increase, the spacing of the counterforts is reduced, and may be as little as one-third
of the wall height. For constructibility considerations, counterforts should not be placed on a spacing less
than about 2.5 m. The toe projection is generally smaller than that for cantilever walls. Photo 4-2 shows
the construction of a conterfort retaining wall used in a cut application.

4.2.3 Buttress Walls

The buttressed wall illustrated in Figure 4-2(c) is another type of CIP semi-gravity wall. It is similar to
the counterfort wall except that the vertical brackets (known as buttresses) are on the outside of the stem
wall, and act in compression rather than tension. This type of wall is not commonly used for retaining
walls because of the exposed buttresses.

4-4
200 mm min ----j
(300 mm preferable) I

IH:48V min
batter 200 mm min.

0.7 m min. Note: The footing level should


be below depth of seasonal
volume change and frost line
II
12

Figure 4-4: Common Proportions of Counterfort Walls. (Adapted from Teng, 1962)

Photo 4-2: CIP Counterfort Wall Construction in a Cut Application

4-5
4.2.4 Other CIP Semi-Gravity Walls

Other types of CIP semi-gravity walls which may be used for special applications are illustrated in Figure
4-5. The cantilever walls shown in Figures 4-5(a) and (b) could be considered for cut and fill applications,
respectively, at sites with tight Right-of-Way restrictions. The U-walls shown in Figure 4-5(c) and Photo
4-3 are often used for construction of depressed roadways, particularly where the clear distance between
the walls is less than about 10m, and for depressed roadways constructed below the groundwater level.

Unless supported on a deep foundation, the cantilever walls shown in Figures 4-5(a) and (b) are structurally
less efficient than the conventional cantilever wall shown in Figure 4-2(a), and typically would be limited
to a lower height than a conventional cantilever wall. The maximum height of a V-wall, however, would
be comparable to that of a conventional cantilever wall.

ROW ROW
I
I
I
I
I
I
I Proposed Road Grade
I

(a) (b)

ROW ROW

Proposed
/RoadGrade

(c)

Figure 4-5: Other Types of Cast-In-Place (CIP) Walls: (a) Fill Wall with Limited ROW, (b) Cut Wall with
Limited ROW, and (c) V-Wall for Depressed Roadway

4-6
Photo 4-3: CIP V-Wall for Depressed Roadway

4.3 WALL DESIGN

The stability of CIP walls founded on a spread footing foundation is evaluated by analyzing the potential
for overturning about the toe, sliding along the base and global stability. In addition, the foundation
bearing capacity and settlement, along with the corresponding wall tilt, must also be evaluated. The
criteria for analysis and design of CIP walls are presented below.

4.3.1 Design Earth Pressures

Lateral earth pressures for design of CIP walls are determined using the procedures presented in Chapter
2. Generally, Coulomb theory is used to compute earth pressures either directly on the back face of the
wall (gravity wall case) or on a vertical plane passing through the heel of the base slab (semi-gravity wall
case) as illustrated in Figure 4-6. Conventionally, it is assumed that the angle of wall friction, 0, is equal
to the backslope angle, p, for semi-gravity walls. For gravity walls, the angle of wall friction can be
determined using Table 2-1 based on the anticipated type of backfill to be used and the condition of the
back face of the wall (i.e., stepped or inclined).

If adequate drainage measures are provided (Section 4.3.6), the hydrostatic pressure due to groundwater
behind the wall generally need not be considered; however, hydrostatic pressure must be considered for
portions of wall below the level of weep holes unless a deeper drainage pipe is provided behind the base
of the wall. When it is a necessary to maintain the groundwater level behind the wall, the wall must be
designed for the full hydrostatic pressure.

4-7
LOAD DIAGRAM DESIGN FACTORS
DEFINITIONS
GRA VITY WALL B = Width of the base of the footing
Tan 0b = Friction factor between soil and base
W =
Includes weight of wall for gravity walls.
Includes weight of wall and soil above footing
for cantilever, counterfort, and buttress walls.
c = cohesion of foundation soil
c. = Adhesion-concrete on soil
o = Angle of wall friction
Pp = Passive resistance
LOCATION OF RESULTANT
Moments about toe (assuming Pp = 0):
Wa+P yg -Phb B
d Eccentricity e = d

.\ W+P y 2

OVERTURNING
Moments about toe:
CANTILEVER WALL
Wa ~ 2.0 for footings on soils
F s P b _p g S 1.5 for footings on rock
II y

Ignore overturning if R is within middle third of footing


on soil and middle half of footing on rock.

;r
For very rigid walls, e.g., U-walls, soil pressure may be
more than active; use at-rest pressure.
b
P
Heel of
Slab RESISTANCE AGAINST SLIDING
Base of
footing (W +P )Tanob+c B
Fs y • ~ 1.5
Ph
I" For coefficients of friction between base and soil see
1----------------------1 Chapter 2.
COUNTERFORT WALL
CONTACT PRESSURE ON FOUNDATION
Compute maximum (qmax) and minimum (qmin)
pressures at base of the footing as follows:
q
max
= w +Pv (1 +~)
B B
q.
mIn
= w +Pv (1 _~)
B B
For allowable bearing pressure for inclined load on strip
foundation, see Module 7. For analysis of deep loads
fA beneath strip foundation, see Module 8. For equivalent
uniform bearing pressure:
W+P v
CIcq=~
Section A-A where, B' = B - 2e
~I
GLOBAL ROTATION STABILITY
For analysis of global stability, see Module 3.
Figure 4-6: Design Criteria for Cast-In-Place (CIP) Concrete Retaining Walls. (After AASHTO,
NAVFAC, 1982)

4-8
Generally, CIP walls with spread footing foundations are designed for active earth pressure since a small
amount of wall rotation or displacement will occur. However, higher lateral earth pressures are
appropriate for unyielding walls such as culverts, tunnels and rigid abutment V-walls (Photo 4-4), where
wall rotation and displacement are restrained. For such cases, the applied lateral earth pressure should be
determined by assuming at-rest conditions using procedures discussed in Chapter 2. For walls with deep
foundations where some small lateral movement is anticipated, common practice is to design for an average
of active and at-rest earth pressure.

In addition to the lateral earth pressure, the wall must be designed for lateral pressure due to surcharge
loads (see Chapter 2). For stability analyses of CIP gravity walls, the surcharge loads are generally
assumed to be applied starting directly behind the top of the wall, unless otherwise considered appropriate.
For CIP semi-gravity walls, the surcharge loads are generally assumed to be located behind the heel of the
wall, and conservatively neglected within the width of the base slab; however, for design of the wall stem,
the surcharge loads within the width of the base slab must be considered.

The resistance due to passive earth pressure in front of the wall is neglected unless the wall extends well
below the depth of frost penetration, scour or other types of disturbance (e.g. utility trench excavation in
front of the wall). Development of the passive earth pressure in the soil in front of the wall Fequires a
rotation or outward displacement of the wall; accordingly, the passive earth pressure is neglected for walls
with deep foundations, and for other cases where the wall is restrained from rotation or displacement.

4.3.2 Analysis of Sliding and Overturning

Figure 4-6 presents criteria for design of CIP walls against sliding and overturning. The dimensions of
a CIP wall are determined by satisfying the following factor of safety (FS) criteria:

Sliding: FS ~ 1.5

Overturning: FS ~ 2.0 for footings on soil


FS ~ 1.5 for footings on rock

The location of the bearing pressure resultant, R, on the base of the wall footing must be within B/6 of the
center of the footing for foundations on soil, and within B/4 of the center of the footing for foundations
on rock, where B is the width of the footing.

For seismic loading, AASHTO (1997) allows reduction in the factors of safety against sliding and
overturning to 75 percent of the factors of safety listed above.

If there is inadequate resistance to overturning, consideration should be given to either increasing the width
of the wall base, or providing a deep foundation. If insufficient sliding resistance is available,
consideration should be given to increasing the base width, providing a deep foundation, or lowering the
base of the wall. If the wall is supported by rock, granular soils or stiff clay, a key may be installed below
the foundation to provide additional resistance to sliding (Figure 4-7).

4.3.3 Global Stability Analyses

Where retaining walls are underlain by weak soils (Figure 4-8), the overall stability of the soil mass must
be checked with respect to the most critical failure surface (see Module 3 - Soil Slopes and Embankment).
A minimum factor of safety of 1.5 is desirable. If global stability is found to be a problem, deep
foundations or use of lightweight backfill may be considered, or measures can be taken to improve the

4-9
Photo 4-4: eIP Abutment with Integral Wingwalls

Note: See Figure 4-6 for list of symbol


Ph definitions.

Cohesive Soils:
F = (W+PV)Tanob+ca(B-alb)+c(alb)+Pp

Granular Soils:
F = (W +P)TanOb+Pp

.1 Factor of Safety FS = F/P h

Figure 4-7: Resistance Against Sliding from Keyed Foundation

~----+/-:-'/'- Wall rotates


SOil bulges
/ bac~ard
; here
~-rr- Segment rotates-L
"\ ?;~?JA~ ~

, , \ ,.. Shallow
Deep
failure -
, . . . . . -.failure

Soft material with


low shear strength

Figure 4-8: Typical Modes of Global Instability. (Adapted from Bowles, 1988)

4 -10
shear strength of the weak soil stratum (see Module 4 - Ground Improvement Techniques). Alternative
wall types, such as an anchored soldier pile and lagging wall, or tangent or secant pile wall, should also
be considered in this case.

4.3.4 Bearing Capacity

The computed vertical pressure at the base of the wall footing must be checked against the ultimate bearing
capacity of the soil. The generalized distribution of the bearing pressure at the wall base is illustrated in
Figure 4-6. Note that the bearing pressure at the toe is higher than that at the heel. The magnitude and
distribution of these pressures are computed using the applied loads shown in Figure 4-6 and discussed in
Chapter 2. The procedures for determining the allowable bearing capacity of the foundation soils are
discussed in Module 7 - Shallow Foundations. Generally, a minimum factor of safety against bearing
capacity failure of 3.0 is required for the spread footing foundation.

For seismic loading, AASHTO (1997) allows a reduction in the factor of safety against bearing capacity
failure to 1.5 for footings on soil or rock.

CIP walls founded on a deep foundation may be subject to potentially damaging ground and structure
displacements at sites underlain by cohesive soils. This may occur if the weight of the backfill material
exceeds the bearing capacity of the cohesive subsoils, causing plastic displacement of the ground beneath
the retaining structure and heave of the ground surface in front of the wall. When the cohesive soil layer
is located at and below the base of the wall, the factor of safety against this type of bearing capacity failure
can be approximated by the following equation (Peck, et.al., 1974):

FS ::::
Sc
(4-1)
(yH + q)

where H = the height of the fill


Y = unit weight of fill
c = shear strength of cohesive soil
q = uniform surcharge load

The computed factor of safety should not be less than 2.0 for the embankment loading. Below this value
progressive lateral movements of the retaining structure are likely to occur (Peck, et al., 1974). As the
factor of safety decreases, the rate of movement will increase until failure occurs at a factor of safety of
unity. For CIP walls founded on vertical piles or drilled shafts, this progressive ground movement would
be reflected by an outward displacement of the wall. CIP walls founded on battered piles typically
experience an outward displacement of the wall base and a backward tilt of the wall face (Figure 4-9).

4.3.5 Settlement and Tilt

Since the vertical pressure distribution under the wall footing is not uniform, tilting of the wall may occur
due to differential settlement. Foundation settlement can be computed by methods discussed in Module
7 - Shallow Foundations, to estimate the settlement and tilt of the wall. CIP walls can generally
accommodate a differential settlement of up to about 1/500 (ratio of differential settlement of two points
along the wall to the horizontal distance between the points). If the computed settlement and tilt exceed
acceptable limits, the wall dimensions can be modified to shift the resultant force closer to the center of
the base and thereby reduce the load eccentricity and differential settlement. In some cases, use of
lightweight backfill material may solve the problem. The use of deep foundations can also be considered.

4 - 11
I

Displaced "
I

Wall -------.;

Clay

Figure 4-9: Typical Movement of Pile Supported Cast-In-Place (CIP) Wall with Soft Foundation

Unless CIP walls are provided with a deep foundation, a small amount of wall tilting should be anticipated.
It is therefore advisable to provide the face of the wall with a small batter to compensate for the forward
tilting. Otherwise, a small amount of forward tilting may give the illusion of instability of the wall.
Figures 4-1, 4-3 and 4-4 provide recommended minimum batter values for CIP walls.

In unusual cases, where the foundation materials are stiffer or firmer at the toe of the base than at the heel,
the resulting settlement may cause the wall to rotate backwards, towards the retained soil. Such wall
movements could substantially increase the lateral pressures on the wall since the wall is now pushing
against the soil similar to a passive pressure condition. This case can be avoided by reproportioning the
wall, supporting the wall on a deep foundation, or treating the foundation soils.

4.3.6 Drainage

It is preferable to provide backfill drainage rather than design the wall for the large hydrostatic water
pressure resulting from a saturated backfill. Saturation of the backfill may result from either a high static
water table, or from rainfall or other wet conditions.

Drainage measures typically consist of the use of a free-draining material at the back face (or within the
entire active zone), with "weep holes" and/or longitudinal collector drains (perforated pipes or gravel
drains) along the back face as shown in Figure 4-10. In addition, geocomposite drains, consisting of a
dimpled sheet of plastic with a geotextile filter cover layer, can be placed against the back face of the wall
to collect seepage and channel it to the weep holes or collector pipes. Weep holes are openings (about 75
to 100 mm in diameter) provided through the wall at periodic intervals to discharge the accumulated water
from the backfill. If weep holes are used with a counterfort wall, at least one weep hole should be located
between counterforts. Where the base of the wall is embedded well into the ground it may only be
practical to locate the weep holes close to the ground surface in front of the wall to drain the backfill to
that level; below this level, it may be necessary to design for hydrostatic water pressure unless a collector
drain is provided behind the wall base. The flow through the collector drain can be discharged at the end
of the wall, or by discharge pipes placed through or beneath the wall.

To guard against the possibility of backfill material washing into the weep holes or drainage pipes and
ultimately clogging them, a filter material in the form of a geotextile or graded filter system is typically
placed behind the weep holes and around the drainage pipes.

4 - 12
Backfill Soil

Prefabricated
Drainage
Element

Figure 4-10: Typical Retaining Wall Drainage Alternatives. (Sabatini, et. al., 1996»

CIP walls designed to maintain the groundwater level behind the wall are typically provided with a
geomembrane or prefabricated clay panels installed on the back face of the wall to prevent undesirable
seepage through the wall.

4.4 WALL CONSTRUCTION

Following is a typical sequence of construction for CIP concrete walls.

• The first stage of construction consists of excavating to the wall foundation grade and preparing
the wall foundation. If stable excavation slopes cannot be maintained during the period of wall
construction, temporary excavation support needs to be provided. Foundation preparation includes
removing unsuitable materials such as organic matter and vegetation from the area to be occupied
by the wall and leveling and proof-rolling the foundation area. Since CIP walls cannot tolerate
significant differential settlement, preparation of the wall foundation to provide a relatively stiff
and uniform bearing surface is particularly important. For walls founded on compressible soils,
foundation preparation may require ground improvement (see Module 4 - Ground Improvement
Techniques) to increase bearing capacity and stiffness, or the construction of deep foundations (see
Module 8 - Deep Foundations) for wall footing support.

• For CIP semi-gravity walls, the footing outline is formed and the reinforcing steel for the footing
is placed and extended into the wall stem. For a counterfort (or buttress) wall, reinforcing steel
is also extended into the counterforts (or buttresses) (Photo 4-2). The footing concrete is then
poured.

• The wall stem is formed, weep hole inserts placed, and concrete poured. For counterfort (or
buttress) walls, reinforcing steel for the counterforts (or buttresses) is placed and the counterforts
(or buttresses) are then formed and the concrete is poured. Typically, concrete is poured in
sections between vertical expansion joints and, wherever possible, it is poured for the full wall
height to eliminate cold joints.

• Drainage systems are then constructed behind the wall. Concurrent with this activity is the
placement and compaction of select fill to finished grade. Care should be taken to ensure that the
backfill soils are not overcompacted just behind the wall face. Overcompaction can induce large
lateral earth pressures which may over stress the wall and, if prefabricated drainage material is
constructed against the back wall face, damage the drainage components.

4 - 13
4.5 EXAMPLE PROBLEM 4-1

Statement

Analyze the elP cantilever wall shown in Figure 4-11 for factors of safety against sliding, overturning
and bearing capacity failure. The backfill and foundation soils consist of clean, fine to medium sand, and
the groundwater table is well below the base of the wall.

E
0.5 m ", "1
Not to Scale
I
3 y=18 k~/m3
Yconc=23 kN/m I 0
<1>=30 •
I E
!CO V"\

~
V"\
I @
(i)
I Ph
I I
y=18 kN/m E I
<1>=36
0 "'!

E A B E
l"
0
~ l"
0

Figure 4-11 : Geometry and Parameters for Example Problem 4-11

Solution

Step 1: Detennine the total height of soil exerting pressure.

H = thickness of base slab + height of stem + (width of heel slab) tan (backslope angle)
H = 0.7 + 5.5 + 2.6 tan 10°
= 6.66 m
Step 2: Compute coefficient of active earth pressure (Figure 2-2), using vertical backface (8=0).

cos 2<1>
Ka = ---:-------'-------:~
Sin(<I>+O)Sin(<I>-p)] 2
cosO[1 +
cosOcos( -p)

where:
<t> = Internal friction angle of soil = 30°
~ = Angle of backfill slope = 10 0
o = Angle of wall friction = ~ = 10°

4 - 14
Substituting values, we obtain:

cos 230 0
Ka = ---------------
Sin(300+100)Sin(300-IOO)] 2
cos I0 0[ I +
cos I0 °cos( -10 0)

K a = 0.35

Step 3: Compute resultant of active pressure, Pa'

I
Pa = -K yH 2
2 a

= .!.(0.35)(18kN/m 3)(6.66m)2
2
= 149.3 kN/m

Resolving Pa into horizontal and vertical components, we get:

Ph = Pa cos P Pv = P sin P
a

= (149.3 kN/m) cos 10° = (149.3 kN/m) sin 10°


= 147.0 kN/m = 25.9 kN/m
Step 4: Determine weights and sum moments about wall toe A.

Moment arm of Ph about point A, b = (0.7+5.5+0.46)/3 = 6.66/3 = 2.22m


Moment arm of Pv about point A, g =4 m
The moments due to the weights of various areas shown in Figure 4-11 are set out in the following table.
Note that the weight of the soil above the footing toe has been neglected.

Area Weight kN/m Moment arm about A m Moment about A kN'm/m


1 (0.5) (5.5) (23) = 63.3 0.7 + 0.2 + (0.5/2) = 1.15 (63.3) (1.15) = 72.80
2 (0.5) (0.2) (5.5) (23) = 12.7 0.7 + (2/3) (0.2) = 0.83 (12.7) (0.83) = 10.54
3 (4) (0.7) (23) = 64.4 4/2 = 2.00 (64.4) (2.0) = 128.80
4 (2.6) (5.5) (18) = 257.4 0.7 + 0.7 +(2.6/2) = 2.70 (257.4) (2.7) = 694.98
5 (0.5) (2.6) (0.46) (18) = 10.8 0.7 +0. 7 +(2/3)(2.6) = 3.13 (10.8) (3.13) = 33.80
Total W = 408.6 Mw = 940.90

4 - 15
Step 5: Check factor of safety against sliding.

where: W = Weight of concrete and soil on AB


0b = Friction angle between concrete base and foundation soil

Use 0b = (3/4) <l>b = (3/4) (36°) = 27°, for friction angle between concrete and clean, fine to medium
sand (see Table 2-1).

= (408.6kN/m +25.9kN/m)tan27° = 221.4kN/m


= 1.51 O.K.
147.0kN/m 147.0kN/m

Step 6: Check factor of safety against overturning.

where: ~MR = Sum of resisting moment about point A = Mw


~Mo = Sum of overturning moment about point A

Substituting values:

940.9kN . m/m 940.9kN-m


FS o = - - - - - - - - - - - - - - = - - - - - = 4.22>2.0 O.K.
(147 .OkN/m)(2.22m) -(25.9kN/m)(4m) 222.7kN-m

Step 7: Check the factor of safety against bearing failure.

(a) Compute the location of resultant at distance d from point A.

= 940.9kN·m/m +(25.9kN/m)(4m)-(147.0kN/m)(2.22m)
408.6kN/m +25.9kN/m

where: ~V = W + Pv

d = 719.2kN . m/m
= 1.65 m
434.8kN/m

4 - 16
(b) Compute the eccentricity of load about center of base.

4m
e = ~-d = --1.65m
2 2

e = 0.35m<- =
B 4m
= O.67m O.K.
6 6

(c) Compute the maximum and minimum pressures under the waH footing.

q max.min = :EBV (1± 6 e)


B

= 434.8kN/m (1 ± 6(0.35m»)
4m 4m

= 108.7kN (1.525 or 0.475)

Le., ~x= 165.8 kN/m2

~in= 51.6 kN/m 2

(d) Estimate ultimate bearing capacity.

Using procedures presented in Module 7 - Shallow Foundations for footing with eccentric and inclined
loading the computed ultimate bearing capacity is:

qull = 1017.0 kN/m2


(e) Check factor of safety against bearing capacity failure.

1017.0kN/m 2
FS bc = = = 6.42 > 3.0 O.K.
158.5kN/m 2

SUMMARY

Factor of safety against sliding FSs = 1.60


Factor of safety against overturning FSo = 3.08

Factor of safety against bearing failure FSbc = 6.42

In addition, the factor of safety against global failure should be evaluated using methods presented in
Module 3 - Soil Slopes and Embankments, and wall settlement including tilting and lateral squeeze should
be evaluated using procedures presented in Module 7 - ShaHow Foundations.

4 - 17
CHAPTER 5.0
MODULAR GRAVITY WALLS

5.1 INTRODUCTION

This chapter describes the principal types of modular gravity wall systems, including the crib wall, the
concrete module wall, the bin wall and the gabion wall (Figure 5-1). Also presented are the specific design
considerations and construction requirements for each of these types of wall.

Modular gravity walls are composed of two major components: modular structural elements and the fill
material placed within these elements. The structural elements, which are often proprietary, may consist
of steel modules or bins, prefabricated concrete modules, timber units, wire baskets or other configurations
and materials. The fill material typically consists of free draining granular soils, gravel, or rock fragments.
Modular gravity walls rely on their own weight and the weight of the fill within and above the wall
elements to resist lateral earth pressures.

Modular gravity walls have been used in a variety of projects including highways, bridges, railroads,
channels, dikes, and others (Figure 5-2). Modular gravity walls may be used where conventional cast-in-
place concrete retaining walls are considered, as discussed in Chapter 4. Because of their modular
configuration, they are easy to assemble, dismantle and transpon; therefore, they are ideally suited for use
in remote areas for construction of permanent or temporary structures.

(a) (b)

(c) (d)

Figure 5-1: Modular Gravity Walls. (a) Metal Bin Wall; (b) Precast Concrete Crib Wall; (c) Precast
Concrete Module Wall; and (d) Gabion Wall. (AASHTO, 1997)

5- I
Highway
...... . ,. .. ' . ' .....
Proper Erosion
Protection Needed~

River

I~ New Location

.·····1 .."." r
. . -C:l~ Highway

Figure 5-2: Examples of Modular Gravity Wall Applications. (Contech, 1997)

5-2
The primary advantages of modular gravity walls are:

• Low cost
• Fast and easy construction
• Low labor requirement (except for gabion walls)
• Erection unaffected by temperature
• "Plant-controlled" quality of prefabricated units
• Flexibility and tolerance to differential settlements
• No need for architectural finish
• Units can be disassembled and re-used economically

The disadvantages of this type of wall are:

• Requires bottom-up construction


• Potential corrosion of exposed metallic elements including metallic bins, and wire used in gabion
baskets
• Potential decay of timber units exposed to water action
• Required storage space at the site for relatively large prefabricated units
• Possible unavailability of required fill material
• Susceptibility of some systems to vandalism (wire baskets can be damaged more easily than
concrete structures)
• Not suitable for use with deep foundations
• Many of the available systems are proprietary

The principal types of modular gravity walls are the crib wall, the concrete module wall, the bin wall and
the gabion wall. Although the same basic design considerations and procedures generally apply to all of
these wall systems, each type of wall has distinctive features and construction requirements, as discussed
below.

5.2 CRIB WALLS

5.2.1 General

Crib walls are built using prefabricated units which are stacked and interlocked and filled with free
draining granular material, (Figure 5-3). The crib units are made of either reinforced concrete or timber.
Timber units require treatment for long-term durability against decay, particularly in walls with extensive
contact with water.

The wall face can either be open or closed. In closed-face cribs, the stretchers are placed in contact with
each other. In open-faced cribs, the stretchers are placed apart but at close intervals so that the infill
material does not escape through the face. When subject to wave or current action, such as in stream or
waterfront applications, a geotextile filter layer is usually placed behind the wall face to prevent the loss
of fines through the spaces between stretchers, even in closed-face cribs.

5-3
(0) (b)

Bock stretcher ..,

Type n

Type I

Type m
(e)

Note: Type I is an open-face wall system. Types II and III are both closed-face wall
systems

Figure 5-3: Crib Walls. (a) Uniform Cross-Section; (b) Stepped Cross-Section; and (c) Typical
Details of a Reinforced Concrete Crib Wall. (After HKGEO, 1993)

5-4
Walls higher than about 2 m are usually built with a batter to improve stability. Crib walls of constant
width are usually used for heights up to about 5 m (Figure 5-3(a». For higher walls, stepped wall cross-
sections (Figure 5-3(b» are often used to increase stability and reduce the cost of the wall.

5.2.2 Wall Design

The design of crib walls, and other types of modular gravity walls, requires an evaluation of: (a) stability
against sliding; (b) overturning resistance; (c) bearing capacity of the foundation soils; (d) global stability;
and (e) wall settlement. Each of these design issues should be evaluated using the same procedures
described in Section 4.3 for cast-in-place concrete walls. Following are additional considerations and
procedures for design of crib walls and other types of modular gravity walls:

• Lateral earth pressures for wall design are determined using the procedures presented in Chapter
2. Generally, Coulomb theory is used to compute earth pressures on a vertical or inclined plane,
as appropriate, along the back face of the wall (Figure 5-4). Where the wall is founded on a
relatively incompressible foundation, the angle of wall friction, 0, may be assumed to be equal to
<1>'/2, where <1>' is the friction angle of the compacted backfill soil. It is important to adequately
compact the infill material for the assumption of 0=<1>'/2 to hold true (HKGEO, 1993). For walls
on relatively compressible foundation, assume 0=0 as the wall may settle relative to the backfill.

• In addition to the lateral earth pressure, the wall must be designed for lateral pressure due to
surcharge loads (see Chapter 2).

• The resistance due to passive earth pressure in front of the wall is neglected.

• The required minimum factors of safety for resistance to sliding are the same as those presented
in Section 4.3.2 for Cast-in-Place Concrete walls.

• Criteria for overturning, and for location of the resultant force on the base of the wall, are the
same as those presented in Section 4.3.2 for Cast-in-Place Concrete Walls.

• Global stability should be investigated using the procedures presented in Module 3 - Soil Slopes
and Embankments.

• The ultimate bearing capacity of the foundation soils should be determined using the procedures
discussed in Module 7 - Shallow Foundations. Generally, a minimum factor of safety against
bearing capacity failure of 3.0 is required.

• The computed compression, shear and bending stresses on the bottom module element should be
compared with the manufacturer's specified structural capacity for that element. If the computed
load exceeds that capacity, a special design may be required for the bottom segments or
alternatively, the wall configuration may be modified to reduce the load on the bottom module.

• For retaining walls with a stepped cross-section, perform sliding and overturning analyses at each
section change.

• Possible erosion of the soil at the wall toe by an act of nature, or removal by human activities,
should be taken into account in the design.

5-5
• Deformations of modular gravity walls built on firm ground usually fall within the service ability
limits of these walls; thus, deformation calculations are rarely performed.

• The possibility of differential settlement along the length of a modular gravity wall should be
evaluated if the soil profile underneath the foundation is variable. Table 5-1 summarizes the
tolerances of a crib wall and other types of modular gravity walls to differential settlement,
expressed a ratio of differential settlement (d V) to horizontal distance (dH). Walls built of precast
concrete units are more sensitive to differential settlement, and problems may arise for walls more
than about 7 m high. These walls are generally not suitable for use on compressible ground, or
for supporting heavy loads.

(a)

(b) (c)

Figure 5-4: Active Earth Pressures for Design of Modular Gravity Walls. (a) Vertical Back Face; (b)
Back Face with Positive Slope; and (c) Back Face with Negative Slope. (Modified from
AASHTO, 1997)

5-6
TABLE 5-1
TOLERANCE OF MODULAR GRAVITY WALLS
TO DIFFERENTIAL SETTLEMENT

Wall Type Tolerance


(!::.V / !::.H)

Crib Wall 11300


Concrete Modular Wall 11300
Bin Wall 11300
Gabion Wall 1150

5.2.3 Materials

Precast concrete should comply with the standard specifications for reinforced concrete. The precast
concrete members should be constructed of Portland cement concrete with a minimum compressive
strength at 28 days of 28 MPa. Timber should be of structural grade, properly treated and should comply
with standard specifications for structural timber. Crib members should be fabricated so that they are fully
interchangeable, without the need to drill, cut or offset the members to correct for non-uniform sections.
Bolts, nuts and miscellaneous hardware should be galvanized.

The infill material should be well graded, free-draining, granular soil which will not sift or flow through
openings in the wall. Infill should have a maximum particle size of 80 rom and should have no more than
15 percent of particles passing through the openings of the No. 200 sieve (0.074 rom). For wall heights
greater than 6 m, not more than 5 percent of the particles should pass through the openings of the No. 200
sieve (0.074 rom).

5.2.4 Wall Construction

The first stage of construction consists of excavating to the wall foundation grade and preparing the wall
foundation. If stable excavation slopes cannot be maintained during this period of wall construction,
temporary excavation support may be needed. Foundation preparation includes removing unsuitable
materials such as organic matter and vegetation from the area to be occupied by the wall and leveling and
proof-rolling the foundation area.

A cast-in-place concrete leveling pad is normally provided at the base of the wall. This slab usually
extends beyond the front and back faces of the crib and, on sloping ground, it is often stepped to follow
the slope.

Timber and concrete crib members are placed in successive tiers in accordance with the design spacing
and arrangement (Photo 5-1). Drift bolts at the intersection of timber header and stretcher members must
be accurately installed to maintain the required minimum edge distances. At the intersection of concrete
header and stretcher members, asphalt felt shims, or other suitable material, are used to obtain uniform
bearing between the members.

After each course of stretchers is assembled, the void within the crib is filled with infill material.
Backfilling progresses simultaneously with the erection of the crib members, and the backfill is carefully
placed and compacted to avoid displacing or damaging the crib members. The infill material should be
placed in uniform layers not exceeding 300 mm in thickness, and compacted with a manual vibratory

5-7
tamping device to at least 95 percent of the maximum density as determined in accordance with AASHTO
T-99.

Unless the infill material can serve as a filter, a geotextile layer is usually provided behind the rear face
of the crib wall to prevent migration of fines from the backfill. Drainage systems similar to those used in
conventional Cast-in-Place concrete walls may be needed, particularly for closed-face cribs. In open-face
cribs, the interspace can be planted with vegetation to help blend the wall with the surrounding
environment.

5.3 CONCRETE MODULE WALLS

5.3.1 General

Concrete module walls use interlocking precast concrete cells erected at the site and filled with compacted
earth (Figure 5-5). A common concrete module (Figure 5-5(b» has a face height of about 1.2 m, a face
length about 2.4 m, and a width ranging from 1.2 m to 6.0 m. The wall units can be assembled vertically
or at a batter. a variety of surface treatments (stains, striations, exposed aggregate, etc.) are available to
meet specific aesthetic requirement. On top of the wall, a parapet module can be placed and held rigidly
by a cast-in-place concrete slab. A reinforced cast-in-place or precast concrete footing is usually placed
at the toe and heel of the wall.

Assembly of the interlocking wall units does not require bolts, nuts, pins, or fasteners. Accordingly,
corrosion problems are minimized. Since all the units are manufactured in the plant and hauled to the site,
their fabrication is standardized and performed to shop-applied quality control procedures.

Photo 5-1: Setting Precast Elements for an Open Faced Crib Wall

5-8
(b)

ost .tN_PlACE
CONCttETE
PAttAVET ttEeAtt

(a.)

(c) VI 11 (b) 'typical Module; (c) Ptec""


ule
(.) 'typical section 01 • concrete l/Iodton .;
p.ra!"''" (""'" pouhl"",.l coq>O,at )
figure 5-5: 5. 9
5.3.2 Wall Design

Except as noted below, the design procedures for a concrete module wall are the same as those discussed
in Section 5.2.2.

The base sliding resistance is obtained by adding the sliding resistance at the bottom of the toe and heel
footings to the sliding resistance from the soil between the toe and heel footings. The sliding resistance
for the footings is determined by multiplying the computed normal loads at the footings by the coefficient
of sliding for concrete in contact with soil (see Table 2-1 for coefficient of friction values); the sliding
resistance between the footings is determined by multiplying the computed normal force for this zone by
the value of tan 4>', where 4>' is the angle of internal friction of the foundation soils.
5.3.3 Materials

Precast concrete should comply with the standard specifications for reinforced concrete. The precast
concrete modules should be constructed of Portland cement concrete with a minimum compressive strength
at 28 days of 28 MPa. The concrete modules should be fabricated so that they are fully interchangeable,
without the need to drill, cut or offset the modules to correct for non-uniform sections.

The infill material should be well graded, free-draining, granular material with a maximum particle size
of 80 mm and not more than 15 percent of particles passing through the openings of the No. 200 sieve
(0.074 mm). For wall heights greater than 6 m, not more than 5 percent of the particles should pass
through the openings of the No. 200 sieve (0.074 m).

5.3.4 Construction

Construction of a concrete module wall is relatively fast and simple, requires minimum labor and can be
accomplished with a small crane.

The first stage of construction consists of excavating to the wall foundation grade and preparing the wall
foundation. If stable excavation slopes cannot be maintained during the period of wall construction,
temporary excavation support may be needed. Foundation preparation includes removing unsuitable
materials such as organic matter and vegetation from the area to be occupied by the wall and leveling and
proof-rolling the foundation area.

Next, pre-cast or cast-in-place concrete footings are installed. The footings set the line and grade,
establish the batter of the wall if it is not vertical, and help distribute the load to the foundation soils.

The concrete modules arrive at the site ready to be installed. A crane equipped with a special handling
device picks up the units and places them in their proper positions within the structure (Photo 5-2(a».
Using a building-block concept, the interlocking modules can quickly be aligned as they are placed.

Between each module course, rubber pads or other suitable material are placed in the front horizontal
joints, and at other bearing areas. Alignment and elevations are corrected by adding or removing pads.
It is critical, however, that full bearing is provided on each pad.

Once the modules are properly set, shims are used to fill any space between the keys to prevent movement
of the modules during backfilling. A geotextile filter layer is placed behind the front vertical joints to
prevent the migration of fines and allow the joints to act as weep holes permitting the passage of water that
might otherwise build up behind the wall.

5 - 10
(a)

(b)

(c)

Photo 5-2: Construction of Concrete Module Wall: (a) Placement of Precast Modules, (b) Placing
Fill Within the Modules, and (c) Compacting the Infill Material (Doublewal Corp.)

5 - II
After each course is set, the modules are filled with select material (Photo 5-2(b», and backfill is placed
behind them. The infill and backfill are carefully placed and compacted to avoid displacing or damaging
the concrete modules. The infill should be placed in uniform layers not exceeding 300 mm in thickness,
and compacted with a manual vibratory tamping device to at least 95 percent of the maximum density as
determined in accordance with AASHTO T-99 (Photo 5-2(c». The backfill behind the wall is placed and
compacted following normal retaining wall construction procedures.

The parapet units, if required, are placed on top of the upper course, and adjusted to the proper alignment
by hardwood wedges. A concrete slab is then cast-in-place to hold the wall parapet in its pennanent
position (Figure 5-5(c».

5.4 BIN WALLS

5.4.1 General

Bin walls consist of adjoining closed- face cells filled with compacted select backfill to fonn a gravity type
retaining structure. The cells are generally constructed of sturdy, lightweight, steel members that are
bolted together at the site. The flexibility of the steel structure allows the wall to flex against minor ground
movements that might damage rigid-type walls.

Figure 5-6 illustrates the typical geometry of a bin wall. Basically, the wall is assembled using modular
panels called stringers and spacers. The stringers constitute the front and back faces of the bin, and the
spacers its sides. Vertical connectors (columns), grade plates, and stronger stiffeners act to hold the wall
together (Figure 5-7). The stringers are usually 3 m long and come in a variety of heights. Shorter
stringers are sometimes used to allow curvature in the wall. The length of the spacer (wall depth) is
variable, and is detennined by stability requirements.

The walls of the bin (stringers and spacers) are formed of S-shaped steel members. As the wall height and
depth increase, the thickness of these members is also increased to resist the internal pressure from the fill
material. Table 5-2 shows the typical relationship between thickness of steel panel members and wall
height, as recommended by one wall manufacturer.

The advantages of bin walls include flexibility, versatility, ease of installation and dismantling for reuse.
Their main disadvantage is the potential corrosion of the steel elements, particularly when in contact with
corrosive media.

TABLE 5-2
TYPICAL HEIGHT-TIDCKNESS RELATIONSHIP
FOR BIN WALLS
(After Contech, 1994)

Wall Height, Steel Thickness,


Gage
m mm
0-3.5 1.62 16
3.5-5 2.00 14
5-8 2.75 12
8-11 3.50 10

5 - 12
Corner Split Rear
Vertical Vertical Panel ""
Stringer
Connector \ Connector JC::::::::::======:::==;;n
~;::::.===== ==:t::::=J]
A-+-

Spacer
Spacer/
""
~==::::::::=.;d!;====t::::::-====LR;;::::=:::::::=;:::=~~
Vertical / A -4- Front t
Connector Panel
,~
Length "I
(a)

(b)

(c)
Figure 5-6: Typical geometry of Bin Wall. (a) Plan, (b) Elevation and (c) Section A-A (Contech,
1997)

5 - 13
Stringer

Stringer ----..

(a)

Column

Column
Splice

(b)

Figure 5-7: (a) Elements of Bin Walls with T-Shaped Vertical Connector for Bin Walls, (b) Channel
Shaped Vertical Connector (Contech, 1997)

5 - 14
5.4.2 Wall Design

Except as noted below, the design procedures for a bin wall are the same as those discussed in Section
5.2.2.

Walls can be designed to be vertical or provided with a slight batter. Battering the wall helps improve wall
stability and avoids a possible outward tilting appearance which may occur for vertical walls subject to
differential settlement.

Positive measures must be provided to drain the material within and behind the bins. Drainage is typically
provided by installing perforated drain pipes behind and below the rear base of the bin wall. The
perforated pipe should be surrounded by pervious backfill wrapped in a geotextile filter material, and
provided with suitable drainage outlets.

5.4.3 Materials

The wall units are usually formed from galvanized steel sheets. Where service conditions are unusually
severe, the wall may be built using fiber-bonded steel. Aluminum steel has also been used, but for
stringers and spacers only. The exposed stringer, or front panel, frequently consists of corrugated steel
or aluminum panels (Photo 5-3(a», but precast concrete panels with variable architectural treatments
(Photo 5-3(b» can also be used.

The infill material should be well-graded, free-draining, granular material with a maximum particle size
of 80 mm and not more than 15 percent of particles passing through the openings of the No. 200 sieve
(0.074 mm). For wall heights greater than 6 m, not more than 5 percent of the particles should pass
through the openings of the No. 200 sieve (0.074).

5.4.4 Wall Construction

The lightweight wall components are easily assembled by a small crew of unskilled labor, using a small
crane and ordinary hand tools.

The first stage of construction consists of excavating to the wall foundation grade and preparing the wall
foundation. If stable excavation slopes cannot be maintained during the period of wall construction,
temporary excavation support may be needed. Foundation preparation includes removing unsuitable
materials such as organic matter and vegetation from the area to be occupied by the wall and leveling and
proof-rolling the foundation area. A gravel working pad can be placed to facilitate bin construction.

Generally, the bins are erected at the site by bolting panels to vertical connector elements. However, on
larger projects, or where unusual working conditions are encountered, panel and transverse sections may
be preassembled at the shop, then transported to the site (Photo 5-4).

The infill material should be placed in uniform layers not exceeding 300 mm in thickness and compacted
using manual tamping equipment to at least 95 percent of the maximum density as determined in
accordance with AASHTO T-99. If the fill has a high percentage of [me sand, the coarser material is
usually placed adjacent to the vertical connectors in bin comer to prevent the loss of [me material through
normal small openings in the comers. Caulking or sealing these critical areas with geotextile filter material
can also be done.

5 - 15
(a)

(b)

Photo 5-3: Bin wall with (a) Corrugated Steel Facing, and (b) Precast Concrete Facing (Contech
Company)

5 - 16
(a)

(b)

Photo 5-4: Construction of a Bin Wall (a) Setting Preassembled Panels, and (b) Filling the Completed
Bins (Contech Company)

5 - 17
As the bins are filled, the backfill material is placed behind the wall and compacted using standard
earthwork construction procedures. At all times, the level of the infIll material should be kept higher than
the fill behind the wall, to maintain stability.

Construction of bin walls on curves requires special attention. Installation of the wall can be accomplished
through one of two methods (Figure 5-8) as follows:

1. Shorter stringers can be used at the front and the rear faces of the wall. These stringers can either
be manufactured in shorter sizes, or can be cut in the field to fit any wall configuration.

2. Shop-fabricated comer plates are used to provide angularity and curvature.

5.5 GABION WALLS

5.5.1 (;eneral

Gabion walls are composed of rows and tiers of orthogonal wire cages or baskets filled with rock
fragments and tied together (Figure 5-9). They are widely used for channel and river bank protection, but
are also used for earth retaining structures on land, particularly in rugged terrain. The finished wall easily
blends with the natural landscape, especially if local rock is used as infill material.

Gabion walls are simple and easy to construct. The construction of gabion walls becomes more economical
for sites where suitable infill rock is available. Gabion walls are free draining and, if the backfill does not
trap water, they will not be subjected to hydrostatic water pressures. Because they are free-draining, they
also are frost resistant.

Since gabion walls are flexible structures they are suitable for construction over compressible soils. They
are also well· suited for remote areas that cannot be easily accessed by heavy machinery. In these
situations, the wire baskets are assembled by hand and filled with local stone. Another advantage of gabion
walls is that they allow penetration by protruding objects such as pipes. With time, the infill stones tend
to collect soil and promote vegetation which improves the wall aesthetics.

Gabion walls are labor intensive; thus, they may not be cost-effective in certain regions. The wire baskets
may be subject to vandalism, and the long-term durability of the wire is questionable, particularly in
corrosive environments.

The front and rear faces of a gabion wall may be straight or stepped (Figure 5-9). A batter is usually
provided for walls higher than about 3 m to improve stability. The wall batters commonly used are 1 in
10, 1 in 6, and 1 in 4.

A variety of cage sizes can be produced to suit the terrain, but the conventional dimensions of the gabion
modules are 2 m x 1 m x 1 m. The longer gabions may be divided into cells by diaphragms made of the
same mesh as the gabion basket itself, and directly joined to the base panel during manufacture. These
diaphragms reinforce the structure and make assembly and erection easier (Figure 5-10).

5.5.2 Wall Design

Except as noted below, the design procedures for a gabion wall are the same as those discussed in Section
5.2.2. .

5 - 18
(a)

(b)

(c)

Figure 5-8: Construction of Bin Walls at Curves. (a) Typical Outside Comer, (b) Typical Inside
Comer and (c) Any Wall Configuration. (Contech, 1997)

5 - 19
h h
- < 1.5 - < 1.5
b b

(a)
(b)

(c)

Figure 5-9: Typical Gabion Wall Cross-Sections. (a) Wall with Stepped Front Face; (b) Wall with
Stepped Back Face; and (c) Sloped Wall.

2.00 m

(a) (b)
Figure 5-10: Gabion Baskets. (a) Module Without Diaphragms; (b) Module with Diaphragms.

5 - 20
• Since the infill and backfill materials are well-drained, no hydrostatic pressure is used in the
design.
• The angle of wall friction, 0, between the gabion structure and the backfill material may be
reduced substantially if a stiff, smooth, geotextile layer is used as a filter on the back face of the
gabion structure.
• For stepped walls, analyses for sliding and overturning should be carried out at each change in
wall section, ignoring the resistance contributed by the wire mesh and the connection between the
baskets.
• The unit weight of the wall infill material used in design analyses is dependent on the type and
porosity of the infill material. The weight of the wire basket is relatively small and is usually
ignored in the design. The unit weight, Yg' of the wall is calculated as follows:

Yg =(l - n r)G s yw (5-1)

where llr is the porosity of the rock fill, G s is the specific gravity of the rock, and y w is the unit
weight of water. The porosity of the infIll material is generally between 0.3 and 0.4 (HKGEO,
1993), and is dependent on the grading and angularity of the rock fragments, as well as the method
of installation and degree of compaction of the infill material.
• The wires forming the gabion basket can be galvanized and coated with a layer of polyvinyl
chloride (PVC) to protect against corrosion.
• Due to the flexibility of the structure, the gabion wall can tolerate substantial settlements and
relatively large horizontal deformations before failure. However, excessive vertical and horizontal
deformations may result in an unattractive appearance.

5.5.3 Materials

Gabion baskets are made from a range of materials including steel, nylon, polypropylene or polyethylene.
The polymer-type materials have the advantage of being lightweight and corrosion resistant. However,
they are susceptible to attack by fire or ultraviolet light. Therefore, if these materials are used in making
the baskets, it is advisable to cover the exposed grids with a non-flammable material such as shotcrete.

The material most commonly used in commercial production of gabions is steel wire-mesh, which is
manufactured in two types: hexagonal woven with an approximate opening size of 80 mm by 115 mm, and
square welded with an approximate opening side dimension of 75 mm.

The hexagonal wire mesh is mechanically woven in a continuous sheet. The wires are double twisted to
form the mesh. The diameter of the steel wires is between 2 and 3 mm. The gabion base, top and sides
are usually formed from a single piece of mesh, with its edges made of a wire with a diameter of about
1.5 times that of the wire mesh to prevent unraveling. High-tensile zinc-coated, steel wires are used, with
a minimum tensile strength of 350 MPa. For permanent applications, the wires should be at least 2.7 nun
in diameter and galvanized before weaving.

The welded wire mesh is manufactured from high tensile steel wire, electrically welded at each
intersection. The welded mesh should be hot-dip galvanized after welding. The fabrication of panels by
welding the wires after they have been galvanized is not recommended as the welds are left unprotected
(HKGEO, 1993). Since the wires are welded at each intersection, the welded wire mesh has less ability

5 - 21
to stretch and contract, and thus the assembled baskets are less flexible than comparable woven wire mesh
gabions.

In highly corrosive environments, such as sea water or polluted water, PVC (polyvinyl chloride) coating
is usually provided for the wires. a black, PVC coat, 0.4 to 0.6 nun thick, is applied to the woven mesh
by hot dipping or by extrusion onto the galvanized wire before weaving. For welded mesh, the PVC
coating is applied electrostatically to the welded panel. The PVC coat should be bonded sufficiently to the
wire core to prevent capillary flow of water between the wire and the PVC, which may cause corrosion.
The durability of the PVC coating against contact with acidic, salt or polluted water, and exposure to ultra
violet light and abrasion, is tested in accordance with ASTM E 42-65 specifications.

All wires used in the construction of gabion walls (lacing wires, etc.) are made of the same quality wire
used in the wire mesh. In waterfront structures, armor units, such as rip rap, are usually provided to
protect the wire baskets from wave action or heavy water-borne material.

The infill material consists of rock fragments that are sound, durable and well-graded, with a maximum
size of 300 nun and a minimum size equal to twice the size of the mesh opening. The preferred size range
is 150 nun to 300 nun (HKGEO, 1993).

The backfill material consists of open-graded granular soil similar to that placed behind other gravity-type
structures. In partially submerged walls, the backfill material should be free-draining to prevent the build
up of water pressure behind the wall when the water level in front of the wall is lowered. Drainage behind
the wall should be similar to that provided for crib walls and other types of modular walls. a filter layer
(soil or geotextile) is usually provided between the gabion baskets and the backfill to prevent migration of
the backfill soil through the relatively large voids of the infill material.

5.5.4 Wall Construction

The first stage of construction consists of excavating to the wall foundation grade and preparing the wall
foundation. If stable excavation slopes cannot be maintained during the period of wall construction,
temporary excavation support may be needed. Foundation preparation includes removing unsuitable
materials such as organic matter and vegetation from the area to be occupied by the wall and leveling and
proof-rolling the foundation area.

a gravel bedding layer, or concrete leveling pad is often placed to provide a working surface and to help
establish the alignment and elevation of the wall.

The hexagonal woven-mesh gabions are supplied folded flat and in a bundle. Each gabion is erected at
the site by folding up the sides and lacing together all vertical edges with the lacing wire. The four comers
of the box are laced first, followed by lacing the edges of internal diaphragms to the sides of the box.

The erected units are placed in their proper locations and adjoining empty gabions are laced along the
perimeter of their contact surfaces to obtain a monolithic structure. The connected units are then stretched
taut along their proper alignment before they are filled. The first gabion in the line is usually partially
filled to provide anchorage during stretching.

The welded wire gabions are either assembled in the factory and shipped flat to the project site similar to
the woven-mesh gabions, or the wire mesh panels are cut in the field to the dimensions of the sides, top
and base of the units in separate pieces which are joined together by spiral binders or tie wires. No
stretching is required for the welded wire gabions.

5 - 22
The one meter high baskets are tightly fIlled in three equal layers to ensure minimum voids (Photo 5-5(a».
Tie wires between the front and back faces of the gabion basket are typically installed at the top of the fIrst
and second infIll layers to restrain bulging of the front face. Adjoining units are fIlled to the same
elevation during each stage of fIlling. The last layers of stone are leveled with the top of the gabion to
allow proper closing of the lid, and to provide a level surface for the following course (photo 5-5(b». The
mesh of the lid is tied down to the sides and ends of the basket as well as to any internal diaphragm panels.
Well packed filling (without undue bulging) and secure connections are essential in all gabion structures.

Gabions are fIlled by any type of earth-handling equipment such as a backhoe, clam shell, etc. Care must
be taken to avoid damaging the wire mesh or its protective coating by sharp particles of crushed rock.
Special care is particularly required when using polymer grids.

The gabion wall can be constructed at curves or angles by cutting and folding the wire-mesh to make units
of special shapes and sizes. When building under water, the gabion units are usually assembled and fIlled
at the surface, then lowered into position by a crane.

Prior to placing backfIll, a geotextile fIlter material or graded soil fIlter is placed to fully cover the back
faces and steps of the gabion wall.

5.6 Example Problem 5-1

Statement

Design a gabion wall 4 m high, supporting a soil backfIll with a uniform surcharge of 10 kPa. Properties
of the backfIll and foundation soil are: y = 19 kN/m3, <I> = 34 0 • The rockfIll within the wire basket will
be compacted to a unit weight of 22 kN/m3 • The ultimate bearing capacity of the foundation soil is 400
kPa.

3m

I' ~
10 kPa

-
S

y= 19 kN/m
3

S ~ ® <I> = 34 0
-.;to

-
S
~ @

-
S

A
(i)
B PI pz
Not to Scale

~
Resultant Force
On Base, R

Figure 5-11: Geometry and Parameters for Example Problem 5-1.

5 - 23
(a)

(b)

Photo 5-5: Gabion Wall Construction: (a) Filling Gabion Baskets with Stone, and (b) Closing
Gabion Lid for Tying.

5 - 24
Solution

Step 1: Select a trial section

Consider using a front-stepped vertical back gabion wall, consisting of 1 m high baskets, placed for a
cumulative base width of 3 m as shown. Check stability of wall, by comparing resisting and driving forces
and moments respectively.

Step 2: Compute lateral earth pressures using Rankine theory, neglecting wall friction.

= K"q = (0.28) (10 kN/m2) = 2.8 kPa

= (0.28) (19 kN/m3) (4 m) = 21.28 kPa

Step 3: Calculate resisting forces (weights) and resisting moment of wall about point A.

Section Volume, m3 Weight, kN Arm,m Moment about A,kN'm


1 3)(1)(1) = 3.0 (3.0 m3)(22kN/m3) = 66 3/2 =1.50 (66) (1.50) = 99.00
2 (2.5)(1)(1) = 2.5 (2.5 m3)(22kN/m3) = 55 0.5 + (2.5/2) = 1.75 (55) (1.75) = 96.25
3 2.0)(1)(1) = 2.0 (2.0 m3)(22kN/m3) =44 1.0 + (2.0/2) =2.00 (44) (2.00) = 88.00
4 (1.0)(1)(1) = 1.0 (1.0 m3)(22kN/m3) = 22 2.0 + (1.0/2) =2.50 (22) (2.50) = 55.00
Total: W = 187.0 Mr = 338.25

Step 4: Calculate driving forces and moments about Point A.

Section Force, kN Arm,m Moment about A, kN'm


5 (2.8 kN/m2) (4 m)= 11.2 4/2 = 2.0 (11.2) (2.0) = 22.4
6 (lh) (21.28kN/m2) (4 m)=42.56 4/3 = 1.33 (42.56) (1.33) = 56.75
Total Ph = 53.76 Md = 79.15

5 - 25
Step 5: Check factor of safety against sliding.

Estimate friction between rockfill and foundation material,

JL = tanQ> = tan 34 0 = 0.675


where: <P = Angle of internal friction of foundation soil (conservative)

Factor of safety against sliding is defmed as:

= ~Resisting Force
~Driving Force

(0.675)(187.0 kN/m) 126.23


= =- - - = 2.35 > 1.5 O.K.
53.76 kN/m 53.76

Step 6: Check factor of safety against overturning.

~Resisting Moment
= --_---:::_---
~Driving Moment

:EM R 338.25 kN.m


= - - - - - = 4.27 > 2.0 O.K.
~M o 79.15 kN/m

Step 7: Check bearing pressure.

(a) Compute the distance of resultant force. R. from point A.

338.25 kN .m/m -79.15 kN .m/m


d = =-------------
187.0 kN/m
= 1.39 m

(b) Compute the eccentricity ofresultant about centerline of base.

e = !-d = 3m -1.39m = 0.11 m


2 2

e 0.11 m
= = 0.037 < -1 O.K.
B 3.00 m 6

5 - 26
(c) Compute the maximum and minimum bearing pressures under the footing.

qmax,min
= ~BV( 1 ± 6e) = 187.~~N/m [1 ± (6)(0.037)]
B

= 62.3(1 ± 0.22)

'lmax = 76.0 kPa


'lmin = 48.6 kPa
Step 8: Check factor of safety against bearing capacity failure.

Using the given value of 400 kPa for the ultimate bearing capacity, the factor of safety against bearing
failure can be determined as follows

400
FS = = = 5.26 > 2.5 O.K.
76

SUMMARY

Factor of safety against sliding FSs = 2.35


Factor of safety against overturning FSo = 4.27

Factor of safety against bearing failure FSbc = 5.26

Since all factors of safety computed are higher than required, this wall can be optimized by reducing the
base width and each step proportionately.

After optimization of the wall section, the factor of safety against global failure should be evaluated using
methods presented in Module 3 - Soil Slopes and Embankments, and wall settlement including tilting and
lateral squeeze should be evaluated using procedures presented in Module 7 - Shallow Foundations.

5 - 27
CHAPTER 6.0
:MECHANICALLY STABILIZED EARTH WALLS

6.1 INTRODUCTION

Mechanically Stabilized Earth (MSE) walls have been used increasingly, worldwide, since the late 1960s.
This type of wall consists of a select backfill material, reinforcing elements within the backfill, and a facing
(Figure 6-1). The reinforcing elements may consist of steel strips or bars, welded wire mats, polymer
grids, or geotextile sheets.

MSE walls are typically constructed in fill situations by placing alternating layers of soil and reinforcing
elements, as shown in Figure 6-1. The resulting reinforced-soil structure is flexible and can generally
accommodate relatively large horizontal and vertical movements without excessive structural distress.
Figure 6-2 illustrates various applications of the MSE wall. Photos 6-1 to 6-5 show a variety of
constructed MSE walls.

Table 6-1 lists many of the MSE wall systems currently available on the market. The primary differences
between these systems are the materials, geometries and arrangements of the reinforcing elements, and the
materials and details of the facing elements.

There have been many publications on MSE walls in recent years. The most recent comprehensive
reference on the subject is the Participants Manual for FHWA Demonstration Project 82 by Elias and
Christopher (1997). Some sections of this manual, including graphs and tables, are included in this
chapter.

Mechanic:ally' Stabilized Finished Grade


Earth Mass

Facing

Foundation
Soil

Figure 6-1: Principal Components of a Mechanically Stabilized Earth Wall. (Christopher, et al., 1990)

6- 1
In this chapter, various MSE wall systems are introduced and their reinforcement, backfill and facing
elements are described. The concept of soil reinforcement is presented and its impact on the wall's design
is discussed in detail. Internal and external stability analyses are outlined and acceptable safety factors are
recommended. Settlement and lateral movement of the wall and its durability and long-term performance
are discussed in detail. Step-by-step construction procedures are presented for differerent MSE wall
applications.

.. " .',. . ..... ' , '

Retaining Wall Access Ramp

(a) (b)

... ' . ',.

Waterfront Structure Bridge Abutment

(c) (d)

Figure 6-2: Examples ofMSE Wall Applications; (a) Retaining Wall; (b) Access Ramp; (c) Waterfront
Structure; and (d) Bridge Abutment.

6-2
Photo 6-1

Photo 6-4

Photo 6-2

Photo 6-5

Photo 6-3

Photos 6-1 to 6-5: Examples of MSE Walls

6-3
TABLE 6-1
SUMMARY OF REINFORCEMENT AND FACE PANEL DETAILS FOR
SELECTED MSE WALL SYSTEMS (Elias and Christopher, 1997)

System Name Reinforced Detail Typical Face Panel Detail


Reinforced Earth Galvanized Ribbed Steel Strips: 4 mm Facing panels are cruciform shaped
The Reinforced Earth Company thick, 50 mm wide. Epoxy-coated strips precast concrete 1.5 x 1.5 m x 14 cm
2010 Corporate Ridge also available. thick. Half size panels used at top and
McLean, VA 22102 bottom.
VSL Retained Earth Rectangular grid of Wl1 or W20 plain Hexagonal and square precast concrete
VSL Corporation steel bars, 61 x 15 cm grid. Each mesh 1.5 x 1.5 m x 14 cm thick. Half size
2840 Plaza Place may have 4, 5 or 6 longitudinal bars. panels used at top and bottom.
Raleigh, NC 27612 Epoxy-coated meshes also available.
Mechanically Stabilized Embankment Rectangular grid, nine 9.5 mm diameter Precast concrete; rectangular 3.81 m
Department of Transportation plain steel bars on 61 x 15 cm grid. Two long, 61 cm high, 20 cm thick.
Division of Engineering Services bar mats per panel (connected to the
5900 Folsom Blvd. panel at four points).
P.O. Box 19128
Sacramento, CA 95819
Georgia Stabilized Embankment Rectangular grid of five 9.5 mm plain Precast concrete panel; rectangular 1.83
Department of Transportation steel bars on 61 x 15 cm grid 4 bar m wide, 1.22 m high, 20 cm thick with
State of Georgia mats per panel. offsets for interlocking.
No.2 Capitol Square
Atlanta, GA 30334-1002
Hilfiker Retaining Wall Welded steel wire mesh, grid 5 x 15 cm Welded steel wire mesh, wrap around
Hilfiker Retaining Walls. ofW4.5 x W3.5, W9.5 x W4, W9.5 x with additional backing mat 6.35 mm
P.O. Drawer L W4, and W12 x W5 in 2.43 m wide wire screen at the soil face (with
Eureka, CA 95501 mats. geotextile or shotcrete, if desired).
Reinforced Soil Embankment 15 cm x 61 cm welded wire mesh: W9.5 Precast concrete unit 3.8 m long, 61 cm
Hilfiker Retaining Walls. to W20 - 8.8 to 12.8 mID diameter. high.
P. O. Drawer L
Eureka, CA 95501
ISOGRID Rectangular grid of Wl1 x Wll four Diamond shaped precast concrete units,
Neel Co. bars per grid. 1.5 by 2.5 m, 14 cm thick.
6520 Deepford Street
Springfield, VA 22150
GENESIS HDPE Geogrid. Keystone* Standard unit (200 mm high
Tensar Earth Technologies, Inc. by 40 mm long face, 600 mID nominal
5775-B Glenridge Drive, Ste 400 depth); OR Keystone International
Lakeside Center Compact* (200 mm high by 450 mm
Atlanta, GA 30328 long face, 300 mID nominal depth).
PYRAMID Galvanized welded wire mesh, size varies Pyramid* unit (200 mID high by 400 mID
The Reinforced Earth Company with design requirements OR Grid of long face, 250 mm nominal depth).
2010 Corporate Ridge PVC coated, Polyester yarn (Matrex
McLean, VA 22102 Geogrid).
Maccaferri Terramesh System Continuous sheets of galvanized double Rock filled gabion baskets laced to
Maccaferri Gabions, Inc. twisted woven wire mesh with PVC reinforcement.
43A Governor Lane Blvd. coating.
Willamsport MD 21795
Strengthened Earth Rectangular grid, W7, W9.5 and W14, Precast concrete units, rectangular or
Gifford-Hill & Co. transverse bars at 230 and 450 mID. wing shaped 1.82 m x 2.13 m x 14
2515 McKinney Ave. cm.
Dallas Texas 75201
Additional facing types are possible with most systems.

6-4
6.1.1 Aldvantages/LUEdtations

MSE walls have many advantages over conventional cast-in-place retaining walls. These include:

• Generally lower cost


• Simple and rapid construction procedures that do not require large construction equipment
• Do not require experienced craftsmen with special construction skills
• Need little space in front of the structure for construction operations
• Construction at all temperatures (no curing required)
• Wide variety of aesthetically pleasing facings
• Flexibility and tolerance to vertical and horizontal ground movements
• Feasibility to heights in excess of 25 m

Their main limitations are:

• Potential corrosion of the metallic reinforcements


• Elongation of the extensible materials, such as geotextiles
• Potential long-term deterioration of geosynthetic materials due to chemical attack and exposure to
ultraviolet light
• Wider space requirement behind the wall as compared to conventional walls
• Requirement of select granular fill in certain systems, which may render them uneconomical if
granular borrow sources are not readily available

MSE walls should not be used under the following conditions (Elias and Christopher, 1997):

• When utilities other than highway drainage must be constructed within the reinforced zone where
future access for repair would require the reinforcement layers to be cut.

• When metallic reinforcements are exposed to surface or groundwater contaminated by acid mine
drainage or other industrial pollutants as indicated by low pH and high chlorides and sulfates.

• When floodplain erosion may undermine the reinforced fill zone, or where the depth of scour
cannot be reliably determined.

Based on current experience, MSE walls are limited to 30 m in height for inextensible steel reinforcements
and 15 m in height for extensible geosynthetic reinforcements.

6.1.2 Cost

The cost of a MSE wall is a function of many factors, including cut/fill requirements, wall size and type,
in situ soil conditions, available backfill material, facing finish, and temporary or permanent application.
It has been found that MSE walls with precast concrete facings are usually less expensive than reinforced
concrete retaining walls for heights greater than about 3 to 4.5 m and average foundation conditions.

6-5
In general, the use of MSE walls results in savings on the order of 25 to 50 percent and possibly more in
comparison with a conventional reinforced concrete retaining structure, especially when the latter is
supported on a deep foundation system. A substantial saving is obtained by elimination of deep
foundations, which is usually possible because the reinforced-soil structures can absorb relatively large total
and differential settlements. Other cost saving features include ease and speed of construction. Typical
total costs of MSE walls range from $160 to $300 per m 2 of wall face, generally as function of height and
cost of select fill.

6.2 REINFORCING ELEMENTS

6.2.1 Reinforcement Geometry

A variety of reinforcement geometries is available, including:

• Linear Unidirectiomd. Strips including smooth or ribbed steel strips, and coated geosynthetic
strips over a load-carrying fiber (Figure 6-3).
• Composite Unidirectional. Grids or bar mats characterized by grid spacing greater than 150 mm
(Figure 6-4).
• Planar Bidirectional. Continuous sheets of geosynthetics, welded wire mesh, and woven wire
mesh. The mesh is characterized by element spacing of less than 150 mm (Figure 6-5).
• Fiber Reinforcement. Reinforcing elements in the form of tensile resistant strands or fibers are
included in the soil mass to form a composite mass with improved mechanical properties. The
fibers being investigated for possible use include natural fibers (reeds and other plants), synthetic
fibers (geotextile threads), and metallic fibers (small-diameter metal threads). Unlike other
reinforcements, fibers can potentially provide reinforcement in three directions. The major
limitations to using fiber reinforcement are the difficulty associated with economically mixing the
fibers uniformly into the backfill, and the lack of experience with this system.

6.2.2 Reinforcement Material

Distinction can be made between the characteristics of metallic and nonmetallic reinforcements:

• Metallic Reinforcements (Photos 6-6 and 6-7). Typically formed of mild steel. The steel
reinforcement is usually galvanized or may be epoxy coated.
• Nonmetallic Reinforcements (photos 6-8 and 6-9). Generally made of polymeric materials
consisting of polypropylene, polyethylene, or polyester.

6.2.3 Reinforcement Extensibility

There are two classes of extensibility:

• Inextensible Reinforcements. The deformation of the reinforcement at failure is much less than
that of the soil.
• Extensible Reinforcements. The deformation of the reinforcement at failure is comparable to,
or even greater than, the deformation of the soil.

6-6
Facing Element

Connection
~,-_[I-
Steel Reinforcing Strip

Low Density Polyethylene ~heath A LA,NE I

Polyester or PoIY~~-.""'··--:::""': I

Geosynthetic Strip

Figure 6-3: Strip Reinforcements. (Adapted from Mitchell and Villet, 1987)

~M;:~:2::~~t7'''----- Facing Elements

~6-9-"'l;:"'-- Compacted Fill

Geogrid Reinforcement

Plan View of Geogrid Reinforcement

Figure 6-4: Grid Reinforcements. (Mitchell and Villet, 1987; Lawson, 1992)

6-7
.\ ... '

.{. '-':' . : . Soil


. '-.: :-
.. :-..: z.: ....... _ . .

Figure 6-5: Geotextile Sheet Reinforcements. (Adapted from Mitchell and Villet, 1987; Lawson,
1992)

6.3 FACING ELEMENTS

The facing elements used in the different MSE systems reflect their aesthetics because they are the only
visible parts of the completed structure. A wide range of fInishes and colors can be provided in the facing.
In addition, the facing protects the backfIll from sloughing and erosion, and provides drainage paths in
certain cases. In walls using geosynthetics for reinforcement, the wall facing provides protection against
potential deterioration of the reinforcement from exposure to ultraviolet radiation. Figure 6-6 illustrates
a variety of wall facings. Major facing types are:

• Segmental Precast Concrete Panels. Summarized in Table 6-1 and illustrated in Figure 6-7, the
precast concrete panels have a minimum thickness of 140 mm and are of a cruciform, square,
rectangular, diamond, or hexagonal geometry. Temperature and tensile reinforcement are
required but will vary with the size of the panel. Adjacent units are usually vertically connected
with shear pins.

• Dry Cast Modular Block (MBW) Units. These are relatively small, squat concrete units
specially designed and manufactured for retaining wall applications. The weight of these units
commonly ranges from 15 to 50 kg, with units of 35 to 50 kg routinely used for highway projects.
Unit heights typically range from 100 to 200 mm. The exposed face usually varies in length from
200 to 450 mm. The nominal width (dimension perpendicular to the wall face) typically ranges
between 200 and 600 mm. The MBW units may be manufactured solid or with cores. Full height
cores are fIlled with aggregate during erection. The units are normally dry-stacked (i.e., without
mortar) and in a running bond confIguration. Vertically adjacent units may be connected with
shear pins, lips, or keys. They are referred to by trademarked names such as Keystone@,
Versalock@, Allen@, etc. Modular block units are illustrated in Figure 6-8 and in Photos 6-10 and
6-11.

6-8
Photo 6-6

Photo 6-7

Photos 6-6 and 6-7: Metallic Reinforcements

6-9
Photo 6-8

Photo 6-9

Photos 6-8 and 6-9: Geosynthetic Reinforcements

6 - 10
." ......
.
~ ~ ;~;;:~ ••• ..•
\"
"'~'''''
-",

~,..\ --..,...c...i\
?... ' .. '
~
' .. ,
" . \

\----,~

Segmental Geotextile
Precast Concrete Wrapped-Facing
with Shotcrete Cover

_.•._......._......_.+._.....
·, .. , ' ...
... :' .
."
' '
"

·... ..... : . " ..


-'
"
' '
"
........ , ...... '"
_.•._......_.+._....... ~

·..............
.. .
", ' ...- -'
,

Modular Block Wall Units Full-Height Concrete Panel

Gabion-Facing Tire-Facing Units

Figure 6-6: Various Types of Wall Facings. (Wu, 1994)

6 - 11
Figure 6-7: MSE Wall Surface Textures. (Various Manufacturers)

6 - 12
Figure 6-8: Various Types of Modular Blocks. (Simac, et al., 1996).

6 - 13
Photo 6-11

Photo 6-10

Photos 6-10 and 6-11: Modular Block Facing

6 - 14
• Metallic Facings. The original Reinforced Earth system had facing elements of galvanized steel
sheet formed into half cylinders. Although precast concrete panels are now typically used in
Reinforced Earth walls, metallic facings may be appropriate in structures where difficult access
or difficult handling requires lighter facing elements. They may also be appropriate for temporary
structures. The primary disadvantage of this wall facing is its greater susceptibility to corrosion.
• Welded Wire Grids. Wire grids can be bent up at the front of the wall to form the wall face.
This type of facing is used in the Hilfiker, Tensar, and Reinforced Earth wire retaining wall
systems.
• Gabion Facings. Gabions (rock-filled wire baskets) can be used as facing with reinforcing
elements consisting of welded wire mesh, welded bar-mats, geogrids, geotextiles or the double-
twisted woven mesh placed between or connected to the gabion baskets.
• Geosynthetic Facings. Various types of geotextile reinforcement are looped around at the facing
to form the exposed face of the retaining wall. Typically the geosynthetic facing is covered with
gunite or concrete to protect the geotextile from ultraviolet light degradation, vandalism and
damage due to fire. Alternatively, a geosynthetic grid used for the reinforcement of the soil can
be looped around to form the face of the completed retaining structure in a manner similar to
welded wire mesh and fabric facing. Vegetation can grow through the grid structure and can
provide both ultraviolet light protection for the geogrid and a pleasing appearance.
• Other Facings. Two other types of facing have recently been used, particularly in construction
of temporary, low-rise or low-cost structures: timber facings made of preservative-treated timbers
stacked and interconnected from behind by plywood boards and nails; and tire facings consisting
of rows of staggered rubber tires filled with compacted backfill material.

Precast elements can be cast in several shapes and provided with facing textures to blend aesthetically into
the environment (Figure 6-7). MSE walls using precast concrete elements as the facings can have surface
fmishes similar to any reinforced concrete structure.

Facings using welded wire or gabions have the disadvantages of an uneven surface, exposed backfill
materials, more tendency for erosion of the retained soil, possible shorter life due to corrosion of the wires,
and more susceptibility to vandalism. These disadvantages can, of course, be countered by providing a
gunite covering or by hanging facing panels on the exposed face and compensating for possible corrosion.
The advantages of such facings, on the other hand, are low cost, ease of installation, good drainage
characteristics (depending on the type of backfill), and possible treatment of the face for vegetative and
other architectural effects. The facing can easily be adapted and well blended with the natural
environment. These facings, as well as geosynthetic wrapped facings, are especially advantageous for
construction of temporary or other structures with a short-term design life.

The recently introduced dry cast modular block (MBW) facings raise some concerns as to their durability
in aggressive freeze-thaw environments because their water absorption capacity can be significantly higher
than that of wet-cast concrete. Historical data provide little insight as their usage history is less than a
decade. Also, because the cement is not completely hydrated during the dry cast process (as is often
evidenced by efflorescence on the surface of units), a highly alkaline regime may establish itself at or near
the face area, and may become an aggressive aging medium for some geosynthetic products that may be
used as reinforcements. Freeze-thaw durability is enhanced for products produced at higher compressive
strengths and/or sprayed with a post-erection sealant.

6 - 15
6.4 THE CONCEPT OF SOIL REINFORCEMENT

6.4.1 Basic Soil Reinforcement Concept

A properly designed reinforced soil structure supports itself as a coherent body. The inclusion of
reinforcement within the soil tends to restrain the soil deformations which, in turn, increases the strength
of the soil and the stability of the composite material.

The stability of a reinforced soil mass is dependent on three fundamental characteristics:


• the soil-reinforcement interaction
• the tensile strength of the reinforcement
• the durability of the reinforcing material

The tensile strength of the reinforcement is discussed in Section 6.5. The durability and long term
performance ofreinforcing materials are discussed in Section 6.9. Following is a discussion of the soil-
reinforcement interaction of MSE systems.

6.4.2 Soil-Reinforcement Interaction

The stress transfer between the soil and the reinforcement takes place through two mechanisms: (1) friction
along the soil-reinforcement interface (Figure 6-9a), and (2) passive soil resistance or lateral bearing
capacity developed along the transverse sections of the reinforcement (Figure 6-9b).

The contribution of each transfer mechanism for a particular reinforcement depends on the roughness of
the surface (skin friction), normal effective stress, grid opening dimensions, thickness of transverse
members, and elongation characteristics of the reinforcement. Equally important for interaction
development are the soil characteristics, including its grain size distribution, particle shape, density, water
content, cohesion and stiffness.

For strip or sheet reinforcement, the interaction between the soil and the reinforcing element is mainly
through friction along the soil-reinforcement interface. Providing ribs on the reinforcing strips results in
the addition of passive soil resistance. In grid-reinforcing systems, the resistance to pullout forces is
provided by both the friction between the soil and the horizontal surface of the grid, and the passive soil
resistance along its transverse members. Passive resistance is considered to be the primary interaction for
rigid geogrids, bar mats, and wire mesh reinforcements.

The soil-reinforcement interaction is simplistically expressed in terms of the pullout resistance developed
by the reinforcement. The pullout resistance of the reinforcement is defined by the ultimate tensile load
required to generate outward sliding of the reinforcement through the reinforced soil mass. The design
of the soil reinforcement system requires an evaluation of the long-term pullout resistance with respect to
three basic criteria as follows:

• Pullout Capacity. 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.
• Allowable Displacement. the relative soil-to-reinforcement displacement required to mobilize the
design tensile force should be smaller than the allowable displacement of the structure.
• Creep. the pullout load should be smaller than the critical creep load that can be tolerated by the
structure. Since soil-reinforcement systems are not generally used with cohesive soils susceptible
to creep, the critical creep load is primarily that of the reinforcing element.

6 - 16
Pull Out Force Normal Pressure

Frictional Force Normal Pressure

(a)

Frictional Resistance
Frictional Resistance
\ Passive Resistance _Pullout
- .=i-J/A
~_Z~~"""'~ff~""'2.,..,~...,,:t ~ .::::.s==~~~==:~~=:::::I
Passive
Resi~tance
L-~~~~~==:::::::D
Normal Pressure
Pullout Force

(b)

Figure 6-9: Mechanisms of Pullout Resistance. (a) By Friction, (b) By Passive Resistance. (After
Christopher, et aI., 1990; Lawson, 1992)

6 - 17
Laboratory and field studies have been conducted to study pullout resistance. A number of different
approaches, methods and evaluation criteria are currently used to estimate the pullout resistance. Herein,
a unified normalized approach presented by Elias and Christopher (1997) is adopted.

6.4.3 Pullout Resistance

The pullout resistance, Pr , of the reinforcement per unit width of reinforcement is given by:

(6-1)

where: F* = the pullout resistance factor


II = a scale effect correction factor
,
avo = the effective vertical stress at the soil-reinforcement interfaces.
Le = the embedment or adherence length of the reinforcement in the resisting
zone behind the failure surface; sometimes the termed resisting length of
the reinforcement (see Figure 6-22).
C = the reinforcement effective unit perimeter, e.g., C=2 for strips, grids and
sheets.

The pullout resistance factor, F*, and the scale effect correction factor, II, can be estimated as follows:

Pullout Resistance Factor, F·

The pullout resistance factor, F*, can be obtained from laboratory or field pullout tests performed on
reinforcing elements embedded within the specific backfill material to be used on the project.
Alternatively, F* can be derived from empirical or theoretical relationships developed for each soil-
reinforcement interaction mechanism and provided by the reinforcement supplier. For any reinforcement,
F* can be estimated using the following general equation:

F* = Frictional Resistance + Passive Resistance


= tan p + Fq • (X~ (6-2)

where: p = the soil-reinforcement interaction friction angle.


Fq = the embedment (or surcharge) bearing capacity factor
II~ = a bearing factor for passive resistance which is based on the thickness per unit width
of the bearing member.

The above equation represents systems that have both the frictional and passive resistance components of
the pull-out resistance. In certain systems, however, one component is much smaller than the other and
can be neglected for practical purposes.

In the absence of site specific pullout testing data, the following semi-empirical relationships may be used
to determine F*. These relationships, when used in conjunction with the standard specifications for backfill
(see Section 6.7), provide a conservative evaluation of the pullout resistance.

Steel Ribbed Reinforcement

(6-3)
1.2 +logC u at top of structure (F· = 2.0 maximum)
F* = tan p ={
tan <I> at depth of 6 m and below

6 - 18
where Cu is the uniformity coefficient of the backfill (°60/°10) and <I> is its friction angle. If the specific
Cu for the wall backfill is not mown at the design time, a Cu of 4 should be assumed for backfills meeting
the requirements of Section 6.7.

Steel Grid Reinforcement

For steel grid reinforcements (bar mat) with spacing between the longitudinal and transverse bars of at
least 150 mm, the pullout resistance factor, F*, is primarily a function of the passive resistance and is
expressed as follows:

20 (D *IS) at top of structure (6-4)


F* = F a. = I
q ~ { 10 (D *1S )
t
at depth of 6 m and below

where 0* is the diameter of the wire or bar and SI is the distance between transverse bars. The transverse
bar spacing, St, shall be uniform throughout the length of the reinforcement rather than having transverse
grid members concentrate only in the resistance zone.

Geogrid Reinforcement

For geogrid reinforcement, the pullout reisistance factor, F*, which is often referred to as an "Interaction
Factor," is primarily frictional and is commonly taken as:

F*=0.8 tan <I> (6-5)

When used in the above relationships, <I> is the peak friction angle of the soil which, for MSE walls using
select granular backfill, is usually taken as 34 degrees.

Geos,ynthetic Sheet Reinforcement

For geosynthetic sheet reinforcement, the pullout resistance factor F* is entirely frictional and is commonly
taken as:

F* = 2/3 tan <I> (6-6)

Scale Effect Correction Factor, a

The correction factor a depends primarily upon the strain softening of the compacted granular backfill
material and the length and extensibility of the reinforcement. In general, a is approximately 1.0 for
inextensible reinforcements, but it can be substantially smaller than 1.0 for extensible reinforcements. The
a factor can be obtained from pullout tests on reinforcements with different lengths, or can be derived
using analytical or numerical load transfer models that have been "calibrated" through numerical test
simulations. In the absence of test data, a=0.6 is recommended for geosynthetic (or extensible)
reinforcements.

6.5 TENSILE STRENGTH OF REINFORCEMENT

6.5.1 Steel Reinforcement

The allowable tensile force per unit width of the reinforcement, Ta, is calculated as follows:

6 - 19
FA c (6-7)
T = FS_Y_
a b

where:

FS = 0.55 for strips and 0.48 for grids


b = the unit width of the reinforcing element (strip, sheet, grid or bar mat)
Fy = yield stress of steel
Ac = design cross section area of the steel, defined as the original cross section area minus
corrosion losses anticipated to occur during the design life of the wall.

For temporary structures (Le., design life of 3 years or less), the allowable tensile stress may be increased
by 40 percent. The global factor of 0.55 applied to F y for permanent structures accounts for uncertainties
in structure geometry, fill properties, externally applied loads, the potential for local overstress due to load
nonuniformities, and uncertainties in long-term reinforcement strength. Safety factors less than 0.55, such
as the 0.48 factor applied to grid members, account for the greater potential for local overstress due to load
nonuniformities for steel grids than for steel strips or bars. The use of hardened and otherwise low strain
steels may increase the potential for catastrophic failure; therefore, a lower allowable material stress may
be warranted with such materials.

Determination of Ac (Design Cross Section)

The parameters needed for determination of Ac for steel strips and grids are shown in Figure 6-10. For
steel reinforcements, the design life is achieved by reducing the cross-sectional area of the reinforcement
used in design calculations by the anticipated corrosion losses over the design life period as follows:

(6-8)

where tc is the thickness of the reinforcement at the end of the design life, to is the nominal thickness at
construction, and ts is the sacrificial thickness of metal expected to be lost by uniform corrosion during the
service life of the structure. For a mildly corrosive backfill material having the controlled electrochemical
property limits (see Table 6-5) the following corrosion rates are suitable for conservative design:

• For zinc coated reinforcements use 15 f-lrnlyear for first 2 years and 4 f-lrnlyear thereafter
• For residual carbon steel use 12 f-lrnlyear.

The designer of a MSE structure should also consider the potential for changes in the reinforced backfill
environment during the service life of the structure. In certain parts of the United States, it can be
expected that deicing salts might cause such an environment change. For such cases, the depth of chloride
infiltration and its concentration are of concern.

6.5.2 Geosynthetic Reinforcement

The initial stress-strain characteristics of geosynthetic reinforcements depend on the geometry and tensile
properties of their load carrying elements. The characteristics of geosynthetic reinforcing elements
manufactured with the same base polymer may vary widely. The tensile capacity of the reinforcement is
further affected by creep, temperature, construction damage, aging and other factors.

6 - 20
Ac = btc where tc is the strip thickness corrected for corrosion loss

1t .2
A = (N ) - D
c b 4
where:
Ac = Cross-sectional area of the reinforcement
Nb = Number of longitudinal bars per unit width b
D* = diameter of bar or wire corrected for corrosion loss.
b = unit width of reinforcement (if reinforcement is continuous count number of bars for
reinforcement width of 1 unit).

where:
Ta = allowable long-term tensile strength of reinforcement (strength/unit reinforcement width)
FS = factor of safety (see Section 6.5.1)
Fy = yield stress of steel
Rc = reinforcement coverage ratio = b/Sh
Use Rc = 1 for continuous reinforcement (Le., ~ =b = 1 unit width)
Tmax = maximum load applied to reinforcement (load/unit wall width)

Figure 6-10: Parameters for Metal Reinforcement Strength Calculations. (Elias and Christopher, 1997)

6 - 21
The ultimate (or yield) tensile strength of the geosynthetic reinforcement is usually determined from the
wide strip tensile strength test (ASTM D 4595). When used in design, however, an allowable tensile
strength is computed by applying reduction factors to the ultimate strength to account for the above-
mentioned factors. The allowable tensile strength is computed as follows:

T ult T u1t (6-9)


T =----:.::..:....-- =-,...----_---=..:.:-----..,.--
a (RF)(FS) ( RF CR X RF D X RF ID ) (FS)

where:

Ta = Allowable (long-term) tensile strength on a load per unit width of reinforcing basis.
Tu!t = Ultimate (or yield tensile strength) from wide strip tensile strength tests (ASTM
D 4595 or GR1:GGl for geogrids), based on minimum average roll value
(MARV) for the product.
RF= Dimensionless product of all applicable reduction factors (RF =RFcR x RFD x
RFID)·
RF cR = Creep reduction factor
RFo = Durability reduction factor
RFID = Installation Damage reduction factor
FS = 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.

The determination of reduction factors for each geosynthetic product requires extensive field and/or
laboratory testing. Elias (1997) described testing requirements and gave guidelines for values to be used
in the absence of product-specific data, as follows:

Creep Reduction Factor, RFCR

This reduction factor is obtained from long term laboratory creep testing. Creep testing is essentially a
constant load test on multiple product samples, loaded to various percentages of the ultimate product load,
for periods of up to 10,000 hours. The creep reduction factor is the ratio of the ultimate load to the
maximum sustainable load within the design life. Typical reduction factors as a function of polymer type
are indicated in Table 6-2:

TABLE 6-2
CREEP REDUCTION FACTORS, RFCR

Polymer TyPe Creep Reduction Factors, RFCB

Polyester 2.5 to 2.0

Polypropylene 5 to 4.0
Polyethylene 5 to 2.5

6 - 22
Durability Reduction Factor, RFD

This factor is dependent on the susceptibility of the geosynthetic to attack by microorganisms, chemicals,
ultraviolet radiation, thermal oxidation, hydrolysis and stress cracking, and can vary from 1.1 to 2.0. In
the absence of test results which document the long-term resistance of these materials to the aging
mechanisms described above, a default reduction factor of 2.0 should be applied.

Installation Damage Reduction Factor, RFID

The placing and compaction of the backfill material against the geosynthetic reinforcement may reduce its
tensile strength. The level of damage for each geosynthetic reinforcement is a variable and is a function
of the weight and type of the construction equipment and the type of geosynthetic material. The installation
damage is also influenced by the lift thickness and type of soil present on either side of the reinforcement.
Where granular and angular soils are used for backfill, the damage is more severe than where softer, [mer,
soils are used.

The installation damage factor, RPm, ranges from 1.1 to 3.0 depending on backfill gradation and product
mass per unit weight. To account for installation damage losses of strength where full-scale product-
specific testing is not available, the following is recommended:

• For geogrid products, a default value of 1.8 is recommended as a reduction factor.


• For geotextiles, a minimum product weight of 270 g/m2 should be specified for reinforcement
applications and a default value of 3.0 is recommended for a reduction factor.

Factor of Safety, FS

This is a global factor of safety which accounts for uncertainties in externally applied loads, structure
geometry, fill properties, potential for local overstress due to load nonuniformity and uncertainties in long-
term reinforcement strength. For ultimate limit state conditions, a FS of 1.5 for permanent walls and 1.2
for temporary walls is recommended (Elias and Christopher, 1997, AASHTO 1997).

6.6 WALL DESIGN

Irrespective of the materials used in construction, the same basic design methods are applied for the MSE
wall systems. In general, the basic design procedure for MSE walls is well established and successfully
applied in practice. The basic design criteria involves satisfaction of both external and internal stability.

6.6.1 External Stability

For external stability, the wall facing and its reinforced backfill are considered as a coherent block
structure with active earth pressure acting on the back face of that block. Figure 6-11 presents a flow chart
of the external stability analysis performed for the MSE walls.

As with conventional gravity and semigravity retaining structures, four potential external failure
mechanisms are usually considered in sizing MSE walls, as shown in Figure 6-12. They include sliding
on the base, limiting eccentricity (or overturning), bearing capacity failure and deep-seated shear failure
(global stability). Due to the flexibility and satisfactory field performance of MSE walls, the adopted
values for the factors of safety for external failure are in some cases lower than those used for reinforced
concrete cantilever or gravity walls described in Chapter 4.

6 - 23
I Define Wall Geometry and Soil Properties I
I
I Select Reinforcement and Backfill I
I
I Preliminary Sizing of the Wall
I
I
I Evaluate External Stability I
I
I I I I
Sliding Eccentricity Deep Seated
Bearing Capacity
FS=1.5 (Overturning) Failure
FS=2.5
FS=2.0 FS=l.3
I I I I
I
I Estimate Settlement and Lateral Deformations I
Figure 6-11: External Stability Analysis Flow Chart

.I
~ (
I
I
I

-
(a) Sliding (b) Eccentricity (Overturning)

(c) Bearing Capacity (d) Deep Seated Stability (Rotational)

Figure 6-12: Potential External Failure Mechanisms for a MSE Wall. (Elias and Christopher, 1997)

6 - 24
The evaluation of external stability proceeds by first performing a preliminary sizing of the wall followed
by the computation of the factors of safety for various potential failure mechanisms. General guidelines
for performing the various aspects of external stability are described below:

Preliminary Sizing

Preliminary sizing of the structure is normally performed to establish the minimum wall height and
reinforcement length for use in the external stability calculations.

The design height of the MSE structure is determined by adding the required embedment to the wall
height. To prevent local bearing capacity failure and damage from frost heave, the minimum embedments
shown in Table 6-3 should be used in the design unless the wall is constructed on a rock foundation. The
wall embedment is determined as a function of the height of the structure, H, above the leveling pad.

A minimum embedment depth of 0.5 m is recommended, except for structures founded on rock at the
surface, where no embedment is required. To reduce the embedment depth and hence the overall wall
height, frost-susceptible soils could be overexcavated and replaced with non frost-susceptible backfill.

TABLE 6-3
MINIMUM EMBEDMENT REQUIREMENTS FOR MSE WALLS

Slope in Front of Structure Minimum Embedment

Horizontal (for walls) H/20


Horizontal (for abutments) H/lO
3H: 1V (for walls) HI 10
2H:1V (for walls) H/7
3H:2V for walls H/5

The minimum reinforcement length should be the greater of O. 7H and 2.5 m, where H is the design height
of the structure (as measured from the top of the leveling pad). Structures with sloping surcharge fills or
other external loads such as abutment footings or surcharges, generally require longer reinforcements for
stability, often on the order of O. 8H to as much as 1.1H. A minimum horizontal bench of 1.2 m should
be provided in front of walls founded on slopes.

Sliding and Limiting Eccentricity (Overturning)

A MSE wall structure must have a minimum factor of safety against sliding of 1.5 and an eccentricity of
the resultant force less than L/6 in soil and L/4 in rock where L is the length of the reinforcement (Elias
and Christopher, 1997).

The maximum permissible eccentricity is an alternative form of the conventional overturning criteria
discussed in Chapter 4. Conventionally, the structure is checked for the possibility of overturning and a
minimum factor of safety of2.0 against overturning is usually required (AASHTO, 1997). For most MSE
walls, this is generally achieved if the eccentricity is within the maximum permissible limits as specified
above. However, there may be specialized cases, such as if a steep back slope is used behind the wall,
or a lightweight fill is used in the reinforced zone, where the factor of safety against overturning may be

6 - 25
lower than 2.0 even though the condition of limiting eccentricity is satisfied. In such specialized cases,
it may be necessary to check the factor of safety against overturning. Since the values (vertical and
horizontal forces and the geometry of the structure) required for determination of the factor of safety
against overturning are determined anyway for computation of other safety factors, it would be an easy
and straightforward exercise to determine the factor of safety against overturning.

Figures 6-13 through 6-15 show procedures for external stability analyses for common MSE wall
configurations, including the computation of safety factors for sliding and overturning (if necessary). The
eccentricity computations for these cases are shown in Figures 6-16 and 6-17. Note that the effect of
external loadings on the MSE mass, which increases sliding resistance, should only be included if the
loadings are permanent. For example, live traffic load surcharges should be excluded. The computations
should be repeated by increasing the reinforcement length until the required safety factors are achieved.

Bearing Capacity

The allowable bearing capacity should be computed using a minimum factor of safety of 2.5 for AASHTO
Group I loading applied to the calculated ultimate bearing capacity. The length of the reinforcement at the
foundation level is considered as the footing width for ultimate bearing capacity calculations. The bearing
pressure is computed using the Meyerhoff distribution, which considers a uniform base pressure
distribution over an effective base width of B' =L-2e, as shown in Figures 6-16 and 6-17. Sometimes "B"
is used in lieu of "L" in the calculations.

The procedures for determining the bearing capacity are described in Module 7 - Shallow Foundations and
AASHTO (1997). FHWA (Elias and Christopher, 1997) recommends that the effect of wall embedment
be neglected in the computation of the ultimate bearing capacity to account for the possibility of future
excavations in front of the wall.

Where soft soils are present, or if the MSE wall is built on sloping ground, the difference between the bearing
stress beneath the reinforced soil zone and that beneath the facing elements should be considered while
evaluating the bearing capacity. This is especially important where heavier concrete wall facings are used.
Furthermore, the differential settlements between the facing elements and the reinforced soil zone could
create concentrated stresses at the connections between the facing elements and the reinforcements. In such
cases, the leveling pad should be embedded adequately to meet the bearing capacity and settlement
requirements, or dimensioned to keep the bearing stresses beneath the leveling pad and the remainder of the
wall as uniform as possible.

In general, the vertical stress can be decreased and the ultimate bearing capacity increased by lengthening
the reinforcements. If an adequate support condition cannot be achieved, or lengthening the reinforcements
significantly increases the cost, improvement of the foundation soil should be considered. Ground
improvement techniques such as dynamic compaction, soil replacement, stone columns, preloading, etc.,
are studied in Module 4 - Ground Improvement.

Deep Seated (Global) Stability

Deep seated or global stability is determined using either a rotational or a wedge analysis, as appropriate,
which can be performed using classical slope stability analysis methods. These methods are described in
Module 3 - Soil Slopes and Embankments. The reinforced soil block is considered as a rigid body and
only the failure surfaces which are completely outside the reinforced mass are considered in the stability
analyses. A minimum safety factor of 1.3 is recommended for global stability. If the calculated safety
factor is less than 1.3, the reinforcement length should be increased, or the in situ soil should be improved.

6 - 26
HQrizQntal Backslo.pe with Traffic Surcharge
q Assumed for bearing capacity

q=
and overall (global) stability comps.
Assumed for overturning
sliding & pullout resistance comps.

Retained Fill
Reinforced CPr Yr Kof
Soil Mass
1~,y,K.
V 1=y,HL F2 = qHK af
H
1 2
FI = -Y fH K af
. .+----......Iot~ 2
H/2
H/3

where: e= Eccentricity
R= Resultant ofvertical forces (Vl+qL)

:1 1 - sin <l>r
K =----'-
of 1 + sin <l>r

FQr Limiting Eccentricity CalculatiQns, see Figure 6-16

FactQr Qf Safety Against Overturning if Necessary (MQments abQut PQint 0):

FS _ ~Moments Resisting (Mr) = V 1(L/2) ~ 2.0


o ~MomentsOverturning(Mo) F 1 (H/3)+F 2 (H/2)

FactQr Qf Safety Against Sliding:

~Horizontal Resisting Force(s) VI (Tan p or Tan <1>*)


FS = ~ 1.5
s ~HorizontaIDrivingForce(s) F 1 + F2

<I> = FrictiQn Angle Qf ReinfQrced Backfill Qr FQundatiQn, whichever is IQwer


where q = Traffic Live LQad
*Tan p is fQr CQntinUQUS SQil reinfQrcements (e.g., grids and sheets). FQr discQntinuQus SQil
reinfQrcements (e.g., strips) use Tan <1>. p is the SQil/ reinfQrcement interface frictiQn angle. Use the
IQwer Qf Tan <I> at the base Qf the wall Qr Tan p at the IQwest reinfQrcement layer fQr CQntinUQUS
reinfQrcements.
NQte: FQr relatively thick facing elements (e.g., segmental CQncrete facing blQcks) it may be
desirable tQ include the facing dimensiQns and weight in sliding and Qverturning calculatiQns (i.e. use
"B" in lieu Qf "L").

Figure 6-13: External Analysis: Earth Pressures/Eccentricity - HQrizQntal BackslQpe with Traffic
Surcharge. (Elias and ChristQpher, 1997; AASHTO, 1997)

6 - 27
Slopin~ Backslope Case

-r-.------ 2/3 L
Retained Fill
h-H ,- cl>fYfK.f

Reinforced
Soil Mass
cl>r Yr K.r
h

h/3

L
B :1
Compute ~ for retained fill using 0 = ~ (see Chapter 2; Figure 2-2)
For Limiting Eccentricity Calculations, see Figure 6-17
Factor of Safety Against Overturning if Necessary (Moments about Point 0):

:EMoments Resisting (Mr) _ V 1(Ll2) +V 2(2L13) +F v(L)


FS o - ~ 2.0
:EMoments Overturning (Mo) F H (h/3)

Factor of Safety Against Sliding:

:EHorizontal Resisting Force(s) (V 1 +V 2 +F )(Tan p or Tan 4>*)


FS = v ~ 1.5
S :EHorizontal Driving Force(s) FH

4> = Friction Angle of Reinforced Backfill or Foundation, whichever is lower.


*Tan p is for continuous soil reinforcements (e.g., grids and sheets). For discontinuous soil
reinforcements (e.g., strips) use Tan 4>. p is the soil/ reinforcement interface friction angle. Use the
lower of Tan 4> at the base of the wall or Tan p at the lowest reinforcement layer for continuous
reinforcements.

Note: For relatively thick facing elements (e.g., segmental concrete facing blocks) it may be
desirable to include the facing dimensions and weight in sliding and overturning calculations (i.e. use
"B" in lieu of "L").
Figure 6-14: External Analysis: Earth Pressures/Eccentricity - Sloping Backfill Case. (Elias and
Christopher, 1997; AASHTO, 1997)

6 - 28
2H

- - -Retained Fill
I3cs

h-H 4>fYf K .f

Reinforced
SoU Mass
4> r y r K. r

h
H j
V 1 =Yr HL

£ h/3

:I
FH = FT cos (I) Fv = FT sin (I) For infinite slope ~cs =~
Compute Kar for Retained Fill Using o=~=~cs (see Figure 2-2); For Limiting Eccentricity
Calculation, see Figure 6-18
K = sin 2 (8
-::__ +"')
"--........:.'I'-'-- _:=__

a ~
sin ( <I> + 0) sin ( <I> _ cs ) ] 2
sin 2 8 sin ( 8 - 0) 1 +
sin(8 -o)sin(8 +~cs)

Factor of Safety Against Overturning (Moments about Point 0):


~Moments Resisting (Mr) V I (Ll2) + V2(2L13) + F (L)
FS o = v ~2.0
~Moments Overturning (Mo) F H (h/3)

Factor of Safety Against Sliding:

~Horizontal Resisting Force(s) (V I + V 2 + F )(Tan p or Tan <1>*)


FS = v ~ 1.5
s ~Horizontal Driving Fqrce(s) FH

<I> = Friction Angle of Reinforced Backfill or Foundation, whichever is lower.


*Tan p is for continuous soil reinforcements (e.g., grids and sheets). For discontinuous soil
reinforcements (e.g., strips) use Tan <1>. p is the soil/ reinforcement interface friction angle. Use the
lower of Tan <I> at the base of the wall or Tan p at the lowest reinforcement layer for continuous
reinforcements .

Note: For relatively thick facing elements (e.g., segmental concrete facing blocks) it may be
desirable to include the facing dimensions and weight in sliding and overturning calculations (i.e. use
"B" in lieu of "L").

Figure 6-15: External Stability for Wall with Broken Backslope. (Elias and Christopher, 1997)

6 - 29
q

Reinforced Retained Fill


Soil Mass <l>t Yt Kat
<1>, Yr Ka,

H/2
H/3

where: q = Traffic Live Load


R= Resultant of Vertical Forces (V 1 + qL)
:1

Summing Moments about Point C:

F I (H/3) + F 2 (H/2)
e =
VI + qL

VI +qL
°v=
L -2e

If additional concentrated dead loads occur, such as those illustrated in figures 6-16 and 6-17, the
external forces resulting from those dead loads should be added to the earth pressure forces shown
above by superposition.

Note: For relatively thick facing elements (e.g., segmental concrete facing blocks) it may be desirable
to include the facing dimensions and weight in sliding and overturning calculations (Le. use "B" in
lieu of "L").

Figure 6-16: External Analysis: Calculation of Vertical Stress at the Foundation Level for Bearing
Capacity Calculations - Horizontal Backslope Condition. (1997 AASHTO Interims)

6 - 30
Retained Fill
<l>f YrK.r

Reinforced
Soil Mass
<I>,Y,K,

h
£
H

1
h/3

Compute Kat for retained fill using 0 = P(see Chapter 2; Figure 2-2)
Summing Moments about Point C:
FT(cosp)h/3 -F (sinp)Ll2 -V (Ll6)
T 2
e =---"--------------
VI +V 2 +FTsinp

VI +V 2 +FTsinp R = Resultant of vertical forces


av=~-~-~-
L -2e
.N.Q!.§:
1. If concentrated dead load surcharges present, such as those illustrated in Figures 6-16 and 6-17,
include effect of those surcharges. If broken backslope condition exists, design using equivalent slope
"I" as shown in Figure 6-15. H is the total wall height at the face.

2. For relatively thick facing elements (e.g., segmental concrete facing blocks) it may be desirable
to include the facing dimensions and weight in sliding and overturning calculations (Le. use "B" in
lieu of "L").

Figure 6-17: External Analysis: Calculation of Vertical Stress at the Foundation Level for Bearing
Capacity Calculations - Sloping Backslope Condition (Elias and Christopher 1996)

6 - 31
z rill. ~I' .1\\l2
2 VL-----!.---~~
/ \
/ \
/ \
/ \
/ \
For z ~ Zl
\
\
D 1 = br +2z
2
= b +
r
Z

For z> zl For strip load: Q


b.O = -.l!
v D1
D = br + z +d
1 2 For isolated footing load:
A
- Qv'
b. v - D1(L+z)
For point load: Q'
b. 0v = 'I)'2 with b r = 0
1

Where: D1 = Effective width of applied load at any depth, calculated as shown above
bf = Width of applied load. For footings which are eccentrically loaded (e.g., bridge
abutment footings), set bf equal to the equivalent footing width B' by reducing it by
2e', where e' is the eccentricity of the footing load (i.e., bf - 2e').
Lf = Length of footing
Qv = Load per linear meter of strip footing
Qv' = Load on isolated rectangular footing or point load
Zl = Depth where effective width intersects back of wall face = 2d1 - b f
Assume the increased vertical stress due to the surcharge load has no influence on stresses used to
evaluate internal stability if the surcharge load is located behind the reinforced soil mass. For external
stability, assume the surcharge has no influence if it is located outside the active zone behind the wall.

Figure 6-18: Distribution of Stress from Concentrated Vertical Load. (Elias and Christopher, 1997)

6 - 32
F I = lateral force due to
earth pressure
Stress F2 = lateral force due to
Distribution traffic surcharge

PHI = lateral force due to


superstructure or other
concentrated lateral loads
L
e'=eccentricity of load on footing

(a) Distribution of Stress for Internal Stability Calculations

L 2=(Cr+b f - 2e') tan(45+4>/2)

br -2e'.1 J
/ /
/ /
' \45{4> 12
/ I P m = lateral force due to
/ / superstructure or other
concentrated later& loads
,/ /
------V / '\45+4>/2
/ Iffooting is located completely outside
active zone behind wall, the footing load
does not need to be considered in the
external stability calculations.

(b) Distribution of Stress for External Stability Calculations

Figure 6-19: Distribution of Stress from Concentrated Horizontal Loads for External and Internal
Stability Calculations. (Elias and Christopher, 1997)

6 - 33
Impact of Surface Loads

In addition to the uniform surcharge loads shown on Figures 6-13 and 6-16, any concentrated vertical and
horizontal loads at the surface may influence both the internal and external stability of the wall. The lateral
stresses from such loads can be calculated as described in Section 2.6. These stresses can also be
approximated using a simplified approach illustrated in Figures 6-18 and 6-19 shown on the previous two
pages. If such loadings occur, the resulting forces should be appropriately incorporated in the stability
calculations.

6.6.2 Internal Stability

To be internally stable the MSE structure must be coherent and self supporting under the action of its own
weight and any externally applied forces. This is accomplished through stress transfer from the soil to the
reinforcement as described in Section 6.4. Basically, the internal stability analsyis involves determination
of (a) the tensile strength and (b) the pullout resistance of the reinforcing element.

Figure 6-20 illustrates the internal failure mechanisms for MSE walls. At each level, the reinforcement
must be sized and spaced to preclude rupture under the stress it is required to carry and to prevent pullout
from the soil mass. The process of sizing and designing to preclude internal failure, therefore, consists
of determining the maximum developed tension forces in the reinforcements, their location along a critical
failure surface and the resistance provided by the reinforcements both in pullout capacity and tensile
strength.
Failure Surface Failure Surface

~~ :A:
.,·.t-------rl --l
",__---'r1 r.
'.~
~;
l:,~----;
~I--_ _ -+ __ J
-
~.~
~~~---f / - -.... -
~i
~~I----+-1 I
J
:"& ~:f.
.... I
;:-'1-----1 /----1 ~!'1----+-+---1
't
it-----I 1 - - - - , t----+-J"-----1
.' I
f:;t----II-----; ,1-----,f-+----1
~~~
t---,<~--_J
I
.
~ '47.,~~--------'

(a) (b)
Figure 6-20: Mechanisms of Internal Failure in MSE Walls. (a) Tension Failure, (b) Pullout Failure.
(Christopher, et al., 1990)

Figure 6-21 illustrates a flow chart of the internal stability analysis. Following is a discussion of the main
steps involved in the analysis:

Step 1: Determine Critical Failure Surfaces

The most critical failure surface in a simple MSE wall is assumed to coincide with the maximum tensile
forces line (i.e., the locus of the maximum tensile force, Tmax , in each reinforcement). The critical failure
surface separates the reinforced soil into two zones: an active zone where the shear stresses in the
reinforcement are directed towards the wall face, and a resistant zone where the shear stresses are directed
away from the wall face.

6 - 34
I Determine Critical Failure Surface I
I
I Determine Lateral Pressure Coefficient, K I
I
I Detennine Maximum Tensile Force in Reinforcement, Tmax I
I
I Detennine Tensile Strength of Reinforcement, Ta I
I
I I
I Inextensible Reinforcement (Metal) I Extensible Reinforcement (Geosynthetic)
I I
I Estimate Corrosion Loss, ts I Determine Reduction Factors
(Creep, Durability, Installation)
I
I Determine Design Cross Section, Ac I I
I Select Overall Factor of Safety, FS
IDetermine Allowable Tensile Force, Ta I I
Detennine Allowable Tensile Force, Ta
I I
I
I Detennine Pullout Resistance, Pr I
I
I I
I Inextensible Reinforcement (Metal) I Extensible Reinforcement (Geosynthetic)
I I
IDetermine Pullout Resistance Factor, F*I Determine Pullout Resistance Factor, F*
I I
I I I I
I Steel Strips
II Steel Grids
I Geogrids
L...---"T',-._---I
II Geotextile Sheets
I
I I
I I

I Determine Scale Effect Factor, a I Determine Scale Effect Factor, a


I I
I Determine Pullout Resistance, Pr
I Determine Pullout Resistance, Pr

I I
I
I Determine Tensile Stress at Connection, To I
I
I Design Facing Element I
Figure 6-21: Internal Stability Analysis Flow Chart

6 - 35
The shape and location of the maximum tensile force line can be estimated for simple structures from a
large number of previous experiments and theoretical studies. This line passes through the toe of the wall
and is approximately bilinear for inextensible reinforcements (Figure 6-22a), and linear for extensible
reinforcements (Figure 6-22b).

Step 2: Detennine Maximum Tensile Force in Reinforcement

The maximum tension in the reinforcing element at a given level can be computed as follows:

T max =Ka' vo s for sheet type reinforcement (6-10)


T
max
=K a'vo s v s h for strip type reinforcement (6-11)

where K is the lateral earth pressure coefficient, a/yO is the vertical soil stress at the reinforcement level,
s is the vertical spacing of sheet type reinforcement, Sy is the vertical spacing of strip type reinforcements,
and ~ is the horizontal spacing of strip type reinforcements. For preliminary analyses, spacings of 0.3 to
0.6 m can be used.

The lateral earth pressure coefficient K varies with the type (Le., extensibility) of the reinforcement
(Figure 6-23). If geosynthetic reinforcements are used, the active earth pressure coefficient K. is
customarily used in the deign. For metallic reinforcements, K decreases in value from (1.7 - 2.5) K. at
the surface to 1.2 K. at a depth of 6 m below the surface. It remains constant below that depth.

The vertical soil stress, a/yO' at each reinforcement level can be computed as follows:

(6-12)

where Yr is the unit weight of the retained fill, z is the depth to the reinforcement layer under
consideration, q is the uniform surcharge load (if present) and !!:..ay is the stress due to the concentrated
surcharge loads (if present). For sloping soil surfaces above the MSE wall section refer to AASHTO
(1997) or Elias and Christopher (1997). For cases where concentrated vertical loads occur refer to Figures
6-18 and 6-19 or Chapter 2 for computation of !!:..ay •

Step 3: Detennine Tensile Strength of Reinforcement

The allowable tensile capacity per unit width of the reinforcement, T., is calculated as discussed in Section
6.5, using Eq. 6-7 for steel reinforcements and Eq. 6-9 for geosynthetic reinforcements.

Step 4: Check for Internal Stability Against Reinforcement Breakage

At each level, the reinforcement must have adequate tensile strength to sustain the tensile force Tmax •
Accordingly, the internal stability with respect to breakage is determined as follows:

T max ~T Rc (6-13)
a

where Rc is the coverage ratio, b/~, with b the gross width of the reinforcing element, and ~ the center-to-
center horizontal spacing between reinforcements (e.g., Rc = 1 for full coverage reinforcement).

6 - 36
I""O.3HI*~1
Zone of maximum stress
or potential failure surface~ H = H + tan13 xO.3H
I 1- 0.3 tan 13

* If wall face is battered,


an offset of O.3H I is still
required, and the upper
portion of the zone of
Active Resistant maximum stress should
H
Zone
be parallel to the wall face

HI
2

I... L
(a)
Zone of maximum stress
or potential failure surface

For vertical face


\\I = 45 +-<I>
2
H

For walls with a face batter angle (6) 10 0 or more from the vertical,
-tan(cj> -13) + .Jtan(cj> -13)[tan(cj> -13) + cot(cj> + 8 - 90)][1 + tan(o + 90 - 8)cot(cj> + 8 - 90)]
tan(\jI - 8) = _ _.:...:.::..._--=-----=-_.:....:....-_ _...:..:..J
_--=....:..--=-.:--....!..---=....:..--=-~---:..:..--=-.:--_.:....:....-

1 + tan(o + 90 - 8)[tan(<j> -13) + cot(<j> + 8 - 90)]


with 0= P
6 = wall batter angle (b)
For wall with a broken backslope, use o=13cs as shown in Figure 6-15.

Figure 6-22: Location of Potential Surface for Internal Stability Design of MSE Walls. (a) Inextensible
Reinforcements, (b)Extensive Reinforcement. (Elias and Christopher, 1996)

6 - 37
o KIKa
o

1.0 1.2
*Does not include,polymer strip reinforcement

Figure 6-23: Design Values of the Lateral Earth Pressure Coefficient, K, for Various Types of Soil
Reinforcement Systems.

Step 5: Determine PuIl-out Resistance

The pull-out resistance of the reinforcement, Pr , is calculated according to Eq. (6-1) in Section 6.4.3.

Step 6: Check for Internal Stability Against Pullout

Stability with respect to pullout of the reinforcement requires that the following criteria be satisfied:

P (6-14)
T < __f _ R
max - c
FS po
where:
Pr = Pull-out resistance (Eq. 6-1)
FSPO = safety factor against pullout
Rc = Coverage Ratio
A minimum safety factor of 1.5 is recommended for stability against pull-out. Note that traffic loads are
not included in Tmax in pullout calculations as indicated in Figure 6-13.

If the pullout criterion is not satisfied for all reinforcement layers, the reinforcement length has to be
increased and/or reinforcement with a greater pullout resistance per unit width must be used.

6.6.3 Required Reinforcement Length

If the internal stability against pull-out is not satisfied, the required embedment length of the reinforcement
in the resistant zone (Figure 6-22) should be increased. The minimum length can be detennined by

6 - 38
substituting Eq. (6-1) in Eq. (6-4) which results in the following equation:

(6-15)
~ 1 m

The total length of reinforcement, Lt , required for internal stability is then determined from:

(6-16)

where La is obtained from Figure 6-22 for simple structures not supporting concentrated external loads
such as bridge abutments.

For construction ease, a fInal uniform length is commonly chosen, based on the maximum length required.
However, if internal stability controls the length, it could be varied from the base, increasing with the
height of the wall to the maximum length requirement based on a combination of internal and external
stability requirements.

6.6.4 Connection Design

Different types of connections are used in MSE walls depending basically on the type of reinforcement
used and the wall facing details. The connection of the reinforcement with the facing must be checked to
ensure that the tensile force at the connection is not greater than the allowable tensile strength of the
connection. The capacity of the connection is usually tested according to standards established by
AASHTO (1992), Article 8.31. Connection details and their strength requirements are discussed by Elias
and Christopher (1997).

6.6.5 Face Design Considerations

In the selection of wall facing, the following issues should be considered:

• The units should be sufficiently durable to survive the design life of the structure.
• The units should be sufficiently flexible to tolerate differential movements without structural distress.
• The compressive strength of the units in a particular layer should be able to sustain the weight of
the units above.
• The choice of shapes of the facing units may be dictated by the shape of the wall, e.g, the batter of
the wall face, and the concave and/or convex curves along its length.
• The units should provide adequate mechanisms to properly anchor the reinforcements.
• In welded wire, expanded metal, or similar facings, the units must be designed to prevent excessive
bulging as the backfIll material is compacted.
• In modular block facings, sufficient inter-unit shear capacity must be available to resist the
horizontal earth pressure at the facing with a safety factor of 2.0.
• The geosynthetic facing units should not be exposed to sunlight (ultraviolet radiation) and to
potential contact with harmful chemicals.
Design of the wall facing is discussed further in Elias and Christopher (1997).

6 - 39
6.6.6 Wall Settlements

MSE structures have significant tolerances, both longitudinally along the wall and perpendicular to its front
face. However, conventional settlement analyses should be carried out to ensure that immediate, primary
consolidation, and secondary consolidation settlements of the wall are less than the performance
requirements of the project. Module 7 - Shallow Foundations, describes various procedures to compute
settlements. Significant total settlements at the end of construction would indicate that the planned top-of-
wall elevations need to be adjusted. This can be accomplished by increasing the top of the wall elevations
during design, but more practically, by delaying the casting of the top row of panels to the end of erection.
The fmal required height of the top row would then be determined with possible further allowance for
continuing settlements.

The allowable differential settlement of MSE walls is limited by the longitudinal deformability of the facing
and the ultimate purpose of the structure. Limiting tolerable differential settlements for systems with
common size panels are presented in Table 6-4.

TABLE 6-4
RELATIONSmP BETWEEN JOINT WIDTH AND LIMITING DIFFERENTIAL
SETTLEMENT FOR MSE WALLS (After Elias and Christopher, 1996)

Joint Width Limiting Differential Settlement

20mm 11100
13mm 11200
6 mm 11300
Note: Table applicable for panel size less than 2.8 square meters

MSE walls constructed with full height facing units should be limited to differential settlements of 11500.
Walls with drycast facing (modular block walls) should be limited to settlements of 11200. For walls with
welded wire facings, the limiting differential settlement should be 1150.

Where significant differential settlements are anticipated (exceeding 11100), sufficient joint widths and/or
slip joints must be provided to preclude panel cracking. This in tum may influence the type and design
of the facing panel selected. Square panels generally adapt to larger longitudinal differential settlements
better than long rectangular panels of the same surface area.

Differential settlement of the wall face with respect to the reinforced soil will lead to an increase in the
reinforcement stress near the wall face. This stress increase can be significant for rigidly connected
reinforcements since bending stresses will develop. Where significant differential settlements perpendicular
to the wall face are anticipated, the reinforcement connection must allow for vertical movement.

Although uniform settlement is of little consequence to the stability of the MSE structure, the total
settlement may influence the serviceability of the wall, such as in the case of a bridge supported by a MSE
abutment. The total settlement is particularly important when a MSE wall interacts with an adjacent
structure that has different load-deformation characteristics, e.g., when an abutment is supported on piles
and the approaches utilize MSE walls. In such cases, downdrag forces on the abutment piles due to
settlement of the adjacent MSE structure need to be evaluated.

6 - 40
6.6.7 Lateral Wall Displacements

No method is presently available to definitely predict lateral displacement of MSE walls. In general, the
lateral deformations in the reinforced backfill usually occur during construction. Post construction
movements, however, may take place due to post construction surcharge loads or long-term settlement of
the foundation soils.

The lateral displacements depend on compaction effects, reinforcement extensibility, reinforcement length,
reinforcement-to-facing connection details, and details of the wall facing. In general, increasing the length-
to-height ratio of reinforcement, from its theoretical lower limit of 0.5H to 0.7H, decreases the
deformation by 50 percent. Also, the deformation of MSE structures constructed with polymeric
reinforcements (extensible) is much greater than if constructed with metallic reinforcements (inextensible).

The probable lateral displacements of MSE walls can be estimated based on empirical correlations (Elias
and Christopher, 1997). For critical structures requiring precise tolerances, such as bridge abutments,
more accurate calculations using the finite element method may be warranted.

6.7 Backfill Material

The backfill material for a MSE wall has a major impact on:

• Short-term stability during construction


• Long-term stability and deformation of the completed structure

MSE walls require high quality backfill for durability, good drainage, constructability, and good soil-
reinforcement interaction. This can be obtained from well graded, properly compacted, granular materials.
Whether the MSE system depends on friction between the reinforcing elements and the soil, or on passive
pressure on reinforcing elements, a fill material with high friction characteristics is desirable.

Considering only reinforcement capacity, lower quality backfills could be used for MSE structures;
however, a high quality granular backfill has the advantages of being free draining, providing better
durability for metallic reinforcement, and requiring less reinforcement. There are also significant
handling, placement, and compaction advantages in using granular soils. These include an increased rate
of wall erection and improved ability to maintain of wall alignment tolerances.

Table 6-5 presents the gradation limits and other backfill requirements recommended by FHWA (Elias and
Christopher, 1997) for MSE walls. For metallic reinforcements, the electrochemical requirements are
particularly important to provide the reinforcing elements with increased protection against corrosion; the
reinforcements for a MSE wall must be designed for the maximum corrosion rate associated with these
material properties. Where geosynthetic reinforcements are planned, the electrochemical criteria varies
depending on the polymer.

The fill material must be free of organic matter and other deleterious substances, as these materials not
only enhance corrosion but also result in excessive settlements. The compaction specifications are
discussed in detail in Section 6.8.3. Lighter compaction equipment is usually used near the wall face
(within 1 m) to prevent the buildup of high lateral pressures from the compaction and to prevent facing
panel movement. Because of the use of lighter equipment, a backfill material of good quality in terms of
both friction and drainage, such as crushed stone, is recommended close to the face of the wall to provide
adequate strength and tolerable settlement in this zone. Additional information on backfill compaction and
its impact on the lateral earth pressures is included in Chapters 2 and 3.

6 - 41
The reinforced fill criteria outlined above represent materials that have been successfully used throughout
the United States and have resulted in excellent structure performance. For MSE walls, a lower bound
frictional strength of 34 degrees would be consistent with the specified fill, although some nearly uniform
fine sands meeting the specifications limits may exhibit friction angles of31 to 32 degrees. Values higher
than 34 degrees may be used if substantiated by laboratory direct shear or triaxial test results on the site-
specific material used or proposed.

TABLE 6-5
BACKFILL REQUIREMENTS FOR MSE WALLS (Elias and Christopher, 1997)

All backfill material used in the structure volume must be reasonably free from organic or otherwise
deleterious materials and shall conform to the following gradation limits:

Sieve Size Percent Passing


102 mm (4 inches) 100
0.425 mm (No. 40 mesh sieve) 0-60
0.075 mm (No. 200 sieve) 0-15

The backfill shall also conform to the following additional requirements:

1. For geosynthetics, and epoxy and PVC coated reinforcements, the maximum particle size shall be
limited to 19 mm unless tests are or have been performed to evaluate the extent of post construction
damage anticipated for the specific fill material and reinforcement combination.

2. The plasticity index (PI) as determined by AASHTO T-90 shall not exceed 6.

3. Soundness: The material shall be substantially free of shale or other soft, poor durability particles.
The material shall have a magnesium sulfate soundness loss (or an equivalent sodium sulfate value)
of less than 30 percent after four cycles when tested in accordance with AASHTO T-104.

4. Following are the recommended electrochemical requirements for backfills when using steel
reinforcements:

Prwerty Criteria Test Method


Resistivity > 3000 ohm-cm AASHTO T-288-91
pH 5 to 10 AASHTO T-289-91
CWorides < 100 PPM AASHTO T-291-91
Sulfates < 200 PPM AASHTO T-290-91
Organic Content 1 % max AASHTO T-267-86

5. Following are the recommended electrochemical requirements for backfills when using geosynthetic
reinforcements:
Base Polymer Property Criteria Test Method
Polyester (PET) pH 3 to 9 AASHTO T-289-91
Polyolefin (PP, HDPE) pH >3 AASHTO T-289-91

6 - 42
6.8 WALL CONSTRUCTION

Construction of MSE walls involves placement of alternating layers of reinforcing elements and compacted
backfill behind a wall facing. Although the construction sequence is basically the same for most MSE wall
types, certain construction details may differ from one system to another due to differences in the
reinforcing elements, the wall facings, the labor/equipment requirement and the experience of the specialty
contractors. Construction of two types of MSE walls are presented herein. One wall type uses strip
reinforcing elements and precast concrete facing panels. The other uses sheet reinforcement and gunite
facing, such as in a geotextile retained earth wall. Construction details for other walls can be obtained
from their vendors.

6.8.1 Construction of MSE Walls with Strip Reinforcements and Precast Facing

Construction of a MSE wall with precast facing involves four major tasks:

• Site preparation including overexcavation


• Construction of leveling pad
• Wall construction including placement of wall facing, reinforcement and backfill
• Concrete coping or traffic barrier

The basic construction procedures are illustrated in Photos 6-12 through 6-17 and discussed below:

• Preparation of Subgrade. This step involves removal of unsuitable materials from the area to be
occupied by the retaining structure. All organic matter, vegetation, slide debris and other unstable
materials should be stripped off and the subgrade compacted.
In unstable foundation areas, ground improvement methods, such as dynamic compaction, stone
columns, preloading, or other foundation stabilization/improvement methods would be employed
prior to wall erection (see Course Module 4 - Ground Improvement).
• Placement of a Leveling Pad for the Erection of the Facing Elements. This generally
unreinforced concrete pad is often only 300 mm wide and 150 mm thick. A gravel pad is sometimes
substituted for the concrete pad in MBW construction. The purpose of this pad is to serve as a guide
for facing panel erection and is not intended as a structural load-support foundation.
• Erection of the First Row of Facing Panels on the Prepared Leveling Pad. Setting the first row
of facing elements is a key detail. Construction should always begin adjacent to any existing
structure and proceed toward the open end of the wall.
Facings may consist of either precast concrete panels, metal facing panels, or dry cast modular
blocks. The first row of facing panels may be full, or half-height panels, depending upon the type
of facing used. The first tier of panels must be shored up to maintain stability and alignment. For
construction with modular dry-cast blocks, full sized blocks are used throughout with no shoring.
The first course of panels should be set directly on the leveling pad. Horizontal joint material or
wooden shims should not be permitted between the first course of panels and the leveling pad.
Temporary wood wedges may be used between the first course of panels and the leveling pad to set
the panel batter, but they must be removed during subsequent construction. Some additional
important details are:
i. For segmental panel walls, panel spacing bars, which set the horizontal spacing between panels,
should be used so that subsequent panel rows will fit correctly.

6 - 43
Photo 6-12

Photo 6-13

Photo 6-14

Photos 6-12 to 6-14: Construction of MSE Wall with Inextensible Strip Reinforcements.
Photo 6-12: Concrete Pad Construction.
Photo 6-13: Erection of Facing Elements.
Photo 6-14: Placement of Reinforcements

6 - 44
Photo 6-15

Photo 6-16

Photo 6-17

Photos 6-15 to 6-17: Construction ofMSE Wall with Inextensible Strip Reinforcements.
Photo 6-15: Placement of Backfill.
Photo 6-16: Spreading of Backfill.
Photo 6-17: Backfill Compaction.

6 - 45
ii. The first row of panels must be continuously braced until several layers of reinforcements and
backfill have been placed. Adjacent panels should be clamped together to prevent individual
panel displacement.

iii. After setting the batter of the first row of panels, the horizontal alignment should be visually
checked with survey instruments or with a stringline.

iv. When using full height panels, the initial bracing alignment and clamping are even more critical
because small misalignments cannot be easily corrected as construction continues.

v. Most MSE systems use a variety of panel types on the same project to accommodate geometric
and design requirements (geometric shape, size, finish, connection points). The facing element
types must be checked to make sure that they are installed exactly as shown on the plans.

• Placement and Compaction of Backf"ill on the Subgrade to the Level of the First Layer of
Reinforcement and its Compaction. (See Section 6.8.3)

• Placement of the First Layer of Reinforcing Elements on the Backf"ill. When the fill has been
brought up to the level of the connection, the reinforcements are placed and connected to the facing
panels. The reinforcements are generally placed perpendicular to the back of the facing panels;
however, structure geometry or the presence of obstructions, such as abutment piles, may require
angled connections. In all cases, overlapping layers of reinforcements should be separated by a 75
mm minimum thickness of wall fill.

• Placement of Subsequent Facing Courses (Modular Facings). Throughout construction of


modular panels walls, facing panels should only be set at grade. Placement of a panel on top of one
not completely backfilled should not be permitted.

Walls using precast concrete panels require bearing pads in their horizontal joints that provide some
compressibility and movement between panels, and preclude concrete to concrete contact. These
padding materials are either neoprene or SBR rubber. The compressibility of the horizontal joint
material should be a function of the wall height. Walls with heights greater than 15 m may require
thicker or more compressible joints in the lower portion of the structure to accommodate the larger
vertical loads due to the weight of the panels.

Vertical joints, if large, may be filled with synthetic foam. All joints are covered with geotextile
strips to prevent the migration of fmes from the backfill. The erection of facing panels and
placement of the soil backfill proceed simultaneously.

• Placement of the Backfill over the Reinforcing Elements to the Level of the Next Reinforcement
Layer and Compaction of the Backf"ill. (See Section 6.8.3)

• Placement of Successive Layers. The above procedures for placement of wall facing panels,
backfill and reinforcements are repeated for each successive layer..

• Construction of Traffic Barriers and Copings. This final construction sequence is undertaken
after the fmal panels have been placed, and the backfill has been completed to its final grade.

6 - 46
6.8.2 Construction with Geotextile or Mesh Facing

Walls with geotextile sheet reinforcements or continuous mesh reinforcements (metallic or synthetic) can
be constructed using the sheet or mesh as the wall facing element. The construction steps for this type of
wall are illustrated in Figures 6-24 and 6-25. and Photos 6-18 through 6-21, and discussed below.

• Site Preparation. Preparation of the site consists of successively clearing and grubbing, leveling
and proof-rolling the subgrade, prior to the placement of the first level of reinforcement.

• Reinforcing Layer Placement. The reinforcement should be placed with the principal strength
direction perpendicular to the face of the wall and secured with retaining pins to prevent movement
during fill placement. To facilitate placement of the reinforcement, it is generally desirable to unroll
the geotextile or grid reinforcement in a direction parallel to the line of the wall. However, if the
geotextile or geogrid roll is not wide enough for the required width of wall, the
reinforcement should be unrolled perpendicular to the line of the wall. A minimum overlap of 150
mm is recommended along the edges perpendicular to the wall face. Alternatively, with geogrid
reinforcement, the edges may be clipped or tied together.

• Backfill Placement. (See Section 6.8.3). Compaction of the backfill is done with conventional
light weight equipment so that the geoxtextile fabric or geogrid is not damaged during construction.

• Face Construction. The reinforcement is turned up at the face of the wall and returned a minimum
of 1 m into the backfill below the next reinforcement layer (see Figure 6-24). Form work is
generally required to support the face for each construction lift to maintain wall alignment during
backfill placement and compaction. For geogrids, a wire mesh or geotextile may be required at the
face to retain backfill materials.

• Placement of Successive Layers. The above procedures for placement of reinforcement and
backfill are repeated for each successive layer.

• Face Protection. After placement of the final backfill lift, the exposed face of the wall may be
covered with a layer of gunite or concrete to protect against deterioration and vandalism. This
surface protection layer is typically lightly reinforced with metallic mesh, and is pinned into the
backfill to provide face stability.

When gunite or similar coating system is applied to the face, groundwater behind the wall may not be
effectively drained, particularly if the backfill material is not open graded. To prevent accumulation of
groundwater behind the wall and the resulting hydrostatic pressure on the face, slotted drain pipes are
installed horizontally behind the face to provide drainage (Figure 6-25).

6.8.3 Backfill Placement and Compaction

Proper backfill placement and compaction is extremely important to the overall success of a MSE wall.
Following are typical requirements for backfill placement:

• Backfill placement should closely follow erection of each course of facing units. The backfill should
be dumped onto or parallel to the rear and middle of the reinforcements and bladed toward the front
face of the wall. Backfill placement methods near the facing must assure that no voids exist directly
beneath the reinforcing elements. In general, the backfill should be placed in such a manner as to
avoid any damage to the wall materials or misalignment of the facing units or reinforcing elements.

6 - 47
Step 1
EBrace} board Form

Step 2 ~"""'4'
. ~"4"(" ... '.Q "':l?'"
0_:," ',<7.,

. Backfill

Step 3 JF~w..."'"
, :'~,
•• ~ •.•'
"<3 ° •••
• • . • C>. ' . '
, <3. .
.... d
, 0 <7 • • "7-

• . : • ' . qf'.

Step 4
h·· ·
. q.. .' . . " :v.. :", : <q '.' •• q : .o'.'q': .

. • -:¥" ~.'. c'· .


. . . . q. . " : '9 .'. <tZ' • <;' •

·.·.·d·.·.·
Step 5 .
.oJ.'. ' . '
'<J'
': <1, .•.. '<r"
<2 •• "0' ",""'"

<1 . . '

Figure 6-24: Construction Procedures: Geotextile Retained Earth Wall with Wrap-Around Facing.

Welded
Wire Mesh ~
:>-1111:""'""_ _ ~ ~ 1

Overlap of GeoteXtile Sheet

Geotextile Sheet .

Gunite Slotted drain Pipe


.Cover
Layer

Figure 6-25: Geoxtextile Retained Earth Wall Details.

6 - 48
Photo 6-18

Photo 6-19

Photo 6-20

Photo 6-21

Photos 6-18 to 6-21: Construction ofMSE wall with Geotextile Sheet Reinforcements.
Photo 6-18: Placement of Backfill Over Geotextile Sheets.
Photo 6-19: Backfill Compaction
Photo 6-20: Placement of Wire Mesh Cover Over Wrapped Geotextile Facing
Photo 6-21: Gunite Facing and Drainage Pipes

6 - 49
• At no time should any construction equipment be in direct contact with the reinforcements because
the reinforcements and their protective coatings can be damaged.

• The soil layers should be compacted up to or even slightly above the elevation of each level of
reinforcement connections prior to placing that layer of reinforcements.

• Moisture and density control is imperative. Even when using high-quality granular materials,
problems can occur if compaction control is not exercised. The backfill material should be placed
and compacted at or within 2 percent dry of the optimum moisture content. If the reinforced fill is
free draining with less than 5 percent [mes passing the No. 200 U.S. sieve, water content of the fill
may be within ±3 percentage points of the optimum. The moisture content during placement can
have a significant effect on the soil-reinforcement interaction. Moisture contents wet of optimum
make it increasingly difficult to maintain an acceptable facing alignment, especially if the fines
content is high. Moisture contents that are too dry result in significant settlement during periods of
precipitation.

• A density of 95 percent of the maximum density as determined by AASHTO T-99 is recommended


for MSE walls. For applications where spread footings are use,j to support bridge or other structural
loads, the top 1.5 m below the footing elevation should be compacted to 100 percent AASHTO T-99.

• A procedural specification is preferable where a significant percentage of coarse material, generally


30 percent or greater retained on tl:1e 19 mm sieve, prevents the use of AASHTO T-99 or T-180 test
methods. In this situation, typically three to five passes with conventional vibratory roller
compaction equipment is adequate to attain the maximum practical density. The actual requirements
should be determined based on field trials.

• The maximum lift thickness after compaction should not exceed 300 mm. The lift thickness may
be decreased, if necessary, to obtain the specified density.

• With the exception of the 1 m zone directly behind the facing elements, large smooth-drum,
vibratory rollers should generally be used to obtain the desired compaction. Sheepsfoot rollers
should not be permitted because of possible damage to the reinforcements. When compacting
uniform medium to fine sands (in excess of 60 percent passing the No. 40 sieve), a smooth-drum
static roller or lightweight (walk behind) vibratory roller should be used. The use of large vibratory
compaction equipment with this type of backfill material will make wall alignment control difficult.

• Within 1 m of the wall, small, single or double drum, walk-behind vibratory rollers or vibratory
plate compactors should be used (Photos 6-22 and 6-23). Placement of the reinforced backfill near
the front should not lag behind backfill placement for the remainder of the structure by more than
one lift. Poor fill placement and compaction in this area has in some cases resulted in a chirnney-
shaped vertical void immediately behind the facing elements. Within this 1 m zone, quality control
should be maintained by a methods specification such as three passes of a light drum compactor.
Higher quality fill is sometimes used in this zone so that the desired properties can be achieved with
less compactive effort. Excessive compactive effort or use of too heavy equipment near the wall
face could result in excessive face panel movement (modular panels) or structural damage (full-
height, precast panels), and overstressing of reinforcing materials.

• Inconsistent compaction and undercompaction caused by insufficient compactive effort or allowing


the contractor to "compact" backfill with trucks and dozers will lead to gross misalignments and
settlement problems and should not be permitted.

6 - 50
Photo 6-22

Photo 6-23

Photos 6-22, 6-23: Lightweight Compaction Adjacent to Wall Facing

6 - 51
• At the end of each day's operation, the last level of backfill should be sloped away from the wall
facing to rapidly direct runoff away from the wall face. In addition, surface runoff from adjacent
areas must not be allowed to enter the wall's construction site.

6.9 DURABILITY AND LONG TERM PERFORMANCE

The service life of a MSE wall depends to a great extent on the durability of the reinforcement and to a
lesser extent on that of the facing elements. The durability of metallic reinforcement is usually measured
by their resistance to corrosion while that of geosynthetics is assessed by the resistance to: (a) hydrolysis
in polyester; (b) oxidation in polyethylene and polypropylene; (c) stress cracking; (d) ultra-violet light
exposure; and (e) bilogical degradation. A detailed discussion of durability issues is presented by Elias
(1996). Herein, only a brief discussion is presented.

6.9.1 Corrosion of Metallic Reinforcements

Corrosion of buried metals is essentially an electrochemical process. To occur, there must be a potential
difference between two points that are electrically connected in the presence of an electrolyte such as water
rich in oxygen and dissolved salts.

The corrosion of metallic reinforcing elements depends on a number of backfill properties (Table 6-7),
including:

• Porosity: Porosity allows the soil to retain moisture over a certain time. High porosity also allows
maximum aeration. Both factors tend to increase the initial corrosion rate.
• Electrical conductivity: Electrically conductive environments can cause loss of metal which is
proportional to the intensity of the electrical current.
• Concentration of dissolved salts: Certain ions, such as chlorides and sulfates are aggressive and
accelerate corrosion while others, such as magnesium and calcium, are inhibitors of corrosion.
• Degree of saturation: Moisture is a necessary agent for corrosion. Usually the rate of corrosion
increases with increasing the degree of saturation of the soil.
• Acidity or alkalinity (pH): No well-defmed relationship exists between metal loss and pH; however,
greater metal losses have been measured at sites with pH values less than 5.

Most MSE walls use galvanized steel for reinforcement. The zinc coating provides a sacrificial anode
which corrodes while protecting the base metal. After the zinc is oxidized, corrosion of the base metal
begins. The deterioration of the reinforcing elements is usually estimated for the life of the structure and
accounted for by using increased metal thickness (Section 6.5.1).

The potential for future change in the corrosive environment should also be taken into account when
determining the rate of corrosion. Deicing salts used on roadways, for instance, may penetrate the backfill
material, increasing the concentration of cWorides in the soil and thereby accelerating the rate of corrosion.
The influence of the deicing salts usually extends 2 to 3 m in depth. For walls directly supporting
roadways where deicing salts are used, it would be advisable to place a geomembrane beneath the road
base for the length of reinforcements to prevent the infiltration of salts into the backfill material.

6 - 52
6.9.2 Durability of Geosynthetics

Geosynthetics are generally made of synthetic polymers manufactured through different processes.
Geosynthetics made of polyester, polyethylene and polypropylene are commonly used. The durability of
these products is usually assessed by studying the hydrolysis of polyester, the oxidation of polyolefins
(polyethylene and polypropylene), stress cracking, degradation upon exposure to ultra-violet light, and
biological degradation.

Hydrolysis

Hydrolysis occurs when water molecules react with the polymer molecules resulting in chain separation,
reduced molecular weight and strength loss. Of the main types of polymers used, only polyester is
susceptible to hydrolysis.

Hydrolysis is a very slow reaction affected by humidity, temperature, polyester structure, external chemical
agents, and external applied loads. To protect polyester from chemical attack, robust coatings of
polyethylene or PVC have been used.

Oxidation

Oxidation occurs due to heat (thermo-oxidation) and exposure to ultra-violet light (photo-oxidation).
Oxidation primarily affects polyethylene and polypropylene. The combination of heat and oxygen tends
to cause breakdown and cross-linking of the molecular chains resulting in embrittlement of the polymer.
Antioxidants are usually added to the material during processing.

Stress Cracking

Semi-crystalline polymers such as high density polyethylene (HDPE) have a potential for stress cracking,
which is a material failure caused by tensile stresses less than the short term mechanical strength. The
failure is characteristically brittle, with no elongation close to the failure stage. Stress cracking can also
be caused by adverse chemical environment.

Certain grades of polyethylene can experience stress cracking under "low" stresses at ambient temperatures
given sufficient time. This mode of failure may limit the lifetime of, and/or the allowable stress levels on,
polyethylene used for critical load-bearing applications, such as landfill linings and reinforcement
applications. Because of the viscoelastic characteristic of polyethylene, failure of this material is time
dependent and can be traced to either stress(creep) rupture or slow crack growth.

Ultraviolet Exposure

All polymers are susceptible to degradation by ultra-violet light. The high energy from the sun's
ultraviolet radiation tends to break the bonds in organic polymers, causing embrittlement. Protection
against ultraviolet light is provided by incorporating an ultra-violet stabilizer into the polymer during
manufacturing and/or by keeping the geosynthetics covered at all times by backfill or a protective facing.
Rolled geosynthetics in storage stockpiles should be kept wrapped in a protective cover. Ultra-violet
stability should be evaluated on a product-specific basis.

Biological Degradation

Microorganisms causing deterioration are found in a wide range of environmental conditions. These

6 - 53
microorganisms require a source of carbon for growth and obtain it from reactions degrading organic-based
materials such as some of the polymers and additives potentially used in geosynthetics. Environmental
factors controlling biodeterioration are temperature, humidity, pH, etc. Microorganisms of importance
in biodeterioration are bacteria, fungi, actinomycetes, algae and yeast. In general, elevated temperatures,
high humidity, and the absence of UV light are required conditions. The net effect of microorganism
attack is a reduction in molecular weight, with ensuing deterioration of physical properties such as weight,
strength, and elongation.

No completely relevant test to measure the resistance of geosynthetics to biological effects in unstressed
states is presently available. ASTM 3083-89 has been used and can be adopted in the interim.
Statistically, significant strength losses measured from this test should disqualify a candidate geosynthetic
material for long-term in-ground applications.

6.10 EXAMPLE PROBLEM 6-1

Statement

Design a 6 m high retaining wall shown in Figure 6-26a. The wall uses inextensible steel linear
reinforcements. The total wall height including embedment is 6.6 m. Consider a traffic surcharge of q
= 10 kN/m2 • A traffic barrier is also provided with the barrier cast integrally to the concrete pavement.

q = 10 kN/m 2

Reinforced Retained Fill


Soil Mass <Pr Yr Kaf
<Pr Yr Kr
£,
e e
~ IC
IC
1\
II
.:
::t:
V1=y)IL
H/2
L - 2e H/3

°v
where: q = Traffic Live Load
R= Resultant ofVertical Forces (Vl+qL)
:1
(a) (b)

Figure 6-26: (a) Geometry of the Problem and (b) External Forces to be Considered in Analysis

6 - 54
The engineering properties of soils are as follows:

Foundation soils:

• <1>' = 30° (clayey sand, medium dense)


• Allowable bearing capacity - 300kPa
• Differential settlements on the order of 1/300 are estimated

Retained and reinforced backfill:

• <1>' = 30°, YT = 18.8 kN/m3 for retained fill.


• <1>' = 34°, YT = 18.8 kN/m3 for reinforced backfIll meeting the
• F* = 2.0 based on Cu > 7. specifications in Section 6.7

The wall should meet the following performance requirements.

• External stability FS.


Sliding = 1.5; Eccentricity 5: L/6; Maximum foundation pressure 5: allowable bearing capacity;
Global stability ~ 1.3
• Internal stability FS
Pullout ~ 1.5; Allowable stress = 0.55 Fy; Design life = 75 years

Solution

The design process for the above conditions is illustrated below in a step-by-step procedure.

Step 1: Choose Facing Type, Reinforcement Spacing and Type.

Based on the urban location a precast concrete facing with an architectural finish is usually required. For
aesthetic reasons select maximum panel dimensions of 1.5 x 1.5 m with joints no greater than 19 mm.
Since the estimated differential settlements along the wall are 11300, panel joints of 19 mm are acceptable.

Because of numerous surface drainage obstructions, linear galvanized ribbed strip reinforcements are
preferable and the base design is prepared accordingly.

Given the panel size, the most efficient vertical spacing is 0.75 m, allowing for 2 rows of reinforcements
per panel.

Step 2: Establish Preliminary Length for Reinforcing Strips.

For horizontal backfill slopes, L = 0.7H is reasonable; therefore:

L = 0.7H = 0.7 (6.6) = 4.62 m, say L = 4.65 m

6 - 55
Step 3: Check External Stability for L = 4.65 m.
Compute K. for retained the fill, with a <I> = 30 0

K. = la.'( 45' - ~) = 0.33

(a) Compute sliding FS at base (see Figures 6-13 6-26b):

VI tan<j>
FS = ---
s ~FH

where:

VI = HLy = (6.6) (4.65) (18.8) = 577.0 kN/m


V 2= qH = (10) (4.65) = 46.5 kN/m
2
FI =0.5 yH K a =0.5(18.8)(6.6)2(0.33) = 136.5 kN/m
F 2= qHK. = (10) (6.6) (0.33) = 22kN/m

Substituting values, we obtain,

577 tan 30°


FS s = - - - - = 2.10 > 1.5 O.K.
158.5

(b) Compute eccentricity at base:

where:
~MR = Sum of resisting moments =V I (L/2) + V2(L/2) = 1449.6 kN/m
~Mo = Sum of overturning moments = F I (H/3) + FlHI2) = 372.9 kN/m

Substituting values:

e = 4.65 _( 1449.6 -372.9)


2 623.5

L
e = 0.60m < - = 0.78m O.K.
6
L
e = 0.60m < = 0.78m O.K.
6

6 - 56
(c) Check factor of safety against overturning (as shown in Figure 6-13):

1341.5
= - - - = 3.6 > 2.0 O.K.
372.9

(d) Compute bearing pressure at base:

~v
av =
L-2e

623.5
av = - - - - - =180.72 kPa < 300 kPa O.K.
4.65 -2(0.60)

Step 4: Detennine Internal Stability at Each Reinforcement Level and Required Horizontal Spacing.

(a) Compute K at each level. e.g., at z = 2.925 m from surface:

K. = tan' ( 45 - ; ) = 0.283 for reioforced fill

From Figure 6-23: K = 0.412

(b) Compute a H at this level per unit width:

avo = yz +q = (2.925) (18.8) + 10 = 65 kPa

aH = avo K = (65) (0.412) =26.8 kPa

The impact barrier will not transfer stress to reinforced volume because it is cast to the concrete pavement
structure for the full width of the roadway. .

The horizontal spacing is determined from pullout considerations by using a tributary area over 2 panel
widths centered by the reinforcements at each level. Therefore, the tributary area At is:

At = Sv Sh = (0.75) (2) (l.5) = 2.25 m2 (Sh = 2x panel width)

The maximum force on the tributary area is:

T= a H At = (26.8) (2.25) = 60.3 leN

6 - 57
(c) Compute number of reinforcement strips to satisfy pullout resistance:

If pullout FS ~ 1.5 then the resistance PR is:

The number of reinforcing strips, N, required to satisfy the minimum resistance can be calculated from:

N =
2bF *L e ay

where:
b = 50rnm
L e = 2.67 m (see Figure 6-21) [L - 0.3H = 4.65 - 0.3(6.6) = 2.67 m]
Oy' = yz (Neglect live load surcharge for pullout)
F*= 1.35 (Obtained by interpolation from 2.0 at z = 0 to tan <p at 6 m, i.e., F* = 2 - 0.221z
with z = 6 m)

Substituting values:

90.4
N = - - - - - - - - - - - - - = 4.56
2 (0.05) (1.35)(2.67) (18.8) (2.925)

N = 5 strips per tributary area for FS > 1.5 placed in 2 rows (3 in upper; 2 in lower)

(d) Check stress in reinforcement based on a section loss of Es due to corrosion:

The basis for the thickness losses per year are as follows:

zinc loss = 15 /lm (first 2 years)


= 4/lm (thereafter)
steel loss = 12 /lm

Service life of zinc coating, of initial thickness of 86 /lm, is:

. 86-2(15)
Llfe = 2 + = 16 years
4

The base carbon steel will lose section for:

75 years - 16 years = 59 years at a rate of 12 ,urn/year/side.

6 - 58
Therefore, the anticipated loss is:

Es = 12 (59) 2 = 1.416 mm and


Ee = 4.000 - 1.416 = 2.584mm
and the section area = 129.2 mm2

If 60 grade steel is used


fy = 413.7 MPa and fall = 0.55 (fy) = 227.5 MPa
The tensile stress in each strip can be calculated from:

fs - .--!...- = 60.3 = 93.34 < 227.5 MPa


NEe 5(0.0001292)1000

(d) Calculate internal stability at each layer and detennine the number of reinforcing strips per tributary
area as shown in the following table:

Depth Vertical K F* Horizontal Effective N strips Tensile FS


z (m) Pressure Pressure Length,Le per trib stress Pullout
kPa kPa m area MPa
0.675 22.69 0.47 1.85 10.55 2.67 6 30.64 1.58
1.425 36.79 0.45 1.69 16.46 2.67 5 57.34 1.63
2.175 50.89 0.43 1.52 21.87 2.67 5 76.18 1.69
2.925 64.99 0.41 1.35 26.78 2.67 5 93.29 1.65
3.675 79.09 0.39 1.19 31.20 3.09 5 108.65 1.81
4.425 93.19 0.38 1.02 35.11 3.49 4 152.85 1.50
5.175 107.29 0.36 0.86 38.52 3.89 5 134.18 1.87
5.925 121.39 0.34 0.69 41.44 4.29 5 144.33 1.77

For ease of construction provide 5 strips per tributary area below the top row.

Finally, the wall should be checked for global stability using appropriate slope stability methods (Refer
Module 3 - Soil Slopes and Embankments).

6 - 59
CHAPTER 7.0
EXTERNALLY SUPPORTED STRUCTURAL WALLS

7.1 INTRODUCTION

Externally supported structural walls rely primarily on the bending resistance of a vertical structural
element to resist the applied lateral loads. The vertical wall elements may consist of discrete elements
(e.g., soldier piles) spanned by a structural facing, or may be a continuous structure (e.g., sheet pile wall
or tangent pile wall). The primary types of externally supported walls are illustrated in Figure 7-1. The
wall configurations include cantilever walls and structures with single or multiple levels of support. Wall
support may be provided by a system of struts or rakers on the exposed side of the wall, or anchors
installed through the wall (Figure 7-2).

The following sections of this chapter provide a discussion of specific types of walls, including sheet pile
walls, soldier pile and lagging walls, slurry (diaphragm) walls, and tangent/secant pile walls. Section 7.6
provides design procedures for these different types of walls; Section 7.7 discusses the design and
construction of ground anchors for support of these walls; and Section 7.8 discusses methods for estimating
wall movements.

75 to 100 mm
Wood or Precast

I~ p
Concrete Lagging
1==];
C
I.. 1.6 to 3 m

(a)
.1
0.6 to l.Orn

(e)
+

p]--ld 1.8 to 3 m
.1
; 0 . 5 to 1 m

(b) (f)

I~I
(c) (g)

(d)
IIIII (h)

Figure 7-1: Primary Types of Externally Supported Structural Walls: (a) Soldier Pile and Lagging
Wall; (b) Soldier Pile and Cast-In-Place Concrete Lagging Wall; (c) Master Pile Wall; (d)
Sheet Pile Wall; (e) Slurry (Diaphragm) Wall; (f) Secant Pile Wall; (g) Tangent Pile Wall;
and (h) Interlocking H-Pile Wall. (After Dismuke, 1991)

7-1
(a) (b)

Anchor
Block
'/""-"'l>-"-"/-
Kicker Block or
Foundation Slab
(c)
(d)

,,,-'-.."l>-"-"::/-
Soil or
Rock Anchors
Bond
Length

(e) (f)

Figure 7-2: Wall Support Systems: (a) Cantilever Wall; (b) Earth Berm Support; © Raker System; (d)
Deadman Anchor; (e) Cross-lot Braced Wall; and (t) Anchored Wall. (After NAVFAC,
1982)

7-2
Externally supported structural walls can be used for both temporary and permanent wall applications.
Typical applications for such walls are listed below:

Walls for temporary excavation support:


• Cofferdams on land and in water for footing construction
• Excavation support for Cast-in-Place wall construction
• Grade separation during staged roadway construction
• Excavation support for cut-and-cover roadway tunnels and culverts
• Excavation support for utility trenches

Permanent wall applications:


• Roadway cuts
• Roadway fill containment
• Abutments
• Cut-and-cover tunnel walls
• Slide stabilization
• Waterfront structures

The primary advantages of externally supported structural walls in comparison to other types of walls are:
• They allow top-down method of construction (Le., wall elements are installed before the start of
excavation); with advantage of reduced ground displacement.
• When used as permanent walls they require:
Reduced quantity of excavation
Reduced quantity of backfill
Reduced work area
• Faster construction time.
• Can be used to provide a seepage barrier (e.g., for a depressed roadway).
• Can effectively support large vertical loads as well as lateral loads.

Some of the disadvantages of externally supported structural walls are:


• They typically require more specialized construction techniques.
• Complicated soil-structure interaction may make analysis and design more difficult.
• The performance of the completed wall is more dependant on the construction method and quality
of the work.
• Metallic components, if in contact with soil, are more susceptible to corrosion.

7.2 SHEET PILE WALLS

7.2.1 General

A sheet pile wall consists of a series of interlocking sheet piles driven side by side into the ground, thus
forming a continuous vertical wall. Sheet pile walls are commonly used for waterfront structures and for
temporary earth support applications, but can also be used as permanent walls for highway structures.
Sheet piling is also used for stabilizing ground slopes and for cellular cofferdam construction. Photos 7-1
through 7-4 show several applications for sheet pile walls.

In certain applications sheet pile walls are termed as "bulkheads" or "cofferdams." A bulkhead is
generally a sheet pile retaining wall used for waterfront construction. A cofferdam is a reasonably water
tight enclosure made of sheet piling, usually temporary, built around a working area for the purpose of
supporting lateral soil and water loads and excluding water during construction.

7-3
Photo 7-2

Photo 7-·1

Photo 7-4

Photo 7-3

Photo 7-1: Sheet Pile Wall for Earth Support Behind a Cast-in-Place Wall.
Photo 7-2: Sheet Pile Cofferdam for Footing Construction on Land.
Photo 7-3: Sheet Pile Cofferdam for Construction of Foundations in Water.
Photo 7-4: Anchored Sheet Pile Bulkhead for Waterfront Structure.

7-4
The sheet piles can be of timber, reinforced concrete or steel. Timber sheet piling is used for short spans,
light lateral loads, and commonly for temporary support of shallow excavations. Concrete sheet piles are
precast members, possibly prestressed, usually with a tongue and groove joint, designed to withstand the
permanent service loads and handling stresses during construction. They are heavy and bulky, and require
heavier equipment to drive and handle. Concrete sheet piles are generally used only for permanent wall
applications.

Steel sheet piling is the most common type used and is preferred over the other material because it:
• is lightweight
• is easier to drive and extract
• has higher bending resistance
• can be reused several times

Some of the limitations in the use of steel sheeting are:


• Cannot penetrate hard layers
• Sheeting length limited to about 29 m
• Wall elements are susceptible to corrosion

Steel sheeting is fabricated either using hot rolling or a cold formed manufacturing process. Interlocking
Z-shaped or V-shaped sections are typically used for retaining wall applications since these sections provide
a higher bending resistance and corresponding greater moment of inertia. Flat sheets are generally limited
to use in cellular cofferdams. Some common steel sheet pile sections used for sheet pile walls are shown
in Figure 7-3.

..
:. :-"':~

"
I.
I•

,:,\ II

~•.!I:"'·
'~. ~====~/I
.. "

(a) (b)

(c)

Figure 7-3: Steel Sheet Pile Sections Commonly Vsed for Retaining Walls and Cofferdams. (a) Z -
Section, (b) V-Section, (c) Cold Formed Section.

7-5
7.2.2 Wall Construction

Procedures

The two general construction conditions for sheet pile walls include (a) driving sheet piles into the ground
and then backfIlling behind the wall, and (b) driving sheet piles into the ground and excavating the soil in
front of the wall. These two general construction conditions can therefore be classified as:

• BackfIlled Structure
• Excavated Structure

The sequence of construction for a backfilled structure is illustrated in Figure 7-4 and generally proceeds
as follows:

Step 1: Excavate the in situ soil in front and back of the proposed structure, if necessary.
Step 2: Drive the sheet piles.
Step 3: BackfIll to the level of the anchor and install the anchor system.
Step 4: Backfill to the top of the wall.

For a cantilever type of wall, only Steps 1, 2 and 4 apply.

The sequence of construction for an excavated structure is illustrated in Figure 7-5 and generally proceeds
as follows:

Step 1: Drive the sheet piles.


Step 2: BackfIll (or excavate) to the anchor level and install the anchor system.
Step 3: BackfIll to the top of the wall.
Step 4: Excavate the front side of the wall.

For cantilever sheet pile walls, Step 2 is not required.

When backfill is required, it should be composed of clean granular material to facilitate placement and
compaction, provide favorable drainage characteristics, and lower the lateral earth pressures. If the backfill
is placed under water, consideration should be given to compaction of the backfIll by deep compaction
methods (see Module 4 - Ground Improvement Techniques) to reduce post construction settlement.
BackfIll within the passive wedge of the anchor block or anchor wall should be compacted to not less than
95 percent of the maximum density as determined in accordance with AASHTO T-99.

Equipment

Sheet piles are typically installed with a vibratory hammer, but impact hammers are also used. The size
of the hammer and the size of the support crane will depend on the length of the sheeting and the nature
of the ground that the sheeting must penetrate. Installation of sheet piles typically requires the use of a
template to help set the piles and maintain them within vertical and horizontal alignment during driving.
A heavy steel H-pile section, driven with an impact hammer, may be required to fracture or displace
subsurface obstructions in advance of sheet pile installation.

7-6
Origimil
ground
surface
,I

i/"
" "I
" " '..... 't
"
.....
"
" .....
.....
..... I
/
" " " .......l
----'---'--
" " " ~\- __u
" I .':'
::::::::..:;::.i,;....~:.'.••;: ~:-,:~:·r.; ;~.z....;.. :~~{

Excavation Excavation
Limits
Step 1 of Soft Soil Step 2
Anchor
rod

Step 3 Step 4

Figure 7-4: Sequence of Construction for a Backfilled Sheet Pile Structure. (After Das, 1990)

Anchor
rod
Original
ground
surface

Step 1 Step 2

Step 3 Step 4

Figure 7-5: Sequence of Construction for an Excavated Sheet Pile Structure. (After Das, 1990)

7-7
7.3 SOLDIER PILE AND LAGGING WALLS

7.3.1 General

Soldier pile and lagging wall systems (i.e., walls with discrete vertical wall elements) are commonly used
for temporary excavation support and in dense or stiff soils where sheet pile walls may not be suitable.
They are also being used more frequently for permanent earth retaining structures.

Soldier pile and lagging walls consist of soldier piles usually set at 1.8 to 3 m spacing, and lagging which
spans the distance between the soldier piles. The lagging is used to retain the soil face and transmit the
lateral earth pressure to the soldier piles. Included in this category of walls is the master pile (king pile)
wall system, shown in Figure 7-1(c), which consists of discrete vertical H-pile sections interlocked and
alternating with steel sheet pile sections. Soldier pile and lagging walls are commonly used to support
excavations (Photo 7-5) and are being used more frequently for permanent walls (Photo 7-6).

The most common soldier piles are rolled steel sections, normally H-pile or wide flange sections.
However, soldier piles can be almost any structural member such as pipe or channel sections, cast-in-place
concrete, or precast concrete. Figure 7-6 shows several types of soldier piles.

For temporary walls, lagging is most commonly wood, but may also consist of light steel sheeting or
precast concrete. Cast-in-place concrete sheeting (or facing) is generally used for permanent wall
applications, but such construction typically requires the use of temporary sheeting to support the soil face
during excavation and concrete placement operations. Lagging may be omitted in hard clays, soft shales
and soils with natural cementation, provided that the piles are installed at relatively close spacing and with
adequate steps taken to protect against erosion and spalling of the face.

t
Anchor
I

~Lagging
~ "C~S~
!wedge behind
front flange

(a) (b)

t
Lag clipped
to angle

welded}
angle section

Welded ST section (c)


Figure 7-6: Several Types of Soldier Piles: (a) Wide Flange Section; (b) Double Channel Section;
(c) Pipe Section. (Xanthakos, et ai. 1994)

7-8
Photo 7-5: Soldier Pile and Lagging Wall for Temporary Excavation Support.

(a) (b)

Photo 7-6: Permanent Soldier Pile and Lagging Walls for (a) Roadway Embankment, and (b)
Roadway Cut.

7-9
The advantages of soldier pile and lagging walls include:
• The availability of a wide range of pile sections to match loading requirements.
• The ability to resist high bending moments.
• Can be installed through hard soils and rock.
• Can support large vertical loads.
• Piles can be spliced.

Some of the disadvantages of soldier pile and lagging walls are:


• They are generally free draining and not suitable for applications where it is necessary to maintain
the groundwater level behind the wall.
• Steel elements in direct contact with soil are more susceptible to corrosion.
• Greater ground displacement in comparison to stiffer wall systems (Le., slurry walls, tangent pile
walls, etc.)
• Excavation for placement of the lagging between soldier piles increase risk of ground loss.

7.3.2 Wall Construction

In this section, only braced (struts or rakers) walls are discussed; anchored walls are discussed in Section
7.7. The procedures presented below follow those described by Goldberg, et ai. (1976).

General

The soldier piles are installed by driving, or by concreting them within pre-drilled holes (photo 7-7). After
installation of the piles the excavation is made to the first support level, placing lagging as the excavation
proceeds. Brackets are then attached to the soldier pile to support the wale, and the wale is then placed
in position and connected to the soldier piles. The brace is cut a few centimeters short to facilitate
placement. This extra space is closed by plates and wedges when the final connection is made. The above
sequence of excavation, and installation of lagging, walls and braces is continued until the required
excavation invert level is reached.

(a) (b)

Photo 7-7: Soldier Pile and Lagging Wall: (a) Drilling for Pile Installation, and (b) Excavation
Between Soldier Piles in Cohesive Soil for Installation of Lagging.

7 - 10
Preloading

Preloading is usually required for installation of bracing members since it results in more reliable load
determination and load distribution within the structural support system, and it effectively reduces ground
displacements adjacent to the excavation. Preloading is particularly important when the excavation is
located near structures or other facilities which may be damaged by settlement or lateral ground movement.
Cross-lot bracing members are commonly preloaded when they are installed. Rackers and comer braces,
however, are not generally preloaded due to the more complicated skew connections required.

Preloading of internal braces is accomplished by loading hydraulic jacks to the desired load followed by
securing the member with steel blocking, steel wedges and welding. One procedure is to jack to the
desired load, and then to drive steel wedges between the member and the wale until the jack load is
essentially zero. A second procedure is to weld the connection tight while maintaining the jack load and
then drop the pressure in the hydraulic jack, thus transferring the load through the connection to the wale.
The second procedure may result in additional wall movement as the load is transferred, although the
magnitude of movement is generally small.

High preloads may cause over stressing of struts because of unforeseen job conditions or temperature
effects. Accordingly, the general practice is to preload bracing members to about 25 to 50 percent of their
design load. This preload removes the slack from the support system and at the same time reduces the risk
of over stressing. Larger preloads, up to as much as 80 percent of the design load, may be desirable to
further reduce ground movements and protect adjacent structures from settlement.

Strut Removal and RebA"acing

Temporary support bracing, or struts, are generally removed during or after construction of the permanent
structure elements (Le., invert and/or roof slabs). Also, it may occasionally be necessary to remove and
reinstall temporary bracing (rebracing) to maintain lateral support while installing the permanent structure
elements. Strut removal (and rebracing) may be an additional source of wall and ground displacement.
Factors controlling the amount of displacement are the wall stiffness, the properties of the retained soil,
the span distance between remaining braces, and the quality and the compaction of backfill between the
final structure and the excavation support wall.

7.4 SLURRY WALLS

7.4.1 (;eneral

A slurry wall (or diaphragm wall) is a structural, cast-in-place concrete wall constructed by tremie
placement of concrete in a pre-excavated, slurry filled trench, as shown in Figure 7-7. The wall is
constructed in a series of panels which interlock to form a continuous structure. Commonly, trenches 0.6
m to 1.0 m wide are excavated in lengths of 3 m to 6 m.

Typically, slurry walls are used at sites where a relatively water tight excavation support wall is required.
This situation may occur when a) groundwater lowering outside the excavation limits may lead to
potentially damaging settlement of nearby structures or other facilities, b) when it is not practical to
dewater a site, i.e., adjacent to an open body of water, and c) where seepage gradients initiated by
dewatering operations may risk migration of existing groundwater contamination plumes. With penetration
into an underlying stratum of low permeability, or with sufficient penetration below the bottom of the
excavation, a slurry wall provides an effective seepage barrier that can preserve groundwater conditions
outside the excavation.

7 - 11
Slurry walls are also used at sites where it is necessary to restrict ground displacements adjacent to the
excavation. This is a particular concern when the excavation is in close proximity to a building or other
structure which is founded above the bottom of excavation level. As a relatively rigid excavation support
system, a slurry wall typically results in considerably smaller ground displacements than a soldier pile and
lagging support system.

Concreted Concreted
ponel ponel Steel tub:"
Slurry level Ground Slurry level
level

Bentonite Bentonite
slurry slurry

(a) (b)

Steel tube
Reinforcement Steel Tremie
cage tube pipes
Coner eted Bentonite
panel slurry
\
r
~
'P~(j
~:~~~
~;Q ...
~:>fb:'"~c
:>0
"
~:~
C)O:'Q'~

.
;:o~"o
.()~
""","0
::1)'9-
~:~
J:;o 9.
;,co~~

~~~ "/~
~
'l~

~~;o{t:::::::::::~ ()
(c) (d)

Figure 7-7: Typical Construction Sequence for a Slurry Wall: (a) Excavation; (b) Insertion of Steel
Tubing (End Stops); (c) Placement of Reinforcement Cage; and (d) Concrete Placement.
(Xanthakos, 1994)

7 - 12
Slurry walls can be used solely for temporary support of excavations, or serve both as a temporary
excavation support and as the permanent structural wall. A significant advantage of using the slurry wall
for permanent structural support is that it eliminates the need for a costly cast-in-place concrete interior
wall. It also allows a reduction in the width of the excavation. The use of slurry walls for permanent
support also facilitates top-down construction of the fmal structure which may be advantageous for projects
which require a minimum duration for surface disturbance (i.e., construction of a depressed underpass at
a busy intersection).

Disadvantages of slurry walls include:


• Use of specialized equipment and construction methods.
• Requires large construction area for fabrication and handling of rebar cages, and for slurry mixing
and treatment.
• Relatively high cost in comparison to other wall systems.

7.4.2 Wall Types

The two primary types of slurry walls are the conventional reinforced concrete wall and the Soldier-Pile-
Tremie-Concrete (SPTC) wall. Other wall types include the Precast-Concrete-Panel wall and the Post-
Tensioned Concrete wall. These walls are briefly described below.

Conventional Reinforced Concrete WaIl

A conventional reinforced concrete slurry wall uses steel reinforcement cages that are placed in the slurry
trench before the concrete tremie pour. The reinforcing bars are proportioned to resist bending principally
in the vertical direction between bracing levels. Supplemental bars may be provided to serve as "internal
walers" or to distribute forces around inserts or openings. This method of reinforcement is currently most
popular and is used extensively throughout the U.S., particularly for permanent walls and where tieback
anchors are to be used (Figure 7-8). Panel end joints are formed by stop end pipes, shown in Figure 7-7,
or other suitable forming devices.

Soldier-Pile-Tremie-Concrete (SPTC) WaIl

SPTC walls are becoming increasing popular, particularly in deep excavations requiring high bending
resistance. These walls are also referred to as "Soldier Beam and Concrete Lagging Walls" owing to the
similarity with soldier pile and lagging walls. SPTC walls are commonly constructed of vertical wide
flange sections set in a slurry stabilized trench with a reinforcing bar cage placed between the soldier piles
to transfer earth and water loads laterally to the soldier piles. They provide a relatively watertight wall,
significant strength in the vertical direction, greater flexibility for moment connections within the
excavation, and relatively easy connection for temporary cross lot bracing and wales (Figure 7-9).

Other Wall Types

A variation of the above wall types is the Precast-Concrete-Panel wall in which precast concrete wall
elements are placed in the excavated trench. This method produces the best quality finished wall but is
limited in use by cost, transportation and handling length limitations, and other specific site constraints.
Wall panels are best cast on site adjacent to the work and installed directly into an oversized trench
excavation. This type of wall requires the use of a cement-bentonite slurry which will eventually harden
in the void between the wall and soil. Rubber water stops are usually required between panels since the
cement-bentonite joints are subject to drying and shrinkage cracking (Figure 7-10). The precast panels are
temporarily suspended within the excavation until the cement-bentonite hardens. Wall installation usually

7 - 13
proceeds in a linear fashion with precast panel installation following closely behind trench excavation. The
use of a Precast-Concrete-Panel wall becomes less practical for deeper excavations and at locations where
utility crossings or obstructions can be expected.

Another wall type used successfully in Europe but which has had limited use in the U.S. is the Post-
Tensioned-Concrete wall (Figure 7-11). In this type of wall, post - tensioning provides increased bending
resistance. However, major disadvantages of this type of wall are its higher cost and the need for
specialized construction techniques.

Bentonite film at joint 75 mm cover


Reinforcing steel (TYP·l
rI)
rI)

.;
CJ
=
.•.= = ==
~

Q
Q Q
Second pour
--
First pour
[:~T
seco::m
-..
\C Q

-
~ etl Q
.2
as tii 0

~ >-
Inside face Plate for bracing
ofwall or tiebacks

Varies 2 m to 8 m

Figure 7-8: Conventional Reinforced Concrete Wall. (Tamaro, 1990)

Bentonite cake Reinforcing steel


waterproofuig (if required)

-
~
rI)
rI)
~

.;
.•.=
CJ

as
Q
Q
=
==
\C
en
.2
=
...
til
Q
Q

..
0-
0
[ -----] [- ---J
First pour
__e . e
Second pour
e
~ >
I '--- Shims as reqd. ~I Bracing
Inside face
I I
Varies 2 m to 5 m
I.. ~I

Figure 7-9: Soldier-Pile-Tremie-Concrete (SPTC) Wall. (Tamaro, 1990)

7 - 14
Cement bentonite Waterstop&
-e
o
:s

cement-bentonite

lPl
grout
grout at joint 1~
T~
ll!l
QI
e e ...-..... oA-&..

.... 0e 0e
.fa
~
.:I
0
First panel Third panel -o
In

--= ·B ----- o
In
e- 00
t"'-

~ >
.9
..................... ......... ............
Precast finish on inside face

I.e Varies I.S m to 3 m ../

Figure 7-10: Precast-Concrete-Panel Wall. (Tamaro, 1990)

High strength steel


Light weight reinforcing
tendon-position varies
to position tendons
to follow moment diagram
lPl
ll!l
QI e e
e .......... • • .......... ..
.....fa 00 0e
G- --
~ First pour Second pour

- - o
.:I 'C 0
e- 0 0 0
-
~
ll!l
U
·C
lIS
>-
0

S ..

I. Varies2mto6m

Figure 7-11: Post-Tensioned-Concrete Wall. (Tamaro, 1990)

7 - 15
·7.4.3 WALL CONSTRUCTION

Materials

The three basic materials used in the construction of slurry walls include slurry, concrete and
reinforcement. A brief description of each follows:

The slurry used to stabilize the excavated trench consists of bentonite, or attapulgite, and water. When
mixed with water, the bentonite (which comes as a powder, chips or pellets) forms a colloidal suspension
(slurry). To date, the use of additives in the bentonite slurry has been limited. Polymer slurries are
generally not used for slurry wall construction since they are not as effective as mineral slurries for
supporting large size excavations in coarse granular soils.

To maintain a fluid slurry until concrete is completed, the slurry must be circulated and agitated.
Desanding devices are typically used to remove a sufficient amount of suspended soil so that the slurry
can be used two or more times prior to disposal.

Tamaro and Poletto (1992) recommend the following specifications for fresh bentonite slurry:

• Minimum specific gravity of 1.03.


• Minimum viscosity of 32 seconds measured by the Marsh Cone Funnel.
• pH between 7 and 11.

Prior to placement of concrete, the bentonite in the excavated trench should meet the following
requirements:

• Sand content not more than 5 percent, measured at about 1.5 m above the bottom of the trench.
• Specific gravity not more than 1.10.
• Viscosity not more than 50 seconds.

Concrete

The concrete must be a free-flowing mix capable of displacing the bentonite slurry and bonding to the
reinforcement. Following is a summary of the concrete mix specifications as recommended by Tamaro
and Poletto (1992):

• A concrete strength of 20 MPa to 35 MPa.


• A well graded aggregrate less than 20 mm in maximum dimension.
• A 200 mm slump.

In addition, Goldberg, et al. (1976) recommend a water/cement ratio less than 0.6, a sand content of 35
to 40 percent of the total weight of aggregate, and a cement content of at least 400 Kg/cubic meters of the
tremie concrete.

Premature stiffening of the cement may negatively affect the tremie operation. Retarders are sometimes
added to the mix to keep the concrete workable during the entire pour. Some of the retarders, however,
may reduce the strength. The retarders most commonly used are discussed by Xanthakos (1979).

7 - 16
Reinforcement

The slurry wall reinforcing can be in the form of a rebar cage, or a combination of a cage and vertical
wide flange sections. The rebar cage can be prefabricated and assembled either in a shop or at the site.

The minimum bar spacing is generally 150 mm for vertical bars and 300 mm for horizontal bars. The
rebar cages, as well as inserts (e.g., sleeves for soil or rock anchors, casing for instrumentation, etc.),
should be secured with tie wire. Welding of the rebar cage should be avoided since welding can change
the metallurgical properties of the reinforcing bars, and welded connections are prone to breaking during
lifting and handling of the rebar cage.

Spacer devices should be used on the outside of the rebar cage to provide a minimum concrete cover of
75mm.

Equipment

A wide variety of trench-excavating equipment is used in slurry wall construction, including backhoes,
draglines, clamshells, bucket scrapers and others. Buckets are raised and lowered by cable or kelly bar,
and are opened or closed hydraulically or by cable (photo 7-8a). To penetrate hard layers, percussive tools
can be used either to assist in clamshell excavation or as independent excavating tools. The cable operated
bucket, hung from a crane, is the excavation equipment most commonly used by American contractors.

Another type of slurry wall excavation equipment is the "hydromill" which was developed by French and
Italian equipment manufacturers (Photo 7-8b). This excavator is basically a grinding device consisting of
two milling heads rotating in opposite directions about axes perpendicular to the trench. The rotating heads
excavate soil and soft rock from the bottom of the trench, and the excavated material is lifted by suction
to the surface where the soil is removed from the bentonite by sand separators.

For excavation in rock, heavy drop chisels, chisel drills, or large diameter roller bits can be used. Pre-
drilling can also be used to facilitate advancement of the clam shell through hard layers.

Procedures

Construction of a slurry wall includes five primary elements: placement of guide walls, trench excavation,
placement of reinforcement, concreting, and the formation of joints.

Guide Walls

The construction of a slurry wall usually begins with the installation of guide walls. The functions of the
guide walls are to (a) prevent caving of the trench wall in the uppermost part of the excavation, (b) align
the trench, (c) contain the slurry, and (d) support suspended precast elements in Precast-Concrete-Panel
walls, or reinforcing steel in Cast-In-Place walls. Figure 7-12 illustrates a cross-section of a guide wall.

The guide walls are usually 150 to 300 mm thick and 1 to 2 m deep. To provide the necessary support
for suspended rebar cages or pre-cast panels, the guide walls should be cast on a stable subgrade.

7 - 17
(a)

(b)

Photo 7-8: Slurry Trench Excavation Equipment: (a) Hydraulically Operated Clamshell Bucket, and
(b) Hydromill

7 - 18
Concrete Wall

Compacted Backfill
f~~~r----=--

(a)

Concrete Wall
Compacted Backfill
. ..
. .. . . .
I-2m
.
-L I
I

I
_..L.-'\j I •

Note: Add cement to backfIll


to prevent
undermining and to
increase stability

(b)

Figure 7-12: Cross-Section of a Guide-Wall (a) Compact Cohesive Soil, (b) Loose Cohesionless Soil.
(Goldberg, et al., 1976)

7 -19
Trench Excavation

In conventional bucket excavation, the bucket brings the material to the surface, discharges its load, and
then is lowered back into the trench. With direct or reverse circulation equipment, the material is broken
up into smaller particles so that it may be suspended in the bentonite slurry, which is circulated to the
surface, screened and desanded. The cuttings are brought to the surface by suction and/or air lift through
suction pipes, or the excavation tool itself.

The stability of the trench is maintained by the slurry pressure on the trench wall, and soil arching. Also,
local penetration of the bentonite into pervious soils will provide some cohesion that helps to prevent
spalling.

The slurry level in the trench is maintained at an elevation at least 1.2 m higher than that of the
groundwater table. A membrane or a "mudcake" is formed against the walls of the trench by a
combination of hydrostatic pressure, osmotic pressure, and electrolytic properties of the colloid. This
mudcake will maintain the pressure against the trench walls, and prevent fluid losses through pervious
materials.

The verticality of the wall is checked as the excavation advances. If the excavation is found to be out of
verticality tolerance, the trench can be backfilled with lean concrete and the excavation operation repeated.
This technique can also be used to fill cavities formed in the side wall of the trench due to caving.

Reinforcement

Prior to placement of the rebar cage in the slurry filled trench, the trench bottom should be sounded by a
weighted tape, or other means, to verify the depth of the trench and assess the cleanliness of the trench
bottom. Also, the slurry in the trench should be sampled and tested to assure that it meets the specification
requirements noted previously. If necessary, slurry circulation and trench bottom cleaning operations
should be continued to meet the specified requirements.

The rebar cage and end stops should be placed in the trench as soon as possible after the trench is cleaned
and inspected (Photo 7-9). Figure 7-13 shows a reinforced panel in cast-in-place slurry wall.

In the Soldier-Pile-Tremie-Concrete wall, the soldier piles, together with the reinforcing cage, are set
within the excavated trench prior to concreting. An alternative approach is to set the soldier piles in pre-
augered holes and then excavate and place tremie concrete between consecutive piles.

Concreting

The concrete should be placed as soon as practical after installation of the reinforcement cage, and concrete
placement should proceed continuously until completion of the slurry wall panel. Concrete placement is
performed through one or more tremie pipes lowered to the bottom of each panel (Photo 7-10). The use
of a single tremie pipe (and shorter panel length) is preferred since the use of two tremie pipes increases
the risk of entrapment of laitance within the panel. The tremie pipe must remain embedded in fresh
concrete a minimum of 2 m and a maximum of 5 m. The tremie concrete displaces the bentonite slurry
progressively as it rises uniformly to the surface.

The concrete should be sampled and tested at the start of placement, and at defined intervals during
placement, to verify that the delivered concrete meets the specified requirements.

7 - 20
(a) (b)

Photo 7-9: Slurry Wall Reinforcement Cage (a) On Fabrication Bed, Showing Styrofoam Knock-Out
Panel, and (b) During Lifting for Installation into Slurry Filled Trench.

Photo 7-10: Slurry Filled Trench with Tremie Pipes Just Prior to Concrete Placement.

7 - 21
Horizontal Bending Steel

Casing for Slope Inclinometer


or Grouting Below Panel

Main Bending Reinforcing

Plate for Bracing or


Tieback Anchor Sleeve
Attached to Cage

Spacer
(TYP)

Figure 7-13: Reinforced P~el in Cast-In-Place Slurry Wall. (Tamaro and Poletto, 1992)

7 - 22
End Joints

A construction joint is provided between two adjacent panels. This joint should allow excavation of the
new panel without significant disturbance to the previously poured panel. It should be watertight and
capable of transferring shear and compressive stresses. The joint can be formed using different
configurations and details, but generally consists of the three basic elements: a steel tube, a steel plate, or
a steel beam (Xanthakos, 1979).

Unless permanent steel piles are used, the end stops are removed after the initial set of the concrete. For
deep walls, it may be necessary to partially extract the end stops while concrete is placed in the upper part
of the wall panel. The end stops should be removed in a smooth and continuous manner.

Types of Supports

Slurry walls can be supported using either internal bracing or soil or rock anchors. Similar to soldier pile
and lagging walls, the support elements are generally preloaded when they are installed to reduce ground
displacements behind the wall, and to obtain more reliable and more uniform loading in the support
elements. Rakers and comer bracing, however, typically are not preloaded due to the more complicated
skew connection.

Cross-lot struts are usually preloaded to 25 to 50 present of their design load. Generally, the lower preload
is desirable to avoid potential damage to the wall, particularly when the slurry wall will be used for
permanent support; higher preloads are used when it is considered necessary to minimize ground
movements behind the wall.

Soil and rock anchors are usually locked off at about 80 percent of their design load. See Section 7.7 for
guidelines for the design of soil and rock anchors.

7.5 TANGENT/SECANT PILE WALLS

7.5.1 General

A tangent or secant pile wall consists of a line of bored piles (Photo 7-11). If the bored piles are
contiguous, or tangent, to each other the wall is called a "tangent pile" wall. In an alternate case, referred
to as a "secant pile" wall, the pile elements overlap so as to form an interlocking wall. Another variation
of this wall type is called an "intermittent wall" in which the piles are installed at a spacing exceeding the
pile diameter; this type of wall can be considered only if the ground is stable or secondary elements, such
as shotcrete or a cast-in-place facing, is used to provide a continuous wall. Various configurations of bored
pile walls are shown in Figure 7-14.

Tangent pile walls and secant pile walls are stiff, continuous walls that are constructed by the top-down
method. Similar to slurry (diaphragm) walls, tangent pile and secant pile walls can be used when it is
necessary to minimize groundwater lowering outside the excavation or to reduce ground displacements.
Also, similar to slurry walls, tangent pile and secant pile walls can be used for either temporary or
permanent ground support. An advantage of tangent pile and secant pile walls, however, is that they can
be constructed using conventional bored pile (or drilled shaft) excavation equipment and procedures with
which more contractors are familiar. Also, this type of wall may be more suitable than slurry walls at
constricted work sites since less area is needed for slurry containment and treatment, and for fabrication
and handling of rebar cages.

7 - 23
(a) (b)

Photo 7-11: Tangent Pile Wall a) With Structural Steel Section as Core Reinforcement, and b) Face
of Completed Wall.

Concrete pile

(a) (d)
Secondary
row

Primary Soil side Precut lagging


i row
Face concrete
(b)

(e)

Figure 7-14: Various Configurations of Bored Pile Walls: (a) Tangent Pile Wall; (b) Staggered
Tangent Pile Wall; (c) Secant Pile Wall; (d) Intermittent Pile Wall with Grouted
Openings; and (e) Intermittent Pile Wall with Lagging.

7 - 24
In comparison to slurry walls, however, there are some disadvantages including:
• Increased seepage through vertical joints
• More difficult connections for bracing members, ground anchors and attached slabs
• Rough and irregular exposed face

7.5.2 Wall Construction

The bored piles for a tangent pile or secant pile wall are usually installed using auger type drill rigs. When
penetrating unstable soil strata which may be susceptible to squeezing or caving it may be necessary to use
temporary casing. A faster and more economical construction method, however, involves rotary drilling
and the use of bentonite slurry to keep the drilled hole stable and remove the excavated material to the
surface by reverse circulation (FHWA 1988).

Figure 7-15 illustrates the typical construction sequencing for tangent pile and secant pile walls. The bored
piles should be installed in a staggered pattern to avoid disturbing the concrete in an adjacent pile that has
not set.

For a tangent pile wall, shown in Figure 7-15a, the direction of pile installation is from the edges of a
section towards its center. This sequence prevents interference between adjacent piles during concreting,
and allows all the piles in the section to be installed along the same alignment except possibly the center
pile which may have to be displaced slightly to fit in the remaining space and still be tangent with the two
adjacent piles.

In secant pile walls (Figure 7-15b), alternate piles (numbered 1,3,5, etc.) are drilled and concreted first
with or without reinforcement. Reinforced piles (numbered 2,4,6, etc.) are cut into these piles about one
day later after the concrete in the first group has achieved its initial set but before it becomes too hard.
Sometimes, reinforcement is provided in every pile, however, this is generally practical only when the
piles are reinforced with steel sections. Due to the difficulty in cutting reinforcing bars, it is usually
necessary to place rebar cages only in alternate piles.

The wall facing may consist of reinforced shotcrete, pre-cast concrete panels, or cast-in-place concrete
(Figure 7-14). A drainage gallery may be provided behind the concrete facing to intercept and channel
any seepage that may penetrate the bored pile wall. Grouting is sometimes performed behind the joints
between piles to reduce water seepage.

1 3 5 7 9 8 6 4 2

(a)
2 1 4 3 6 5 8

(b)

Figure 7-15: Construction Sequence (a) Tangent Pile and (b) Secant Pile Walls. (Xanthakos, 1994)

7 - 25
7.6 WALL DESIGN

7.6.1 Sheet Pile Walls

Cantilever Wall

A cantilever wall derives its support from the structural stiffness of the wall sheeting and the passive soil
resistance below the exposed base of the wall. This type of wall is limited to heights up to about 5 m and
is generally suitable only in granular soils or stiff clays. For cantilever walls greater than about 5 min
height the maximum bending moment in the wall may exceed the capacity of commonly available sheeting,
or ground displacement behind the wall may become excessive. Typically, the penetration of the sheeting
will be 1.0 to 1.5 times the exposed height of the wall, depending on soil and groundwater conditions,
ground slope, and surcharge loads.

The lateral displacement of a cantilever wall penetrating a sand layer is shown in Figure 7-16. The wall
rotates about point 0, resulting in the reversal of active and passive earth pressures in the three different
zones shown. The corresponding net earth pressure distribution on the wall is shown in Figure 7-16(b)
and 7-16(c). Based on this net pressure diagram, a cantilever wall in granular soils can be analyzed
following the procedure shown in Figure 7-17.

Two other general cases for sheet pile walls include a) sheeting penetrating cohesive soils but retaining
granular soils, and b) sheeting penetrating cohesive soils and retaining cohesive soils. For these cases, the
strength of the clay changes with time and consequently the earth pressure changes with time also.
Immediately after the sheet piling is installed, the lateral earth pressures can be calculated based on the total
stress method of analysis, i.e. using undrained shear strength parameters. Figure 7-18 presents the earth
pressure diagrams for the end of construction condition. For the long term condition, sheet piling in clay
should be analyzed using effective strength parameters, c' and <1>', obtained from triaxial shear strength
tests. Assuming that the effective cohesion value is small, the value of c' can conservatively be taken as
zero. The long term condition can then be analyzed based solely on the effective friction angle, <1>', of the
clay using the procedure illustrated in Figure 7-17 for cantilever walls in granular soils.

\
,
I
/:::"
\
\ .' '. ."'"
." 0:. 'j.:."
.' 1- .
I
\
I
I
\
I
Zo~e ..\
\

=~ I 50"
\
I
I
I
Dredge \
line y

(a) (b) (c)

Figure 7-16: Typical Deformation Conditions and Pressure Distribution for Cantilever Walls: (a)
Yielding Pattern of Cantilever Wall Penetrating a Sand Layer; (b) Net Actual Earth
Pressure Distribution; (c) Simplified Net Earth Pressure Distribution. (After Das, 1990)

7 - 26
B

I yD~-y(H+D)Ka y(H+D)Ka I yDKa


I I
yD~ y(H+ D)I(,,-yDKa
y(H+D)~

1. Assume a trial depth of penetration, D, using the following approximate correlations:

Standard Penetration Resistance, N (Blows/300mm) 0-4 5-10 11-30 31-50 +50


Depth of Penetration, D, in terms of H 2.0H 1.5H 1.25H 1.0H 0.75H

2. Determine the active and passive lateral pressure using appropriate coefficients. If the
Coulomb method is used, it should be used conservatively for passive pressure.

3. Satisfy the requirement of static equilibrium as follows:


(a) Sum of horizontal forces ~FH=O. In terms of pressure areas, this is written as follows
Area (EAIA~ - Area (FBAz) - Area (EeJ) = 0 (7-1)
Solve this equation for the distance Z. For a uniform granular soil,
Z={KpDZ - KA(H+D)Z} I {(Kp - KA)(H+2D)} (7-2)
(b) Take moments about point F. If sum of moments is other than zero, readjust D.
Repeat calculations until the sum of moments about F is zero.

4. Compute maximum moment at point of zero shear.

5. Increase D by 20 to 40 percent to result in approximate factor of safety of 1.5 to 2.


Alternatively, it may be preferable to use a factor of safety of 1.5 to 2 on ~ in Step 2 and
thus use a reduced value of passive resistance; in such a case the computed depth should not
be increased.

~: In addition to the earth pressures shown, add water pressures due to differing water levels,
and any pressures due to surcharge loads.

Figure 7-17: Analysis for Cantilever Wall in Granular Soils. (Adapted from USS, 1975 and Teng, 1962)

7 - 27
2C=~

Sand backfill
qjyc z
H COHESIVE
SOIL

COHESIVE
SOIL
D

2"~ J
li"Il..;..-_2_q_.+_y_'H_......
Passive
pressure
k I
C

I.
pp=y'(z-H)+q. e
2q.-y'~ 2q.+y'H •

= Vertical pressure behind exposed qu = Unconfined compressive strength


base of wall due to backfill of clay
(submerged weight below water level) y' = unit weight (effective) of soil

a) Embedment in Cohesive Soil Retaining b) Embedment in Cohesive Soil Retaining


Granular Soil Cohesive Soil

1. Select point d and the depth D to satisfy the conditions of static equilibrium; that is sum of
horizontal forces equal zero and the sum of the moments about any point equal zero.
2. The computed depth, D, should be increased by 20 to 40 percent to obtain the total design depth
of penetration. Alternatively, it may be preferable to use the factor of safety 1.5 to 2.0 on the
unconfined compressive strength and thus a reduced value of passive resistance; in such a case
the computed depth should not be increased.

These procedures are based on the following restrictions:


• The ratio of overburden pressure to undrained shear strength (Le., stability number N =y'H/c)
must be less than 3.
• The active earth pressure shall not be less than 0.25 times the effective overburden pressure at
any depth.

Notes:
1. In addition to the earth pressures shown, add water pressures due to differing water levels, and
any pressures due to surcharge loads.
2. The procedure and the lateral earth pressure diagram presented above are based on undrained
shear strength parameters, and should therefore be used only to analyze the end of construction
condition. The long term condition for the wall should be analyzed using effective shear
strength parameters and the procedure illustrated in Figure 7-17.

Figure 7-18: Earth Pressure Distributions and Design Procedure for Cantilevered Walls Embedded in
Cohesive Soils - End of Construction Case. (Adapted from USS, 1975, and Teng, 1962)

7 - 28
In developing the design lateral pressures, the lateral pressure due to traffic, and other surcharge loads,
differential water pressure and backfill compaction, if necessary, are added to the lateral earth pressure in
accordance with procedures described in Section 2.6 of Chapter 2.

Figure 7-19 presents the AASHTO procedure for analysis of cantilever walls. This procedure is a
simplified version of the procedure presented in Figures 7-17 and 7-18. For granular soils, the
simplification is achieved by assuming the passive resistance increases linearly with depth on the front side
of the wall and by substitution with a concentrated force, F, on the back side near the bottom of the wall.
For cohesive soils, the simplification is achieved by assuming a rectangular passive resistance distribution
on the front side and by a concentrated force on the back side. These simplifications may result in some
error, but reduce the computation effort. It is recommended that the more rigorous solutions illustrated
in Figures 7-17 and 7-18 be used for final design of major wall structures and those with limited extra
bending capacity or embedment.

Anchored Walls

When the height of the retained soil behind a cantilever sheet pile wall exceeds about 5 m, it is generally
necessary to provide a lateral support near the top of the wall to reduce the wall bending moments and to
limit ground displacements behind the wall. The lateral support may consist of drilled soil or rock anchors;
rods tied to anchor blocks, an anchor wall, or anchor piles; or internal strut or raker braces. This
configuration is referred to as an "anchored wall".

Anchored sheet pile walls derive their lateral support from the passive pressure on the front of the
embedded portion of the wall, and the pull out resistance of the anchors. Anchors reduce the required
depth of penetration and also permit wall heights to be increased to about 11 m, depending on the soil and
groundwater conditions. For higher walls the use of high-strength steel sheet piling, reinforced sheet
piling, or additional levels of anchors may be necessary. The methods of analysis presented in this section
apply for walls with a single anchorage level; refer to Section 7.6.2 if multiple levels of anchors are used.

Anchored walls with one level of anchors may be designed using triangular active and passive earth
pressure distributions as discussed in Chapter 2. Using these applied pressures, there are two basic
methods for design of anchored sheet pile walls: (a) the free earth support method, and (b) the fixed earth
support method. In selecting the method to be used for wall design, consideration must be given to a
number of factors such as the relative stiffness of the sheet piles, the depth of pile penetration, the relative
compressibility and strength of the soil, and the amount of anchor yield. Figure 7-20 shows the assumed
deflection and bending moment distribution for the two methods.

Typically, the fixed earth method is used in granular soils and stiff cohesive soils and for walls of relatively
lower stiffness; the free earth method is used in cohesive soils, for walls of relatively higher stiffness, and
for situations where the sheeting penetration may be restricted because of obstructions, a shallow rock
surface, or sheeting length limitations. The fixed earth method generally results in smaller tie rod loads
and smaller wall bending moments, but a greater depth of sheeting penetration than the free earth method.

Free Earth Support

This method is based on the following assumptions (Teng, 1962; USS, 1975):

• The soil into which the piling is driven is incapable of producing effective restraint from passive
pressure to the extent necessary to induce negative bending moments.

7 - 29
SIMPLIFIED DESIGN PROCEDURE
1. Determine the active earth pressure on the wall
due to surcharge loads, the retained soil and
differential water pressure above the exposed wall
base (refer to Figure 2-3 for determination ofK.).
2. Determine the magnitude of active pressure at the
exposed wall base (P*) due to surcharge loads,
retained soil and differential water pressure, using
the earth pressure coefficient Kaz
3. Determine the value of x=P ·'[(Kpz -K 12)'Y'z] for
the distnbution of net passive pressure in front of
the wall below the exposed wall base (refer to
Figure 2-3 for determination ofK. and~.
4. Sum moments about the point of action of F to
determine the embedment (DO> for which the net
passive pressure is sufficient to provide
equilibrium.
5. Determine the depth (point a) at which the shear
0
0 in the wall is zero (i.e., The point at which the
Do N areas of the driving and resisting pressure
- diagrams are equivalent).
'"
0 6. Calculate the maximum bending moment at the
point of zero shear.
a
F 7. Calculate the design depth D= 1.2Do to 1.4Do for
a safety factor of 1.5 to 2.0.

(a)
2S u1
M .-rT'lrl-----"-
COt£SIVE

T\ SOIL I
ly;. SUI)

..'
COHESIVE
COHESIVE SOlL2
SOIL IYi,Suz)

.
Q
.
Q

(b)

NOTES:
(1) Surcharge and water pressures must be added to the above earth pressures.
(2) Forces shown are per unit horizontal length of wall.
(3) The ratio of overburden pressure to undrained shear strength (Le. stability number,
N=y'H/c) must be less than 3; otherwise wall failure may occur.
(4) For Case (c) the active earth pressure shall not be less than 0.25 times the effective
overburden pressure at any depth.

Figure 7-19: Simplified Earth Pressure Distributions and Design Procedure for Cantilevered Walls. (a)
for Granular Soils; (b) for Walls Embedded in Cohesive Soil Retaining Granular Soil; and
(c) for Walls Embedded in Cohesive Soil Retaining Cohesive Soil. (AASHTO, 1992)

7 - 30
, ;.;oJ-

I Anchor I Anchor
I I
I I
Deflection I
Deflection I
I I
I Groundline I
I ~~J- '

Groundline \
'~J- '\

\
" ---- --
(a) (b)

Figure 7-20: Variation of Deflection and Bending Moment. (a) Free Earth Support Method; (b) Fixed
Earth Support Method. (Das, 1990)

• The piling is driven just deep enough to assure stability, assuming that the passive resistance is fully
mobilized.

• No pivot point exists along the embedded portion of the wall, Le., no passive resistance develops on
the back side of the piling.

• The piling is free to rotate but is not permitted lateral movement at the anchor level. At the ultimate
capacity, the piling fails by outward movement rotating about the anchor level.

• The sheet piling is perfectly rigid as compared to the surrounding soils.

With the above assumptions, the problem can be solved by considering static equilibrium. Figure 7-21
presents analysis by the free earth support method for two common cases.

The free earth support method assumes a rigid piling and therefore uses lateral pressures according to
hydrostatic distribution as shown in Figure 7-21. However, in reality, steel sheet piling is quite flexible
causing the earth pressure to redistribute and thus differ from the classical hydrostatic distribution. With
increasing flexibility, the embedded portion of the wall rotates about its lower edge (Figure 7-20a) causing the
center of the passive pressure to move closer to the exposed wall base. This in tum decreases the maximum
bending moment. Figure 7-22 presents a procedure to reduce the maximum design moment obtained from the
free earth support method based on wall flexibility considerations. In practice, it is recommended that the
moment reduction factor be used only for cases where a factor of safety was applied to the soil shear strength
parameters prior to computing the penetration depth. Also, it is recommended that the reduced bending
moment not be lower than that computed by the fixed earth method.

7 - 31
T

L
H

y
Horizontal anchor load per unit horizontal
length of wall
Pa = Force from earth pressure above point a
+ other horizontal forces (except T)
yH1+y'H w
Yeq = H + H
1 w
Kal = Active pressure coefficient for the soil
below the exposed base of wall

1. Compute appropriate values for active and passive pressures. If the Coulomb method is used
to compute the passive soil pressure, the penetration should be computed using the more
conservative values presented in Step 9, below.
2. Compute weight of overburden and surcharge load at exposed wall base, YeqH.
3. Locate the point of zero pressure:
y = YeqHK al
(p p - Pa)

where Pp and Pa are the passive and active earth pressures, below the exposed base of wall,
respectively. (7-4)
4. Compute P a, the resultant force due to active earth pressure (and surcharge, if any) above point
'a', and its distance, L, below the anchor level.
5. Take moment about the anchor level as follows and solve for D1•
(7-5)

6. Compute the horizontal anchor load: (7-6)

7. Determine the maximum bending moment at the point of zero shear.


8. Select pile section for the maximum moment or use the moment reduction theory, Figure 7-22.
9. Add 20 to 40 percent to D1 to provide for safety margin, or divide Pp by a factor of safety of
1.5 to 2.0 in above e uations.
Figure 7-21: Analysis by Free Earth Support Method: (a) Sheet Piling in Granular Soils. (Adapted from
Page 1 of 2 Teng, 1962 and USS, 1975)

7 - 32
H

T= Horizontal anchor load per unit


D' horizontal length of wall

Pa = Force from earth pressure above point a


+ other horizontal forces (except T)

1. Oetennine the immediate and long-tenn strength of the soil by undrained tests (<\>"'0) and
drained tests (c"'O), respectively·. Check if the· design height is less than the critical height,
H <2q)Ycq, in order that a net passive pressure will result; otherwise failure can occur.
2. Compute Pa , the resultant force due to active earth pressure (and surcharge, if any) above
point 0, and its distance, L, below the anchor level.
3. Compute the weight of overburden and surcharge load at the exposed base of the wall, YcqH.
4. Take moment about the anchor level as follows and solve for Of.
LP a - (2q u - Ycq H) D '(H I + ~ D ') = 0 (7-7)

5. Compute the horizontal anchor load:


T = P a - (2q u - Ycq H)D' (7-8)

6. Oetennine maximum bending moment at the point of zero shear.


7. Select pile section for the maximum moment or use the moment reduction theory, Figure
7-22.
8. Add 20 to 40 percent to Of to provide for safety margin, or divide qu by a factor of safety of
1.5 to 2.0 in above equations.

Figure 7-21: Analysis by Free Earth Support Method: (b) Sheet Piling Backfilled with Granular
Page 2 of2 Soil and Embedded in Cohesive Soil. (Adapted from Teng, 1962 and USS, 1975)

7 - 33
FOR GRANULAR son.s

1.0 r------~---___,r__---__,----._,

0.8 ........,~-...:.--+---~-+---

Mil
0.6 t-------t-=~---_+-""o::_--___ir__ 1
MinaI
0.4 r-----T-----r----"'-....::::::::lt-~c::::::::::==:::t
Unsafe
section
0.2 t------+------+-----I-----~
Stiff Flexible
piles piles -
OL- ..L- ~ ....&.- ....

-4.0 -3.5 -2.5 -2.0

Plot of log p against MiMmax for sheet piles penetrating sand. M d is design or reduced
moment; ~ is theoretical maximum moment computed as per Figure 7-21; Md<~

1. Choose a sheet pile section.


2. Find the section modulus, S, of the selected section per unit length of the wall.
3. Determine the moment of inertia, I, of the section per unit length of wall.
4. Obtain the total height of the sheet pile, H +D, and calculate the relative flexibility of the wall,
p, as follows:
p = 10.91 X 10-7 [(H +Dt/(EI)] (7-9)

where (H +D) is in meters, E is the modulus of elasticity of the pile material in MPa and I, the
moment of inertia, is in m4/m of wall.
5. Find log p.
6. Find the moment capacity of the pile section chosen in Step 1 as Md=0allS where 0811 is the
allowable stress in pile material.
7. Determine MiMmax •
8. Plot log p (Step 5) and Md~ in the above Figure.
9. Verify that the plotted point for the selected pile section falls above the curve (loose sand or
dense sand, as appropriate). If the plotted point falls below the curve, select another sheet pile
section and re eat the above Ste s 2 throu h 9.

Figure 7-22: (a) Moment Reduction for Anchored Sheet Pile Wall Analyzed by Free Earth Support
Page 1of2 Method for Granular Soils. (After Rowe, 1952, 1957; Das, 1990)

7 - 34
FOR COHESIVE SOILS
1.0
Log p = -3.1

0.8
Mil a=0.8
M m..
0.6 0.7
0.6
0.4 ~_.....L. _ _..L..-_--L._ _...L-._--L_ _ ....L-_---'

1.0

,\ ~
I
Log p = -2.6

0.8

~
Mil
M m..
0.6
\'r'---
~ "'-
.=::---
a
0.7
0.8

0.6-
0.4
1.0
Log p = -2.0

0.8
Mil
M mu
0.6

0.4 ~_.....L. _ _..L..-_--L._ _...L-._--L_ _....L-"';:""---I

o 0.5 1.0 1.75


Stability number, S.

Plot of Sn against MiMmax for sheet piles penetrating clay.

1. Compute the Stability Number, Sn' as follows:


Sn = (c/yeqH)(1 +c/cu)'Iz (7-10)
where cu( =q/2) is the undrained cohesion below the exposed base of wall, ca is wall adhesion
(See Table 2-1).
For design purposes, Sn = 1.25(c/yeqH) may be used. Any sheet piling driven into cohesive
soils should have a minimum Sn of 0.3 times a desired factor of safety; otherwise failure may
occur.
2. Compute ex as ex=H/(H+D).
3. Compute p as per Equation (7-9) in Figure 7-22a.
4. For the magnitudes of ex and Sn obtained above, determine the Md~ for various values of
log p from above charts and plot Mi~ against log p. Values of log p can be interpolated.
5..Verify that the plotted point for the selected pile section falls above the curve developed in
Step 4. If the plotted point falls below the curve, select another sheet pile section and repeat
the above Ste s 1 throu h 5.

Figure 7-22: (b) Moment Reduction for Anchored Sheet Pile Wall Analyzed by Free Earth Support
Page 2 of2 Method for Cohesive Soils. (After Rowe, 1952,1957; Das, 1990)

7 - 35
Fixed Earth Support

The fixed earth support method is based on the assumption that the toe of the wall is restrained from
rotation as shown in Figure 7-20b. The deflected shape reverses its curvature at the point of contraflexure,
c, and becomes vertical at the toe of the wall.

The problem may be solved using the theory of beam on elastic foundation, but this procedure is laborious.
In practice, a procedure known as the equivalent beam method proposed by Blum (see Tschebotarioff,
1951) is used. This method utilizes a theoretical relationship between the angle of internal friction, <1>, and
the distance below the exposed wall base to the point of contraflexure. Figure 7-23 presents the equivalent
beam method, which is limited in its use to granular soils. For cohesive soils, use methods based on the
theory of beam on elastic foundation.

Anchor Systems

General types of anchor systems used for sheet pile walls are shown in Figure 7-24. Figure 7-25 and 7-26
presents criteria for design of a deadman anchor system. Figure 7-25 can be used to determine the
additional lateral load acting on the wall, if any, due to the location of the deadman anchor, and Figure 7-
26 can be used for design of the deadman anchor. The pile anchor system may be designed using
procedures presented in Module 8 -Deep Foundations. The design of soil or rock anchors is discussed in
Section 7.7.

A deadman anchor, if used, must be located outside the potential active failure zone developed behind the
sheet pile wall as shown in Figure 7-25. If the anchor system must be located closer to the wall than the
friction angle slope shown in Figure 7-25, anchorage resistance is decreased and an additional passive
reaction is required for stability of the wall base.

The tie rods must be protected to resist corrosion. Typically, the rods are coated with one or more
applications of tar-base paint, and spirally wrapped with a durable fabric or fiber glass tape after the
application of the first coat of paint. The tie rods may also be enclosed in a rigid casing or supported
vertically to eliminate sag for cases where the backfill will settle significantly or unevenly.

In determining the feasibility of using ground anchors at a particular location, consideration must be given
to obtaining the necessary underground easements, identifying the location of buried facilities, and the
suitability of subsurface conditions.

Global Stability Analysis

The global stability of the entire wall system is analyzed by using an appropriate method discussed in
Module 3 - Soil Slopes and Embankments. For sheet pile walls embedded in cohesive soils, it is also
necessary to determine the limiting height of the wall by Eq. 7-10 in Figure 7-22b; walls exceeding this
limiting height may fail due to over stressing of the cohesive soils below the exposed base of the wall.

7 - 36
x
~ .~. ~ ::::' Tie O.25H
V' f:";
::':.:
rod
O.20H
\
Dredge
line
." Granular
backfill
H
O.15H
\
D
b

d e C c
L e C
O.lOH

O.05H
o
20
"
25
'\

30
r\.
"-
35 "
(Pp'-Pa') Degrees
(al lbl (cl (dl (el

The equivalent beam method assumes a hinge at the point of contraflexure (point b), since the
bending moment there is zero. The part above the hinge can then be treated as a separate, freely
supported beam with an overhanging end as shown in Figure (d). The reactions R and T and the
bending moments can be determined from statics and simple beam theory. The lower portion,
below the point of contraflexure, can also be analyzed as a separate, freely supported beam on two
supports, R and C. Based on these assumptions, the sheet piling in granular soils may be designed
by the following steps:

1. Compute the active and passive lateral pressures using appropriate earth pressure coefficients.
2. Determine the distance, y, from
y=(YJIK,.')/(Pp'-Pa') (7-11)
where YJI = weight of backfill and surcharge load above the exposed base of wall, using
buoyant weight for soil below the water level
K,.' = coefficient of active earth pressure for the soil below the exposed base of wall.
pp' , Pa' = passive and active earth pressures in the soil below the exposed base of wall.
3. Locate the point of contraflexure, b, by the chart shown in Figure (e).
4. Determine the reaction R at the point of contraflexure. R is the horizontal reaction at point b
obtained by treating the piling above b as a simple beam supported at b and at the anchor level
as shown in Figure d.
5. Treat the lower portion of the piling, eb, as a simple beam and determine the dimension eb by
equating the moment about the base e to zero.
6. The depth of penetration, D, is equal to the sum of the dimensions eb and x. To provide a
margin of safety, either add 20 to 40 percent of the calculated depth of penetration, D, or reduce
the value of Kp' by dividing it by a safety factor of 1.5 to 2.0.
7. Determine the maximum bending moment at the point of zero shear, and size the piling.

Generally, the point of contraflexure and the point of zero pressure are very close and for design
purposes the value of x may be taken equal to y. Therefore, the depth of penetration may be
expressed in a simple equation as follows:
D=y+{(6R)/(Pp'-Pa')}~ (7-12)
where y = distance from the expose base of wall to the point of zero pressure
R = horizontal reaction at point 0, obtained by assuming the piling is simply supported
at point 0 and at the anchor level

Figure 7-23: Analysis by Equivalent Beam Method. (After USS, 1975; Teng, 1962)

7 - 37
Original ground
~~ \ I
\\ Y/ Excavattono
. f
\ / trench to construct

~
., adeadman
Anchor rod "~': '.:.:
,": '~,

(a)
Final ground Final ground
~ ~ Cast anchorage
Anchor rod +--......,A1 Anchor rod ~--

Original
ground

Piles

(b) (c)

Steel. wood. or
concrete wall ;::;vC4SS'~~' _ _

Backfill
Wale
Anchor
rod
Line of short sheet
I piles
: Pairs to greater depth
Iat frequent Intervals

(d) (e) Grouted anchor


in soil or rock

Figure 7-24: Sheet Pile Anchorages. (a) Cast-in-Place Deadman; (b,c) Pile Anchorages; (d) Steel
Sheet Pile Wall Used as Anchorage; (e) Soil or Rock Anchor. (Bowles, 1988)

7 - 38
a c d f
-r----r-

uH
/ I /
Tie Rod
/ -p. / Jil
t---------.,.;----~- _a_/-/-_ ~
Rupture Surface -to; e ~
of Active Wedge I / / /Possible Position
H / / / / of Anchor Block
/ w / / ./

l
/-Slopeat
/ / / Friction Angle
/ / /
/ / /
= 45 .!. / / / .>/
2/ / /
.. / / cI>"
~

y
//
////
"
Reaction
~}
to P .-..,I;---,......f-----t.-
b Estimated Point of Zero
Moment in Wall

Anchor block left of bc provides no resistance.


Anchor block right of bf provides full resistance with no load transferred to wall.
Anchor block between bc and bf provides partial resistance and transfers load app to wall.

.tt.P

w
F

Vector diagram for free body abed where


PA = active force on back of de at anchor block
W = weight of soil wedge abed
F = resultant force on surface be
aP p = increased load on wall

Figure 7-25: Effect of Anchor Location Relative to the Wall. (NAVFAC, 1988)

7 - 39
CONTINUOUS ANCHOR WALL LOCATED BETWEEN RUPTURE
SURFACE AND SLOPE AT FRICTION ANGLE

a
-If---~ ~ h
I e 'c. Ap "
" . PASSIVE waxi CJ'A~~/h/3
H
!£.. " 'ZICTrJE WEDGE OF BULKHEAD-.s-7
45- 2,' Forces Per Linear Foot of Anchor Wall
I Anchor Wall Right ofCC' Anchor Wall Left ofCC'
I
I Forh! ~ h/2 Forh! ~ h z
......... ,' Pp =1/2K pyh 2 Pp =1/2K pyh 2 -(p'-p')
p.
p. = 1/2 K. yh 2
Pp = 1/2 K pyh 2 - (1/2 K pyh: -1/2 K. yh:)
~ obtained from Figure 2-5 p. =1/2k. yh 2
Using - 0/<1> = 0.5 Ka is obtained from Figure 2-2

EFFECT OF DEPTH AND SPACING OF ANCHOR BLOCKS

Anchor Resistance For hi ~ h/2


1. Continuous Wall:
A Ultimate ~/d = P: ,P: where ~/d is anchor resistance andp: ,P:
~2 taken per linear foot of wall.

~-L
2. Individual Anchors:
If d > b + h, ultimate ~ = b( P: ,P: ) + 2 Po tan<l> where Po =
~+ Resultant force of soil at rest on vertical area cde or c"de.
If d = h + b, ~/d is 70% of ~d for continuous wall.
SECTlONA-A SECTION A-A
CONTIMJOUS INDIVIDUAL L for this condition is L' and L' = h.
WALL ANCHORS If d < h + b, ~/d = ~/d - L/L' (.3 ~d), L' =h
Anchor Resistance for hi < h/2
Ultimate ~/d or ~/d equals bearing capacity of strip footing of
width hi and surcharge load y (h - h/2), see Module 7 (Shallow
Foundations)
Use friction angle <1>': where tan <1>' = 0.6 tan <1>.

General Requirements:
1. Allowable Value of ~ and ~ = (Ultimate Value)/2, factor of safety of 2 against failure.
2. Values ofKA and ~ are for cohesionless materials. If backfill has both c and 4> strengths,
compute active and passive forces according to Chapter 2.
3. Tie rod is designed for allowable ~ or~. Tie rod connections to wall and anchorage are
designed for 1.2 times the allowable ~ or ~.
4. Tie rod connection to anchorage is made at the location of the resultant earth pressures
acting on the vertical face of the anchorage

Figure 7-26: Design Criteria for Deadman Anchor Systems. (NAVFAC, 1988)

7 - 40
7.6.2 Soldier Pile and Lagging Walls

Earth Pressure Distributions

The earth pressure distributions applicable for different stages of top-down construction of a soldier pile
and lagging wall are illustrated in Figure 7-7 and are discussed below.

Stage 1 Install soldier piles to required depths below the proposed bottom of excavation.
Excavate in front of the wall and between soldier piles to a level slightly below the first
support row. Install lagging as excavation proceeds. This is the cantilever condition
during which the top of the wall moves outward and the classical active and passive
earth pressures with triangular (hydrostatic) distribution are developed in the back and
front of the wall, respectively.

Stage 2 Apply a support (strut or anchor) force, excavate to the next level of support and place
lagging. As the excavation proceeds downward, the wall tends to rotate around the level
of previously installed supports, with the top moving into the soil face and the bottom
tending to deflect out. Walls with greater penetration would deflect as shown in Figure
7-27b, and increased wall base movement would occur as the penetration decreases
(Figure 7-27c). For this condition, which is similar to an anchored wall, the pressure
distribution may be assumed to remain triangular.

Stage 3 Install subsequent support levels, if necessary. As the excavation proceeds downward,
the wall portion below the last installed support tends to deflect outward, while the wall
portion above the last installed support has essentially no further deflection. The earth
pressure is redistributed to the upper wall section as each lower section deflects.

For stable ground masses (See Figures 7-31 and 7-32), the final distribution and magnitude oflateral earth
pressure on a completed wall with two or more levels of anchors or braces constructed from the top down
may be computed using the apparent earth pressure distributions shown in Figure 7-28.

For unstable or marginally stable ground masses, the design earth pressure may exceed those shown
in Figure 7-28; for such cases, the wall design loads should be estimated using methods of slope
stability analysis which incorporate the reactions at the anchors or braces. In developing the design
earth pressures, particular consideration should be given to wall displacements that may occur and
their impact on adjacent structures or utilities.

Following are several key conditions which formed the basis of the empirical pressure distribution
diagrams presented in Figure 7-28 (Goldberg, 1976; Canadian Geotechnical Manual, 1992):

• The empirical data was obtained from strut load measurements, not ground anchors.

• The pressure diagrams were obtained as an envelope of the maximum pressures found from
several projects. Thus, they do not represent the actual loads that might occur.

• The pressure diagrams reflect the variability that normally occurs in construction, including
load variations caused by construction sequencing, preloading operations, lock-off losses,
temperature effects, etc.

7 - 41
Resultant Wall Applied Resultant Wall Applied
Earth Forces Displacement SoH Pressure Earth Forces Di§placement SoH Pre§§ure

(a) (b)

Resultant Wall Applied Resultant Wall Applied


Earth Forces Displacement SoH Pressure Earth Forces Displacement SoH Preswre

\ \

\ \

- 1----;\

1----; "
\
'----'-- ~

(c) (d)

Figure 7-27 Various Stages of Top-down Construction. (Cheney, 1988)


(a) Stage 1 - Cantilever Condition (Fixed Earth Support),
(b) Stage 2a -Intermediate Condition with One Level of Support (Fixed Earth Support),
(c) Stage 2b -Intermediate Condition with One Level of Support and Excavation to
Second Level of Support (Free Earth Support),
(d) Stage 3 - Multi-level Support (Free Earth Support)

• The empirical pressure diagrams do not include water pressure. When the groundwater
immediately behind the wall is not drained, the empirical earth pressures should be computed
using submerged unit weight for the soils below the groundwater level, and the hydrostatic
water pressure should be added to the empirical earth pressure. In cohesive soil, if there is
a potential for shrinkage cracks becoming filled with water, hydrostatic water pressure must
be added within the depth of the cracks.

• Earth pressures due to surcharges such as traffic loads, construction equipment, adjacent
structures, or other loads must be added to the empirical pressures.

7 - 42
EMPRICAL PRESSURE DISTRIBUTION TOTAL FORCE

(a) SAND PI =0.65 K A yH 2


Fl--~&-'\-----' -'-.

F2 - - + I "
"
Ph =0.65K A yH
Pa
=050K
' A
yH 2

~)
'\.

~_ F3_ - + I "" H where K A =tan


2
( 45 - PI
-=1.30
Pa
'------I' where:
"
" PI =total horizontal force from

" '\.
trapezoidal pressure distribution.
P a = total horizontal force from
Rankine (triangular) distribution.

P/Pa is a ratio representing


the increase in total horizontal
force for excavation support
systems with multiple levels
of bracing.

b) SOFT TO MEDIUM Form = 1.0:


H'( 1-:.1
CLAY (No> 6)
For clays, base the P,"O.87S Y
selection on

p. "O.sOYH'( 1- :.1
K =1-m 4c = 1- 4m
A yH No

loE m = 1 except where cut


is underlain by soft
normally consolidated
clay, then m = 0.4

(c) STIFF CLAY (No < 4)


~
o
lI'I

o For 4 < No < 6, use larger of


~ diagrams (b) and (c).
lI'I
("'l
o Ph =0.2yH; Ph =O.4yH
I 2

Use lower value when


movements are minimal and
construction period is short

Figure 7-28: Pressure Distribution for Braced Walls. (Modified After Terzaghi and Peck, 1967)

7 - 43
Layered Soils

The idealized soil conditions shown in Figure 7-28 may not be readily applicable for cases involving
stratified soil conditions, an irregular ground surface or irregular surcharge loading. Under such
circumstances, the following approach may be taken (Goldberg, et al., 1976):

1. Determine the total lateral force from classic active earth pressures; for complicated stratification,
an irregular ground surface or irregular surcharge loading the lateral force may be determined by
trial wedge stability analysis.

2. Once the lateral force from active earth pressure is determined, it should be increased by an
appropriate value of P/PA (ratio of force from the empirical diagram to the force from active earth
pressure or wedge equilibrium). The right side of Figure 7-28 provides P/Pa values for simplified
soil conditions

3. Select an appropriate pressure distribution diagram from Figure 7-28 and compute the maximum
lateral pressure, Ph' such that the total lateral force from the selected pressure distribution diagram
equals the lateral force determined in Step 2 above. Complicated cases may need field
measurements during construction to confirm the assumed distribution.

Effect of Support Type

Wall support elements may consist of bracing (struts), soil or rock anchors, or rakers as shown in Figure
7-2. The empirical pressure diagrams in Figure 7-28 are directly applicable for support by bracing. The
applicability of these empirical pressure diagrams for ground anchors and raker systems is discussed below.

Soil and Rock Anchors

• Soil and rock anchors are usually proof-tested to at least 133 percent of design load; then the load
is reduced and locked off at about 80 percent or more of the design load. This process removes
some of the load variations inherent with the installation and preloading of internal braces. Also,
soil and rock anchors are generally insulated from temperature effects. The use of soil and rock
anchors therefore results in better load control and more uniform load distribution.

• Similar to internally braced walls, the soil or rock anchors restrain the wall from lateral
displacement. As the excavation proceeds downward, the wall portion below the last installed
anchor level tends to deflect outward, rotating about the last installed anchor level.

Although soil and rock anchor systems provide better control and more uniform distribution of the support
loads than internal bracing systems, walls supported by multiple levels of soil or rock anchors are
commonly designed using the same lateral earth pressures (Figure 7-28) as used for internally braced walls.
If the top anchor level is near the top of the wall, a rectangular rather than trapezoidal pressure distribution
is used.

Rakers

Raker-braced walls usually yield more at the top than walls supported by other methods because additional
earth is removed in front of the wall prior to positioning the bottom of the raker. Also, rakers are typically
not preloaded. As a result, a triangular active pressure distribution is generally appropriate.

7 -44
Soldier Pile Spacing

Above the bottom of excavation, the soldier piles must be designed for the total lateral force in the span
distance between adjacent soldier piles. The appropriate pressure magnitude obtained from Figure 7-28
is multiplied by the spacing of the soldier piles, s, to obtain the design load per unit length on a given
soldier pile. Below the bottom of the excavation, the active earth pressure is assumed to act only on soldier
pile width, b, while the passive earth pressure develops over an effective element width equal to three times
the soldier pile width, i.e. 3b, or the center-to-center spacing of the soldier piles, whichever is less (Figure
7-29).

Methods of Analysis

Use methods described in Section 7.6.1 to analyze the cantilever and single level of support load cases.
For two or more levels of support, follow the procedures presented in Figures 7-30 through 7-33.

Support Spacing and Forces

Support struts and braces are normally spaced 4.5 to 6 m apart horizontally, and 3 to 4.5 m vertically.
Smaller spacing is typically required for soil and rock anchors. A horizontal wale is typically used to
transfer loads from a pair of soldier piles to a support anchor or strut, or the anchor supports can be
installed through each of the soldier piles (Figure 7-6).

Once the support forces are obtained using the methods shown in Figures 7-30, the wales and the braces
(struts or rakers) are designed based on structural considerations. The controlling design criterion for
braces is the column-action of combined axial and bending stress. In that regard, a pipe section is more
efficient than a wide flange section.

Connections and details are of critical importance. Improper connections between strut and wale or
between wale and the wall are perhaps the most common causes of difficulties in support systems. They
can lead to twisting, buckling, and rotation of members.

Soldier Pile Design

Flexural Capacity

The soldier pile is sized based on the maximum moment from all the load cases described above as well
as any applied vertical loads. If the maximum moment corresponds to an intermediate load case which is
short-term, design codes may allow an over stressing of the pile member.

vertical Capacity and Embedment

In addition to the lateral pressures, a significant vertical load may be imposed on the soldier pile by the
vertical component of the support force (i.e., inclined anchors), or by dead and live loads supported by the
soldier pile. This vertical force must be resisted by the embedded length of the soldier pile below the fmal
grade. The embedded length is determined by using methods described in Module 8 - Deep Foundations.
The embedded length may also be governed by global stability considerations as discussed in Module 3 -
Soil Slopes and Embankments, and base stability as illustrated in Figure 7-33.

7 - 45
H

Active Pressure
D

(a)

Anchor
Force

Bottom of
Excavation

~-- Active Pressure


D

(b)

Final Grade
Temporary
Excavation

Active Pressure

(c)

Figure 7-29: Pressure Diagrams for Soldier Pile and Lagging Walls. (a) Cantilevered During
Excavation for Top Row of Supports; (b) Intermediate Excavations for Subsequent
Support Installation; (c) Final Constructed Condition Assuming Excavation for
Utilities in Front of Wall. (After Cheney, 1988)

7 -46
•.1,
'I"'I"l:"'-~~""""~­
~"v.n."'"

Support Forces
/ "
'II+ ~
(Struts/Anchors) ,
, II+----I~
-Appropriate Pressure
Wale~
Diagram Including Surcharges
-'II+---~
Sheeting or Soldier--4
Beams with Lagging ,
I
J•

Deflected P sition....J
.~~I...--tr/
,,
,"'-"AWf'

1. Compute lateral pressure above base of cut using the appropriate pressure distribution from
Figure 7-28. For walls supported by raker braces use active pressures. Add lateral pressure
due to surcharge loads and water pressure, if appropriate.
2. Determine stability of the base of cut by methods shown in Figures 7-31 and 7-32 for granular
and cohesive soils, respectively. If base is stable, determine pile penetration from Figure 7-
33. If base is unstable, piles penetrate as shown in Figure 7-33 and the unbalanced force, Ph'
acts on the buried length. Verify the computed pile penetration is adequate to support the
applied vertical loads (i.e, vertical dead and live loads, or the vertical component of anchor
loads).
3. Determine the bending moments in soldier piles and the reaction at each bracing level by
conventional methods of structural analysis, Tributary Area Method (Figure 7-34 (a», or by
treating the soldier pile as a continuous beam with rigid support at the bracing levels and at
the point of zero net pressure below the bottom of excavation. This Continuous Beam Method
is illustrated in Figure 7-34(b). Alternatively, the soldier pile can be analyzed using the Beam
on Elastic Foundation Method (Hetenyi, 1946), using elastic springs to model each of the
support levels and the passive soil resistance below the bottom of the excavation. This later
method, shown in Figure 7-34(c), provides an estimate of inward wall movement.
4. For computing the bending moment in soldier piles and wales, use 80 percent of the pressures
shown in the diagrams in Figures 7-28. For rigid walls, such as slurry (diaphragm) walls, use
the full pressure (100 percent) shown in the diagrams in Figures 7-28 for structural design of
the walls.
5. For determining the loads in internal bracing elements and ground anchors, use the full
pressure (100 percent) shown in the diagrams in Figures 7-28.
6. The moment in wales supporting concentrated loads (from soldier piles or anchors) should be
calculated on the basis of statics.
7. Determine the bending moment in the wall elements and loads in braces and anchors for each
stage of excavation. Often the maximum moment and the maximum support load occurs
during the construction stage when excavation to a support level is complete, just prior to
installation of support.

Figure 7-30: Design Procedure for Walls with Multiple Levels of Supports. (Modified from
Goldberg, et al., 1976)

7 -47
Sheeting or soldier piles and lagging

B
~I
.. y+~

i~
I
I
i~P.tan~,
H

I P a = Resultant
I Active Pressure

I~
I

Base stability is independent of H and B, but varies with 1, cP and seepage condition. The Safety
Factor is given by:

(7-13)

where Ny2 = bearing capacity factor (see Module 7 - Shallow Foundations)

If groundwater is at a depth of (B) or more below base of cut:


11 and 12 are taken as moist unit weight

If groundwater is static at base of cut:


11 = moist weight, 12 = submerged weight

If seepage is moving upward to base of cut:


12 = (saturated unit weight) - (uplift pressure)

For upward moving seepage, check resistance to piping.

Figure 7-31: Stability of Base for Braced Cuts in Cohesionless Soils. (NAVFAC, 1988)

7 -48
Cut in Clay, Depth of Clay Unlimited (h > O.7B>
L = length of cut
q

B;..j2

c
Load on buried length of sheeting: f,S.=Nc - - > 1,5
'YH+q
28 where:
If d> r:;
3",2 d = embedment below excavation
p' •. 7 hHB - 1.4cH - '/fcB) Nc = bearing capacity factor
2B
c = unit strength of clay
Ifd<- H = excavation depth
3J2 y = unit weight of soil
p' "' 1.5dB hHB - 1.4cH - '/fcB) q = surcharge
If F,S, < 1.5 drive sheeting deeper and reanalyze

Cut in Clay. Depth of Clay Limited by Hard Stratum (h ~ O.7B)

/'""'7--..-...,...-...,----, 10 Effect of H
HIB NcDINc
9 0 1.00
0.5 1.15
Value of hlB
I 1.24
0.3 2 1.36
:h----------~7 3 1.43
0.4 4 1.46
:.--------~~--__t6
L::::t==:::t:=:l::::=t;:=:;t::=-0~.5~=:l::::::;15
.33
1.4 1.6 1.8 2.0 2.2 2.4 2.6
Hard stratum
Ratio c2/c,

For BtL = 0:
C
1
FS =N cD ---~1.5 (7-14)
yH+q

where Nco = bearing capacity factor


c1 = unit strength of soft stratum

Note: In each case, friction and adhesion on back of sheeting is disregarded. The clay is assumed to
have a uniform shear stren th (c) throu hout the failure zone,

Figure 7-32: Stability of Base for Braced Cuts in Cohesive Soils. (NAVFAC, 1988)

7 - 49
Penetration in Relatively Uniform Competent Soil Conditions

\ Ph(From Design Earth Presssure


-r--..JIJ------'\_ Figures 7-28)

Bottom of Cut
1-'::' /: .

Competent Soils
(Medium Dense to
Dense Granular Soils v •
and Stiffto Hard Clays)
V = Shear at tip of sheeting or soldier pile = 0
R d= Horizontal component of load at
bottom brace or anchor level
L d...,= Vertical distance from bottom brace
or anchor level to bottom of cut

1. Compute equivalent reaction at base of cut: Re = 0.5PhLd_e


2 Compute depth, x, such that: Pp = R. + Pa
Use minimum F. S. = 1.5 to compute passive pressure.
3. Check Mmax ~ Allowable moment of sheeting or soldier pile.
4. Drive to depth D = 1.2x.
5. Verify that the computed penetration is adequate to support applied vertical loads (Le. vertical
dead and live loads, or the vertical component of anchor loads).

Penetration in a Weak Underlying Layer

Use the following procedure to determine the depth of penetration when the excavation is underlain
by weak soils, consisting of loose granular soils and soft to medium stiff clays, and in cases where
the computed active pressure below the bottom of excavation exceeds the computed passive
pressure:

1. Use nominal penetration of 0.2H or 1.5 m, whichever greater.


2 Check base stability (see Figures 7-31 and 7-32 and Global Stability).
3. Design sheeting or soldier pile for cantilever condition below bottom brace or anchor level.
Since the passive resistance of the weak soil layer is small, a large load may be developed at
the bottom brace or anchor level.
4. Verify that the computed penetration is adequate to support applied vertical loads (Le. vertical
dead and live loads, or the vertical com onent of anchor loads).

Figure 7-33: Procedure for Determining Depth of Penetration. (Goldberg, et ai., 1976)

7 - 50
L,
Area A
:;~rA
L /2 2
L2

AreaB
~R.
L/2
L3

AreaC

~RCI·
~RDlL s
AreaD

AreaE
.."':%: ...~/../,
I.. Ph ·1

(a)

'//. "/// I
'//A:. /'//

~
I
Supports Ins1alIed
Deflected Wall Position
II ,
I

/
/
I
>s_rts Loading ~
I
\
>EI.stiC Suppo",
Loading ~
\
I
I
r-- Deflected Shape
'''''
Deflected Shape

I Subgrade

/-\''''\'-
, . '\/>, \
~Passive Limit
. Subgrade Reaction
~Springs

(b) (c)

Figure 7-34: Analytical Models for Walls with Multiple Levels of Supports by (a) Tributary Area
Method. (After Goldberg, et aI, 1976) (b) Continuous Beam Method. (Tamaro,
1990); (c) Beam on Elastic Foundation Method. (b & c After Gould, 1993)

7 - 51
Lagging Design

Because of soil arching (Figure 7-35), lagging design is based primarily upon experience or semi-empirical
rules. Two commonly adopted semi-empirical rules are as follows (Goldberg, et ai., 1976):

• Vary the amplitude of the design pressure diagram with maximum pressure at the soldier piles and
minimum pressure midway between the soldier piles.

• Reduce the basic pressure distribution diagram (Figures 7-28) by applying a reduction factor.
Typical reduction factors range from 0.5 to 0.8.

The required section of lagging is then computed based on flexural considerations assuming the lagging
as a beam simply supported at the soldier piles. Table 7-1 can be used for determining the allowable
flexural stress for some common types and grades of woods used for wood lagging. Table 7-2 presents
recommendations for selecting wood lagging thickness based on soil type, depth of excavation and clear
distance between soldier piles.

~ ~ _---:.- _ ~_-
_

,
- ili - .'Yt'- - -;It -
'"
.......
,I, I
,,-: _ - - - - - _
",""" - :
~I I I 2 < I I I I ? I I ?? I
",""'.iIIIo.-
I
~
, I I I
,- .
I I I I 2
' ili .111"".,- ••.. Z.$
-2'1lo. 2S
2 I 2 2 2

Soil Outside Arch

Plan of Arching Action

Displacement

.,.,...-- ....... /r-' i""'--I"-. ....-


, i-
• i-
'"
Stress

Figure 7-35: Arch Action at Soldier Pile and Lagging Walls. (Goldberg, et ai., 1976)

7 - 52
TABLE 7-1
STRENGTH PROPERTIES FOR TYPICAL GRADES OF TIMBER
(American Institute of Timber Construction, 1974)

Allowable Flexural Stress Modulus of Elasticity


Wood Type and Grade fb, MPa E,MPa

Douglas Fir - Larch, surfaced dry or


surfaced green used at max. 19 %
moisture content
Construction 8.3 10,300
Select Structural 14.1 12,400

Douglas Fir - South, surfaced dry or


surfaced green used at max. 19%
moisture content
Construction 7.9 7,600
Select Structural 13.4 9,700
Northern Pine, surfaced at 15 % moisture
content, used at max. 19% moisture
content
Construction 7.2 8,300
Select Structural 12.1 10,300
Southern Pine, surfaced at 15 % moisture
content K. D., used at 15 % max.
moisture content
Construction 9.0 10,300
Select Structural 15.5 13,100
Southern Pine, surfaced dry, used at
max. 19% moisture content
Construction 8.3 9,700
Select Structural 14.1 12,400

7 - 53
TABLE 7-2
RECOMMENDED THICKNESS OF WOOD LAGGING (AFTER GOLDBERG, et al., 1976)

Unified Depth Recommended Thickness of Lagging (roughcut) for Clear Spans of:
Soil Descriotion Classification (m) 1.5 m 1.8 m 2.1 m 2.4 m 2.7m 3.0m

-
~
0
Vl
Silt or fine sand and silt above water table

Sands and gravels (medium dense to dense)


ML, SM-ML

GW,GP,GM,
GC, SW, SP,
0-8 50mm 75mm 75 mm 75mm l00mm l00mm

~ SM

~
~
Clays (stiff to very stiff); non-fissured CL,CH 8 - 18 75mm 75mm 75mm l00mm l00mm 125 mm

~ CL,CH
0 Clays, medium consistency and yH <S
U c

Sands and silty sands (loose) SW, SP, SM

..... Clayey sands (medium dense to dense) below


VI ~ water table SC 0-8 75mm 75mm 75mm l00mm l00mm 125 mm
.J::>.
~~
~O
Clays, heavily over consolidated, fissured CL,CH 8 - 18 75mm 75mm l00mm l00mm 125 mm 125 mm
~V)
Cohesionless silt or fine sand and silt below
Q
water table ML, SM-ML

-
~
~O
....:IV)
Soft clays yH >S
c
CL,CH 0-5 75mm 75mm IOOmm 125 mm ----_.._--- -------_..-
~;g
Z~
Slightly plastic silts below water table ML 5-8 75mm l00mm 125 mm 150mm ---------- -- ...-------

~~ Q
Clayey sands (loose), below water table SC 8 - 11 l00mm 125 mm 150mm ---------- _....-----_.. -_........_-_ ..

Notes: 1) In the category of "potentially dangerous soils", use of lagging is questionable; a continuous type of wall (slurry wall or tangent/secant pile
wall) should be considered. 2) The values shown are based on construction grade lumber.
7.6.3 Slurry Walls

The design procedure presented in Section 7.6.2 for Soldier Pile and Lagging Walls is generally applicable
for the design of slurry walls. However, there are important differences in the design of slurry walls since
a) they are generally continuous structures for their full depth of penetration, b) they are considerably
stiffer than soldier pile and lagging systems, and c) they are typically designed for full hydrostatic water
pressure behind the wall. The design considerations specific to slurry walls are discussed below.

Earth Pressure Distribution

For analysis and design of slurry walls use the following lateral pressures:

• For cantilever walls and walls with a single support level, use active earth pressure on the back
face of the wall.

• For walls with two or more levels of supports, use the pressure distribution diagrams presented
in Figure 7-28.

• For walls supported by rakers, use active earth pressure on the back of the wall.

• Determine the lateral earth pressure below the bottom of the excavation using the pressure
diagrams presented in Figure 7-29 except that the active and passive pressures shown are
continuous along the length of the wall.

• To the lateral earth pressures add the pressures due to surcharge loads, if any.

• Add the water pressure acting on the wall.

• Analyze each stage of excavation using the appropriate earth pressure distribution identified above.
It should be noted that the governing load or bending moment in individual members may occur
at intermediate construction stages rather than at the fmal excavation depth.

• In determining the lateral earth pressure, ignore wall friction.

Methods of Analysis

Use the methods described in Section 7.6.1 to analyze the cantilever and single level of support load cases.
For cases with a single level of support, slurry walls should be considered as stiff wall elements:
accordingly there should be no reduction in the computed free earth bending moment.

For two or more levels of support, follow the procedure presented in Figure 7-30 through 7-34.

Conventional reinforced concrete slurry walls are typically analyzed as a beam spanning vertically between
support levels. For SPTC walls, however, the reinforced concrete wall is typically analyzed as a beam
spanning horizontally between the steel piles, and the steel piles are used to transmit the applied loads
vertically to the support levels.

In addition to the methods of analysis presented in Figure 7-34, slurry walls can also be analyzed using the
Finite Element Method (FEM). FEM can be used to model displacement of supports, soil changes with
time, and varying properties at the soil structure interface, as well as drained and undrained soil conditions.

7 - 55
A particular advantage of the FEM is that it can provide an estimate of horizontal and vertical soil
movements on both sides of the structure and at each stage of the excavation. However, FEM is difficult
to use because it requires a definition of the elasto - plastic and time dependent behavior of the soils, and
a determination of the detailed sequencing of construction. FEM requires considerable experience in the
collecting, processing and evaluation of the data. Also, the results may vary significantly with variation
in the assumed soil and structure properties. At present, the use of FEM is generally limited to complex
structures that require an assessment of soil and structure displacements for the different stages of
construction.

A discussion and comparison of various methods of analysis for slurry walls are provided by Tamaro and
Gould (1993) and Kerr and Tamaro (1990).

7.6.4 Tangent and Secant Pile Walls

The earth pressure distribution diagrams, types of wall support, methods of analysis for design of tangent
pile and secant pile walls are the same as those discussed in Section 7.6.3 for slurry walls.

When a steel reinforcing cage or steel structural section is placed in all of the bored piles, the design load
for each pile is determined by multiplying the appropriate soil and water pressure diagrams by the center-
to-center spacing of the piles. When only alternate piles are reinforced, the design load for the reinforced
piles is determined by multiplying the appropriate pressure diagrams by the center-to-center spacing of the
reinforced piles.

7.7 ANCHOR DESIGN AND CONSTRUCTION

7.7.1 Anchor Components

Ground anchors, sometimes referred to as tiebacks, are structural tension elements that receive their
support in soil or rock. They are used to resist lateral loads on earth retaining structures and for other load
resisting applications. Figure 7-36 presents a schematic of a typical ground anchor. The basic anchor
components include:

• Tendon: a prestressing steel element consisting of steel wires, strands or bars.

• Anchorage: a bearing plate and an anchor head (or threaded nut) used for prestressing and lock-off
of the tendon.

• Grout: a Portland cement based mixture which allows load transfer from the tendon to the ground,
and provides corrosion protection for the steel element.

The tendon is divided in two sections: (a) the bond (fixed or anchorage) length through which the load is
transferred to the ground, and (b) the unbonded (free) length which is free to elongate elastically and
transmit the resisting force from the bond length to the structural facing element.

Anchors may be used for either temporary or permanent support applications. The design and construction
of both are similar, except that permanent anchors generally are designed for lower stresses in the tendon,
and are provided with added protection against corrosion.

7 - 56
- - - Structural Element

;- Protective Plastic Cap and Nut

- - - - Anchor Head

- - - Antk:orrosion. Grease

.;----- Bearing Plate

; - - - - - Prestressing Steel

- - - Secondary Grout
- - - Sheath

Figure 7-36: Schematic of a Typical Ground Anchor. (Cheney, 1988)

7 - 57
The use of ground anchors offers the following advantages:
• Eliminates internal bracing that may impair construction operations.
• Generally easy installation.
• Reliable load capacity, since each element is tested..
• More uniform load distribution.

Disadvantages of ground anchors are:


• May not be suitable for some soils (See 7.7.2).
• Requires penetration through the retaining wall, creating potential path of groundwater seepage.
• The vertical load component of the anchor is an added vertical load on the wall.
• Potential interference with utilities, basements and other underground facilities.
• Corrosion potential of the steel elements.

7.7.2 Suitable Ground Conditions for Anchors

Ground anchors can be used in essentially all cohesionless soil deposits. Good performance has also been
observed for anchors in stiff to hard cohesive soils with N-value greater than 9 (Weatherby 1982). Soil
deposits not generally suitable for the anchor bond length (Cheney, 1988) include:

• Organic soils
• Cohesive soils with an average liquidity index greater than 0.2; and
• Cohesive soils with an average liquid limit greater than 50.

In addition to these criteria, caution should be exercised with soils which have a liquid limit greater than
20 percent as they may exhibit excessive creep (Weatherby 1982).

Soil tests should be performed to measure the aggressiveness of the anchor environment, especially if field
observations indicate corrosion of existing structures. The most common and simplest tests are electrical
resistivity, pH, chlorides and sulfate. Critical values and relevant ASTM and AASHTO standard tests for
these parameters are listed in Table 7-3. Anchors placed in environments where the results from anyone
of these tests indicate critical values must be provided with additional protection in both the free and bond
lengths.

Ground anchors can be used in almost any type of rock, although the allowable bond stress varies
considerably with the type of rock. If no adverse jointing system or plane of weakness exists, an ultimate
bond stress may be determined from the rock strength, anchor installation procedure, and grout strength.
However, for an adversely jointed rock mass or weakened planes containing weathered material, the rock
mass strength should be determined by an engineering geologist based on a review of rock cores and testing
of samples of joint fill material (Cheney, 1988). Particular care should be exercised in shales and other
types of rock which may be subject to softening during anchor installation operations (Munfakh, et ai. ,
1987).

Physical constraints which may preclude anchor installation are utilities, basements, or other underground
facilities which cannot be relocated, and right-of-way restrictions.

7 - 58
TABLE 7-3
CRITICAL VALUES AND TEST STANDARDS TO EVALUATE SOIL AGGRESSIVENESS
(Cheney, 1988)

Test Test Designation Critical values

Resistivity ASTM G 57, AASHTO T-288 < 2000 ohm centimeters


pH ASTM G 51, AASHTO T-289 < 4.5
Chlorides ASTM D 512, ASTM D 4327, AASHTO T-291 > 100ppm
Sulfates ASTM D 516, ASTM D 4327, AASHTO T-290 > 3000 ppm

7.7.3 Determination of Minimum Spacing, Free Length and Inclination

Figure 7-37 provides a procedure for determining the minimum vertical and horizontal spacing and free
length of the anchors. The criteria illustrated in Figure 7-37 are considered appropriate for walls with
multiple levels of anchors, and for walls with a single level of anchors when wall embedment is determined
by the free earth support method (see Section 7.3.1).

The free length should also be checked based on external stability of the wall-anchor-soil system by using
conventional circular or wedge methods of analysis described in Module 3 - Soil Slopes and Embankments.
The anchors must develop their resistance behind any possible zone of slippage. The stability analysis
considers the wall and soil mass as a gravity wall using a simplified cross-section as illustrated in Figure
7-38. A minimum safety factor of 1.3 is used for non-critical applications, and 1.5 for critical applications
where adjacent structures are present above the wall or soft soils exist below the wall.

Anchors must be inclined at least 10 degrees below horizontal to facilitate tendon installation and grouting.
The flattest possible inclination is desirable to reduce the vertical component of the anchor load transferred
to the wall and to increase the horizontal resistance component to the maximum possible. Steeper
inclinations may be dictated by buried utilities, adjacent foundations, right-of-way constraints, or
unfavorable soils. The inclination of any particular anchor or group of anchors can be varied within the
system to avoid obstructions or account for anomalies in subsurface conditions. If a steeper inclination is
needed in an area, the anchor load can be increased by lengthening the bond zone to provide the same
horizontal design component, or the same anchor bond length can be used with closer horizontal or vertical
spacing.

7.7.4 Selection of Anchor Tendons

The load in an anchor can be computed using the procedure and lateral earth pressure distribution presented
in Sections 7.3 through 7.6. Since the anchor is inclined, the anchor load will always be greater than the
horizontal component determined from the pressure distribution diagrams. The required tendon bar size,
or number and size of strands or wires can be selected from Table 7-4 based on steel design procedures
from current prestressing steel codes. Both the test load and the permanent load should be considered in
anchor size selection. Suggested limits for temporary test loads are between 75 percent and 80 percent of
the Guaranteed Ultimate Tensile Strength (GUTS), while for anchor design loads the limits are between
50 and 60 percent of GUTS (Cheney, 1988). The design load for permanent anchors should be limited to
50 percent GUTS.

7 - 59
e ....-.d
. e
V')

~ '-'

H CI'.l
>

Sv = 2.5 to 3.5 m (commonly used)


Minimum Free Length = 4.5 rn

(a)

1---------1
.--_bl-
T
/
Wall Face

Sh > minimum 3b, or 1.2 m,


whichever is greater
= 2 m to 3m (commonly used)

Figure 7-37: Determination of Minimum Spacing and Free Length (a) Vertical View, (b) Plan View
(Cheney, 1988)

7 - 60
A E

w 84>

84>
a; = cI> on Failure Plane
T
F.S.=~
T dcs

(a) (b)

Tenninology:

W Weight of soil mass within the failure surface.


Pa Design force acting on the surface DE. A driving force due to water must be considered when
below the water table. While Pa has been drawn horizontally, it could have been an inclined force.
S. Frictional component of soil resistance. This force is applied at an angle, a; = cI> (full obliquity),
to the normal base of the soil mass. It should be noted that a cannot be greater than the internal
friction angle of the soil. Mobilized shear resistance acting along the plane is (S. cos cI» tan cI>.
Sc Component of soil resistance due to cohesive soil strength.
PA Active earth force between point A and point C. Point C is the point of zero shear.
T Ground Anchor force.

Notes:
• The wall is not considered part of the free body diagram shown.
• The entire anchor load is assumed to be transmitted between points D and F, and the presence of
the bond length between points D and G is neglected.
• Ifwater is present, the pore water forces acting on the free body must also be considered.

Figure 7-38: Stability Analysis of Anchored Wall Systems: (a) Free Body Diagram, and (b) Force
Polygon (After Cheney, 1988)

7 - 61
Anchor sizing and selection should be done prior to design of the bond length because the required hole
diameter varies as a function of the tendon size. Table 7-5 can be used for estimating hole diameter
required for strand or bar anchors.

TABLE 7-4
TYPICAL STEEL PROPERTIES AND DIMENSIONS FOR TENDONS (Cheney, 1988)

Diameter Area Ultimate Stress Ultimate Load


Type of Tendon (rnm) (rnm2) GUTS (MPa) (kN)
(I) Wire (ASTM A-421) 6.4 31.6 1655 52.5

(2) Cables or Strands 6.4 31.6 1862 45.8


(ASTM A-416/A-779) 12.7 126 1862 183.8
15.2 183 1862 260.8

(3) Bars or Rods 12.7 126 1103 151.7


(ASTM A-722) 15.9 198 1586 314.2
25.4 506 1034 567.4
25.4 506 1103 608.8
31.8 794 1034 834.4
31.8 794 1103 890.0
34.9 955 1034 1054.7
(1)
Many wires are used in an anchor to obtain load carrying capacity.
(2)
Several cables or strands are used in an anchor.
(3)
There are many bar or rod types and manufacturers. The data presented here are typical and
are not meant to indicate the only bar types available.

TABLE 7-5
RELATIONSIDP BETWEEN TENDON SIZE AND DRILLED HOLE DIAMETER (Cheney, 1988)

*Minimum Suggested Hole Diameter (rnm)

Tendon Type Single Corrosion Protection Double Corrosion Protection


Number of Strands
4 100 150
7 115 165
9 130 180
11 140 190
13 150 200
17 165 215
Bar Diameter (rnm)
25.4 65 90
31.8 70 95
34.9 75 100
*Includes allowance for centralizers.

7 - 62
7.7.5 Design of the Bond Length

Due to the inherent variability of soil deposits, and the variety of anchor types and installation methods
currently used, the design of a ground anchor bond length is based on a number of relationships that have
been developed from field results and empirical studies. Some important points which the designer must
consider in determining the bond length are listed below (Cheney, 1988):

1. A minimum bond length should always be specified in the contract documents. Suggested values
are 3 m for rock and 4.5 m for granular or cohesive soils. The specifications should give the
contractor the responsibility for determining the actual bond length to be used, provided that the
actual length exceeds the specified minimum length.

2. Bond lengths exceeding the customary values of 12 m in soil and 7.5 m in rock do not efficiently
increase the anchor capacity.

3. Set the bond length a reasonable depth below the ground surface to permit high pressure grouting
without damage to existing facilities and to ensure adequate overburden pressure to mobilize the
full friction between the soil and the grout. A 4.5 m minimum overburden cover over the bond
zone is recommended for anchors of average capacity, Le., 700 kN or less.

4. The bond length may be designed to mobilize capacity in more than one deposit providing:
a. All deposits are suitable to sustain bond zone loading.
b. Weaker bond zone deposits overlie stronger; Le., clay soil over granular soil over rock.
c. The total bond length is computed based on the weaker deposit.
d. A significant portion of the total bond length is in the upper deposit.
e. Preproduction test anchors are used to establish production acceptance criteria.

5. The minimum horizontal spacing of anchors should be either three times the diameter of the
bonded zone or 1.2 m, whichever is larger. If smaller spacings are required, consideration can
be given to differing anchor inclinations between alternating anchors. Typically, horizontal
spacings of 2 to 3 m are commonly used.

Anchor Capacity in Soils

The load-carrying capacity of a soil anchor is dependent on the type of soil surrounding and above the bond
length of the anchor. The ultimate anchor capacity per unit length may be preliminarily estimated using
guidelines presented in Table 7-6. These guidelines are for preliminary design of straight shaft anchors
installed in small diameter holes using a low grout pressure. Other anchor types and installation procedures
could provide other ultimate anchor capacities.

When possible, the load capacity of a soil anchor should be established by a pull-out test. The allowable
anchor load capacity, Ta , is determined by dividing the test load capacity, Tl , by a factor of safety FS.
Where no pull-out tests are carried out, the maximum allowable anchor load is commonly estimated by
multiplying the bond length by the ultimate transfer load and dividing by a factor of safety of 2.5.

Anchor Capacity in Rock

The ultimate load transferred from the bond length to competent sound rock deposits may be estimated
from the rock type in Table 7-7. Lower values may be recommended after a review by a geologist,
especially if the rock mass strength is controlled by discontinuities. The maximum allowable anchor design

7 - 63
TABLE 7-6
PRESUMPTIVE ULTIMATE VALUES OF LOAD TRANSFER
FOR PRELIMINARY DESIGN OF ANCHORS IN SOIL
(Modified after Cheney, 1988)

Relative Density/Consistency(1) Estimated Ultimate Transfer


Soil Type (SPT Range) Load (kN/m)
Sand and Gravel Loose (4-10) 145
Medium Dense (10-30) 220
Dense (30-50) 290

Sand Loose (4-10) 100


Medium Dense (10-30) 145
Dense (30-50) 190

Sand and Silt Loose (4-10) 70


Medium Dense (10-30) 100
Dense (30-50) 130

Silt-clay mixture with minimum Stiff (10-20) 30


LL, PI, and LI restrictions, or Hard (20-40) 60
fme micaceous(2) sand or silt
mixtures
(I)
Values corrected for overburden pressure.
(2)
The presence of mica tends to increase the volume and compressibility of sand and silt deposits
due to bridging action and subsequent flexibility under increased pressures.

TABLE 7-7
PRESUMPTIVE ULTIMATE VALUES OF LOAD TRANSFER
FOR PRELIMINARY DESIGN OF ANCHORS IN ROCK
(Modified After Cheney, 1988)

Rock Type Estimated Ultimate Transfer Load (kN/m)


Granite or Basalt 730
Dolomite Limestone 580
Soft Limestone 440
Sandstone 440
Slates and Hard Shales 360
Soft Shales 150

7 - 64
load in rock may be estimated by multiplying the bond length by the ultimate transfer load and dividing
by a safety factor of 3.

Anchor design in rock is also based on an allowable grout-to-rock bond stress acting over the bond length.
The allowable bond stress should be smaller than 1/30 times the unconfined compressive strength of the
rock and 1/30 times the unconfined compressive strength of the grout, but not more than 1300 kPa.

Whenever possible, the capacity of an anchor in rock should be established by means of a pull-out test.

7.7.6 Corrosion Considerations in Design

The long-term performance of an anchored wall is influenced to a great extent by the corrosivity of the soil
or rock and groundwater as well as the corrosion potential of the anchor tendon. Corrosion of steel
structural members (soldier piles, steel lagging, wales, reinforcement cages, etc.) and decay of wood
lagging also influence the long-term performance of an anchored wall.

All permanent ground anchors must be protected against corrosion. Depending on anticipated corrosion
rates and the design life of a temporary earth retaining structure, corrosion protection measures may also
be required for temporary anchors. Head connections, couplers, and other junctions between the different
parts of the anchor are particularly vulnerable to corrosion.

Most of the tendon corrosion occurs within 2 m of the anchor head in the zone of aeration because air
reacts with moisture to generate corrosive solutions. In addition to air, tendon corrosion is influenced by
many variables including chemical composition of groundwater, temperature, rate of groundwater flow,
and stress concentration in the tendon. Stray electrical currents accelerate corrosion due to establishment
of an electric current potential along the tendon.

Anchor Protection Against Corrosion

The extent of corrosion protection measures depends on such factors as level of aggressivity, consequence
of failure and cost. In general, protection measures consist of one or more physical barriers between the
tendon and the surrounding ground. The use of sacrificial steel in lieu of physical barriers is not
recommended as corrosion is rarely uniform and is typically concentrated at localized pits or surface
irregularities (Littlejohn, 1990).

Following are basic guidelines (Cheney, 1988) for selection of corrosion protection measures:

1. All anchors used for permanent applications require corrosion protection in the free stressing
length and at the anchor head.

2. For routine applications, only a single degree of corrosion protection (grease and sheath) will be
required in the free stressing length with only grout used to protect the bond zone, i.e., a minimum
grout cover of 13 mm in thickness.

Protection of Free Length

The free length is protected by a variety of means since the protection is not required to transfer stresses
from the tendon to the ground. The protection must:

a. Accommodate elastic movements during testing and stressing.

7 - 65
b. Withstand potential damage during storage, handling, and construction.
c. Resist attack in aggressive environments.
d. Allow elastic movements after lock-off.

The most common methods of protecting the free length include the use of smooth sheaths filled with anti-
corrosion grease, heat shrink. sleeves, and secondary grouting after stressing. Except for secondary
grouting, the protection is usually in place prior to inserting the tendons in the hole.

Most frequently when "secondary grout" is mentioned, the procedure referred to involves filling the
annular space between the ground and the smooth sheath in the free length. Secondary grouting should
be done by gravity or low pressure methods to prevent damage to the sheath and load transfer to the ground
in the free length. In addition, the grout should terminate before the bearing plate area to prevent load
transfer up the grout column to the structure. Such secondary grouting, when done properly, provides an
extra insurance against corrosion and is recommended for most installations.

Secondary grout, in the true sense, refers to grout placement after stressing and lock-off around a tendon
with an unsheathed free length. After set, the grout surrounding the free length becomes bonded to the
ground and the tendon over the initially unbonded length. Figure 7-39 illustrates such an anchor. These
anchors are only normally recommended for semi-permanent, low-risk applications. When secondary
grouting is used for these anchors, extreme care must be taken to insure no voids in the grout exist beneath
the anchor bearing plate. Such areas are the most critical to protect against corrosion and require complete
encapsulation by cement grout.

Protection of Bond Length

Figures 7-39 through 7-41 show several schemes for corrosion protection of the tendon bond length. These
are briefly described below:

a. Simple Protection: The use of simple protection (Figure 7-40) relies on Portland cement grout to
protect the tendon, bar or strand, in the bond length. Steel will not corrode in a high pH
environment such as Portland cement which possesses pH values up to 12.6. When Portland
cement is used for protection, it is assumed that the pH will not be lowered with time. When the
simple protection is used, care should be taken that the tendon has at least 13 mm of grout cover
and the anchor grout extends 600 mm over the bottom of the free length sheathing.

b. Double Protection: Complete encapsulation of the anchor tendon is accomplished by uniformly


corrugated plastic or steel tube (Figure 7-41). The tube must be capable of withstanding the
deformations associated with transportation, installation, stressing and testing of the anchor, and
transferring the load applied to the tendon. Regardless of the encapsulating medium, the annular
space between the corrugated tube and tendon is usually filled with neat cement grout containing
admixtures to control bleed of water from the grout. Shorter tendons are grouted before insertion
in the hole.

Protection of the Anchor Head

The most critical area to protect from corrosion is in the vicinity of the anchor head connection. Below
the bearing plate, the corrosion protection over the free length is usually terminated to expose the bare
tendon. Above the bearing plate, the bare tendon is gripped by either wedges, nuts, or deformed in the
case of the wires. Regardless of the type of tendon, the gripping mechanism creates stress concentrations

7 - 66
at the connection. In addition, a very aggressive corrosion environment may exists at the anchor head
since oxygen is readily available. The vulnerability of this area is demonstrated by the fact that most
anchor failures occur within a short distance of the anchor head.

Figure 7-42 illustrates a typical protection system for the anchor head. A "trumpet," usually of steel or
strong durable plastic is used to overlap with the free length corrosion protection, and to protect the short
exposed length of the tendon below the anchor plate. One end of the trumpet is fastened and sealed to the
bearing plate and the other fitted with a deformable seal that fits tightly around the protective tube, but
allows free tendon movement within the trumpet. The annular space between the trumpet and the tendon
is filled with anti-corrosion grease.

The anchor head, including exposed tendon and friction grips or locking nuts above the bearing plate, is
protected by a cap filled with anti-corrosive grease or by embedment in concrete. Covering the anchor
head with a grease-filled cap allows future lift-off tests and/or load adjustment. When filling the trumpet
and cap, care is required to ensure that the grease or grout fills the entire space.

Protection of the bearing plate is accomplished by painting it with bituminous or other corrosion-resistant
materials. Care should be taken to ensure that the applied coating is compatible with the materials selected
for both inner head and outer head protection (Littlejohn, 1990).

Unbonded Length Bonded Length


(Secondary Grout) (Primary Grout)

................................................................................. .
••••••••••••• ••
,"

LEGEND:
1. Anchorage Cover 6. Secondary Grout
2. Anchor Head 7. Primary Grout
3. Anti-corrosion Grease 8. Tendon
4. Bearing Plate 9. Centralizer
5. Grout Tube 10. Spacer

Figure 7-39: Bonded Tendon (Cheney, 1988)

7 - 67
Unbonded Length Bonded Length

• ::.: .. °.:: 0
• •: . . . . . .

(a)

Unbonded Length Bonded Length

(b)

LEGEND:
1. Anchorage Cover 8. Grease Filled Plastic Tube
2. Anchor and Wedges or Nut 9. Grout
3. Anti-corrosion Grease 10. Individually Greased & Sheathed Strands
4. Bearing Plate 11. Anchor Grout
5. Trumpet 12. Spacer
6. Seal 13. Centralizer
7. Anti-corrosion Grease or Grout 14. Tendon

Figure 7-40: Simple Corrosion Protected Tendons. (a) Bar Tendon; (b) Strand Tendon (Cheney,
1988)

7 - 68
Unbonded Length Bonded Length

(a)

Unbonded Length Bonded Length

............:..... :. "':-., -

(b)

LEGEND: 9. Individually Greased & Sheathed Strands


1. Anchorage Cover 10. Spacer or Internal Centralizer
2. Anchor Head or Nut 11. Strand Tendon
3. Anti-corrosion Grease 12. Corrugated Plastic
4. Bearing Plate 13. Centralizer
5. Trumpet 14. Internal Grout
6. Seal 15. External Grout
7. Anti-corrosion Grease or Grout 16. End Cap
8. Plastic Tube 17. Coupler

Figure 7-41: Encapsulated Double Corrosion Protection. (a) Bar Tendon; (b) Strand Tendon
(Cheney, 1988)

7 - 69
Anti-corrosion
Grease

Anchorage ~~~
Cover

Anchor ---'-.·.·.·:.':.1.
Head

Figure 7-42: Anchor Head Details. (Cheney, 1988)

7.7.7 Construction

The method of anchor construction is greatly influenced by the type of ground at the site. Figure 7-43
illustrates the four main types of anchor construction in current practice, distinguished mainly by the
method of grouting used (Littlejohn, 1990).

Type A anchors are tremie or packer grouted, straight shaft boreholes that may be cased or uncased
depending on hole stability. This type of construction is used mainly in rock and hard clays.

In Type B anchors, grout is injected under low pressure. Usually, the hole is drilled using a hollow stem
auger or a casing. As the auger or casing is withdrawn, the grout is injected into the hole at a pressure
between 350 and 1000 kPa. This method of construction is most suitable for coarse granular soils, but is
also popular for fine grained cohesionless soils, stratified soils, or weak fissured rocks. The cement-based
grout penetrates the pores of these materials and, under pressure, compacts them resulting in larger
effective diameter and higher shear resistance around the anchor.

Type C anchors use injection pressures higher than 2000 kPa. The effective diameter of the bond section
is enlarged mainly by hydrofracturing of the surrounding ground, resulting in a fissured-like mass.
Sometimes, a two-step procedure is followed to create type C anchors. In this method, the high grout
pressure is applied during a secondary injection after the initially injected low pressure grout sets. Type
C anchors are usually used in fine cohesionless soils. They are also successfully used in stiff clays.

In Type D anchors, bells or under reams are created to provide both end bearing and side shear resistances
to pull-out. These anchors are usually constructed using drilled shaft rigs and rotary drills with tremie
grout. They are most suitable for firm to hard cohesive deposits, but have also been used in granular soils.

7 -70
tal Type A Ibl Type B leI Type C Idl Type 0
Figure 7-43: Main Types of Cement Grout Injection Anchorage. (Littlejohn, 1990)

Materials

The cement grout consists of sand, cement, water, and fly ash. Under normal conditions, most specialty
contractors prefer not to use chemical additives in the grout (Munfakh, et ai., 1987). Portland cement with
low sulfate content is normally used. For borehole anchorage, the grout mix should attain a minimum
cube strength (AASHTO T 106) of 24 MPa at 7 days (Cheney, 1988). The water/cement ratio of tendon
bonding grouts is usually in the range of 0.35 to 0.60.

The tendons usually consist of steel bars, strands or wires used singly or in a bundle (Table 7-4). The
tendon materials commonly used include non-alloy steel wire (6.4 mm in diameter), non-alloy 7-wire
strand (12.7 to 15.2 mm in diameter), or low alloy steel bar (25 to 35 mm in diameter).

Centralizers used to ensure that the tendon is centered in the grout column, should be made of materials
having no deleterious effect on the tendon itself.

Construction Sequence

Following are specific details and recommendations regarding the different steps used in construction of
ground anchors. Except where noted, these recommendations are from Littlejohn (1990).

Drilling

A variety of drilling equipment is available for installation of ground anchors, including hollow-stem
augers, rotary drills with casing, and down-the-hole hammers with casing (Photo 7-12). Irrespective of
the drilling method used, care should be taken not to use high pressures with any flushing medium to
minimize the risk of hydrofracturing the surrounding ground, particularly in built-up areas. For this
reason, an open return within the borehole may be beneficial to limit the pressure and give the driller an
indication of the type of ground through soil cuttings so that the fluid pressure may be adjusted
accordingly.

7 -71
(a) (b)

Photo 7-12: Drilling Equipment for Installation of Ground Anchors: (a) Hollow Stem Auger, and
(b) Down-the-Hole Hammer.

The drilled diameter of the hole should not be less than that specified, and allowance for swelling may be
necessary when drilling over consolidated clays or marl. A maximum angle tolerance of ±3 degrees is
usually specified (Cheney, 1988) and the anchor alignment should be carefully evaluated so no interference
of anchor zones takes place (closely spaced anchors may be staggered to avoid interference). An overall
deviation of alignment of 1 in 30 should be anticipated during drilling.

After drilling is completed and the hole is thoroughly flushed out, the drilled hole should be probed to
verify that no collapse of material has occurred before installation of the prestressing element. This
installation and the subsequent grouting should be carried out on the same day as drilling to avoid potential
ground deterioration.

Tendon Installation

Steel tendons should be stored indoors, ifpossible. If left outdoors, however, they should be stacked off
the ground under a waterproof cover that allows air circulation. Neither bare nor coated tendons should
be dragged across abrasive surfaces.

Over the bond length, centralizers should be placed at 3 m intervals to achieve the required minimum grout
cover of 13 mm (Cheney, 1988).

The tendon is lowered in the borehole manually or using mechanical equipment (photo 7-13). Immediately
before its installation, the tendon should be carefully inspected for damage and corrosion.

Grouting

The grouting operation involves ~ecting cement grout at the lowest point of the bore hole so the hole will
fill evenly without air voids. A grout pipe is often tied to the tendon before inserting it in the bore hole.

7 -72
Photo 7-13: Installation of Anchor Tendon

The cement grout is prepared by batching the dry materials by mass and mixing them mechanically, with
water added, for at least two minutes in order to obtain a homogeneous mix. High speed colloidal mixers
(1000 rpm minimum) or paddle mixers (150 rpm minimum) are used for grout preparation. The colloidal
mixers are preferred when grouting water-bearing ground since grout dilution is minimal. After mixing,
the grout is kept in continuous motion until it is pumped to its [mal position. Grout should not be used after
a period equivalent to its initial setting time. Prior to grouting, all air in the pump and the line should be
removed. An air tight system should be maintained at all times during grouting.

Prestressing

The stressing operation involves fitting the jack assembly on the anchor head, loading and/or unloading
of the anchors, locking off the load by anchor nuts or wedges, then removal of the assembly from the
anchor head. The equipment used for stressing operations need to be calibrated. A lock off load equal
to the design load is typically specified.

Stressing should not begin until the grout strength has reached a crushing strength of 24 to 30 MPa.
Cheney (1988) recommends a waiting period of 7 days before stressing can take place.

7.7.8 Testing and Stressing of Anchors

The load testing of ground anchors should be considered an integral part of the design. The typical types
of load tests include:

• Preproduction Tests: These tests are used to verify the design load safety factor and establish
the allowable anchor load.

7 -73
• Perfonnance Tests: These are conducted at the beginning of construction and periodically during
construction to verify short and long-tenn perfonnance of the anchor under the design load.

• Proof Tests: These are used to detennine the behavior of each production anchor after installation.

• Lift-Off Tests: These are used to confirm the load in the tendon after completion of installation and
lock-off of the applied load.

Photo 7-14 shows a typical set-up for anchor testing.

Preproduction Tests

Preproduction tests may be specified when unusual conditions are identified during the design stage.
Situations prompting such tests include: a) soil deposits in which no previous experience exists, (b) very
long anchors, c) difficult drilling conditions, or d) creep susceptible soils in the bond length.

The preproduction tests can be performed under a separate test contract awarded during the design stage,
or at the beginning of the construction contract. The anchors used for preproduction testing are usually
not incorporated in the final structure because of the potential damage that may be induced by the high test
loads. Also, the bond length for preproduction test anchors is typically shorter than that of the production
anchors.

Cheney (1988) provides recommended testing procedures for preproduction tests.

Photo 7-14: Testing Arrangement for Strand Tendon (Cheney, 1988)

7 -74
Performance Tests

Performance tests provide the necessary information to verify that the production anchors will be able to
hold the design load without excessive movement or creep. Cheney (1988) recommends performance tests
be run on the fIrst two anchors installed, then on 2 percent of the remaining anchors in rock or cohesionless
soils and 10 percent in cohesive soils.

The performance test is a cyclic test made by incrementally loading and unloading the anchor until the
maximum test load is reached. The typical loading and unloading sequence (PTI, 1997) is as follows:

AL, 0.25P, AL, 0.25P, 0.50P, AL, 0.25P, 0.50P, 0.75P, AL, 0.25P, 0.50P, 0.75P,
LOOP, AL, 0.25 P, 0.50 P, 0.75 P, LOOP, 1.20 P (1.25P), AL, 0.25 P, 0.50 P, 0.75P,
LOOP, 1.20P (1.25P), 1.33P (1.50P), 1.00 P, LL

where: AL = Alignment Load


P = Design Load
LL = Lock-Off Load

The maximum test load should be 1.33 times the design load for cohesionless soils and 1.5 times the design
load in cohesive soils. Each load or unload increment is held constant just long enough to obtain the
movement reading but no longer than one minute. The maximum load should generally be held for one
hour in cohesive soils to determine the long-term creep potential. Coarse granular soils and rock do not
generally exhibit creep; creep tests in such deposits may be terminated if negligible creep, i.e., less than
1.0 mm movement, is observed between the 1 minute and 10 minute readings of the test. The deflection
measurements at the maximum load level are taken at the following intervals: 1, 2, 3,4, 5, 6, 10, 20, 30,
40,50 and 60 minutes. At each load increment, the total moment of the pulling head should be recorded
to the nearest 0.03 mm.

Proof Tests

Every production anchor is proof tested. The proof test is a single cycle test in which the test load is
applied in increments until the maximum load used in the performance tests is reached. For granular soils
and rock, the maximum test load should be 1.33 times the design load (1.20 times the design load for
temporary anchors); for cohesive soils, the maximum test load should be 1.50 times the design load. The
maximum load is held constant for at least one minute, or until the measured deflection is negligible. The
typical loading and unloading sequence is as follows:

AL, 0.25 AL, 0.50 AL, 0.75 AL, 1.00 AL, 1.20 AL, 1.33 AL, AL (optional), LL

or: AL' 0.25 AL, 0.50 AL, 0.75 AL, 1.00 AL, 1.25 AL, 1.50 AL, AL (optional), LL

The maximum test load is maintained constant for 10 minutes, and total movement readings are recorded
at 1, 2, 3,4,5,6 and 10 minutes. If the total creep movement between 1 and 10 minutes exceedes 1 mm,
the maximum test load should be maintained for an additional 50 minutes, and the movement readings are
recorded at 20, 30, 40, 50 and 60 minutes.

When the results of the performance tests cannot be compared directly to the proof tests, the anchor should
be returned to the alignment load after the lO-minute hold at the maximum test load and raised again to the
lock-off load. This will permit the determination of permanent and elastic movements at the maximum test
load.

7 -75
Supplementary Extended Creep Tests

Extended creep tests should be made on permanent anchors in soils having Plasticity Index greater than
20. The creep test is conducted by incrementally loading and unloading the anchor in accordance with the
schedule of the performance test, except that at each new load maximum, the load is held constant in
accordance with the following schedule:

Observation Period
Load (minutes)
AL
0.25 DL 10
0.50 DL 30
0.75 DL 35
1.00 DL 45
1.20 DL 60
1.33 DL 300

The times for reading the creep movements are 1, 2, 3, 4, 5, 6, 10, 15,20, 25, 30,45,60, 75, 90, 100,
120, 150, 180, 210, 240, 270 and 300 minutes (where appropriate).

If the creep rate exceeds 2 mm per logarithmic cycle, the observation period may be extended in an attempt
to determine if the creep rate will diminish to the 2 mm per logarithmic cycle of time.

Extended creep tests are not normally performed on rock anchors since they do not exhibit time dependent
movements. However, anchors installed in very decomposed or argillaceous rocks may exhibit significant
creep behavior.

Acceptance Criteria

Three criteria are commonly used to determine the acceptability of anchor tests (PT!, 1997). These are
as follows:

• Creep
• Movement
• Lock-off Load

The creep amount shall not exceed 1 mm at the maximum test load during the period of 1 to 10 minutes.
If this value is exceeded, then the total creep movement within the period of 6 to 60 minutes shall not
exceed 2 mm.

The creep behavior of epoxy filled strand itself is significant and the measured anchor creep movements
must be adjusted to reflect the behavior of the material. At a maximum test load of 80 percent of the
specified minimum tensile strength of the tendon, creep movements of epoxy filled strand are
conservatively estimated to be 0.015 percent of the apparent free stressing length during the 6 to 60 minute
log cycle, but may be higher than this value (PTI, 1997). For a maximum test load of 75 percent of the
specified minimum tensile strength of the tendon, this percentage can be reduced to 0.012 percent. These
correction factors are based on limited laboratory tests but appear to be reasonable based on field
observations. For epoxy coated bars, these considerations do not apply.

7 -76
Movement

a) Residual Movement

Residual movement is defmed as the non-elastic (non-recoverable) movement of an anchor obtained during
load testing. The residual movement «\) is the difference between the total movement (oJ measured
during the load testing and the elastic (recoverable) movement (oe), as illustrated in Figure 7-44.

There is no absolute criterion for the amount of residual movement which is acceptable. Measurement of
this residual movement is,. however, essential to determine the elastic movement. From that, the apparent
free length of the anchor can be calculated for which the acceptance criteria are described below.

b) Minimum Apparent Free Length

The minimum apparent free length is calculated to verify that the anchor load is being transferred beyond
any potential failure or slip plane in accordance with the overall stability requirements of the anchor-
structure system. The minimum apparent free length at the maximum test load, as calculated on the basis
of elastic movement, should be not less than 80 percent of the designed free tendon length plus the jack
length. If this criterion is not met, the anchor should be reloaded up to two times more from the alignment
load to the maximum test load, and the calculation repeated on these cycles. If the criterion is not met,
then (i) the cause of this inefficiency in load transfer should be investigated, and (ii) the anchor may be
rejected or derated.

A limit higher than 80 percent of the designed free length should be set in cases where later movements
occurring as a result of redistribution of the free length friction would cause unacceptable structural
movement.

c) Maximum Apparent Free Length

The apparent free length as determined at the test load also provides additional information on load transfer
characteristics within and around the bond zone. Apparent free lengths longer than the maximum apparent
free length criterion may be caused by: a) tendon debonding, b) installing the bond length in variable
ground where the more competent ground surrounds the lower part of the bond length, or c) the anchor
approaching or having reached its ultimate load carrying capacity.

The maximum apparent free length at the maximum test load, as calculated on the basis of elastic
movement, should be less than 100 percent of the free length plus 50 percent bond length plus the jack
length. However, anchors with longer apparent free lengths should not be rejected if the cause of the
behavior has been investigated and satisfactorily explained.

d) Acceptability Based on Total Movements

The criteria for the minimum and maximum apparent free length, as described above, are not strictly
relevant if only total movement data are available. However, it is conventional to apply these criteria also
to total movement data when, from past experience or previous tests in the same conditions, the magnitude
of the residual movements is well known, and elastic movements can, therefore, be estimated. In such
cases, the criteria listed above should be applied. Otherwise, only the criterion for the minimum free
length should be used as a basis for acceptance, even though it is based on total movements.

7 -77
Load
o

DL

8,

Total
Movement

6
-I.--L---J..--L------------------------"'.. 10 min.

(a)

Load

DL

8,

Total
Movement

(b)

Figure 7-44: Typical Plots of Tendon Movement for a) Performance Test, and b) Proof Test (PTI,
1997)

7 -78
Lock-Off Load

After the design load of a production anchor has been verified by testing, the anchor load is immediately
transferred to the structure. The magnitude of this initial transfer load must be determined based on
structural design assumptions and the mechanical losses associated with the tendon type selected.

The mechanical losses associated with transferring load to various anchorage systems and long-term
relaxation or creep should also be considered in determining the fmal transfer load. Seating losses may
vary between 2 mm for bars to 6 mm for strand. Long-term load loss will vary with stress level but may
be estimated at 2 percent for bars and 4 percent for strand at typical design stress levels.

The magnitude of the transfer load is generally specified, and should not exceed 70 percent of the specified
minimum tensile strength of the tendon. In practice, transfer loads of 80 percent of the design load are
commonly specified. Higher transfer loads are sometimes used when it is required to minimize long-term
wall deflection. In selecting the transfer load, an evaluation should be made to verify that the transfer load,
particularly at the top anchor level, will not exceed the ultimate passive resistance of the soil behind the
wall since this could cause ground displacement and heave which might damage existing facilities.

Lift-off tests are performed either during construction to check the magnitude of seating and transfer losses
or after construction to determine if long-term load losses are taking place. The test is performed by
applying load gradually until the tendon begins to elongate. When a sudden deflection is observed on the
dial gauge, the jack extension should be immediately terminated, and the load required for lift-off recorded.
This load should be approximately equal to the design load plus an allowance for long-term losses. If the
lift-off load varies more than 5 percent from this value, the tendon load is adjusted and the lift-off test is
repeated. When the load in a strand tendon is more than 5 percent above the desired lock-off load, and
where no shims have been prepositioned under the wedge plate for later extraction, then it is preferable
to accept this load and so avoid the danger of having wedge marks below the wedge plate as a result of
strand/wedge regripping.

Failure Purine Testing

If an anchor does not reach the maximum test load as a consequence of bond failure, subsequent actions
depend on whether the anchor can be postgrouted or not. Regroutable anchors should be postgrouted and
then subjected to all the original acceptance criteria. Anchors without a postgrouting system should be
either rejected (and replaced) or locked off at not more than 50 percent of the maximum load attained. In
this event, no further acceptance criteria are applied.

If an anchor fails the creep test at the maximum test load, then the anchor should be postgrouted and
subjected to an enhanced creep criterion, assuming the other acceptance criteria are met. This enhanced
criterion requires a creep movement of not more than 1 mm between 1 and 60 minutes at the maximum
test load. Anchors, which cannot be postgrouted may be rejected or should be locked off at 50 percent of
the maximum test load. In this event, no further acceptance criteria are applied.

7 -79
7.8 WALL MOVEMENTS

The amount of wall movement is a primary consideration in the selection of the type of wall to be used for
any project. For earth retaining structures installed using top-down construction methods, numerous
factors influence the amounts of wall and ground movements that may occur. These factors, identified by
Clough, et al., 1989) are listed below:

Soil conditions & stratigraphy Construction surcharges Support spacing


Support stiffness Wall integrity Wall stiffness
Depth of excavation Weather Installation of wall
Installation of supports Preloading level Over excavation
Lagging installation Construction vibrations Construction inside excavation
Groundwater table location Dewatering Time excavation open
Excavation sequence Level first support placed Shape of excavation

Some of the above factors are not under the control of the designer, but are governed by the contractor's
selected construction procedure and on the quality of their workmanship. Nevertheless, some semi-
empirical guidelines have been developed by combining field experience with analytical tools, as discussed
below.

Figure 7-44 presents correlations for estimating vertical soil movements behind externally supported walls.
Several key points associated with this figure are as follows (Clough and O'Rourke, 1990):

• Only the basic excavation and support process has been considered. Other factors listed above
have to be evaluated separately and could result in more movements.

• Excavations in stiff to very hard clays show variable behaviors as they are influenced by the in situ
horizontal stress, degree of fissuring, degree of weathering, and plasticity. Heave may also be
possible for some conditions. For these materials, the dimensionless diagram in Figure 7-45(b)
should be used as a conservative estimate, provided that the wall is stable and not affected by poor
construction practice. In making judgments about stiff to very hard clays, it often is valuable to
refer to local construction experience.

• The family of curves shown in Figure 7-45(c) is based on numerical studies and assumes average
conditions, good workmanship, and that cantilever defonnation of the wall represents a small
fraction of the total movement. If this is not true, then cantilever movement should be added to
those determined from Figure 7-45(c); for example, this may occur when the first level of support
is not located close to the top of the wall. The cantilever stage movements can be idealized
assuming a point of fixity at an appropriate depth below the ground surface (see Module 8 - Deep
Foundations).

• Use Figure 7-45(c) with caution, especially where the factor of safety against basal heave is below
1.5. In these conditions, construction variables can cause significant increases in movements.

For excavation support systems which have complex soil conditions, or those which will potentially
influence a critical structure, the designer may choose to use the finite element, beam on elastic foundation
or other analytical methods to estimate movements. In any such analysis, care has to be taken to insure
that all inherent assumptions are understood. Calibration of the program. predictions using local project
performance is advisable.

7 - 80
-E 200 d

• Soldier Pile a Lag
or Sheetpiles

-,£e 160
E o Diaphram Walls
6 Drilled Pier Walls

t-.:'
zw
~
w 120 o·
fJ°'fl
.-.-
-J \~'\ ..
w ~~~
80
Cf)

-J
• 5%
0 /\,\,sO'\
Cf) ~~,,«'

X
40
<:{
:E •
0
0
• 40

(a)

-~ 200 • Soldier Pile a Lag

-e
co
J:
or Sheetplles
ODlaphram Walls

~160 05011 Nail Walls


zw 6 Drilled Pier Walls
~ 05011 Cement Walls
w
~ 120 •
:E
..J
..J
~
r-:
«
-J
. •
X
«
:e

~·o

10 20 30 40
DEPTH OF EXCAVATION H, (m)
(b)
Figure 7-45: Evaluation of Externally Supported Wall Movements. (After Clough and O'Rourke, 1990)
Page 1of2

7 - 81
0.0 "'---".'i=""T--::=~IF:::::::-:-:-::-::--"-----'
--][

0.3
~
0.51-----'
6i=
z ~
I-
< 10
.
wo
~x
ww
~~
Cl)j!:
fu
o 2.01o---.-J

0.0 .75 1.0 2.0 3.0 4.0


DISTANCE FROM EXCAVAnON
DEPTH OF EXCAVAnON
Curve I = Sand
Curve II = Stiff to very hard clay
Curve III = Soft to medium clay, factor of Safety against
basal heave 5.1 Sg Equal to 2.0
(= )
yH + q
Curve IV = Soft to medium clay, factor of Safety against
basal heave 5.1 Sg Equal to 1.2
(= )
yH + q
Where: Su = Undrained shear strength of the soil
q = Applied surcharge load
(c)
~
~I
~
SheetDlIe Wolls I 1 m Thick Slurry Wolls

-
4

~
.z= -I

~=t
h=3.5m 1
a.
~ 2.5
\ \
\
-
c:
.2
o
~ 2.0
I\. \ \1\ I -, I
#'~
-.::::::. 1.5 1\ \ 1\
q.~ "09
qC> /
I I I
Factor of Safety Against
Basal Heave

~
~ "
i' r.,~.......... .
~ ~
t-... .~ ....... /,0 .... r--... l"- '--
~ 1.0 ........ '" ~# ....-,.,--
r--
E I'----.... l f;:o~I"""'" I-..

-
-
1'-/.<1 ___
-E
..J
X
0.5
~
""'-
, t-- 2.0-
3.0
- I--

~
- 10 30 50 70 100 300 500 700 1000 3000
(EI}/(Ywh~)~lncreosingSystem Stiffness
(d)

Figure 7-45: Evaluation of Externally Supported Wall Movements. (After Clough and O'Rourke, 1990)
Page 2 of2

7 - 82
7.9 EXAMPLE PROBLEMS

7.9.1 Example Problem 7-1

Statement

A cantilever sheet pile wall was selected by the designer to retain an excavation in granular soil with Y =
17 kN/m 3 and Ysat = 20 kN/m3 as shown in Figure 7-46. Total depth of excavation is 6 m below the
ground surface. Assume no construction surcharge loading. Ground water table is at a depth of 6 m
below the existing ground level. Determine the depth of penetration of the wall and the maximum moment
in the wall structure.
/ Existing Ground Surface
':
Not to Scale
T I
I
. . ,'.:. :'
. .' '.': ..
.... ; .. :.": . -."
....... :.': '.' Y= 17 kN/m 3 .: •......
: ... : ': .' '" - 34 ° . ..... '.
' .
. .,' .

6m . . . .' .. '.. 't'- . . . , . ......


'. . .
. . ,"

3 · ..... :.-, .. : .
,' ,
". . . .
Yw= 10 kN/m
.. , .. ., .'

· .. . '. .: '.'"=?:. . '. . :. . . : .... .: ....


· ',- . :'.:. : : . :." . . ....":.-. "::": .::.:-. :.7,.. . :.... :.....
,"
:.
' .. :.":'· :.' ...
. . . . : ..... . .' . ·...... :.':'. . Ysat =20 kN/m~ .
.. "; ..:
. . .
", .. · . ., ': ., '" 34 ° . .... ' ..
· . . . : .'. : 'I' = ... , . : '''.:. :
.' .... '.
"
· . . .: ..:: .... . '

· ." ... . : ... : :: .


' ", . . .. :.': " :: ':', :.. ". ' .
. . , .. · .' ....
'
. .. ' ..

Figure 7-46: Geometry of Example Problem 7-1

Solution

Step 1: Calculate lateral earth pressure coefficients.

• Active pressure coefficient, Ka = tan2 (45 - 34°/2)


• Passive pressure coefficient

For <I> = 34°, P = 0°, and e


= 0°, from Figure 2-5 in Chapter 2, where
<I> = Angle of internal friction of sand
o = Friction angle between sheetpiling and sand
P = Backslope of backfIll
e = Angle of wall with vertical
.~ = 9.5
Take wall friction into consideration. For <I> = 34° and 0/<1> = -0.5, reduction factor
R = 0.688 (Refer to Figure 2-5).

~ =(9.5) (0.688) = 6.536

7 - 83
Use FS = 1.5 on ~

~ = 6.536/1.5 = 4.36

Therefore, ~ -KJ = 4.36 - 0.28 = 4.08

Step 2: Compute earth pressures as shown in Figure 7-47.

Br------:~~~~-----

a
\0
II
::r:

d
~----Jt-Location of zero shear

Not to Scale

Figure 7-47: Pressure Diagram

The pressure, p, at locations AI' A z, Band J shown in Figure 7-46 is computed as follows:

PAl =~yH = (0.28) (17 kN/m3) (6 m) = 28.6 kPa

PAZ = PAl + ~y/D = 28.6 kPa + (0.28) (10 kN/m3) D = 2.8D + 28.6 kPa

PE = ~-KJY/D-PAI = (4.08) (10 kN/m3) D - 28.6 kPa = 40.8D - 28.6 kPa

PI = ~yH + ~-KJY'D = (4.36) (17 kN/m3) (6 m) + (4.08) = 40.8D + 444.7 kPa


(10 kN/m3) D

Step 3: Determine depth of penetration of wall for a safety factor of 1.5 on ~

The depth of penetration of the pile should be sufficient to satisfy statical equilibrium. In other words the
following conditions of static equilibrium should be satisfied.

Sum of horizontal forces, ~H = 0


Sum of moments, ~M = 0

(a) The summation of horizontal forces can be expressed in terms of various areas in Figure 7-47 as
follows:

Area (BAzF) + Area (EeJ) - Area (EA IAz) = 0

7 - 84
(P
E - PAl) D - PAlH
z =- ------
(P E + PI)

Now: (PE - PAl) D = (40.8D - 28.6 - 28.6)D = 40.8D2 - 57.2D

= (28.6) (6) = 171.6

= 40.8D - 28.6 + 40.8D + 444.7 = 81.6D + 416.1

Therefore:

40.80 2 -57.2D + 171.6


z = --------- (1)
81.6 + 416.1

(b) Sum the moments (~M) about point F as follows:

1 ( H)
- pHD + -
2 al 3
+P
02
- + (p +p) -
Al 2 E I 6
Z2
- (p +P ) -
E Al 6
02
=0

1 (H)
-(28.6)(6) D +- D 2
Z2
+(28.6)-+(81.6+416.1)- -(40.8D-28.6+28.6)- D =0
2

2 3 2 6 6

85.60 + 171.6 + 14.302 + (13.6D + 69.4)r - 6.803 = 0 (2)

7 - 85
(c) Solve equations (1) and (2) simultaneously to satisfy static equilibrium and detennine the depth of
penetration. Use the following procedure:

1. Assume a depth of penetration, D

2. Calculate Z

3. Substitute D and z into ~M about F to verify IlM about F = 0

4. Repeat above steps using different values of D until IlM at F = 0

D,m z,m IlM, kN.m Remarks

6.5 1.61 -126.4 Try D> H, say 6.5 m

6.0 1.43 41.0

6.13 1.48 '" 0.0

Therefore, an embedment of D = 6.13 m will provide a factor of safety of approximately 1.5 on~. This
depth can be increased to get a higher factor of safety if desired. For convenience, round-up the required
depth to 6.15 m.

Step 4: Compute maximum moment in sheeting.

Maximum moment occurs at the point of zero shear. Assume that the point of zero shear lies between
points 0 1 and 0, say at a depth of d below point A (see Figure 747). The shear equation written in terms
of d is as follows:

(112) (28.6) (6) + 28.6d + (112) (0.28 - 4.36) (lOd2) = 0

d2 - 1.4Od - 4.21 = 0

Therefore, d = 2.87 m

Thus, maximum bending moment occurs at d=2.87 m. The moment equation at this depth is written as
follows:

M max =-p
1
2 Al H (d H)
+-
3
+P d
Al
(d)
-
2
1 Y d -d)3 - -K
-K
+
2 a
1 Y d -d)3
'2 (
2 p
'2 (

7 - 86
~ = (112) (28.6) (6) (2.87 + 6/3) + 28.6 (2.872/2) + (112) (0.28) (10) (2.872) (2.87/3) - (112)
(4.36) (10) (2.872) (2.87/3)

~ = 417.85 + 117.79 + 11.03 - 171.78


= 374.88 kN'm say 375kN.m

SUMMARY

For the configuration of the sheet pile retaining wall shown in Figure 7-47, the required depth of
penetration is 6.15 m. For this depth of penetration, the maximum moment induced in the sheeting is 375
kN.m. The next step is to select a sheet pile wall section with sufficient section modulus to resist the
maximum bending moment.

7.9.2 Example Problem 7-2

Statement

A cantilever sheetpiling wall as shown in Figure 7-48 was selected by the designer. The sheet pile will
be driven in a silty clay soil with Ysat = 20 kN/m3 and an undrained shear strength of 50 kPa. The clay
behind the wall will be replaced with granular backfill to the excavation level, which is at 4.5 m below
ground level. Neglect effects of construction surcharge. Determine the depth of penetration of the wall
and the maximum bending moment in the wall structure.

----------~---
-,--
",:. :
.'
;
".': ..
. :."
': '.

e ·.··· .. :··:·.·Y=17kN/m 3 : .
It')
: : .. cIl- 34° . '.
''It
II . .,' . -. ..' .,
"
. : .,'.:. :
::r::
"
......
' . ',' .
3
9 Yw = 10 kN/m

y sat =20 kN/m 3


Cu =50 kPa

Figure 7-48: Geometry of Example Problem 7-2

7 - 87
Solution

Step 1: Calculate lateral earth pressure coefficients.

For active earth pressures behind the wall use Rankine theory and determine the active earth pressure
coefficient using Equation 2-5 as follows:

K, = tan' ( 45' - ~)
For 4> = 34°, the value of Ka is calculated as follows:

Step 2: Compute earth pressures as shown in Figure 7-49.

The procedure presented in Figure 7-18 is used to develop the earth pressures. The net pressure diagram
is shown in Figure 7-49. The pressures PCb' Pcd and Pgf are as follows:

a
y=y'= 17 kN/m 3
4>= 34
0

'H = (17)(4.5)= 76.5 kPa

Ysat= 20 kN 1m 3
c= 40 kPa

+---e

. L . - . . L -_ _ ... -

h
t!::======::::::::::_ f
g

Figure 7-49: Earth Pressure Diagram

The pressures PCb' Pcd and Pgf are as follows:

Pcb = Tl"y'H
..~ = (0.283) (17 kN/m3) (4.5 m) = 21.65 kPa
= 4 (50 kN/m2) - (17 kN/m3) (4.5m) = 123.50 kPa
Pfg = 4cu + y'H = 4 (50 kN/m2) + (17 kN/m3) (4.5 m) = 276.50 kPa

7 - 88
Step 3: Determine depth of penetration of the wall

The depth of penetration of the pile should be sufficient to satisfy statical equilibrium. In other words the
following conditions of static equilibrium should be satisfied.

Sum of horizontal forces, ~H = 0


Sum of moments, ~M = 0

(a) The summation of horizontal forces can be expressed in terms of various areas in Figure 7-49 as
follows:

Area (a b c) + Area (e f h) - Area (c d h g) = 0

or (112) (Pcb) (H) + (1/2) (z) (Prg + Ped) - (Ped) (d' + z) = 0

(112) (21.65) (4.5) + (112) (z) (123.5 + 276.5) - 123.5 (d' + z) = 0

or d' = 0.62z + 0.39 (1)

(b) Sum the moments (~M) about point g as follows:

1
-(p
2 cb
)(H) [ -H +d '+z
3
] +-(p
1 . )(z) ( -
+p
2 fg cd 3
z) -(p cd
, ( d' +z
)(d +z) -
2
-) =0

or .!(21.65)(4.5)[4.5 +d' +Z]+.!(400)(Z 2) 123.5 (d,2+2zd' +Z 2) =0


2 3 6 2

73.07 + 48.71 d' + 48.71z + 4.92 Z. - 61.75 d12 - 123.5zd' = 0 (2)

(c) Solve equations (1) and (2) simultaneously to satisfy static equilibrium and determine the depth of
penetration.

Substitute Equation 1, i.e., d' =(0.62z + 0.39) in Equation 2:


73.07 + 48.71(0.62z + 0.39) + 48.71z + 4.92 Z. - 61.75(0.38z2 + 0.48z + 0.15) -
123.5z(0.62z + 0.39) = 0

Simplifying, we get

8.28 + 1.lOz-95.12z2 =0

Solution of this quadratic equation gives:

z = 0.30 m, d' = 0.58 m and D' = 0.88 m

Increase D' by 40% to get an approximate safety factor of 2.0,

D = 1.4 (0.88) = 1.23 m Choose D=1.25 m

7 - 89
Step 4: Compute maximum moment in sheeting.

Maximum moment occurs at the point of zero shear. Assume point of zero shear within depth dl (see
Figure 7-49). The shear equation written in terms of dl is as follows:

(1/2)(Pcb)(H) - Pcd (d') =0


(1/2) (21.65) (4.5) = 123.5d'
dl = 0.39 m
Thus, maximum bending moment occurs at d=2.87 m. The moment equation at this depth is written as
follows:

M max = (.!.)
2 (p cb)(H)( d' +.!!.)
3 - (p (d I)(~)
cd 2

M mu = ( ~) (21.65)(4.5>( 0.39+ 4~5) -(123.5)(0.39)( 0:9 )


= 92.07 - 9.39

= 82.68 kN'm Say 1\1- = 8S kN.m


SUMMARY

For the configuration of the sheet pile retaining wall shown in Figure 7-48, the required depth of
penetration is 1.25 m. For this depth of penetration, the maximum moment induced in the sheeting is 85
kN.m. The next step is to select a sheet pile wall section with sufficient section modulus to resist the
maximum bending moment.

7.9.3 Example Problem 7-3

Statement

To construct a depressed roadway, cuts up to 10 m deep are needed. A soldier pile timber lagging wall
with two levels of ground anchors has been proposed. The data obtained from subsurface investigation
indicates that the the soils at the site are composed of medium-dense to dense, non-plastic micaceous sandy
silt with <I> = 30° and y = 18 kN/m3• A pullout test conducted on an adjacent site with similar
subsurface conditions provided an estimated ultimate transfer load for anchor design of 115 kN/m.

Design the complete soldier pile lagging wall system using the following assumptions:

• A uniform surcharge of 10 kPa


• Steel soldier piles of width, b = 0.33 m at center to center spacing, s = 2.5 m.
• Two levels of ground anchors at a vertical center to center spacing = 5 m.
• A safety factor of 1.5 for computing passive resistance.

7 - 90
Solution

Step 1: Compute the lateral earth pressure coefficients

• Active earth pressure coefficient, Ka = tan2 [45° - 30°/2)] = 1/3


• Passive earth pressure coefficient, ~ = tan2 [45° + (30°/2)] = 3,
Use a safety factor of 1.5 for~. Thus, use KIp = 3/1.5 = 2

Step 2: Set construction sequence

The wall will be constructed in the following sequence:

(a) First install soldier piles. Excavate to a depth of 3 m which corresponds to a depth of 0.5 m below
the first level of ground anchors. During excavation place lagging.
(b) Install first level of ground anchors at a depth of 2.5 m.
(c) Excavate to a depth of 8 m which corresponds to a depth of 0.5 m below the second level of
ground anchors. Place lagging as excavation proceeds.
(d) Install second level of ground anchors at a depth of 7.5 m.
(e) Excavate to the final grade which is at a depth of 10 m below the ground surface. Place lagging
as excavation proceeds.

The wall should be analyzed at various stages of the above construction sequence. The following stages
are analyzed here:

Stage 1: Cantilever stage which occurs before the installation of the first level of ground anchors.
Stage 2: Stage where one level of ground anchors is installed and the excavation is at a depth of 8 m.
Stage 3: Stage where both levels of ground anchors have been installed and excavation is 10 m deep.

Step 3: Analyze the cantilever condition (Stage 1)

The geometry for the cantilever stage is shown in Figure 7-50.

e.,----
II
.d
First anchor not installed

STAGE 1
Final Grade Not to Scale

Figure 7-50: Geometry for Cantilever Condition (Stage 1)

7 - 91
e

Ps = 6 + 2d l08+36d
Pp = 36d---~r-t:~....a.;,,;;.,,-.--:-------.....
Psq = 1.11
@
Not to Scale
l08+36(d+z)
N
l08+36D

Figure 7-51: Pressure Diagram for Stage 1 (cantilever condition).

The earth pressures for this cantilever condition are shown in Figure 7-51. The earth pressures at various
locations in Figure 7-51 are computed as shown below:

Earth pressure on lagged depth above excavation

= K,.yhs = (1/3) (18 kN/m3) (3) (2.5 m) = 45 kN/m

Below lagged depth (below bottom of excavation), pressure acts only on soldier width (b).

Ps = K,.ydb = (1/3)(18 kN/m3)(dm + 3m)(0.33 m) =.2d + 6 kN/m


Passive pressure on soldier acts on width of 3 b or c/c distance between soldiers, whichever is less.

= (3) (0.33) = 1 m < c/c spacing = 2.5 m, use 3b.


= (3b) ~'dy = (lm) (2) (18 kN/m3) d = 36d kN/m

Lateral pressure due to surcharge q = 10 kPa in lagged depth (above bottom of excavation).

Plq = Kaqs = (113) (10 kN/m3)(2.5) = 8.33 kN/m

Below lagging (below bottom of excavation) up to pivot point (neglect effect of surcharge on passive side).

Psq = K,.qb = (113) (10 kN/m3)(0.33 m) = 1.11 kN/m

These pressures are shown in Figure 7-51. For static equilibrium the summation of horizontal forces (~H)
and the summation of moments (~M) about a given point (say f, the bottom of wall) should be equal to
zero. To facilitate computations, the pressure diagram is split in to simple triangular and rectangular areas
and are labelled areas 1 through 10. The horizontal forces (P), their moment arm about point f and the
moments in the various areas of the pressure diagram are tabulated below.

7 - 92
Moment Ann
~ FOrce. kN about t m Moment. kN-m
1 PI = (Ih) (45) (3) = 67.5 (3/3) + d + z = 67.5 ( 1 + d + z)
MI
2 P2 = (6) d = 6d (d/2) +z M2
= 6d (0.5d + z)
3 P3 = (Ih) (2d) (d) = d2 (d/3) + z = d2 (0.33d + z)
M3
4 P4 = (36d + 108)z = (36d + 108)z z/2 = 0.5z2 (36d + 108)
M4
5 Ps = (Ih) (36z) z = 18z2 z/3 M s = 18z2 (z/3)
6 P6 = (-Ih) (36d) d = -18d2 (d/3) + z M6 = 18d2 (0.33d + z)
7 P7 = 2d (z) = -2dr z/2 M7 = 2dz (z/2)
8 Ps = (Ih) (2z) z = _Z2 (z/3) r
M s = (z/3)
9 P9 = (8.33) (3) = 25 (3/2) + d + z M9 = 25(1.5 + d + z)
10 PIO = (1.11) d = 1.l1d (d/2) + z MIO = 1.11d(0.5d + z)

Substituting values from above table, we get

67.5 + 6d + d2 + (36d + 108)z + 18r - 18d2 - 2dz - r + 25 + 1.l1d = 0

Simplifying, we obtain a quadratic equation as follows:

(92.5 + 7.11d - 17d2) + (34d + 108)z + 17r = 0

The above quadratic equation can be written in tenns of z as follows:

z = [-34d -108 ± V934d -108)2-68(92.5 +7 .11D -17D 2)]_1 (1)


. 34

Substituting values from above table, we get

~M=O = 67.5(1 + d + z) + 6d(0.5d + z) + d2(0.33d + z) + z2(18d + 54) + 6z3+


25(1.5 +d + z) + 1.11d(0.5d + z) - 18d2(0.33d + z) - dr - (1I3)z3

o= 105 + 925(d + z) + 7.11d(0.5d + z) - 17d2(0.33d + z) + (17d + 54)r + 5.67z3 (2)

Substituting Eq. (1) in Eq. (2) and solving for d by trial and error we obtain:

d = 4.47m

Substituting d=4.47 m in Eq. (1) gives z = 0.79 m

Substituting d=4.47 m and z = 0.79 m in Eq. (2) which gives ~M = 4.44 '" 0 which satisfies condition
of static equilibrium

Total depth of penetration, D, needed for the cantilever case = d + z

7 - 93
D = 4.47 + 0.79 = 5.26 m

Substituting d = 4.47 m and D=5.26 m Eq. (1) which gives :I:H = -0.56 '" 0 which satisfies condition
of static equilibrium

Therefore required minimum length of soldier piles is:


3 + 5.26 = 8.26 m i.e. Elevation = 20 - 8.26 = 11.74 m

Check maximum moment in soldier pile.

Maximum moment occurs at the point of zero shear. Assume point of zero shear is distance x below
excavation level. The shear equation is the summation of all horizontal forces from the top of the soldier
pile to distance x below the excavation level. Accordingly, the shear equation written in terms of x is as
follows:

or

67.5 + 6x + x2 - 18x2 + 25 + 1.llx = 0

Simplifying, we obtain a quadratic equation as follows:

92.5 + 7.11x - 17x2 = 0

Solving the quadratic equation yields x = 2.55 m


Thus, the maximum moment occurs at a depth of 2.55 m below excavation level. The moment equation
at this depth is written as follows:

Mmax = (112)(45)(3)(3/3 + x) + 6(x)(x/2) + 0.5(2x)(x)(x/3) - 0.5(36x)(x)(x/3) + 8.33 (3) (3/2 +


x) + 1.ll(x)(x/2)

= 67.5(1 + 2.55) + 6(2.55) (0.5) (2.55) + (2.552) (0.33) (2.55) - 18(2.552) (0.33) (2.55) +
25(1.5 + 2.55) + 1.11(2.55) (0.5) (2.55)
= 239.625 + 19.507 + 5.527 - 99.488 + 101.25 + 3.609
= 270.03 kN'm

Step 4: Analyze the one-level. anchor condition (Stage 2)

Figure 7-52 shows the case of one anchor installed and excavation down to 0.5 m below second anchor
level. Figure 7-53 shows the earth pressure distribution for this condition.

7 - 94
e q = 10 kPa
II'l
N
E1.20 m

e
f
II'!
II'l
Second anchor not installed
::\: STAGE 2
Final d Not to Scale

Figure 7-52: Geometry for Condition of One Anchor (Stage 2)

8.33
e
II'l Free Earth Method
e N
00 - ' - - - - - - + --~It-
II
::r:

r=- ~120 8.33


1.11
I
I
~I
I
I
6d Not to Scale
- - -.......~1-!616+2d 1.11

Figure 7-53: Pressure Diagram for Stage 2

The earth pressures at various locations in Figure 7-53 are shown below:

Earth pressure on lagged depth,

Pd = I5h kN/m
Below lagged depth,

Ps = 2 (d + 8) kN/m
Passive pressure on soldier piles,

Pp = 36dkN/m
Lateral surcharge pressure on lagged depth,

7 - 95
Plq = 8.33 kN/m

Below lagging up to pivot point,

psq = 1.11 kN/m

These pressures are shown in Figure 7-53. For static equilibrium the summation of horizontal forces (~H)
and the summation of moments (~M) about the anchor head) should be equal to zero. The ~M condition
will help determine the depth of penetration, d, while the ~H condition will help determine the anchor
force, T. To facilitate computations, the pressure diagram is split in to simple triangular and rectangular
areas and are labelled areas 1 through 6. The horizontal forces (P), their moment arm about the anchor
head level and the moments in the various areas of the pressure diagram are tabulated below.

Moment arm about


~ Force kN anchor head. m Moment. kN.m
1 PI = (.5)(120)(8) = 480 (2/3)(8) - 2.5 = 2.83 M I =(480)(2.83) =1358.4
2 P2 = (16) d = 16d 5.5 + (d/2) M2 =16d (5.5 + 0.5d)
3 P3 = (lh)(2d) d = d2 5.5 + (2/3)d M3 =d2 (5.5 + 0.67d)
4 P4 = (lh)(36d) d = -18d2 5.5 + (2/3)d M4 =-18d2(5.5 +0.67d) = 100
5 Ps = (8.33)(8) = 66.67 (8/2) - 2.5 = 1.5 Ms =(66.67) (1.5)
6 P6 = (1.11) d = 1.11d 5.5 + (d/2) M6 =1.11d(5.5 + 0.5d)

Substituting values from above table, we get

1358.4 + 16d(5.5 + 0.5d) + d2(5.5 + 0.67d) + 100 + 1.11d(5.5 + 0.5d) - 18d2(5.5 + .67d) = 0

Simplifying, we obtain a third order equation as follows:

Solving by trial and error, we fmd d = 3.77 m Le. Elevation = 20 - 8 - 3.77 = 8.23 m

Substituting values from above table, we get

480 + (16) (3.77) + 3.772 - (18) (3.772) + 66.67 + (1.11) (3.77) - T = 0

Solving for T, we obtain

T = 369.5 kN

Check maximum moment in soldier pile.

Maximum moment occurs at the point of zero shear. Assume point of zero shear is distance h below ground
surface but above the excavation level (see Figure 7-53). The shear equation is the summation of all
horizontal forces from the top of the soldier pile to distance h. Accordingly, the shear equation written in

7 - 96
terms of h is as follows:

or

(1/2) (15) h2 + 8.33h - 369.5 = 0

Solving the above quadratic equation yields h = 6.49 m.

Thus, the maximum bending moment occurs at a depth of 6.49 m below the top of the soldier pile. The
moments within this depth can determined as follows:

Force, kN Arm,m Moment, kN-m

PI = (0,5)(15)(6.486)2 = 315.51 kN 6.486/3 = 2.162 (315,51) (2,162) = 682,1

Ps = (8,33) (6.486) = 54,03 kN 6.486/2 = 3,243 (54.03) (3,243) = 175.2

T = 369.55 kN =-369,55 kN 6.486- 2.5 = 3.986 (-369.55)(3,986)= -1473

Therefore, Mmax = 682.1 + 175.2 - 1473 = -615.7 kN'm


Step 5: Analyze the Final Condition with Two Anchors (Stage 3)

Figure 7-54 shows the case of the excavation at its final grade with the second anchor installed. Figure 7-55
shows the pressure distribution for this case.

q
e
If')
h
~'-t-----::..l ...~ ~

STAGE 3

Not to Scale

Figure 7-54: Geometry for Final Condition with Two Anchors (Stage 3)

7 - 97
1-1I!_ P o --I -rq~
96.5 8.3

--r-------.J96.5 8.3
--~------ 1.09

Not to Scale

36d "--_ _I--_...a 1.98d Plq = 1.09


19.8
Ill! Pp _Ill! PI -I

Figure 7-55: Earth Pressure Distribution for Stage 3.

The pressure distribution shown in Figure 7-55 is based on the following computations.

For prestressed anchors assume an empirical rectangular distribution for earth pressure over the height of
the excavation. This is computed as follows:

For <I> = 30°, Ka = (1 - sin 30°)/(1 + sin 30°) = 0.33

PO = 0.65KayHs where s is the center to center spacing of the soldier piles

= (0.65) (0.33) (18 kN/m3) (10 m) (2.5 m) = 96.5 kN/m

Surcharge pressure on lagged depth,

Pq =Kaqs = (0.33) (10 kN/m 2


) (2.5 m) = 8.3 kN/m
Surcharge pressure below lagged depth,

P1q = Kaqb = (0.33) (10 kN/m2) (0.33 m) = 1.09 kN/m


Active pressure below lagged depth due to overburden within d,

Ps = Kaydb = (0.33) (18 kN/m3) (10 m) (0.33 m) = 19.8 kN/m


Active pressure at depth d below fInal grade,

Ps = Kaydb+ 19.8 = (0.33) (18 kN/m3) d (0.33 m) + 19.8 kN/m = 1.98d + 19.8

7 - 98
Passive pressure below lagged depth,

Pp = ~yd3b = (2) (18 kN/m3) d (3) (0.33 m) = 35.64<1 say 36d

Compute Maximum Bending Moment in Soldier Piles

The maximum bending moment in the soldier piles can be estimated using the assumption of a continuously
simply supported beam between T 1 and T2 (Figure 7-56).

~104.8kN,m

Sm
side view Not to Scale

Figure 7-56: Pressure Diagram for Continuously Supported Span for Stage 3

Thus,

M max =

where: W = Po + Pq = 96.4 + 8.3


= 104.8 kN/m
Substituting values:

(104.8) (5)2
= -'-----'-....;....;-
10

M max = 262.0 kN·m

Compute the Anchor Loads

Using the tributary area method, tieback T 1 will have to resist a larger horizontal lateral pressure on the upper
5 meters of the lagged depth. So use the force on T1 for the design of both tiebacks for conservatism:

Total load on lagged depth, P, is given by:

TH =T 1 = (Po + Pq) (2.5 + 512)

= (96.5 + 8.3) (5)


= 524 kN (horizontal)

7 - 99
Compute the De.pth of Embedment. d. of the Soldier Pile

Calculate required embedment to balance forces as shown in Figure 7-57. Assume uniform lateral pressure
over a depth midway between the lowest anchor level and the bottom of the excavation will be resisted by
the embedded soldier pile, in addition to the active pressure below the excavation (shaded area).

The forces in various shaded areas shown in Figure 7-57 are presented below:

Shaded
Area No. Force. kN
1. (-96.5 kN/m) (1.25m) = -120.63
2. (-19.8 kN/m) d = -19.8d
3. (-lh) (1.98 kN/m2) d2 = -0.99d2
4. (lh) (36 kN/m2) (d) d = 18d2
5. (-8.3 kN/m) (1.25 rn) = -10.38
6. -1.09 kN/m d = -1.09d

1-..._ P o -~I -1 r-
q

96.5

Not to Scale

36d 19.8 PIq = 1.09

I... Pp ~ I... Ps ~I

Figure 7-57: Pressure Diagram for Determination of Pile Penetration

7 -100
For static equilibrium the summation of horizontal forces (~H) should be equal to zero. Thus,

~H = -131.01 - 20.89d + 17.01d2 = 0

Solving the quadratic equation, we get

d = 3.46 m, Use 1.2d $ 4.2 m

Therefore, Elevation = 20 - 10 - 4.2 = 5.8 m


The embedment depth required to resist the vertical loads on the soldier pile (due to vertical load from
anchors) should be evaluated using pile design methods from Module 8 (Deep Foundations).

Summary of Construction Stages

The following table summarized the various forces and bending moments from the above analysis for various
construction stages. The wall should be designed so that it has the maximum depth of penetration and can
resist the maximum bending moment.

Stage Condition Horizontal Force Maximum Bending Soldier Pile Tip


Moment, M max

T)kN T2 kN kN·m Elevation, m

1 Cantilever 270 +11.74

2 One anchor installed 369.6 615.7 +8.23

3 Both anchors installed 524 524 262.0 +5.80

Step 6: Lagging Design

Use pressure diagram as shown in Figure 7-58 (with a reduction factor of 0.5 m for lagging design).
10 kPa 3.3 kPa
El.20

Retainin Surcharge
Rectangular Empirical
Wall Pressure
Pressure
=0.65K.yH = K.q
= 0.65(1/3)(18)(10) = (1/3) (10)
=39 kPa = 3.3 kPa

E1.10

Figure 7-58: Pressure Diagram for Lagging Design

7 - 101
Using 0.33 m flange width, the span length in plan for lagging is:

~ = 2.5 - 0.33 = 2.17 m

For uniform pressure distribution condition shown in Figure 7-58 the maximum bending moment in the
lagging can be calculated as follows:

M max =

where: w = (0.65KayH + Ka q) per unit depth near the base of the excavation

= [(0.65) (113) (18 kN/m3) (10 m) + (113) (10 kN/m2)]


= 42.3 kN/m2

R = Reduction factor for lagging design = 0.5

Substituting values:

= (0.5){42.3){2.17)2 =
M max 12.5 kN·m/m
8

From Table 7-1, try Douglas Fir-Larch with an allowable flexural stress of 8300 kPa.

S = 12.5 = 1.51 x 10-3 m 31m


rcqd 8300

Try timbers of width b=300 mm and depth, d=76 IllIl1;

S = bd 2 = (0.3 m)(0.076m)2( 1.00 boards/m) = (0.076m)2


6 6 0.3 6

< Sreqd Not Acceptable

Try timbers of width b =300 mm and depth, d = 100 mm;

(0.lm)2
S
6

= 1.70 X 10-3 m3/m > Sreqd Acceptable

Use 100 mm thick lagging boards.

7 - 102
Step 7: SOU Anchor Design

El. 140.0
e ~o
Not to Scale
\I')

~ El. 137.5 Failure Plane

e
-
0

II
:x: ~
e
\I')
~o
El. 132.5
e
\I')

~ El. 130

Figure 7-59: Layout of Soil Anchors

The final configuration of the problem is shown in Figure 7-59. The design procedure for the soil anchors
shown in Figure 7-37 is presented below.

Each soil anchor has to resist a horizontal load of 524 kN. Assuining the soil anchor will be inclined at an
angle e with horizontal of 30 0 • the actual soil anchor load is:

T = = 524kN = 605 kN
cos30°

To determine required bond length use a safety factor of 2.5

Therefore. Tul1 = 2.5 (605 kN) = 1513 kN


For medium dense to dense sandy silt. estimated ultimate transfer load is 115 kN/m as determined from soil
investigations.

Required bond length. L. is given by:

T ult 1513 kN
L = = = 13.2 m
115 115 kN/m

7 -103
The free length of anchors should be at least:

1. Rankine plane from bottom of excavation + H/5 or


2. Rankine plane from bottom of excavation + 1.5 m or
3. Minimum 4.5 m measured along the anchors

In our case, H = 10 m
H/5 = 10 m 15 = 2 m > 1.5 m

Therefore, use Rankine Plane +2m


Rankine Plane is inclined at an angle of 45° - <1>/2 with the vertical face of the waH
i.e., 45° - 30°/2 = 30°

Free length envelope should be as shown in Figure 7-37a. The free length, x, can be computed graphically.
Alternatively, the free length shown in Figure 7-37a can be mathematically represented as follows:

y sin(45 0 - <1» H
2 5
x = ------ +
sin(45°+ <I> +8) sin(45°+ <I> +8)
2 2

Therefore, for T):

7.5 sin ( 45 0 - 300) +2.0


2
x = = 5.75 m say 6 m
0 300 )
sin ( 45 + 2 +30

x = = 3.25 m say 3.5 m

Figure 7-59 shows the above computed layout of the soil anchors

Check depth of bond zone for T) below ~round surface

Depth of beginning of bond zone = 2.5 m + 6 m (sin 30°) = 5.5 m > 4.5 m O.K.

7 -104
Tendon Design for Soil Anchors

Assuming 15.2 mm diameter strands, the Guaranteed Ultimate Tensile Strength (GUTS) is 1862 MPa and
ultimate load is 260.8 leN (from Table 7-4)

As recommended use about 50% of GUTS for design load or (0.50) (260.8) = say 130 leN per strand
Therefore, number of strands = 605/130= 4.7 strands Use 5 strands.

SUMMARY

A complete analysis and design of a soldier pile lagging wall system with two levels of anchors has been
presented. The analysis is based on the various construction stages. In addition to the analysis presented
herein, a global stability analysis must also be performed. The fmal depth of the soldier piles may be
governed by the global stability. Global stability methods are discussed in Module 3 (Soil Slopes and
Embankments).

7 -105
CHAPTER 8.0
IN SITU REINFORCED WALLS

8.1 INTRODUCTION

In situ reinforced walls are used in top-down construction applications to support temporary and permanent
excavations. Construction of these walls involves inclusion of reinforcing elements in the in situ soils to
create a composite earth structure. This chapter will present the design and construction aspects of two
types of in situ walls: (1) soil nail walls (Figure 8-1), and (2) micro-pile walls (Figure 8-2).

The main components of an in situ reinforced wall are the in situ material, the reinforcing inclusions and
the wall facing. The reinforcing inclusions usually consist of metal bars, small diameter cast-in-place
reinforced concrete piles, and grouted in place, small diameter steel pipe piles. Shotcrete, welded-wire
mesh, cast-in-place concrete, or precast concrete panels are typically used for the facing. In situ
reinforced walls can be used in a wide range of ground conditions.

In situ reinforced earth walls have been successfully used for a variety of applications including:
• Temporary and permanent walls for excavations in urban areas.
• Cut-slope retention for roadway widening and construction of depressed roadways.
• Support of existing bridge abutments for roadway widening projects.
• Support of cuts at tunnel portals.
• Landslide protection and stabilization of natural slopes.
• Repair or reconstruction of existing retaining walls.

8.2. SOIL NAIL WALLS

8.2.1 General

This section discusses the design procedures and construction requirements for soil nail walls. The
material in this section was obtained primarily from the FHWA Manual for Design & Construction
Monitoring of Soil Nail Walls (1996). Additional material was obtained from the FHWA publication SQ.il
Nailing Field Inspectors Manual (1994). Both of these publications were used as reference material for
FHWA Demonstration Project 103.

Soil nail walls use horizontal to subhorizontal reinforcements to improve the shearing resistance of the soil.
The reinforcements, known as nails, are closely spaced and are not prestressed. The shear stresses in the
ground are transferred as tensile forces in the nails through friction (or adhesion) mobilized at the ground-
nail interface. The nails develop tension as the ground deforms in response to continued excavation. The
nails may also develop some bending and shear forces. The effects of bending and shear are generally
considered secondary and are not included in the recommended design method presented later in this
chapter. Nails can be installed by (a) driving, (b) drilling and grouting, (c) jet grouting, and (d) firing.
In the U.S., the "drilling and grouting" technique is the common practice. Figure 8-3 illustrate the typical
installation sequencing for a soil nail wall. Wall construction is further discussed in Section 8.2.7.

Figure 8-1 illustrates four common applications for soil nail walls, including a) temporary shoring for
construction of CIP walls or other permanent facilities, b) a permanent wall for roadway widening at a

8- 1
80lL EXCAVATIOfl7
AND IACKFLL

r:- i, .
CONVENmNAL ~' -\ ...

AETAMIG WALL ,.,', \:..

l~
. T······
I'
~
----,11> '-
(a) TEMPORARY SHOTCAET! I'ACIIQ - '
\ W!!P HOLIS lTYP.l

PEI'lMANENT SOL -
-tr SOL I&RM TO ~
HAL lTYP.l \ '-r-""'!O'
[ , BE REMOVED

-=::;;;:;!~'~~1nf1 :", PEIlMANENT EXlSTN3 PER


.. ' ,
... , SOL HAL WALL
...... ,
. ...
.......
,,
. :....,:>,
' ..,. I ,....EXlSTING
ROADWAY

~~~~~
EXISTING
(b) AIUTMENT PLEa (c)

GUTTER

TEMPORARY SPIOTCAETE FACNi

GEOCOMPOSITE DRAIN STAP.


IF SPECIFIED

CAST-IN-PLACE REIlFORCED
CONCRETE F;NJSH FACE

/
/
/
EXISTING ROADWAY
/
/
/
/
/

' ':;AllOITlONAL ROADWAY


J
FOOTING DRAIN ....J
(d)

Figure 8-1: Soil Nail Wall Application. (a) Temporary Shoring; (b) Roadway Widening Under
Existing Bridge; (c) Slope Stabilization; and (d) Roadway Cut. (Porterfield, et ai., 1994)

8-2
RETAINING STRUCTURE

(a)

Vertical Cross-section
c::.
r
I

.
:~:~ : .~ .~ ~: ::
.- ..... - . .........
<:Sand/Gravel Formation
With Boulders &
'Old Masonries
.~
Bottom of . J

Excavation °Oy .
-(
.. <
(b)

(e)

Figure 8-2: Examples of Micro-Pile Walls. (a) Reticulated Wall for Slope Stabilization; (b)
Reticulated Wall for Excavation Support; and (c) Type "A" wall for slope stabilization.
[(a) & (b) (Lizzi, 1982); (c) (pearlman, et al., 1992)]

8-3
Step 1. Excavate Small Cut Step 2. Drill Hole for Nail

Step 3. Install and Grout Nail Step 4. Place Drainage Strips,


Initial Shotcrete Layer
Bearing Plates/Nuts

Step 5. Repeat Process to Step 6. Place Final Facing


Final Grade (on Permanent Walls)

Figure 8-3: Typical Nail Wall Construction Sequence

8-4
bridge abutment, c) a pennanent retaining wall for a roadway cut, and d) slope stabilization. Examples
of each of these applications are shown in Photo 8-1.

The primary advantages of soil nail walls are as follows (Juran and Elias, 1991; Byrne, et ai. ,1996):
• Can be used for both pennanent and temporary support.
• Permits top-down construction (less excavation and backfIll, and narrower work area).
• Rapid construction.
• Requires only small construction equipment.
• Eliminates need for internal bracing.
• Eliminates need for piles at wall face.
• Greater tolerance to deformations.
• Redundancy of reinforcement (more support elements than anchored walls).
• Adaptability to different soil conditions.
• Easily accommodates changes during construction.

Following are some disadvantages of soil nail walls:


• May not be suitable for certain ground conditions.
• Somewhat greater horizontal displacements when compared with prestressed ground anchors.
• Corrosion of the metal inclusions in aggressive environments.
• Requires underground easements behind the wall.
• Potential interference with nearby utilities, particularly iiI urban areas.

Soil nail walls are not considered applicable or cost effective in the following conditions (Byrne, et al. ,
1996; Elias and Juran, 1991):

• Loose granular soils (SPT N-values less than 10; Relative Density less than 30 percent).
• Uniformly graded granular soils.
• Soils with excessive moisture or seepage.
• Organic soils or clay soils with Liquidity Index greater than 0.2, and shear strength less than 50
kN/m2 •
• Highly frost susceptible and expansive soils.
• Highly fractured rock with open joints or voids, or open graded coarse granular material
(excessive grout loss).
• Rock with weak or adversely oriented discontinuities.
• Soils with high corrosion potential, including cinder, ash or slag fills, industrial wastes, or acid
mine wastes.
• When very limited wall defonnations are allowed.

8-5
(b)

(a)

(c) (d)

Photo 8-1: Soil Nail Wall Applications: (a) Temporary Shoring; (b) Roadway Widening Under
Existing Bridge, (c) Roadway Cut; and (d) Slope Stabilization

8-6
Although soil nail walls appear to be similar to anchored walls and mechanically stabilized earth walls,
there are considerable differences between them. Comparisons of these wall systems are presented below.

Comparison with Anchored Walls (Bruce and Jewell. 1986: Byrne. et ai.. 1996)

• Ground anchors are prestressed after installation and thereby limit any structural movement. In
contrast, soil nails are not prestressed and require a fInite soil deformation to mobilize their action.

• Nails are in contact with the ground over essentially their entire length, whereas ground anchors
transfer load only along the bond length. As a result, the distribution of stresses in the retained
mass is different for each type (see Figure 8-4).
2
• Since the nails are installed at a far higher density (typically one nail per 1 to 4 m ), with the
design load of each nail lower than that of a ground anchor, the consequences of failure of one unit
are lower. However, there is greater reliability in the capacity of individual ground anchors since
all anchors are tested after installation.

• Ground anchors are generally longer (typically 15 to 30 m, versus 3 to 10 m for soil nails) and,
therefore, may necessitate larger installation equipment and wider Right-of-Way. Also, an anchor
system is normally used in combination with the construction of soldier piles, tangent/secant piles,
or a slurry wall system, which itself necessitate large construction equipment.

• Soil nails result in lower loads applied to the structural facing, and therefore a thinner facing
element is typically required.

• Anchored walls typically must extend some depth below the [mal excavation level to develop the
necessary resistance to the vertical component of the anchor force; soil nail walls typically
terminate at, or a shallow depth below the [mal excavation level.

Comparison with Mechanically Stabilized Earth (MSE) Walls (Schlosser. 1982)

• Although at the end of construction the two structures may look similar, the construction sequence
is considerably different. A soil nail wall is constructed by staged excavation from "top down"
while an MSE wall is constructed "bottom up," (see Figure 8-5). This has an important influence
on the overall deformation patterns and distribution of the forces in the reinforcement.

• Mechanically stabilized earth (MSE) walls use a select granular backfill compacted to a specified
density, while soil nail walls must be designed for existing in situ ground conditions.

• In soil nail walls, grouting techniques are usually employed to bond the reinforcement to the
surrounding ground, and load is transferred along the grout to soil interface. In MSE walls,
friction is generated directly along the steel or geosynthetic reinforcement surface.

• MSE walls typically use precast concrete face elements, whereas soil nail walls use a shotcrete
facing, with or without a cast-in-place or precast outer facing.

8-7
(a) (b)

Figure 8-4: Load-Transfer Mechanism in: (a) Ground Anchors; and (b) Soil Nails. (Juran and
Elias, 1991)

Existing Ground

/
\ t----
\t----
Proposed _I
Typical
Deformation
=-11---
.,1-_ _
Excavation
_._._.-1
1 2 3 4

,._
(a)

..
Typical :--+\ ...... _
I Deformation .
.~ Proposed .
I'~--

~~~~or
.
I
Exca~ation Bottom
Fac~ P:;,[ I . Ii
< Rem orcemenlj
, ~
I
\I~--
$.'Z:'.....,,-"',. 11...-.$ ;c.;; "''' _" ",,,~<-"'W:'r.'~vO'"?,- //"'v//,.,,..
Placed Fill
1 2 (b) 3 4

·gure 8-5: Comparison of the Construction Sequencing for: (a) a Soil Nail Wall; and (b) a
Mechanically Stabilized Earth Wall. (Bruce and Jewell, 1986)

8-8
8.2.2 Design Issues

The fundamental considerations for the design of a soil nail wall include (a) resistance to external forces,
and (b) resistance to internal forces. Other important design considerations include corrosion protection
measures, wall deformations and drainage requirements.

The potential failure modes which must be considered include: (a) internal failure of either the reinforcing
tendons or the facing, or both (Figure 8-6(a»; (b) external shear surfaces that do not specifically intersect
the reinforcements themselves (Figure 8-6(b»; and (c) so-called mixed modes that involve shear surfaces
that extend through the reinforced zone and beyond the physical limits of the reinforced ground (Figure

(a)

~ ..."", , ,
I
I
I
/ ~ External Failure Surface
/
I
I
/
/

_\ _..... ..... ./
/

--_.---
(b)

\\-

(c)

Figure 8-6: Potential Failure Modes to be Analyzed for Soil Nail Walls. (a) Internal Failure;
(b) External Slope Failure; (c) Mixed Failure.

8-9
8-6(c». Both internal and mixed failure modes involve consideration of the following:
• Yield or rupture of the nail.
• Pullout of the nail.
• Failure of the wall facing or the connection between the facing and the nail.

Current Soil Nail Wall Design Methods

The various methods of analysis which are currently used for soil nail wall design, are summarized in
Table 8-1. Most of the methods shown in Table 8-1 are based on limit equilibrium design concepts and
examine the stability of free body blocks defmed by failure slip surfaces that are of circular, log-spiral,
parabolic or bi-linear shapes. They make no assumptions on how each of the installed nails contributes
to the overall required stabilizing force and, except for the GOLDNAIL program, do not consider the
influence or effect of the facing.

As in traditional slope stability analyses, limit equilibrium conditions are used to search for the most critical
failure surface, which is the failure surface with the lowest factor of safety. Most approaches consider only
the tensile capacity of the nails as an addition to the shear resistance of the soil. A few methods consider,
in addition, the effects of shear capacity and bending stiffness of the nails on the overall structure stiffness.

All design methods compute the required total nail reinforcing force to achieve a specified factor of safety,
either based on global stability or based on local stability. An initial nail configuration is first assumed and
analyzed, and is modified as necessary to ensure that the required nail force can be provided without the
nail either breaking or exceeding the bond strength between the nail and the soil. The nail/soil bond
strength is estimated for design and is validated during construction by field pull-out tests.

Recently, an improvement on the limit equilibrium methods taking into account the structural role of the
facing has been developed under FHWA Demonstration Project 103 (Byrne, et ai., 1996). The design
concepts and a step-by-step design procedure for this method are presented below.

Basic Design Concepts

The limit equilibrium approach to soil nail wall design is summarized in Figure 8-7. In this figure, the
global factor of safety is defined as the ratio of the resisting to driving forces along a potential planar slip
surface.

The equilibrium of an unreinforced block of ground and the corresponding expression for the global factor
ofsafety, F, are presented in Figure 8-7(a). The expression for the global factor of safety is a conventional
factor of safety for an unreinforced slope.

Next, a single nail reinforcing element is introduced to examine the manner in which the reinforcement
».
improves the factor of safety or the stability ofthe sliding block of ground (Figure 8-7(b As indcated, the
effect ofthe reinforcement is to improve stability by both a) increasing the normal forceand hence the shear
resistance along the slip surface in frictional soils, and b) reducingthe driving force along the slip surface
in both frictional and cohesive soils.

Of particular importance is the shape of the nail strength diagram, shown in Figure 8-7(b) and further
presented in Figure 8-8 for clarity. For any partiCUlar sliding wedge, the reinforcing contribution of the

8 - 10
TABLE 8-1
SOIL NAIL DESIGN METHODS
(Elias, et aI., 1996)

Method I Analysis Type I Failure Surface I Failure Mode I Output I-Curre;{Jsag; I Computer Code

1. German Method Limit Force Bi-linear Pull-out Global F.S.; Germany Bauer Program;
(1979) Equilibrium, Critical failure proprietary.
Global Stability surface

2. Davis Method Limit Force Parabolic Pull-out or nail Global F.S.; U.S. In public domain.
(1981) Equilibrium. yield stress Critical failure Modified codes by
Modified Davis Global Stability surface various
(1988 -1990) contractors;
proprietary.

00 3. French Method Limit Moment Circular, or of Pull-out or nail Global F.S.; France TALREN;
I
...... TALREN (1983) Equilibrium, any shape yield stress Critical failure proprietary.
...... Global Stability surface

4. Kinematical Working Stress Log Spiral Pull-out or nail Nail forces; U.S. In public domain.
(1988) Analysis, Local yield stress Critical internal
Stability failure surface

5. Caltrans-SNAIL Limit Force Bi-linear Pull-out, nail Global F.S.; California SNAIL; in public
(1991) Equilibrium, yield stress, Critical failure (U.S.) domain.
Global Stability punching shear surface; Average
nail stress

6. Golder Assoc. Limit Force Circular Pull-out, nail Global F.S.; U.S. GOLDNAIL;
GOLDNAIL Equilibrium, yield stress Critical failure proprietary.
(1991) Global Stability surface;
Considers
influence of
facing
W ,tan-1( tan ~y )
. F

Resisting Force S
F = Driving Force
= Wsine

= cul+ N tan ~y For slip surface shown


Wsin9
= Cyl+W cosetan~y
Wsine Expressions for Factor of
Safety as given in (b).

(a) (c)

TERMINOLOGY:

w =service load (F)


c =ultimate soil cohesion (FIL2)
W 4> =ultimate soil friction angle (0)
F =global factor of safety applied
/ tan- 1( ta~~y) to soil shear strength
/ =nominal nail head strength IF)
=cxFTFN
=allowable nail head load IF)
=nominal nail tendon strength (F)
Resisting Force = S =cx NTNN
F = =allowable nail tendon load IF)
Driving Force fY'I sin e - T cos P)
cyl+ N tan ~y
=ultimate pullout resistance (F/L)
= fY'I sin e - T cos p) =cxaOu
=allowable pullout resistance (F/L)
= cyl + fY'I cos e + T sin p> ta~u T =allowable nail load IF)
(see Figure 8-7)
(W sin e - T cos P) 5 = resisting shear force (F)
N =normal force (F)
(b)

Figure 8-7: Soil Nail Design - Basic Concepts and Tenninology. (a) Unreinforced Slope;
(b) Reinforced Slope - Single Nail; and (c) Reinforced Slope - Multiple Nails.
(Byrne, et al., 1996)

8 - 12
ZOne A ZoneB ZoneC

x
..4 - - - - y - - -.......

....4 ----------- Nail Length --------------~~.


Nail Head
Nail Support to Slip Surfaces Intersecting the Nail in Zone A at Point X=TF+Qx
Nail Support to Slip Surfaces Intersecting the Nail in Zone B =TN
Nail Support to Slip Surfaces Intersecting the Nail in Zone C at Point Y=Qy
TF =strength of nail head-facing connector(allowable)
TN =nail tendon tensile strength(allowable load)
Q =nail-ground pullout resistance(aJlowable)

Figure 8-8: Nail Support Diagram. (Byrne, et al., 1996)

nail is a function of the location at which the associated slip surface intersects the nail. The contribution
of any nail to the stability of a particular sliding block will be the least of a) the tensile strength of the nail,
b) the pullout resistance of the length of nail beyond the slip surface, or c) the strength of the facing/nail
connection plus the pullout resistance of the length of the nail between the slip surface and the face of the
wall.

Finally, multiple nails are considered (Figure 8-7(c» as a simple extension of the single nail problem, to
demonstrate the basic design methodology for a soil nail wall. Although the methodology is demonstrated
for only a single planar slip surface, all potential slip surfaces must be examined to identify the minimum
factor of safety. Slip surfaces of other shapes (e.g., circles, log spirals, bilinear wedges, etc.) are
preferred in examining limit equilibrium states since a) they generally provide lower calculated factors of
safety, and b) the planar slip surface can be closely approximated by these more general shapes.

Wall Design Considerations

Following are some factors to be considered for wall/nail layout and dimensioning.

Wall Dimensions

• Walls on tangent or circular radius are preferred over spiral wall lines for easier constructibility,
particularly for field survey control of wall line and nail locations.
• Stepped walls can be used for high walls, or where desired for aesthetic reasons.

Nail Inclinations

• The holes must be inclined downward. An angle of 15 degree is commonly used.


• From a constructibility standpoint, nail inclinations should be as uniform as possible.

8 - 13
• Steepened nail inclination may be required to avoid underground utilities.
• The bottom row of nails is sometimes inclined at greater than 15 degrees because of drill rig access
restrictions .
• In applications where there are headroom limitations, such as beneath a bridge deck, the upper
nails may be installed with a flatter inclination.
• Nail inclination angles should not be less than 5 degrees.

Nail Spacinis

• Typical vertical and horizontal soil nail spacings are within the range of 1 to 2 meters, with a 1.5
meter nail spacing being most common.
• For constructibility reasons, nail spacings should be as uniform as possible.
• Vertical nail spacings may be controlled by the availability of standard wire mesh sizes used in
constructing the shotcrete facing. Typical roll widths in the U. S. are 1.5 meters, except for the
west coast where typical roll widths are 2.1 meters.

Nail Layout Locations

• Nail head columns can be vertical or offset row by row. The offset pattern improves the
excavation face stability during construction through enhanced arching effect.
• To limit the height of the upper cantilever section of the facing, the top row of nails should be
placed at less than about 1.0 meter below top of the cut.
• At bridge abutments, the location of the top row of nails should allow sufficient headroom for
drilling and working beneath the deck.
• The length of the bottom cantilever section of the wall should generally be no greater than about
two-thirds of the average vertical nail spacing.
• Nail heads should be positioned so as not to coincide with the shotcrete construction facing
horizontal joints or the permanent cast-in-place facing vertical expansion joints.

Nail Lengths

• Nail lengths tend to increase with lower soil strengths, lower nail-ground pullout resistances,
steeper face and backslope angles, higher surcharge loadings, and deeper cuts.
• For common applications with modest backslopes and minimal surcharge loadings, the nail lengths
are usually in the range of 0.6 to 1.0 times the height of the wall.
• Common design and construction practice is to use a uniform length of nail and a uniform size
steel tendon over the full height of the wall.
• Because of the top-down method of construction, the nails located in the upper two-thirds of the
wall are most effective in controlling wall displacements, and are more heavily loaded. It may be
possible at the detailed design stage to shorten the nails located in the lower part of the wall.
• For simple cut slope configurations and uniform ground conditions, preliminary nail lengths and
strengths can be obtained from simplified design charts presented in Section 8.2.4.

8 - 14
TABLE 8-2
LOAD COMBINATIONS IN AASHTO SPECIFICATIONS
(AASHTO, 1~ Ed., 1992)

Group D ~RST[§Q] %
I 1 1 1 1 0 0 100

IV 1 1 1 1 1 0 125

VII 1 0 1 1 0 1 133

Notes:
D = Dead load. B = Buoyancy.
L = Live load. RST = Rib shortening, shrinkage, temperature.
E = Earth pressure. EQ = Earthquake.
8.2.3 Recommended Design Procedure

The recommended soil nail wall design procedure presented below is based on a recent research study
funded by FHWA (Byrne, et al., 1996). Two different design approaches were developed in the FHWA
study: service load design, and load and resistance factor design (LRFD). Only the service load design
procedure is presented below.

Step 1 - Set Up Critical Design Cross-Section(s) and Select a Trial Design

Select a trial design for the design geometry and loading conditions. The initial design geometry should
include a proposed trial design nail pattern, including nail lengths, tendon sizes and trial vertical and
horizontal nail spacings. Information presented above for wall/nail layout and dimensioning can be used
as a guide in developing initial design geometry.

The loads and load combinations shown in Table 8-2 illustrate the AASHTO requirements considered most
relevant to soil nail wall design. Specific designs may require consideration of additional loads and load
combinations.

Step 2 - Compute the Allowable Nail Head Load

The three critical failure mechanisms that must be considered for the connection of the soil nail to the wall
facing are: (1) flexure; (2) punching shear; and (3) tension in headed-stud connection system (for
permanent wall facing). The allowable nail head load should be determined using the following procedure:

a) Determine the nominal nail head strength for each potential failure mode of the facing and
connection system using the methods presented below (Byrne, et ai., 1996):

Flexural Stiffness of Facing:

The flexural stiffness of the facing increases with increasing thickness and reinforcement ratio and
decreases with increasing nail spacing. As the facing flexural stiffness decreases with respect to
the soil subgrade reaction modulus, the pressure distribution behind the facing will become highly
non-uniform, with large pressure concentrations occurring behind the nail heads (Figure 8-9).

8 - 15
r Reduction in Pressure
¥ Between Nail Heads

Facing ------... / Pressure Build-Up at .


¥ Nail Head Location
1----_......::::.....

Bearing Plate

Grout Column

Figure 8-9: Typical Facing Pressure Distribution. (Byrne, et aI., 1996)

It has been found that for the common facing configurations employed in soil nail wall
construction, the critical nominal nail head strength, TFN' associated with the flexural capacity of
the facing may be presented by the following relationship:

(8-1)

where IIlv'NEG and IIlv'PQs are the vertical nominal unit moment resistances at the nail head and mid-
span locations, respectively, and SH and Sv are the horizontal and vertical nail spacings. The
pressure factor for facing flexure, C F, is determined from Table 8-3. "CF " may be considered to
account for the non-uniformity of the contact pressure between the facing and the subgrade, with
a value of 1.0 corresponding to a uniform contact pressure, and increasing values being reflective
of increasing concentrations of contact pressure in the vicinity of the nail head.

• Equation 8.1 is applicable to typical soil nail wall facing construction practice. For facing systems that involve
either larger horizontal nail spacings than vertical nail spacings, or when the horizontal unit moment capacities are
less than those in the vertical direction, then Equation 8.1 should also be checked with unit moment capacities
corresponding to the horizontal direction and with the vertical spacing substituted for the horizontal spacing and
vice versa.

8 - 16
TABLE 8-3
FACING PRESSURE FACTORS
RECOMMENDED VALUES FOR DESIGN
(Byrne, et al., 1996)

Temporary Facings Permanent Facings

Nominal Facing Flexural Pressure Shear Pressure Flexural Pressure Shear Pressure
Thickness (mm) Factor C", Factor C" Factor C", Factor C"

100 2.0 2.5 1.0 1.0

150 1.5 2.0 1.0 1.0

200 1.0 1.0 1.0 1.0

Historically, full plastic moment development in slabs at the ultimate state has been considered
only when the reinforcing steel is distributed according to where the moments develop under an
elastic condition. For soil nail wall facings, one method of achieving that distribution is through
the placement of horizontal waler bars and vertical bearing bars. Because soil nail wall
construction practice in many parts of the U.S. includes only waler bars in temporary facings and
neither waler nor bearing bars in permanent facings, the CF factor has been reduced to 1.0 for
permanent wall facings. It is required that continuous horizontal waler bars be incorporated at
each row of nails in a temporary shotcrete facing as the walers passing beneath the nail head
bearing plate provide an element of ductility in the event of a punching shear type of failure.
Vertical bearing bars at each nail head are optional, and may be used as an alternative to
increasing the mesh size if additional vertical moment resistance is required from the facing.

In addition, for reinforcement percentages greater than about 0.35 percent, the available analytical
data at this time indicate that Equation 8.1 may over-predict the available nail head strength. As
the amount of flexural reinforcement is increased, the actual value for CF tends to decrease because
the facing is actually stiff or with respect to the soil. With the unit moment capacities increasing
and CF decreasing, the net result may be little change in the nail head strength. It is therefore
recommended that until further information becomes available, a reinforcement ratio (based on
gross area) of no more than 0.35 percent should be considered when calculating the nail head
strength using Equation 8.1 arid using a CF value greater than 1.0 (Le., for temporary facings with
a thickness of less than 200 mm).

Punching Shear Strength at Facing:

Primarily two types of connection systems are currently used in soil nail wall construction (Figure
8-10). For temporary shotcrete construction facings (Figure 8-IO(a)), the bearing-plate connection
is most commonly used, wherein a steel plate bearing against the front or excavation side of the
facing is connected to the nail tendon using a washer and nut. For permanent CIP concrete facings
cast over the temporary shotcrete construction facing, or for full-thickness permanent shotcrete
facings, a headed-stud connection is most commonly used. As shown on Figure 8-10(b), the
headed-stud connection consists of four headed studs welded to the bearing plate, with the entire
connection system embedded within the permanent wall facing. In either connection system,
failure develops by the punching of a truncated cone through the facing.

8 - 17
~ Ii; ~ ofInternal
W I bars

)
aer

Fa f1ure Surface ~
-.J1.
b PL Facing Component
Resistance. V N

-l 1tpL
I I
40", .
J', , I T
""
, ,"-....
... D'e
, he
"
" Dc L
"
iIJi ill -.......J
'- ...... V
I~
Vv
V
V
~I.--
Jill j j IJ
l< I' 101 1/
Critical Diameter for
Strength Calculation
Located at Mid-depth
"\. n
n
yo

K-
1/
1/

of Cone '\ r/ Soil Component

!
of Resistance
'<~~ k:::l.t:'--

TFN Note:
j--DGc--l D'e = De·he
=bpL+hc

(a)

I te al Fac'ng Component
Failure Surface ~ II / n rn I
I, Dc of Resistance, VN

~
.L /--SHS-l /
4~'l ... L
, J'" /
,
LHS J
'
,~D'e
I
rt-h
I
l', ,,
1
llli 1III '- .... V
WJlllJ
/ :.--I--'L-
.....
~ /'.
Critical Diameter for
Strength Calculation
Lo cated at Mid-depth
"\ t)r--.
l">o 1/
of Cone \ ~ Soil Component
of Resistance

Note:
Temporary faCing not
...... ~~
!
TFN
rt[.e~

Note:

shown for clarity. I---DGc--l Ac =7t 06/4 O'e = De-he


AGe = 7t 0 ~c/4
(b)

Figure 8-10: Punching Shear of Nail Head Connections. (a) Temporary Bearing Plate Connection;
and (b) Permanent Headed - Stud Connection. (Byrne, et al., 1996)

8 - 18
In addition to the strength afforded to the nail head by the internal resistance of the facing, there
exists a soil reaction component of resistance. The soil reaction component of the punching shear
strength develops because 1) the base diameter of the punching cone can be significantly larger
than the nail grout column diameter, and 2) as indicated on Figure 8-10, the soil pressures that
develop in the vicinity of the nail head can be quite large, depending on the flextural stiffness of
the facing.

Based on laboratory test results, the nominal internal punching shear strength of the facing, V N,
is computed by considering a nominal shear stress acting across an effective perimeter area. The
perimeter is defmed by an effective punching cone diameter, Of c, and an effective cone depth,
he, as shown on Figure 8-10. The complete relationship is presented in Equation 8-2:
(8-2)

The effective cone diameter is defined by a rupture surface that begins at the edge of the plate and
extends toward the soil side of the facing at an angle of 45 0 from the plane of the facing (Figure
8-10(a). Therefore, at the center of the facing, the effective cone diameter is defined to be
Ofc=bpL +hc·

As a conservative approximation, the punching shear strength, TFN' of a bearing plate connection
can be taken as the internal punching shear strength, VN' of the facing. If such a conservative
approach does not provide sufficient capacity, then the soil reaction contribution to nail head
resistance can be included as described below. This will usually result in increases in the
calculated punching shear strength of less than 20 percent for typical shotcrete construction
facings.

The soil reaction component of nail head resistance is computed by considering force equilibrium,
the failure cone diameter illustrated on Figure 8-10, and an increased pressure behind the nail
head. The resulting expression for the nominal nail head strength associated with punching shear
of the bearing plate connection with the soil reaction contribution included is shown in Equation
8-3:

(8-3)

where Ac = area of punching shear cone base at back of facing,


~c = cross-sectional area of soil nail borehole.

Punching Shear of Headed-Stud Connections:

The approach for computing the internal punching shear strength of headed-stud connections is
similar to that for temporary construction facings. The same nominal shear stress is applied over
a slightly different perimeter area. The cone diameter is defmed by extension of a line at 45 0 from
the centers of the stud heads (Figure 8-10). Furthennore, the effective cone depth he is defmed
from the top of the headed studs as shown on Figure 8-10. The resulting expression is the same
as Equation 8.2, where the effective punching cone diameter Ofc = SHS + he, where SHS is the stud
spacing as indicated on Figure 8-10.

8 - 19
As with a bearing-plate connection, the soil reaction component is computed by considering an
increased pressure beneath the punching cone, outside the perimeter of the grout column. The
resulting expression for the nominal nail head strength is the same as Equation 8-3. As for the
bearing plate connection, the punching shear strength of the headed-stud connection can be taken
as only the internal punching shear strength of the facing (Equation 8-2). Because of the relative
stiffness of typical pennanent facings with headed stud connections, inclusion of the soil-reaction
contribution to the nail head resistance will generally increase the calculated punching shear
strength by only a few percent.

For headed-stud connections in which the length of the headed-stud is short in relation to the stud
spacing (e.g., on the order of half the stud spacing), the pullout of individual studs may govern
the strength. In such cases, the capacity of the individual studs should be evaluated.

Headed-Studs Tensile Strength:

Because the headed studs resist the nail head force through direct tension, the strength criterion
corresponding to ultimate tensile stress for bolts is most applicable, as indicated by Equation 8-4:

(8-4)

The headed studs must also be checked for bearing on the concrete beneath the heads. To meet
th~ requirement, two geometric criteria must be checked: (1) the cross-sectional area of the stud
head must exceed 2-1/2 times the cross-sectional area of the stud body; and (2) the head thickness
must exceed Ih the difference between the head diameter and the body diameter.

b) For each possible nail head failure mode, detennine the allowable nail head load as a fraction (see
Table 8-4) of the corresponding nominal nail head strength. The fIrst column of nail head strength
factors of Table 8-4 applies to the Load Combination Group I (Table 8-2). For other Load
Combination Groups, the Group I nail head strength factors of Table 8-4 are increased in
accordance with the percentage factors of the last column of Table 8-2. For soil nail wall
applications, Load Combination Groups I, IV, vn are likely to control the design most of the time.
The corresponding nail head strength factors for Load Combination Groups I, IV and vn are,
therefore, also shown in Table 8-4. The allowable nail head load is the lowest calculated value
for the various failure modes.

TABLE 8-4
NAIL HEAD STRENGTH FACTORS
(Byrne, et al., 1996)

Nail Head Nail Head


Strength Factor Nail Head Strength Strength Factor
Failure Mode (Group I) Factor (Group IV) (Group VII)
a;F (Seismic)

Facing Flexure 0.67 1.25(0.67)= 0.83 1.33(0.67)= 0.89

Facing Punching Shear 0.67 1.25(0.67)= 0.83 1.33(0.67)= 0.89

Headed-Stud Tensile Fracture


ASTM A307 Bolt Material 0.50 1.25(0.50)= 0.63 1.33(0.50)= 0.67
ASTM A325 Bolt Material 0.59 1.25(0.59)= 0.74 1.33(0.59)= 0.78

8 - 20
Step 3 - Minimum Allowable Nail Head Service Load Check

Perfonn a minimum allowable nail head load check for the trial facing design. This empirical check is
perfonned to ensure that the computed allowable nail head load exceeds the estimated nail head service
load that may actually be developed as a result of soil-structure interaction. The nail head service load
actually developed can be estimated by using the following empirical equation:

(8-5)

where: = nail head service load factor


= active earth pressure coefficient
= soil unit weight
= vertical height of soil nail wall at wall face

The value of FF is not well defmed at this time and appears to vary from values as low as 0.30 to values
on the order of 0.70 (or higher), with a mean value in the range of 0.40 to 0.45. Unless the designer has
site specific monitoring information from walls constructed in similar soils, it is recommended that a value
of FF equal to 0.50 be adopted for design purposes.

For simple configurations (i.e., unifonn soil conditions, no surcharge, etc.), the active earth pressure
coefficient, KA , can be determined directly from published equations and charts (see Chapter 2). For
more complex configurations that are not tabulated in the published literature (e.g., variable soil layers,
complex wall geometries and surcharge load distributions), the nail head service load can be estimated
from Equation 8-6:

(8-6)

The total active load, PA> per unit length of wall can typically be detennined using a Coulomb-type slip
surface (i.e., slope stability) calculation. If the computed allowable nail head load is less than the
empirically estimated nail head service load, the trial facing/connector design should be modified and Step
2 repeated.

Step 4 - Defme the Allowable Nail Load Support Diagrams

Defme the allowable nail load for each of the nails as a function of location along the nail length. As
discussed earlier (see Figure 8-8), the allowable nail load will vary along the length of the nail and will
depend on the allowable nail head load, the allowable nail tendon load, and the allowable nail-ground
pullout resistance.

The allowable nail head load is detennined as discussed in Step 2.

The allowable nail tendon load is taken as the nail tendon strength factor times the nail tendon yield
strength, TNN , as shown in Table 8-5. The nail tendon yield strength TNN = ABFy where ABis the nominal
cross-sectional area of the bar and Fy is the yield strength of the steel). It is recommended that the
minimum bar size used for soil nails be No. 19 (Soft Metric designation) - corresponding to a No.6
standard bar size.

The first column of the nail strength factors of Table 8-5 apply to Load Combination Group I (Table 8-2).
For other Load Combinations Groups, the Group I nail strength factors of Table 8-5 are increased in

8 - 21
TABLE 8-5
STRENGTH FACTORS AND FACTOR OF SAFETY
(Byrne, et aT., 1996)

Strength Factor Strength Factor Strength Factor


Element (Group I) (Group IV) (Group VII)
a (Seismic)
Nail Head Strength aF, seeTable 8-4 See Table 8-4 See Table 8-4
Nail Tendon Tensile Strength aN = 0.55 1.25 (0.55) = 1.33 (0.55) = 0.73
0.69
Ground-Grout Pullout aQ = 0.50 1.25 (0.50) = 1.33 (0.50) = 0.67
Resistance 0.63
Soil F = 1.35 (1.50*) 1.08 (1.20*) 1.01 (l.13*)
Soil-Temporary Construction F = 1.20 (1.35") NA NA
Conditiont

Notes:
Allowable Nail Head Load (TF) = aF(Nominal Nail Head Strength) = aFT FN
Allowable Nail Tendon Load (TN) = aN (Tendon Yield Strength) = aNT NN
Allowable Pullout Resistance (Q) = aQ(Ultimate Pullout Resistance) = aQQu
Minimum Required Global Factor of Safety "F" (Group I) = 1.35 (= 1.50 for critical structures).
Minimum Required Global Factor of Safety "F" (Group IV) = 1.35/1.25 =1.08 (=1.20 for critical structures).
Minimum Required Global Factor of Safety "F" (Group VII) = 1.35/1.33 = 1.01 (= 1.13 for critical structures).
Minimum Required Global Factor of Safety "F" - Temporary Construction Condition = 1.20 (= 1.35 for critical
structures) .
Critical structures include walls constructed adjacent to high volume roadways or walls in front of bridge abutments.
It is imperative that field pullout testing be done during construction to verify the estimated pullout resistance used
in design (see Section 8.2.9).

* Soil Factors of Safety for Critical Structures.


t Refers to temporary condition existing following cut excavation but before nail installation. Does not refer
to "temporary" versus permanent wall.

accordance with the percentage factors of the [mal column of Table 8-2. The corresponding nail strength
factors for Load Combination Groups IV and VII are also shown in Table 8-5.

The allowable nail pullout resistance will determine the rate at which the allowable nail load can change
along the length of the nail and is taken as the nail pullout resistance factor (see Table 8-5) times the
ultimate soil-grout pullout resistance. Nail pullout resistance estimates should be based on prior experience
in similar types of soil or rock, and with similar methods of nail installation. If adequate experience does
not exist to provide a conservative design value, then a pre-contract test nail program should be considered
to determine the appropriate design values, particularly on large projects.

Guideline ultimate grout-ground bond stress values for different soil and rock materials are discussed below
and presented in Tables 8-6 through 8-8.

8 - 22
TABLE 8-6
ULTIMATE BOND STRESS - COHESIONLESS SOILS
(Byrne, et al., 1996)

Construction Unit Ultimate Bond


Soil Type
Method Stress kN/m2

Open Hole Non-plastic silt 20 - 30

Medium dense sand and silty 50 - 75


sand/sandy silt

Dense silty sand and gravel 80 - 100

Very dense silty sand and gravel 120 - 240

Loess 25 -75

TABLE 8-7
ULTIMATE BOND STRESS - COHESIVE SOILS
(Byrne, et al., 1996)

Construction Unit Ultimate Bond Stress


Soil Type
Method kN/m2

Open Hole Stiff Clay 40 - 60

Stiff Clayey Silt 40 - 100

Stiff Sandy Clay 100 - 200

TABLE 8-8
ULTIMATE BOND STRESS - ROCK
(Byrne, et al., 1996)

Construction Unit Ultimate Bond Stress


Rock Type
Method kN/m2

Rotary Drilled Marl/Limestone 300 - 400

Phillite 100 - 300

Chalk 500 - 600

Soft Dolomite 400 - 600

Fissured Dolomite 600 - 1000

Weathered Sandstone 200 - 300

Weathered Shale 100 - 150

Weathered Schist 100 - 175

Basalt 500 - 600

8 - 23
Cohesionless (Granular) Soils

For tremie or low pressure grouted nails in dry cohesionless soils, data reported in the literature suggest
ranges of ultimate pullout resistance as presented in Table 8-6.

Cohesive Soil

For tremie grouted nails in cohesive soils, the ultimate pullout resistance can be estimated as 0.25 to 0.75
times the average undrained shear strength, with the lower factors associated with the stiffer or harder
clays. For augered holes, a lower factor may be warranted because it is influenced by the care taken in
cleaning the drillhole. For sandy and silty clays, the factor is somewhat higher than the range cited above.
Typical values of ultimate pullout resistance for cohesive soils are presented in Table 8-7.

The ultimate pullout resistance for tremie grouted nails in competent massive rock may be taken as 10
percent of the uniaxial comprehensive strength of the rock up to a maximum value of 4000 kN/m2 •
Estimated pullout resistance for different rock types are presented in Table 8-8. Additional information
on the ultimate pullout resistance of anchors installed in rock is provided by Littlejohn and Bruce (1977).

Step 5 - Select Trial Nail Spacings and Lengths

The following empirical constraints on the design of nail length pattern are recommended for use when
performing the limit equilibrium design calculations:

(a) Nail with heads located in the upper half of the wall height should be of uniform length.
(b) Nails with heads located in the lower half of the wall height should be considered to have a
reduced length in accordance with the recommendations given in Figure 8-11.

The purpose of these recommendations is to ensure that adequate nail reinforcement (length and strength)
is installed in the upper part of the wall. Performance monitoring of several instrumented soil nail walls,
in which both nail loads and wall movements were measured, has demonstrated that the top-down method
of construction of soil nail walls generally results in the nails in the upper part of the wall being more
significant than the nails in the lower part of the wall in developing resisting loads and controlling the
displacements.

It should be noted that the above recommendation for the nail length pattern to be assumed for design
calculation purposes does not imply that the installed nail pattern must correspoild exactly to the design
nail pattern. It is most common to install nails of uniform length, often to simplify construction. Provided
the appropriate stability checks are performed (see Step 8), it may be possible to install shorter nails in the
lower part of the wall.

Step 6 - Derme the tntimate Soil Strength

Define the ultimate soil strength for analysis. Appropriate soil strengths for design are determined using
the procedures discussed in Chapter 3 of this Manual and in Course Module 1 - Subsurface Investigation.
Accurate soil strength characterization is an important part of the design process and should be performed
only by qualified experienced geotechnical personnel.

8 - 24
1.0

H
2'

Note:
lOr" values determined by
linear interpolation between
a value of 1.0 at wall midheight
and lOR" at base of wall.

2.0 T'""'--.......,.-..--.......r -.......~r-- ............-r--,......"'"T'........-....


... ,.-
-~..,-
• •
-,.,- ..-,.-
I' •
-r-,.~-'-. .,.... ..- "·"-T-r-
I •• I., I •
-~...,.~."
I " .,
~-
I I • t I I

-}--:-i-t- -r·:--:-~- -:--:-i-f- ·:--:·~-i- -:--:-~-+- -:-i-f-r-


.. _~ _ ..
_~_ .. _ .. _ ""' 4 _ .
~_

• ••
..... I ••
.. ,.
.a • ""
... _ ~
••••
~ _ .. _ ."4.<1._.
'" I
_• • • • .

-_·.t_~.:~_:_t~_·~~~t.-_i_:.i_: :~+~:f: :~~:i:J~: +i:~:f:


• • ,",:, t
*~:f:~:
·,··,-i-:- ..--:-"-:-
•• . •• _,--:-,.- -'--,_ •• -.-
... • _--..
-~-:-~-~- -~
_~- _ -
1.5 -4-~.;...;...;'...;.-+...;'...;.•....;I~.+..;'...;'...;•...;..-+
_ ...
~.~- -~-:.~-~.
_
-~-:.~.~-

'...;.
7~-~-~·

• ..;.~...;.•....;..~.;...;'~ .........;.'~ '. .


-:-~-~.~

-~..:-~-~- -~-~..:-\..- -~..:.~-~. -~..;....:-~- .:-..:.~-~-


, I I ,

-:--:-;-t- -:--'--:-~' ·'--:-i-:'· -:--:--:-7- -:--:-~-+- ... .........


"-~·T·"·
·:,oi-f·~· L =Maximum nail length

~~~~~~f~ ~~~~~j~I~~~:~~l~ ~gII~ f~~~~f~


~

.......... H = Wall height


..
•• I I

I
.,.~.,,-

I
-r· ..

I
·"-r·
I

I

I =
QD Dimensionless pullout resistance
-:·-:·~ .. t· -~.: .. -:.. ~- ...
.:- ..: '-f- .;--:.-:-~- -:.-:.~-!,. -:-~·t-t· = ClaQu!(ySHSV)
Co
_ ......... _£_
_.. _..... - ..
• • '.
..... -'_J_~_
.... ..-.. .. . ... -.. - .......

-~_.~
I • I
_ ""
~_ .... _"' __'• .J_~_ ..... J.~_ ... ~-~.~ ~.
"
__ r 1 ., 1 • .
1.0 ~.::~•..;.:..;.:+..;.:...:.:...;.:..;::..f...;.•..;.•..;.'...1t+-...;...;';...;.~'-+...;.: ~:;...;:...;:-+...;., ..;:~,...;.:~
-~ _ -
(UH) where:
:~~:~:~:
.. :~:::~:~: :~~:. ~:;: i·t~:~:~:
.~-:-~ ~- . . .. :::~:i:~:
..
-:--:-~-~-
~:i:i:~:.. -:--:.~-~- ~.-: ~ -:-~ ~-~- -:-~.~-~
=
Cla pUllout resistance strength
_ 4 __ 4_04 .. __ ~_~ oo~ _".4_ __ 4 •• _ .. factor (from Table 8-5)
It.
oo_.....-.- I
oo.....
••• I
__ . _......
••
--.,o.-,of'"
r I I
~ ...•oo.oo
• 1'1'
...... ·11-.- ,.
~ ..... _.. I •

=
Qu ultimate pullout resistance
oo,.• ....--,-.oo
I 1.
.......
••••
.roo.,--
I I
..--.,-
••
oo..• .,..I • ..,_I
...
.,..
.. _, •
-,-
•••
•• ,.
=
.,_~_ ~_

-;.-:.~-;- -r·:--{-i- -:--l-i·;- .:.. -:-., i· ·:-~·;-i- ,o:.-!,o;'-~. y unit weight


-:--:·~-i-
__ _ ·~·:-~-i-
oo__ _ -:-~oo~-~-_ _ -~-:--!- ..- - -:-~·~-t-
_ -6
~·~·f"~·.
SH = horizontal nail spacing
, • I' ••• ,.,. I I I' • I Ii' I I

0.5~•.;:.~~.;,'.-•.;.'.~.~,~_1-•.;,,:.~.;.:.-~;._~~~_;-_~:.~.~.••~..~••~.~,.+~.:.~.:.~~-.~;-.flt"~~;..~~~.~~~_+-:~~_~~:-_": ...._~:.-. Sv = vertical nail spacing


I
oo,.~-.,.-,-

.,._. -,- -t'" ... - , - , . • • .--roo..,.,.
•• • I •• ,. t, -..'"
I'
.,-,.- •.•••
I'
~-,,·,,"r·

.1._'.. J_l. -!--:-{.~.. •:-.:-i-f- -:...:.-:.~oo .:- I i-t· ·:·~·f·t·


.~.:.~_.i._ .,-.I.J.J • • 'ooJ • .I,o£ . . .'-.J.J.J • • '.J _&.. J .... £.L.
I
_ _ .. _
I I I
• • • " _ 4• • •
I I I _
I .' "
._ . . . . _
'I
...... 4 ... _
" _
I' ~.
I• "
__ 4
t
•••••
I

I •
__ .... _ ..__.... .
I I I.,. I' I. ".
_ _ ....... _ ......
I ., I I • t ••

."..,. .,.-,.. -,.-,..,-, .."" ... .,.,.- ..,.,- .,. .... ,., .,-, ..,-,.
~_~_._ .~ ~ ~"'·.·4 ~.

• I I' I' I I I... •• t I I. I • I I f


.,..~

·I··:""-~· .~-:--:-~. -:··:-4-+· .~.:.-:.,- ·:··:·i-;'· ·:".;·~·r·


-:-··.~-t· .~.t.~.". -:.-:-~.~. ,o:--:.~-~- ·:·~·~·t- ,J.-i.~-~.

o 0.1 0.2 0.4 0.5

Figure 8-11: Nail Length Distribution Assumed for Design. (Byrne, et al., 1996)

8 - 25
Step 7 - Calculate the Factor of Safety

Calculate the limit equilibrium factor of safety for each potential slip surface using the approach illustrated
in Figure 8-7, taking into account the additional stabilizing forces provided by the trial pattern of nails.
Table 8-5 provides recommended minimum global factors of safety for both critical and non-critical
structures.

It will also be necessary to check the stability of the soil nail wall for the temporary construction condition,
corresponding to the situation in which the next lift has been excavated, but prior to the installation of the
nails for that lift. In these circumstances, because of the temporary nature of such conditions, it is
recommended that the nail strength factors applied be the same as those shown in Table 8-5, but that the
required global factor of safety be reduced to a value of 1.2 (1.35 for critical structures), as shown in Table
8-5. In general, for most applications and typical construction conditions, construction stability
requirements will not control the design. However, in certain circumstances, such as significant existing
surcharge loadings adjacent to the wall during construction, construction conditions may be more critical.

Step 8 - External Stability Check

Perform stability analyses for potential "external" failure modes. The potential external failure modes that
require consideration include overall slope failure external to the nailed mass, and foundation bearing
capacity failure, overturning, and sliding beneath the laterally loaded soil nail "gravity" wall. The methods
of analysis for these failure modes are equivalent to those used for any gravity retaining structure (see
Chapter 4).

The requlred1'actor of safety for overall slope stability is 1.3 (or 1.5 where abutments are supported above
the soil nail wall) for Group I loading (ref. Section 5.2.2.3, AASHTO, 15th Edition, 1992). When
performing overall slope stability checks, consideration should be given to potential slip surfaces that pass
beneath the base of the wall and exit downslope or in front of the wall toe. This condition will be more
critical if the groundwater table is located close to the base of the wall. For this reason, it is recommended
that the nail lengths should not be shortened in the lower part of the wall (reference Step 5, above) unless
stability analyses confirm that deep-seated failure modes will not control.

It is recommended that the factor of safety against bearing capacity failure for Load Combination Group
I should have a minimum value of2.5. For other Load Combination Groups, the required minimum factor
of safety can be decreased in accordance with the percentage factors of the final column of Table 8-2 (Le.,
factor of safety of 2.5/1.25 = 2.0 for Load Group IV, and 2.5/1.33 = 1.9 for Load Group VII).

Step 9 - Check the Upper Cantilever

The upper cantilever section of a soil nail wall facing, above the top row of nails, will be subjected to earth
pressures that arise from the self weight of the adjacent soil and any surface surcharge loadings or internal
forces acting upon the adjacent soil. The magnitude of these earth pressures will depend not only on the
strength of the soil but also on factors such as the method of fill placement (if any) behind the cantilever.
If there is no fill placement and associated compaction-induced stresses behind the upper cantilever
following its construction, an active earth pressure coefficient can be assumed for the upper cantilever
portion of the wall. The cantilever is then checked for moment and shear at its base.

8 - 26
Step 10 - Check the Facing Reinforcement Details

Check waleI' reinforcement requirements, minimum reinforcement ratios, minimum cover requirements,
and reinforcement development and splices, as described below.

WaleI' Reinforcement

For temporary shotcrete construction facing, it is common practice and recommended that a minimum of
two No. 13 (Soft Metric designation, corresponding to a No.4 standard bar size) deformed horizontal
waleI' bars be placed continuously along each nail row and located behind the face bearing plate at each
nail head (i.e., between the face bearing plate and the back of the shotcrete facing). The main purpose of
the waleI' reinforcement is to provide additional ductility in the event of a punching shear failure, through
dowel action of the waleI' bars contained within the punching cone.

If a permanent facing is placed over the shotcrete construction facing, no waleI' bars are required.

Minimum Reinforcement Ratios

Minimum reinforcement ratios are specified in Section 8.17 and 8.20 of the AASHTO Standard
Specifications for Highway Bridges, 15th Edition (1992). These provisions are intended to ensure that
flexural failure mechanisms remain ductile (Section 8.17) and to provide a minimum amount of resistance
to shrinkage and temperah,lre distress (Section 8.20). However, soil nail wall facings are inherently ductile
even at extremely low retltforcement ratios. Therefore, the provisions in Section 8.17 do not apply and
are not necessary .

In addition, for temporary shotcrete facings, shrinkage and temperature cracking are not significant
concerns and the minimum steel ratio requirements of Section 8.20 may be waived for the temporary
facing.

In summary, only the shrinkage and temperature reinforcement requirements of Section 8.20 of the
AASHTO Standard Specifications for Highway Bridges (15th Edition, 1992) must be checked, and only
for a permanent facing system.

Minimum Cover Requirements

Concrete or shotcrete cover is necessary to provide bond resistance for the reinforcing steel and corrosion
protection for the reinforcing steel, the bearing plate and headed studs, as well as any non-encapsulated
or non-epoxy coated nail steel. In permanent applications, corrosion protection is a vital component of
the design. On the front side of the permanent facing exposed to the weather, the minimum cover required
is 50 mm. For permanent shotcrete facings, the minimum cover required on the side in contact with soil
is 75 mm. For permanent CIP facings, on the side of the facing cast against the temporary shotcrete, the
minimum concrete cover required for the reinforcing steel is 38 mm (Le., minimum distance from the CIP
concrete reinforcing steel to the concrete-shotcrete interface). For the temporary shotcrete construction
facing, corrosion protection is not a concern, and it is adequate to place the reinforcing steel near the
center of the facing.

Development and Splices of Reinforcement

Check that the splices and cutoff locations of all mesh reinforcement, deformed reinforcement bars,
horizontal waleI' bars, and vertical bearing bars (if used) are sufficient to develop the yield stress of the

8 - 27
reinforcement at all locations at which it is needed. All splices and development lengths shall be
proportioned in accordance with the AASHTO Standard Specifications for Highway Bridges, 15th Edition
(Sections 8.24, 8.30, and 8.32).

Step 11 - Serviceability Checks

Check the wall for excessive deformation and cracking (Le., check the service limit states). The following
issues should be considered:

Service Deflection and Crack Width of the Facing

Since the span-to-depth ratios for both temporary and permanent soil nail wall facings never exceed about
20, the structural deflections that occur at service load levels are insignificant and therefore not an issue.

The upper cantilevers of permanent soil nail walls are essentially one-way cantilevered slabs and may have
larger effective span-to-depth ratios than interior two-way spans. Therefore, the service crack widths (steel
stresses) must be checked in the same manner as for the stem of a conventional cantilever retaining wall.
The provisions of Section 8.16.8.4 of the Standard Specifications for Highway Bridges, 15th Edition
(1992) are used to check crack widths. For most temporary facings (whether as part of a temporary
shoring system or as the construction facing of a permanent wall system), serviceability requirements are
not imposed because the deflections ~ot pose any significant aesthetic or durability concerns.

Overall Displacements Associated with Wall COnstruction

During the top-down construction of soil nail walls, the reinforced soil zone tends to rotate outwards about
the toe of the wall as part of the process of mobilizing tensile loads within the nails. Hence, maximum
horizontal movements occur at the top of the wall and decrease progressively towards the toe of the wall.
Settlement of the facing also occurs, and tends to be on the same order of magnitude as the horizontal
movements at the top of the wall. Displacements of the facing depend on the following factors (Clouterre,
1991):

• Construction rate
• Nail spacing and excavation lift heights
• Nail and soil stiffness
• Global factor of safety
• Naillength-to-wall-height (LIH) ratio (more tilt and greater displacements when the wall height,
H, is large in relation to the nail length, L)
• Nail inclination (greater displacements for greater inclinations because of less efficient reinforcing
action)
• Bearing capacity of the foundation soils
• Magnitude of any surcharge loadings

For vertical soil nail walls with typical naillength-to-wall-height ratios, negligible surcharge loadings, and
designed with reasonable factor of safety, the peak displacement at the top of the wall tend to vary from
O.OOIH or less for weathered rock and very competent and dense soils (e.g., glacial tills), to O.OO2H for
granular soils, and up to O.OO4H for fme-grained clay type soils.

8 - 28
The construction displacements tend to decrease in magnitude with distance back from the wall facing and
are typically on the order of several tenths of the wall facing displacement at a surface location over the
ends of the nails. Figure 8-12 presents the Clouterre (1991) recommendations for estimating the manner
in which surface displacements decrease with distance from the facing for soil nail walls in various soil
types.

Post-construction monitoring of wall displacement indicates that some ongoing movements may occur with
time, depending on the ground type, and that some additional tension in the nails, particularly those near
the base of the wall, may develop. In many instances, however, ongoing straining of the nails is associated
principally with a redistribution of nail load from the creeping grout annulus to the steel tendon, as
inclinometer or wall survey data may indicate zero to minimal time dependent deformations of the wall.

Facing Vertical Expansion and Contraction Joints

Vertical joints are not required in the temporary shotcrete construction facing.

Per AASHTO Section 5.5.6.5 (1992), contraction joints spaced at intervals not exceeding approximately
10m and expansion joints spaced at intervals not exceeding approximately 30 m, as required for
conventional concrete retaining walls, can be used in permanent CIP or permanent shotcrete fmal fmish
facings.

Seismic Design

For seismic loading conditions, the structural strength factors and the soil factor of safety should be
modified by the percentages given in Table 8-2, for Load Combinations Group VII, in accordance with
the recommendations discussed previously. The construction facing need not be considered for seismic
loading conditions, as it will have limited life. The seismic loading is accounted for by applications of a
seismic coefficient as a psudostatic inertia force. The following guidance is recommended in defIDing the
appropriate design seismic coefficient:

• Select the appropriate design earthquake peak ground acceleration, ~k. In the absence of site
specific data or local seismic map'~k can be taken off the AASHTO Division 1A (1992) map
of Horizontal Acceleration, A.
• For slip surfaces that are primarily "internal" in nature (Le., intersect the nail reinforcement),
define a design seismic coefficient, A = (1.45-<\J<\k in accordance with that recommended for
MSE walls. This design seismic coefficient is applied to "internal" slip modes as a psudostatic
earthquake acceleration where the defmition of "internal" is given on Figure 8-13.
• For slip surfaces that are primarily "external" in nature, the design psudostatic seismic
coefficient, A, will vary depending on the permanent displacements that the retaining wall can
tolerate during the design event. For example, if the wall can tolerate a permanent displacement
of up to 250<\k mm (where <\k is the design earthquake acceleration as a fraction of gravitational
acceleration), then a design seismic coefficient equal to 0.5<\k can be assumed. For other
tolerable permanent displacements, the appropriate seismic acceleration coefficient can be
determined in accordance with AASHTO (Figure 37, Seismic Design Commentary Section, 15th
Edition, 1992).
• For assessment of seismic bearing stability of the reinforced soil block, a design seismic
coefficient equal to 0.5~k is recommended.

8 - 29
Existing Structure

A.
II
=H[1-tan TlllC I

ah:j -.
~

~ L J
L
.-11-------- '"'--'

oJ: -
... '
Displacement_
of Facing I ,

TI ---j , -
H
'I
,I,
II
- =Horizontal top of wall deflection (L)
L' - 5h
5v =Vertical top of wall deflection (L)
=Horizontal distance behind top of wall
"
- A
which may be influenced by construction
d.
- induced wall deformation (L)
=
11 Face batter (degrees)
L = Nail length
K = Damping coefficient(as taken from Clouterre)

Type of Soil Weathered Rocks Sandy Soils Clayey Soils


Stiff Soils

0h - Ov H 11000 2H/1000 3 H 11000


coefficient lC 0.8 1.25 1.5

Figure 8-12: Deformation of Soil Nail Walls (Clouterre, 1991)

8 - 30
/
/
/
/
I
/--+----I.J,/~_· Typical'lntemal'
r-------,.f../_J I
I Slip Surface
/ /----Typical'External'
/ / Slip Surface
/
/
II )/
r~~--,..::/:J'
/
'Internal' Slip Surfaces • intersect ground surface
/ at < 0 from top of wall
/,./' I
D'/~-~~--.._, 'External' Slip Surfaces· intersect ground surface
---~-- at > D from top of wall

Figure 8-13: DefInition of "Internal" and "External" Slip Surfaces for Seismic Loading Conditions.
(Byrne, et aI., 1996)

8.2.4 Simplified Design Charts for Preliminary Design of Cut Slope Walls

Design Variables

For simple cut slope geometric confIgurations in relatively homogeneous soil conditions, the design charts
shown in Figures 8-14 through 8-17 from Byrne, et ai., (1996) can be used to obtain a preliminary soil
nail wall design suitable for scoping or cost estimating purposes. The charts were developed using the
design methodology presented above. The design charts have been prepared for the common nail
declination of 15°. The design charts are presented in dimensionless format, with the following variables:

Geometric Variables

Backslope Angle, ~:

Four sets of design charts are presented (three charts per set), with each set of charts corresponding to a single
backslope angle of 0, 10, 20, or 34 degrees. For intermediate backslope angles, interpolate between the charts.

Face or Batter Angle, cS:

For each backslope angle, design information is presented for two face or batter angles of 0 and 10 degrees
from the vertical. For intermediate face or batter angles, interpolate between the charts.

Strength Variables

Factored Friction Angle, <Po:


The factored friction angle of the soil is defmed by the following relationship:

(8-7)

8 - 31
0.5 . . . . . . . . . , . .

,:
r
......... .. _.j .
---·r-1--- ····1·-··,··+-- ·_-I···j·-··
·~4·.·.4 '."-"'-'

H
iii .,
j If; jf ..J..
04 .. J••••.•••••••• ~ ••••••••••••••• '" ••••••••
. :
I I
I :'·
I :
... :

r
........
p"
:
: --TT- -:-I--r[-
t .
.....
.~.:~ - -
:
I
, ..
• I
t
f , I

-1+
I I
---1-- --1- :--f ...1-1
• I . . .······-i·· • I
tan +u
tan+o= -
I • I • I I I I ~~..:. I I ~I I : ~ I I I I I I I .. I I I I •
F+
0.3

To
00
I
w
N
I· · · !·-·I·-··I-·I-.--i·-I-·--I·-·I--·~ :·--i--··l·-l~I~~l-·I-:N· i.·-ri-J.l-····.i-I..· · 1- .1·. ·.·1 I Co = cJ(Fc'YH)
0.2

I Qo = aaQJ(ySVSH)
0.1 Face Baner • O·

---- Face Baller", 10·

.,
0.0 I I ! I I i I I I ! i . I I ' i . I I 'i! ' , , i ' . ! ( I To = aNTNN/('YHSVSH)

0.2 0.3 0.4 0.5 0.6 0.7 0.8


tanto

Figure 8-14(a): Preliminary Design Chart IA. Backslope = 00 (Chart A).


1.8
II I I•. i ...•
IO~-

•· I
• . I' I

:j!
I .: . I I

!: :
..1 j ·-::::·:.!::i.:~!.- -- :.j'... :l::~: ,'-'.: ·:::·-··1·······....
1.6 -J_-~-·_L~._ I :! •
.-.--~----------
••
••
:/j ••

v r··· r- .,•
·-I···l ,•
1.4
orr
..I'·· r
;.
14 -_.:"

r-·--:--1-",-
I !
~I"
•.•• -.•••

· •
--~--~-_._-----~
••
••
..-;_J,i·~·.[~'f·:::' ~:. i·'
I

: I :
·,•

.•...
1

;Z.~·~~-~.-t~-tl-~·~··t~I"11 ---- CCoO==0.03


1.2 0.01 ..•.........

00 I"·:'··:··:··,:-i·l-'ll·~.::. i' . Co =0.05


i

---~:~:::::1:+¥:f- ~~:1:: -- J::/--- -1:T-:--------


w
UH
w 1.0

1··1·'·····[- . . I·· · · .. "1 II


·· I I I:
'. .
.
:
---.--~•• i. __ L__ -····.:-·J.- 1--
: I
0.8 -- ••• --r--}--r--.------..
+,1·:··
1 ..
·,,••
0.6 ~.+_. _ ._._-
I 'I
Iii ·: ,. I !: I Ii il
j .. "r:· I··
• i ' i
I : ,I

0.4 I
, ! I I
I
1 I ,
I
. i !
i
.....
I I i ; , iI .···1· i·I··I·l
I"
i ·i··l
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
ToI°o

Figure 8-l4(b): Preliminary Design Chart lB. Backslope = 0°; Face Batter = 0° (Chart B).
1.6

00
I
W
~
UH

0.2 0.4 0.6 0.8 1.0 1.2 1.4


ToIOo

Figure 8-14(c): Preliminary Design Chart IC. Backslope = 0°; Face Batter = 10° (Chart C).
0.5 : I I :
.........1 1.1 , .. ··1·

...-1 ··1··.·:.:::: L.. . . . .. ! ......


I I
" I

I
I
H
0.4

tan
tan.o= ~
.u
0.3

To
00
I
VJ
VI
1··.. I···I·..·I···I···i . I··I···I·.. I.. ~,I·_·I . ·1~kT~I-··:N . . 1-·r~J··l . ··I·l.. . . 1 Co = ctl(Fc-yH)
0.2

00 = aa0tl(ySVSH)
0.1
Face Balter .. 00

---- Face Balter .. 100

0.0
'j'
I~--
i 'I'lj I
. To = (XNTNN/(-yHSVSH)

0.2 0.3 0.4 0.5 0.6 0.7 0.8


tan.o

Figure 8-15(a): Preliminary Design Chart 2A. Backslope = 10 0


(Chart A).
1.8 I I I I I! I ! , j, iii I ! I I! I I I I ! i I

1.6

1.4

1.2 ............ CD = 0.01


- - - - CD =0.03
CD = 0.05
00

w
0\
UH
1.0
··
.,
f

._._-~--~----------
.. ··,f .-
f
,
... 1

,f

·• f

0.8 -,- I--~-·~----------


Ii
0.6

0.4 I I ! I Ii! I iIi! iiI I I I i'i ! I I I i" ! I i"


0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
Tot°D

Figure 8-15(b): Preliminary Design Chart 2B. Backslope = 10°; Face Batter = 0° (Chart B).
I.
1.81-1·
I
I
I
I
I
! .
.. I
I '.

I;1.ii '·1I
f
i
II
I
i
II
I
II
I
;
:
I
I
I
I
I III :
- ~ . •
I
I

·J··I ",'
l .... _
·····__·:-·,······1·.. :.. :.._. : i
I I I"
......._._.L .,.. t··l· . L: .. "" '." ..- .. : I

_.
• .
.:
.: . j
I--t" -I'T:"
. ··r f :..:__
I

I I I
: : ! J__ I : .
I I ·
I : :
1.6 ---t __ __ ._ '- • • • •I._~

II'
--------~--.-_I.-

: I :. : : I' ·I
.
I

I .

. I'
.
I

II
.

I
• .
.
I

I .

.
I
/ I I •

. . =.'-~~.~.J Lj'.- ~ ~: L~". :~- '.: ' . ':"1: ...... :·:~:.·:~: . t~:[I: .
_.J .1-:; ..1:.
J
I.

--~--:-t--I---- -- -t- -'~.- -- -t-~---------+ -I


~ ! I
- -1"- --
I I I I I
1.4

i--I-· -...1--i.....- ·····1- -1--· - .... -_ ,-...':---.L _-. ·i··L


I
·_..._...i·_·t--
I L.......
._ .•.. _ .. .. I-.J
I I •
. ' I I

1.2 ._~._~_
II
.. ~- --- .. _.:-.
I
I
I
_. _. -- •I
-~.-
I
•••••••••••• Co = 0.01
*_ "'P_'
I
~_
I
- - - - Co =0.03
PM_ _ •••.•• _ •

• , I
_ _ _ I•.... !• ".._.._ .N.. I
• •...
I


:
I
I
Co = 0.05
00 .. :·_·; .. l·..· ·.. · · · : ..
I
I t · I
UH I

~--~-- --~--_.~--~--
I
: I" .: I
IN
-....I 1.0 ••
- - , .... t- _

.i I •
• I·
•I
I
··1
·
I
I
I

I
I I
I I
• I

0.8 __ 1.
I .. _ .......

I
....

.. I•
I
I
I
I
I

I
I
I

0.6 I

[..
I •
I

--r--j----------
:
II I l:
0.4 I ' • I I I I " I I I I I ·-rr·l· l I I i I

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6


To/CD

Figure 8-15(c): Preliminary Design Chart 2C. Backslope = 10°; Face Batter = 10° (Chart C).
0.51 I
I
I .
i
I· .
I I ;
: I
.•..
I'I i I I
Ln ' n In
I
I' i·1 " : , n-l
l.t i II ~J.: . l. i.LI I I I

.......... ·1··
H
0.4

tan +u
tancllo= ~

0.3

00
To
I
w
00 I ! ··'··j·I·!· . / ../ ·11··f'l···I··I:··'L· :-··I··~~ I··-~I I·· l:~ ~P~Y;':'~ I I Co = cJ(Fc)'tf)
0.2

,.
if · I I • • 1 f • •

. Q o = aaQJ(ySVSH)
0.1 I Face BaUe, =O·
I.
..... 1.. .... 1 - - - - FaceBaUer .. 10·

0.0 IlL
0.2
~";: T 0.3
I I i
0.4
I • • • i
0.5
I ! • • i
0.6
! •. I
0.7
I I· I
0.8
To = aNTNN/()'tfSVSH)

tan +0

Figure 8-16(a): Preliminary Design Chart 3A. Backslope = 20° (Chart A).
1.8 I! I :


I
I
:
II II I' I I·:
I
" '
III: •I I
I :
JI
i.'· : I II ,.:
!I
II I i
I ..'•
I'''' I•

~ ~
L
I

I
.
w_~
!
I

,
!I .: !I '.'J!
I
!
I I'
I .

I
I t·! '.

I
i, .
~
. ! 1.-'.••••
I

, I t. i )~.)
~ " I. . t
I

I
.
.. ~ ..
••• .. •• : . • :. I .:.. :. _ .: ...... : . ... :

~.+ !.• J•• • _~ •••• _•••• -:.-J. •.._••.•.• ~~t..~~..1 ._ ~_.~_ .._.._.._
• • I I ' • I • I •
1.6-"--
·:•!
I

.
:.• I. "

.'1


I

!- .I,'l J .! .• ,.
:t
I
:•
• . . . I.'
! ~< {1 .[¥fl!.
••f ,'~'
••. •. ,
I: ,
.
. /. i
:

1.....•....
• •. r I I • _. , • . •

.. •,
1.4 ~ .. --~--

,,
I

.. ,,,
.,•,

•••••••••••• Co =0.01
- - - - C o =0.03
C o =0.05
00

VJ
10
UH
1.0 lI
-·~·T······_·_-

·····lr·
,
. .

0.8 i I
--:--j----------
, ,

··,•
,•
•,,
0.6 ! I ,. ·-r··~··········
r
, I ! I

!. "j.' ··r·!· . ····i ', ', !


I : !! : I '
: I .,:
I

I. 'I :; : I· :

i·· . , . . i
.! I •
0.4' I ! ! I ii' i! I I , i ' I I , i i " ' l ! Ii I I
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
ToI°o

Figure 8-16(b): Preliminary Design Chart 3B. Backslope = 20 0


; Face Batter =0 0
(Chart B).
1.B I I I I I
I
I. . I I'
I
I
I ,
o'. .•
i
. I.I .I
' I' I
'•
I I I

i I:
I . , •
: •. : f : : I- :
,_ ..··1 .... - --i I -, .] i- - i" .__...} "I··+.~;· , . · · i
:-.. ..---- :.\ I' -~ j -: . .- . . :- . ._.
~.... _···i -l~·i·A· I . r ~ ....-.-
M ·. . . . i
t I: :
. r---- -- j--:-. -- -- .. ·-r-i.. . . --j---:-- -- -- -- --r--
I I:
1.6 -·---~-·r- ---------i;,.;;-~-

. . :! II I : !
:
. •••~:•••:; I
.,,: . . . : :
I I I I I I . ;.I 0\ I I

' • • •: .-. -. .:
. ··1-· :. ·l-· --·r····t· ·······1--···:I ~ ......~:i# -.. ~ ··········~-l···!..·
! I -t-I--- ! - - L•• ..;; ':':;~~'r
·r
I··
Ii' ••• I

1.4
!.-
-- --,i"....
"1
I
:,
... t--- ,,I
:
1.2 , Co = 0.01
- - - - Co =0.03
Co =0.05
00 tan "0 DO.
~
UH
1.0 '-.L~
,
-I··
, .
.
!
I

O.B
, : I .. ·I

--t--~--L-----. _
"
·,
I
I
•,
I

,
I

0.6 --r--~
,
I

. _·_ . _-----
.~.
,I
I

,
I

,,
I

0.4 I I ' • I , I I 'I I 'I I 1-1.. 1· . i


0.0 0.2 0,4 0.6 O.B 1.0 1.2 1.4 1.6
TelQo

Figure 8-16(c): Preliminary Design Chart 3C. Backslope = 20°; Face Batter = 10° (Chart C).
0.5 'i!

!l···l·!Ii··
I' l

." i··I·,!,
,. l f I I l II l Ii I '
I· i· i"j. .
I l I
I '. +··1··1······: . .1: I,.. -!-: i I I.:
I-j···j·-r-j· . . . . . .j. ... .......[- II ! .
I I.

0.4 --~---t----.·j---r-- ---r-- J-- --- ---r-- --- ------ ---r-- --- __ ---r-- ---I---f--l--- r-- ---/---------.

I-'---1--
L ----~i------ --1---- --1-- -1'-1
I·.· · "1 I :
-1-1-
_ :........ .. ' _ :.......
I
tan eIlu
I , "

I
~~ .~~~ ~_ •. -:.~ -~~- .•-.
,
~ •.• ~.~_•. ,. - --. "'--"':'--- _.~-~. --.--_••.
I
~._. •-.- ••..• ;._.. _.... _. • _, ". __ "'M"

: _
tan eIlo = Fell
0.3 --~······-~--·---r-· -.- ,---.- -'-r" --- --, -- '--r'- -·-,·--I--I---r-·~---+---~--~---
"
" "
• I ,

" "
To 'I":: .
"

l
i ..... :
l
.:.
I I :: ,

l
00
,.. . !..... .....! _ .. - .. !.. . Co = 0.01!·
-
• l •
Co = cJ{FclH)
~
0.2
----l: -. :l~ -~: • :.: ·-T- :~-:.-.- -: -l +-J-:-b,-~t,--~1~;r~---:---------'
• , . ' I ,

:I :I : CD =I 0.05 II
I :. : :
I I

CD = 0.01
I

: 00 = aaOJ('}'SVSH)
~..... I I I :
0.1 Face BaUer. 00 - _e ••• ·,-t-- ~.
, l

----
H'_"_" •••.•-.. "1 ' "f ..•. , , ....• , o.
l •
Face Baner _100 l
••·····......_.H...h.H.. t. k •• .. . •_• • . . H..
l
,

r·· ··-T·· ···l·r·"!····· •




. l
l

~
I

0.0 -f-...l-...l-...l-----r-.-L--'-.-L-.-L--r-.-L-.-L-...l-...l--t-...l-...l-...l-...l--t--l-.-l-.-:..--:..--t-.J-..J-.-l-....l.-f To = aNTNN/(lHSvSH)

0.4 0.5 0.6 0.7 0.8 0.9 1.0


tan .0

Figure 8-17(a): Preliminary Design Chart 4A. Backslope = 34° (Chart A).
1.8 I I i ,
: : . I':' . I.


I

I
I
I' I \. I ' I
,.

, .
: .
I

I

I' .
I

i·· .- -- .-! ·1,1---1---, i I - . .- .-... 1


.:
.. ~--~-~_.~__ ~
... !'Il- ! --: .
·_··_~··~_·_·_I' __ JI__ ~ __ ; __ -- --
... . -.+,
-- -- -- --~--~----------
1.6
, I I • I I
-1- -~J -IJ1-J .1: t··--t······ ..... - . . ~
wt-'c·l--- =j:-:-· -~=-nTj::ll:+-
. .
:-: :.
: ...... - ·-----·--1 ." I
.- -: ..". _. --~:------. : -- -I
1'- ,: .- ,-.
- -: --.'.._-
- :.I-lL:-:-:-:
..-.
: .
"1""':' . I- .

1.2 I I i
... -'--
.i i rII i ............ Co = 0.01
I
I
•...
, _..
.... _.....
• _...
-or - - - - Co =0.03
I Co =0.05
00
.·._···_,'----1--..
I

I
I I I
I

--1--lh1nf-t--- mh '
I
LJH . I

~ 1.0 --~--

'.·_I~.I=: .._
I
I
I ..
I
I
I

.·.
I
I I
I
I

·, .---------
I
I
0.8 ·-~--r·
.. ..
"

I
I
I
I

I
I
·,,
0.6 •
--r--~-·--·-----
I
.,I
I
I
I

0.4 i I I I I i I . -1··,·1.··:
..........·1·· I: I
I'" ' - l ' r ' lI- '"I'!
I :
.I I'
'!i I

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6


To/Q D

Figure 8-17(b): Preliminary Design Chart 4B. Backslope 34 0 ; Face Batter = 0 0 (Chart B).
" • I
1.8
r •

.
I
I ,.. I
I

l.



II·
,'.,
,
I
I "
I I
I I
..•. ...._. "'-'~I--" •....
• ._._ .. .. ~ -, ....•••..
,

1.6 -- -- -- -- •••. --
.
!"
!
••
.
I.


.:

..... - ....•
I
;...

..........- .-_
·
.•....
--.--
I

1.4 -- ~...
-. --
.-,".
. I

I
I


o .p. ..- •... ~.. :.. I..
1.2 -- Co =0.01
- - - - Co =0.03
CD =0.05
00

~
w
UH
1.0 ·l
·I
I •

--f--. -----_.. -_..


I

r'r'
.

0.8
rr
--t._l. ... _
·••
I

··••
I

I
I

0.6 I ..... to .... _-- ...
--,
I
_

" ..
I

I
I
I
I

0.4 I I I I 'i! I i I I , 'i' I I I iii I I I I i I I I I J I· I
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
To/Qo

Figure 8-17(c): Preliminary Design Chart 4C. Backslope = 34 0


; Face Batter = 10 0
(Chart C).
where <l>u = ultimate soil friction angle
F~ = global factor of safety applied to soil friction strength, from Table 8-5

Dimensionless Cohesion, CD:

Co is the soil cohesion normalized with respect to the soil unit weight and the vertical height of the cut:

(8-8)

where Cu = ultimate soil cohesion (kN/m2)


Fe = global factor of safety applied to soil cohesive strength, from Table 8-5
The dimensionless cohesion is shown as a parameter for each slope geometry for three values of 0.01,
0.03, and 0.05. Interpolate for intermediate values of the dimensionless cohesion.

Dimensionless Nail Tensile Capacity,To :

The dimensionless nail tensile capacity is the factored nominal nail tensile strength normalized with respect
to the soil unit weight, the vertical height of the slope, and the nail spacings:

(8-9)

where aN = nail tendon tensile strength factor, From Table 8-5


TNN = nominal nail tensile strength (kN)

The dimensionless nail tensile capacity, To, is obtained from the vertical axis of Chart A of each chart set.

Dimensionless Pullout Resistance, Qo :

The dimensionless pullout resistance is the factored ultimate pullout resistance (expressed as a force per
unit length of nail), normalized with respect to the soil unit weight and the nail spacings:

(8-10)

where aQ = nail pullout resistance factor, from Table 8-5


Qu = ultimate pullout resistance (kN/m)

The dimensionless pullout resistance is shown as being incorporated into the ratio, To/Qo, on the
horizontal axis of Charts B and C of each chart set.

Design Chart Procedure

The procedure for using the design charts in conjunction with the dimensionless variables discussed above
consists of the following:

Select the design chart set corresponding to the appropriate backslope angle. If necessary, interpolate
results for intermediate backslope angles to those given in the charts.

8 - 44
Compute the factored soil friction angle, <1>0' and the dimensionless factored soil cohesion, Co, as defmed
above (Equations 8-7 and 8-8). From the appropriate Chart A, determine the dimensionless nail tensile
capacity, To.

The required nominal nail tensile strength, TNN' can then be determined from the relations presented
above (Equation 8-9), and from knowledge of the dimensionless nail tensile capacity, To (from Chart A),
the soil unit weight, the vertical height of the slope, the vertical and horizontal nail spacings, and the nail
tendon strength factor.

Compute the dimensionless nail pullout resistance, Qo (Equation 8-10). Divide the calculated
dimensionless nail tensile capacity, To, by the computed dimensionless nail pullout resistance, Qo, and
determine the required nail length from the appropriate Charts B or C, depending on the batter of the face
of the wall.

8.2.5 Corrosion Protection

The long-term performance of permanent soil nails requires that they be able to withstand corrosive attack
from their local environment. Soil tests should be performed to measure the aggressiveness of the soil
environment, especially if field observations indicate corrosion of existing structures. The most common
and simplest tests are for electrical resistivity, pH, chloride, and sulfate. In general, if the electrical
resistivity of the soil is greater than 5000 ohm-cm and the pH between 5 and 10, the soil may be
considered to be non-aggressive and additional corrosion testing is unnecessary. If the electrical resistivity
is between 2000 and 5000 ohm-cm, sulfate and chloride tests are required. The designations for these
tests, and the critical values defining whether an aggressive soil environment exists, are presented in Table
8-9. The ground is considered aggressive if anyone of these indicators show critical values.

Nail Tendon Corrosion Protection

For permanent applications, soil nail corrosion protection should consist of the following:

• In non-aggressive ground, the nail section should be resin-bonded and epoxy-coated using an
electrostatic process to provide a minimum epoxy coating thickness of 0.3 mm in accordance
with AASHTO M-284 (1993). The intact epoxy coating will prevent tendon corrosion by
isolating the tendon from the surrounding environment. In addition, the recommended minimum
thickness of coating will generally prevent normal handling and construction-induced damage.
A minimum grout cover of 25 mm is recommended throughout the length of the nail.
Centralizers should be placed at distances not exceeding 2.5 meters center to center, and the
lowest centralizer should be placed a maximum of 0.3 meters from the bottom of the grouted drill
hole. The centralizers should be made from a plastic material, be attached to the nail in a way
that will not impede the free flow of grout, and be sized to position the nail tendon within
approximately 25 mm of the center of the drill hole.
• In aggressive ground or for critical structures (e.g., walls adjacent to high volume roadways or
walls in front of bridge abutments) or where field observations have indicated corrosion of

8 - 45
TABLE 8-9
GROUND AGGRESSIVENESS INDICATORS
(FHWA-RD-89-198, 1991)

Property Test Designation Critical Values

Resistivity ASTM G 57, AASHTO T-288 below 2,000 ohm-cm

pH ASTM G 51, AASHTO T-289 below 5

Sulfate ASTM D 516M, ASTM D 4327 above 200 ppm


AASHTO T-290

Chloride ASTM D 512, ASTM D 4327, above 100 ppm


AASHTO T-291

Notes:
1) User should check test standard for latest updates.
2) Individual transportation agencies may have limits on critical values different than
tabulated above.

existing structures, encapsulated nails should be used. Encapsulation is generally accomplished


by grouting the nail tendon inside a corrugated plastic sheath. A neat cement grout containing
admixtures to control water bleed from the grout is usually employed to fill the annular space
(typically 5 mm minimum) between the plastic sheath and the tendon. For this type of protection,
the minimum grout cover between the sheath and the borehole wall can be reduced to 12 mm.

For temporary application (of less than 36 months duration) in non-aggressive ground, the soil nail grout
is considered adequate protection.

Nail Head Corrosion Protection

If the nail is encapsulated or is an epoxy-coated deformed bar with machine threads at the upper end, the
corrosion protection is terminated to expose. the bare tendon at the head of the nail in order to allow
attachment of the bearing plate and nut. This area may be more susceptible to corrosion than the
remainder of the nail since oxygen is more readily available. For this type of nail tendon the following
approach has been most commonly used for providing corrosion protection to the nail head. First, the
bearing plate assembly is embedded in the permanent facing with the normal depths of cementitious cover
to control steel corrosion. Second, the nail tendon protection (epoxy coating or encapsulation) is extended
into the shotcrete construction facing to ensure a minimum depth of shotcrete/nail grout cover of 75 mm.

Figure 8-18 shows examples of the types of acceptable corrosion protection systems for permanent soil
nails.

8.2.6 Wall Drainage

Groundwater is a major concern in both the construction of soil nail retaining walls and in their long term
performance. Soil nail walls are best suited to applications above the water table. Excess water at the face

8 - 46
Embed Nail Encapsulation to
Provide Minimum of 75 mm
Shotcrete or Nail Grout Cover

Epoxy Coated Bar

CIP

/
1~
/
a) Epoxy Coated Soil Nail Detail (Upper 1
Meter Encapsulated) With Temporary
Shotcrete Construction Facing and
Permanent CIP Facing (Caltrans)

Embed Nail Encapsulation to


Provide Minimum of 75 mm
Shotcrete or Nail Grout Cover

Epoxy Coated Bar

CIP

b) Epoxy Coated Nails With Machine Threads


Detail, With Temporary Shotcrete Construction
Facing and Permanent CIP Facing

Figure 8-18 (a) and (b): Alternative Soil Nail Corrosion Protection Details.

8 - 47
Embed Nail Encapsulation to
Provide Minimum of 75 mm
Shotcrete or Nail Grout Cover

CIP
Thickness Typically
100mm

c) Encapsulated Soil Nail With Bare Steel at the


Top Detail, With Temporary Shotcrete
Construction Facing and Permanent CIP Facing

Embed Nail Encapsulation to


Provide Minimum of 75 mm
Shotcrete or Nail Grout Cover

.
. .....
..
.:

:..
~.'
~.,
..
Thickness Required
to Achieve Minimum
Cover Requirement
Shotcrete LocallyThickened in VICinity of Bearing
Plate in Lieu of Providing Greater Shotcrete
Thickness Over Full Facing

d) Permanent Shotcrete Facing Detail With


Encapsulated Nail

Figure 8-18 (c) and (d): Alternative Soil Nail Corrosion Protection Details.

8 - 48
can result in face stability problems during construction together with an inability to apply a satisfactory
shotcrete construction facing. In addition, long-term face drainage is required to prevent the generation
of localized high groundwater pressures on the facing.

A commonly adopted design for controlling surface runoff consists of a surface interceptor ditch, excavated
along the crest of the excavation and lined with concrete applied during the shotcreting of the first
excavation lift. The ditch should be contoured to drain away from the working area, with collector drain
pipes installed at appropriate locations, if necessary. Where larger graded slope areas exist above the wall,
installation of plastic film slope protection sheeting above the interceptor ditch provides another quick and
inexpensive means of controlling surface water during construction. Similar permanent surface drainage
measures are generally required to prevent surface waters from infiltrating behind the facing, or flowing
over the top of the wall, during the operational life of the structure. For stepped or benched walls
vegetation can also be used to inhibit infiltration and lower soil water contents by evapotranspiration.

Long-term groundwater drainage measures may include the following:

• Face Drains: These typically consist of 400-mm-wide prefabricated geocomposite drains that are
placed in vertical strips down the excavation face, on a horizontal spacing corresponding to the
nail horizontal spacing, and discharging either into a base drain or through weep holes at the
bottom of the wall.
• Shallow Drains (Weep Holes): These are typically 400 mm long, 50 to 100 mm diameter PVC
pipes discharging through the face and located where heavier seepage is encountered.
• Horizontal Drains: If it is determined that the retained soil will be subjected to groundwater
pressures and there is no serious impediment to construction of the wall, then the wall must be
designed to support the anticipated driving and uplift groundwater forces. Usually in such cases,
however, deep horizontal drains, typically consisting of 50 mm diameter slotted or perforated
tubes and inclined upward at 5 to 10 degrees to the horizontal, are installed to relieve the
groundwater pressures imposed on the retained soil mass. The design spacing and depth of these
drains are site specific, but they will typically be longer than the length of the nails and with a
density of approximately one drain per 10 square meters of face. Deep horizontal drains may
also be used to control unanticipated water flow during construction. For aesthetic reasons, the
drain outlets may have to be plumbed and carried down the wall face between the shotcrete
construction facing and permanent CIP facing, and then outlet at the wall base. Horizontal drain
flow exiting directly out through and flowing down the exterior permanent face is unsightly.

8.2.7 Construction

Following are the typical sequence and procedures used to construct a soil nail wall by the drill and grout
method of nail installation, which is the most common method used in North America (Figure 8-3).

A. Excavate Initial Cut

Before commencing excavation, measures should be taken to control all surface water. This is usually
done by the use of collector trenches as discussed in Section 8.2.5. The initial cut is excavated to a depth
slightly below the first row of nails, typically about 1 to 2 m depending on the ability of the soil to stand
unsupported for a minimum period of 24 to 48 hours. If face stability is a concern for these periods of
time, a stabilizing berm can be left in place until the nail has been installed, and final trimming then takes
place just prior to construction of the facing. Another method of dealing with face stability problems
includes placing of a flash coat of shotcrete. Generally, face stability problems are likely to be most severe

8 - 49
during the first one or two excavation stages, because of the presence of near-surface weathered and
weakened materials or, in urban environments, the presence of loose fills or voids often associated with
buried utilities.

Mass excavation is done with conventional earth moving equipment (photo 8-2(a». Final trimming of the
excavation face is typically done with a backhoe or hydraulic excavator. Usually, the exposed length of
the cut is dictated by the area of face that can be stabilized and shotcreted in the course of a working shift.
Ground disturbance during excavation should be minimized, and loosened areas of the face should be
removed before the shotcrete facing support is applied. The excavated face profile should be reasonably
smooth and regular to minimize subsequent shotcrete quantities.

A level working bench on the order of 10 m width is typically left in place to accommodate the drilling
equipment used for nail installation. Smaller tracked drills which work on bench widths as narrow as 5
m and with headroom clearance as low as 4 m are available. Larger bench widths may be necessary
depending upon the equipment to be used during nail installation.

B. Drill Hole for Nail Installation

Nail holes are drilled at predetermined locations to a specified length and inclination using a drilling
method appropriate for the ground. Typically nail spacings are 1 to 2 m both vertically and horizontally.
Typical nail lengths are 60 to 100 percent of the wall height and nail inclinations are generally on the order
of 15 degrees below horizontal to facilitate grouting. Drillhole sizes range from about 100 mm to 300 mm
in diameter. This usually permits a grout annulus of at least 30 mm thickness around the reinforcement.

Most soil nailing is undertaken using small hydraulic, track-mounted drill rigs. These rigs are mostly of
the rotary/percussive type that use sectional augers or drill rods. Large hydraulic-powered track-mounted
rigs with continuous flight augers have been used to install nails up to 28 m in length. These rigs have the
advantage that they can drill the entire length of the nail in a single pass without having to stop to add
additional sections of auger. Production rates can therefore be very high. Their main disadvantages are
the high mobilization costs and the large space requirements for their operation.

For ground conditions that are well suited to soil nailing, the rate of nail installation will be increased (and
the costs correspondingly reduced) by drilling the holes without casing. This is commonly referred to as
the "open-hole" method. Open hole methods suitable for soil nailing are augering (photo 8-2 (b», rotary
drilling with a drag bit or roller-cone bit, and rotary-percussive drilling. Air is the preferred medium to
remove drill cuttings, since flushing with water in uncased holes can significantly lower the bond stress
developed during subsequent grouting, particularly in moisture sensitive soils such as silts and clays.

Where ground conditions are such that the drillhole wall will not stand unsupported for the length of time
required to insert and grout the nail, the drilling method must provide casing for support of the drillhole.
It should be noted that such conditions will often have serious adverse impacts on the cost effectiveness
of soil nailing. Such methods include single tube and duplex rotary methods with air or water flush, and
hollow stem augers. Flushing with air should be performed at moderate velocities and volumes to ensure
that ground fracturing does not occur. This precaution is especially important in residual soils or badly
weathered rock where fissures or fracture planes exist. Based on experience with ground anchors, the
temporary support of drillholes by bentonite or other mud suspensions is not recommended, as "smear"
on the drillhole walls can significantly reduce the grout-to-ground bond.

8 - 50
(a)

(c)

(b)

Photo 8-2: Soil Nail Construction: (a) Excavation; (b) Drilling with Auger; and (c) Shotcrete
Construction Facing.

8 - 51
C. Install and Grout Nail

Plastic centralizers are commonly used to center the nail in the drillhole. However, where the nails are
installed through a hollow stem auger, centralizers are generally ineffective and a stiffer (200 mm or lower
slump) grout mix is used to maintain the position of the nail and prevent it from sinking to the bottom of
the hole. The nails, which are commonly 20 to 35 mm bars, are inserted into the hole and the drillhole
is fIlled with cement grout to bond the nail bar to the surrounding soil. Grouting takes place under gravity
or low pressure from the bottom of the hole upwards, either through a tremie pipe for open-hole
installation methods or through the drill string (or hollow stem) or tremie pipe for cased installation
methods.

D. Place Drainage Systems

As noted in Section 8.2.6,400 mm wide geocomposite drainage strips are commonly installed against the
excavation face before shotcreting occurs to provide drainage behind the shotcrete face. The drainage
strips are extended down to the base of the wall with each excavation lift and connected either directly to
a footing drain or to weep holes that penetrate the final wall facing. These drainage strips are intended to
control seepage from perched water or from limited surface infiltration following construction. If water
is encountered during construction, short horizontal drains are generally required to intercept the water
before it reaches the face.

E. Place Construction Facing and Install Bearing Plates

The construction facing typically consists of a mesh-reinforced wet mix shotcrete layer on the order of 100
mm thick (photo 8-2(c», although the thickness and reinforcing details will depend on the specific design.
Following placement of the shotcrete, a steel bearing plate (typically 200 mm to 250 mm square, and 19
mm thick) and securing nut are placed at each nail head and the nut is hand wrench tightened sufficiently
to embed the plate a small distance into the still plastic shotcrete.

The shotcrete facing may be placed at each lift prior to nail hole drilling and nail installation, particularly
in situations where face stability is a concern.

F. Repeat Process to Final Grade

The sequence of excavate, install nail and drainage system, and place construction facing is repeated until
the [mal wall grade is achieved.

G. Place Final Facing

The elP concrete facing (Photo 8-3(a» is typically structurally attached to the nail heads by the use of
headed studs welded onto the bearing plates. Under appropriate circumstances, the final facing may also
consist of a second layer of structural shotcrete (Photo 8-3(b» applied following completion of the [mal
excavation.

Precast concrete facing panels (Photo 8-3(c» have also been used as [mal facings, and can be attached to
the construction facing in a variety of ways. The precast panels can consist of smaller modular units or
of full-height tilt-up panels and provide the means of integrating a continuous drainage blanket behind the
facing. One disadvantage of this type of facing is the difficulty of providing adequate long-term corrosion
protection to all attachment/connection devices. A further disadvantage of the smaller modular panels is
the difficulty of attaching the panels to the nail heads. Some systems are also restricted by patent.

8 - 52
(b)

(a)

(c)

Photo 8-3: Soil Nail Wall Final Facing: (a) Cast-in-Place (CIP) Facing; (b) Shotcrete Facing; and
(c) Precast Concrete Facing.

8 - 53
A disadvantage of the full-height precast panels is that they are practically limited to wall heights of about
8 m because of weight and handling limitations.

8.2.8 ~aterials

The materials required include the nails themselves and the associated corrosion protection system, the nail
grout, the drainage materials, the shotcrete/concrete facing materials, and the connection system between
the nail head and the facing.

A. Nail Tendons

Nail tendons typically consist of deformed steel bars conforming to AASHTO M311ASTM A615, Grade
420 or 520, or ASTM A722 for Grade 1035. The bar tendons must be continuous without splices or
welds, straight and undamaged. They can be bare or epoxy coated or encapsulated, as required based on
corrosion protection considerations. The bars should be threaded a minimum of 150 mm at the exposed
head end to allow proper attachment of the bearing plate and nut. Threading may be continuous spiral
deformed ribbing provided by the bar defonnations (e.g., Dywidag or Williams continuous threadbars)
or may be cut into a reinforcing bar. If threads are cut into a reinforcing bar, the next larger bar number
designation should be used to accommodate the loss in cross-sectional area.

B. Fusion Bonded Epoxy Coating

The fusion bonded epoxy coating should conform to the requirements of ASTM A775. The coating should
have a minimum 0.3 mm thickness electrostatically applied. Bend test requirements can be waived. The
coating at the wall anchorage end of epoxy coated bars may be omitted over the length provided for
threading the nut against the bearing plate.

C. Encapsulation

Encapsulation material should consist of a minimum 1 mm thick corrugated HDPE (High Density
Polyethylene) tube conforming to AASHTO M252 or corrugated PVC (Polyvinyl Chloride) tube
conforming to ASTM D1784, Class 13464-B. Encapsulation must provide at least 5 mm of grout cover
over the nail bar and be resistant to ultra violet light deterioration, normal handling stresses, and grouting
pressures. Factory fabrication of the encapsulation is preferred. Field fabricated encapsulation must be
done in strict accordance with the manufacturer's recommendations.

D. Nail Grout

For conventional drill and grout nail installation, the nail grout consists typically of a neat Portland cement
grout with a water-cement ratio of about 0.4 to 0.5. Where a stiffer consistency grout is required (e.g.,
to centralize the nail when no centralizers are used in a hollow stem auger installation, or to control
leakage of grout into the ground such as in highly permeable granular soils or highly fractured rock), a
lower slump sand-cement grout may be used. Sand-cement grout may also be used in conjunction with
large nail holes for economic reasons.

The grout should have a minimum 3-day compressive strength of 10.5 MPa and a minimum 28-day
compressive strength of 21 MPa in accordance with AASHTO T106/ASTM C109.

8 - 54
Admixtures must be compatible with the grout, and mixed in accordance with the manufacturer's
recommendations. Accelerators should not be permitted. Expansive admixtures should only be used in
grout for filling sealed encapsulations.

E. Drainage Systems

Typical soil nail drainage systems are discussed in Section 8.2.6.

F. Wall Facings

The facing consists of two component parts and is defmed primarily in terms of the timing of construction.
The "construction facing" is the facing erected during excavation and initial construction of the wall. It
usually consists of a minimum 100 mm thick mesh-reinforced wet mix shotcrete, as this system provides
a continuous, flexible surface layer over the excavated soil face. The wet mix shotcrete process is
preferred to the dry mix process and is used almost exclusively for soil nail wall facings. The advantages
of the wet mix process include better quality control of the water content (water-cement ratio of about 0.45
to 0.50), the ability to air-entrain for improved freeze-thaw durability, and the availability from local
ready-mix plants. The shotcrete mix should be capable of attaining a minimum 14 MPa compressive
strength in 3 days and a minimum strength of 28 MPa in 28 days. Additional requirements for shotcrete
are provided in Appendix C 1 of the FHWA manual for Design & Construction Monitoring of Soil Nail
Walls (Byrne, et ai., 1996).

Final facings are usually installed following completion of the excavation to fmal grade. The most
common fmal facing used to date on permanent walls is CIP reinforced concrete (typically 200 mm
minimum thickness), as this type of facing can be readily adapted to satisfy a variety of aesthetic and
durability criteria. A second layer of shotcrete or precast concrete panels can also be used as the final
facing.

8.2.9 Nail Testing

The design of a soil nailing system is usually tested in the field to verify that: (a) the design loads can be
carried by the nails without excessive movements, (b) the contractor's equipment and installation
procedures are adequate, and (c) the long-term behavior of the nails is as anticipated. Four types of tests
are usually performed: an ultimate test, a verification test, a proof test and a creep test. Table 8-10
presents a brief discussion on various aspects of these tests (Portfield, et at., 1994; Clouterre, 1991).

Testing Equipment

Figure 8-19 illustrates a typical set-up for a soil nail test. A hydraulic jack and pump are used to apply
the load to the nail. A jacking frame or reaction block is usually installed between the shotcrete (or the
excavated face) and the jack. Once the jack is centered, an alignment load is applied to the jack to secure
the equipment.

Movement of the nail head is measured by one or preferably two dial gages attached to a rigid support
independent of the jacking set-up. The dial gages are zeroed after the alignment load has been applied.
These gages should be capable of measuring movement to the nearest 0.02 mm.

The load is applied to the nail by a calibrated hydraulic jack. The jack should have a minimum travel of
150 mm. A calibrated load cell, which should be aligned with the axis of the nail and the jack, is used to
detect small changes in load and allow maintenance of constant load during creep testing.

8 - 55
TABLES-to
TYPES OF SOIL NAIL TESTS

Test Purpose Procedure Acceptance Criteria


Ultimate • To determine the ultimate adhesion capacity • Incrementally load the nail until pullout takes • Total movement at the maximum test load (typically
(if carried to pullout failure). place along the grout-ground interface (pullout 200 percent of the design load) must exceed 80
• To verify the design adhesion factor 0 failure is the inability to maintain constant test percent of theoretical elastic movement of the
safety. load without excessive movement). unbonded length.
• To determine the soil nail load at which • Perform test on non-production (sacrificial) soil • Creep movement between 6 and 60 minute readings,
excessive creep occurs. nail. at a specified test load (typically 150 percent of
design load), must be less than 2 mm.
Verification • To verify that installation method will • Incrementally load the nail to a value equal to
provide a soil nail tendon capable of design load times the factor of safety for design • Pullout failure must not occur at the maximum test
load (typically 200 percent of the design load).
achieving the specified design adhesion with soil adhesion (typically 2.0). This value mayor
a specified factor of safety and design may not be equal to that causing pullout.
adhesion capacity. • Perform test on non-production (sacrificial) soil
• To verify capacities for different soil/rock nail.
conditions and/or drilling/installation
methods during production.
00
Proof • To provide information necessary to • In a single cycle, apply increments of load until • Total movement at the maximum test load (typically
VI evaluate the ability of production nails to a maximum test load (typically 125 to 150 125 to 150 percent of the design load) must exceed
0\
safely withstand design loads without percent of the design adhesion capacity) is 80 percent of theoretical elastic movement of the
excessive structural movement or long term reached. unbonded length.
creep over the structure's service life. • Perform test on a specified number (typically 5 • Creep movement at maximum test load (typically
percent) of the total number of production soil 125 to 150 percent of the design load) should be:
nails installed. < 1 mm between 1 and 10 minute readings, and
< 2 mm between 6 and 60 minute readings.
Creep • To ensure that the nail design loads can be • Perform test as part of ultimate, verification or • Follow the above acceptance criteria for creep.
safely carried throughout the structure's proof test.
service life (typically 75 to 100 years) • Under constant load, record movement at
without movements that could damage specified time intervals. Plot deflection versus
the structure. log time results on a semi-log graph and check
acceptance criteria.
Notes:
1. Testing procedures and acceptance criteria have not yet been standardized and vary among different agencies. Check specifications for applicable test acceptance
criteria.
2. The maximum test load, the load increments, and the time that each load increment is held are determined by the type of the test being performed.
3. In no case should the soil nail tendon be stressed to more than 80 percent of its minimum ultimate tensile strength for grade 1035 steel, or more than 90 percent of the
minimum yield strength for grade 420 steel. Otherwise, an explosive failure of the steel can occur.
4. The strength data for each size of soil nail tendon based on diameter and grade of steel should be obtained from the nail tendon manufacturer.
5. Personnel should always remain clear of the testing apparatus/nail tendon, since bar failure could occur at any time, particularly at higher test load.
• 4

.

REFERENCE
PLATE

.. -

DIAL GAUGES
ATTACHED
TO SUPPORTS
ON GROUND
INDEPENDENT
OF WALL

TEST NAIL

..'"
..
"

.
"

• - 4

1 - - - - - - SHOTCRETE FACING.
BULKHEAD OR GROUND

Figure 8-19: Typical Test Nail Detail. (Porterfield, et ai., 1994)

8.3 MICRO-PILE WALL

8.3.1 General

This Section 8.3 describes two types of micro-pile walls, including the reticulated micro-pile wall and the
Type "A" insert wall. In addition, this section presents an overview of the design procedures and
construction requirements for these types of walls. Much of the material presented in this section was
obtained from the FHWA report titled Drilled and Grouted Micropiles: State-of-Practice Review (4
volumes). Due to the limited application of micro-pile walls, this manual does not attempt to provide a
thorough discussion of design procedures. The reader is referred to the above publication for a more
complete discussion of the different design methods currently in use for micro-pile walls.

Micro-piles are small diameter (less than 300 mm) drilled piles constructed with some form of steel
reinforcement, and bonded to the ground with grout that mayor may not be placed under pressure.
Micro-pile walls are earth walls formed by a cluster of micro-piles, usually installed at various angles, to
provide excavation support or to stabilize slopes, roadways or foundations. Figure 8-2 illustrates typical
examples of micro-pile walls for earth retaining applications.

The retaining structures shown in Figure 8-2(a) and (b) are created by a reticulated, or interlocking,
network of micro-piles. These systems are designed to create in situ a coherent, composite, reinforced soil

8 - 57
gravity structure. The design concept used for these reticulated micro-pile (RMP) walls is similar to that
used for gravity retaining wall structures. The wall system shown in Figure 8-2(c) consists of micro-pile
groups installed in an "A" shaped arrangement, often called a Type "A" insert wall in the United States.
The Type "A" insert walls are not reticulated as much and depend less on the network (knot) effect
between the soil and the piles. Type "A" insert walls do not behave as gravity walls and, therefore, a
design procedure different from that used for the RMP walls has been developed. General design
procedures for both types of micro-pile walls will be presented in this chapter.

Advantages/Limitations

The main advantage of a micro-pile wall is that it can be constructed on steep slopes using relatively small
and light-weight equipment. The walls can also be constructed in areas of low headroom such as under
bridge decks.

Other advantages of the micro-pile walls are similar to those for the soil nail walls presented previously.
They include:

• Reduction in the amount of excavation.


• Elimination of backfilling behind the wall.
• Reduction in disruption due to construction operations.
• Adaptability to different site conditions.

The limitations of the micro-pile wall include the requirement of underground easements for installation
of the micro-piles, and the relative difficulty in analyzing the system, which currently relies primarily on
experience and rules of thumb established from limited case history studies.

Photo 8-4: Exposed Piles of a Reticulated Micro-Pile (RMP) Wall (Fondedile Company)

8 - 58
8.3.2 Design

Reticulated Micro-Pile (RMP) Wall

A RMP wall can be considered analogous to a reinforced concrete wall where the ground presents a mass
analogous to concrete while the micro-piles correspond to steel reinforcements which allow the overall
structure to resist compression, tension and shear. By varying the number, spacing and arrangement of
the micro-piles, the desired load carrying capacity of the wall can be achieved.

To design the RMP wall to act as a monolithic unit, it is necessary to assume that the stresses acting on
the wall are distributed to both the soil and the piles rather than the piles alone as is usually considered in
conventional pile foundation design. To achieve this, a "knot effect" is created by which the stress acting
on the pile is partially transferred to nearby piles due to the interaction of the piles and the ground
generated by the high bond between them. This "knot effect" has been confirmed by both model and field
tests (Lizzi, 1978; Plumelle, 1984; Korfiatis, 1984). Photo 8-4 shows exposed piles of a RMP wall and
the resulting reinforced soil mass.

The capacity of a RMP wall is affected by the density, strength and degree of saturation of the in situ
ground as well as the type, size, number, arrangement and method of installation of the micro-piles. For
a relatively shallow slide surface, the cap beam provides added resistance. If the potential failure surface
is deeper, the resistance of the cap beam is not normally considered in the design (Pearlman, et al., 1992).

The design of a RMP wall depends on experience and intuition, as much as on analytical procedures. The
behavior of this system depends to a great extent on the knot effect concept and the reinforced concrete
analogy. Once the knot effect is assured, the RMP wall behaves as a coherent body and is analyzed as a
gravity retaining structure. This gravity wall design concept has been adopted for analyzing several micro-
pile walls in the U.S. Using this concept, the following steps can be followed for preliminary design of
a RMP wall (see Figure 8-20):

1. Estimate the earth pressure against the wall using conventional earth pressure theories (see
Chapter 2) or slope stability analysis (see Module 3- Soil Slopes and Embanlanent), as illustrated
in Figure 8-20(b).

2. Assume a pile configuration that is dense enough to assure a knot effect and minimize the
possibility of plastic flow between piles. A preliminary configuration can be based on a density
of 6 to 7 piles per linear meter along the wall alignment. Stability related to the plastic flow of
soil between piles can be verified by comparing the horizontal thrust exerted by the earth mass
with the limit resistance developed by the arching effect between two adjacent micro-piles. The
procedure developed for this evaluation (Ito and Matsui, 1975) will be discussed later.

3. Assuming the wall acts as a coherent gravity structure (Figure 8-2(c)), check for sliding and
overturning in conformance with the design of a gravity wall (Chapter 5).

4. Check for sufficient shear elements to resist sliding. Determine the safety factor against sliding
along a potential or existing slip surface using the principles of slope stability analysis. In general
terms, the factor of safety, FS, can be expressed as follows:

(8-11)
FS =

8 - 59
.,Rock
,~
Surf;ace

(a) (b)

A -A
~ ---CONC
E CONC + A STL -E.m....
~ E S01L
E S01L

..L
T
Unit Width

Resultant
I~ b
~I

(c) (d)

Figure 8-20: Design Model for RMP Walls - Gravity Wall Concept: a) Reticulated Pile Structure, b)
Calculation of Earth Pressure, c) "Gravity Wall" Model, and d) Transformed Section
(Bruce and Juran, 1997)

8 - 60
where R is the horizontal component of the soil shear resisting forces along the critical slip
surface, n is the number of micro-piles intersecting the critical slip surface, R I is the additional
shear resistance due to the micro-piles, and D is the horizontal component of the driving force
along the critical slip surface. R I depends either on the shear resistance of the piles or the lateral
resistance of the soil in contact with the piles, whichever is smaller.

5. Using the gravity retaining wall design concept, the design should verify that the resultant of the
earth pressure and dead load forces acts in the middle third of the foundation base (see Figure
8-20(c» to resist overturning moment and maintain compression stress in the soil and in all piles.
The recommended procedure for determining the extreme fiber stresses at the base of the gravity
wall section is illustrated in Figure 8-20(d) and (e), which is an analogy to the design of a
reinforced concrete beam. In this procedure, the transformed section area of each pile is
calculated based on the ratios of Young's moduli of the concrete and steel reinforcement relative
to the soil.

The transformed section area is given by:

A = A E eone + A E stl
trans cone E sll E
soil soil

where Aeone = Area of the concrete or grout


Astl = Area of the reinforcing steel
Aeone = Young's Modulus of the concrete or grout
Aeone = Young's Modulus of the reinforcing steel
The moment of inertia of the base of the structure, Itrans , is computed by assigning equivalent
areas of soil to the concrete (or grout) and steel based on the ratios of Young's moduli. Extreme
fiber stresses are computed as:
p
CJ = - -

where P, e and bare defmed in Figure 8-20.

The design should check to ensure the extreme fiber stresses are kept in compression at the heel
of the wall.

6. Using the extreme fiber stresses computed in Step 5, determine the maximum compression
stresses in each component of the composite section discussed above (Le., concrete, steel
reinforcement and soil). Check these compression stresses against allowable design values. This
step also allows calculation of the compression force for each pile.

7. Design each micro-pile to resist the forces determined from Step 6.

8. Adjust pile configuration and repeat above steps as needed.

The procedure presented above for the RMP wall design has adopted a very conservative assumption in
that tension is not allowed to develop in any of the piles to resist the overturning moment. This often

8 - 61
results in a large number of micro-piles required to reinforce (knot) the increased soil mass due to a wider
wall base. A more cost effective design can be achieved if the piles are allowed to carry uplift forces by
deeper embedment in the firm and stable strata below the critical failure plane. If this approach is adopted
the design procedure should also include the determination of tension as well as compression forces in the
piles. This can normally be achieved by using a simplified structural frame analysis. If warranted, a
detailed pile group analysis, taking into consideration the soil-structure interaction, can be performed.

Type "A" Insert Wall

Recent research studies (pearlman, et ai., 1992; FHWA, 1995) have indicated that micro-pile walls with
configurations similar to those in Figure 8-2(c) (often referred to as Type"A" insert walls due to their
shape) were not behaving as gravity walls. A new approach has been developed for this type of wall. In
general, the design procedure involves the following:

• Conduct stability analyses to determine the required increase in resistance along a potential or
existing slip surface to provide an adequate factor of safety.
• Check the potential for structural failure of the piles due to loading from the moving soil mass.
• Check the potential for plastic failure (Le. flow of soil around the pile).

Typically, movement of marginally stable, non-creeping soil slopes occurs within a relatively thin zone
which is subject to large shear strains. Generally, the materials above and below the zone of failure
experience negligible shear strains. The purpose of the micro piles is to connect the moving zone (above
the slip surface) to the stationary zone (below the slip surface), and thus to increase the sliding resistance
along the slip-plane.

Because the micro-piles are relatively flexible, the maximum bending moments in the piles tend to develop
relatively close to the slip-plane. Fukuoka (1977) devised a theory to evaluate the bending moments which
develop in a pile oriented perpendicular to a slip-plane.

Figure 8-21 is a chart for preliminary design of Type"A" insert walls. The chart was developed using
the method described in Fukuoka (1977) and considers four sizes of pile elements. It should be noted that
the ultimate horizontal resistance designated in Figure 8-21 is either the load which causes yield stresses
to develop on the outer edges of the steel pipes (Le., Pile Types 1, 2, and 3) or the load which causes
crushing of the concrete surrounding the centralized reinforcing bar (Le., Pile Type 4).

Figure 8-21 shows that the ultimate horizontal resistance of the piles is a function of the coefficient of
subgrade reaction, Ks, of the soil or rock above and below the slip-plane.

Plastic failure of soil around the piles is analyzed using a procedure developed by Ito and Matsui (1975).
The method is based on the fundamental consideration that soil deformation is restricted to a plane strain
condition. Typically, this type of failure occurs if the soil above the slide plane is relatively soft and the
piles are stiff and spaced far apart. This is usually not the case with relatively flexible micro-piles, but may
govern when stiffer pipe elements are employed. Based on the theory proposed by Ito and Matsui (1975),
the predicted results for various pile spacings and soil conditions are plotted in Figure 8-22.

Pile Design

Due to its small base, the end bearing capacity of an individual micro-pile is minimal and the load is

8 - 62
' ,OJ I J ' , , , , Pilc Holc
Typc Dia.
Pipc O.Dor
Q{ehar Dial11~
Pipe Wall
Thickncss
Steel Yicld
Strcngth
Concrctc
Strcngth I I I I r=!= -~
.J------t\ - -
----., (1111ll ) (111111 ) (nun) Fy (MPa) Fe' (Mila)
I 2(1) 171> 13 552 21{
315 t: Top Stmtul11 -
2
3
152
152
1411
IlI2
III
9.5
552
552
2R
2R --e;
"" 'dc rhmc 4 152 (2R.7) N/A 414 2R

~
Bottolll
Stratum ---Pilc
10
I '" \.0
'f- \J'f-\\01 '" 0.\
'f-~o\J~
-
.£• 2711 I - - - I I

00

0\
w

Cocfficicnt of Horizontal Subgradc Reaction ofTop Stratum, K Tup' (MN/m 3)

Figure 8-21: Preliminary Design Chart for Ultimate Horizontal Resistance of Piles. (Pearlman, et ai., 1992)
SSO 70()
DiD, ~ 0.67

SOO DiD, ~ 0.75


DiD, ~0.88 600
~ 4S0 DiD, = 0.95 '2
'"
c..
~
c..
~ Note: Curves Show Values for
~" 400 .e"
c.. 500
Depth = 3m
~ ~ y= 19kN/m~
ii: ii: For Other Depths, Multiply by d/3m
B IDepth = 15n B
'5
Vl
350 '0
Vl
E
0
E
o oS-I
....
0.0
<l:: <l:: 400 rl--- ~'-'
.!! 300
... ---Pile
~

~
V>
s:: Slide Plane
~
l- I Depth =9m V>
V>
00 a1 250
l:: ~
Vl Vi 300 rl- - -
~ 3s:: 1ils::
0 o
.t! 200 N
'1::
0 o
:c :c
~ ~
2:!

:> 200 I ,I - - - V"..----:A'" I :7.......1
.§ 150
5
I
100 Depth = 3m
0.0
oS-I""'"
~~ ---Pile
I. DI _I

100
I 1
IV - ~--
\ :~.o~
Slide Plane
50 ~dk~

o I I I I I I I o I I I I I I

o 25 50 75 100 125 150 15 20 25 30 35 40


Undruined Shear Strength ofTop Stratum Soil, c (kPa) Angle of Friction ofTop Stratum Soil, ell Top' (Degrees)

Figure 8-22: Ultimate Stress Transfer from Soil to Pile Versus Shear Strength of Soil. (Pearlman, et ai., 1992)
transferred to the soil mainly by skin friction. In some cases, such as piles into rock, it may become
important to take into consideration the end bearing effect. Skin friction allows the pile to resist both
compressive and tensile forces. Only small bending moments, however, can be accommodated because
the pile diameter is small.

The design of a micro-pile involves determining both internal (structural) and external (ground related)
load-carrying capacities. The internal pile capacity is calculated according to conventional structural
design procedures using the forces (compression or tension) determined from analysis. There are various
methods for the geotechnical evaluation of the external axial loading capacity of micro-piles, depending
on parameters such as the techniques used in drilling and grouting, the initial state of stress, engineering
properties of the soils, and others. It is generally recognized that the method of grouting and the soil
characteristics are the two main factors governing the soil/grout bond development and, therefore, the pile
capacity. Micro-piles can be either tremie (gravity) grouted, pressure grouted, or post-grouted. For
micro-pile wall construction, the tremie grouting and pressure grouting techniques are commonly used.
The recommended methods (Bruce and Juran, 1997) for the geotechnical evaluation of the axial loading
capacity of individual micro-piles are presented below with regard to:

• Grouting technique (Le., tremie grout and pressure grout)


• Soil type (Le., cohesionless, cohesive soils and rock)

Tremie Grouted Micro-Piles in Soil

The unit skin friction, fs ' along the micro-pile in cohesionless soil, is estimated using the ~ method:

fs = RO
p vz '
(8-12)

~ = Ktano (8-13)

where ovz' is the vertical effective stress at depth z; K is the coefficient of earth pressure at the wall of
the pile, and {) is the angle of wall friction for the soil in contact with the pile.

The unit skin friction, fs , in cohesive soils is determined using the a method:

fs = as u (8-14)

where Su is the undrained shear strength of the soils, and a is an empirical parameter that depends on
several factors including drilling disturbance on the soil, roughness of interface between concrete and soil,
the method used to assess Su' and others. For micro-pile design, it is generally satisfactory to use an a
value between 0.6 and 0.8 (Bruce, 1994).

Another simple empirical approach developed by Lizzi (1980) for tremie grouted micro-piles in soils is
given by Equation 8-15:

Pu = 1tDLkI (8-15)

where Pu is the ultimate pile capacity in kg; D is the nominal diameter, in cm, of the pile (Le., the drilling
diameter); L is the pile length in cm; k is an empirical coefficient representing average pile-soil interaction

8 - 65
TABLE 8-11
VALUES OF k AND I
(Lizzi, 1980)

Values of k (kg/cm2) Values of I (non-dimensional)

Soil Type k Pile Diameter, D, cm I


Soft Soil 0.5 10 1.00
Loose Soil 1.0 15 0.90
Soil of average compactness 1.5 20 0.85
Very stiff soil, gravels, sands 2.0 25 0.80

in kg/cm2 ; and I is an empirical factor related to the pile diameter. Table 8-11 presents the empirical
values of k and I as proposed by Lizzi (1980).

Tremie Grouted Micro-Piles in Rock

The ultimate bond strength at the rock/grout interface can be estimated as one tenth of the uniaxial
compressive strength (DCS) of intact rock samples up to a maximum of 4.0 MPa:

Des (8-16)
10

For soft or highly weathered rock, Littlejohn and Bruce (1977) suggest the use of SPT-N values as follows:

fs = 0.007N + 0.12 (MPa) (8-17)

fs = O.OlN (MPa) (8-18)

where Equation 8-17 was based on micro-piles installed in weathered granite and Equation 8-18 was
established for stifflhard chalk.

Pressure Grouted Micro-Piles in Cohesionless Soils

The p method used for tremie grouted micro-piles discussed above has been generalized for pressure
grouted micro-piles in cohesionless soils (Littlejohn, 1970) as follows:

(8-19)

where K 1 is the earth pressure coefficient typically ranging between 1.4 and 1.7, and K2 is a parameter
reflecting the effect of increased pile diameter due to grouting pressure. K2 varies from 1.2 to 1.5 for
dense [me sand, 1.5 to 2 for medium sand, and 3 to 4 for coarse sand and gravel.

Pressure Grouted Micro-Piles in Cohesive Soils and Rock

In general, the application of low grouting pressures at the installation stage of the micro-piles in cohesive
soils and rock is considered to be of little practical value in enhancing the micro-pile capacity. Therefore,

8 - 66
design procedures recommended for the tremie grouted micro-piles can be used for pressure grouted
micro-piles in these materials.

Post-Grouted Micro-Piles

As noted previously, the magnitude of skin friction developed along the pile-ground interface is higWy
dependent on the grout pressure, particularly in less competent materials. Figure 8-23(a) illustrates the
influence of grouting pressure on the ultimate load capacity of ground anchors (Littlejohn and Bruce,
1977). If post-grouting is used it may result in significant improvement in the load-carrying capacity of
micro-piles. Figure 8-23(b) illustrates the impact of post grouting on skin friction (Herbst, 1982). For
cases where post-grouting is used, the method presented above may underestimate ultimate pile capacities.
The recommended procedures for estimating axial load capacity of post-grouted micro-piles are presented
in a FHWA publication by Bruce and Juran (1997).

8.3.3 Construction

Figure 8-24 illustrates the construction sequence of a micro-pile wall. The construction steps are as
follows:

1. Excavate an area wide and deep enough to accommodate the cap beam.

2. Install the formwork for the cap beam including steel reinforcement.

3. Place corrugated plastic sleeves for installation of the micro-piles through the cap beam.

4. Pour the concrete cap.

5. Install the micro-piles through the plastic sleeves, applying grout pressure, if necessary.

6. Excavate and install the wall facing, if required.

Pile Installation

Micro-piles are installed basically following the drilling - reinforcing - grouting sequence. Although there
is a wide range of details, dependent on local practice and contractor preference and experience, the
general sequence of construction is similar to that described previously for soil nail or soil anchor
installation and is briefly presented below.

A cased hole is first drilled in the ground to the required tip elevation (Photo 8-5(a». The drilling fluid
is used to remove the excavated soil. Reinforcement, in the form of a steel bar, pipe, or cage, is then
placed inside the casing and the hole is filled with cement grout. (In some countries, e.g. Britain and Italy,
sand is a common additive to the grout.) While the casing is lifted, additional grout is tremie placed and,
if required, a grout pressure is applied to increase the contact stress between the soil and the pile and
produce a pile of larger diameter than that of the drilled hole. Sometimes, the casing is left in place to act
as the pile reinforcement.

For soil drilling, a duplex method using inner and outer drilling elements rotating in opposite direction is
common (Bruce, 1989), but single casings and hollow stem augers have also been used (the hollow stem
auger is preferred when drilling in clays or sensitive soils). Rotary, percussion or core drilling is used
depending on the type of ground and the contractor's equipment and experience.

8 - 67
450
420
390
.... 360

~ 330
....
0300
.s=
g
co
270
'5 240 0.30
:5co 210 N'
E 0.25 I In post grouted
section
li

l6" 180 .§
u 2; 0.20
~
co 150 c:
.2 I
~ 120 U 0.15
;S
5 90 II
~ 0.10
60 en :::
... i=-
In non-post grouted
0.05
30
iirrllllill
0.7 1.4 2.1 2.8 3.5 o 200 400 600 800 1000
Grouting Pressure (N/mm 2 ) Load (KN)
(a) (b)

Figure 8-23: (a) Influence of Grouting Pressure on Ultimate Load Capacity. (Littlejohn and Bruce,
1977); (b) Effect of Post-Grouting on Skin Friction. (Herbst, 1982)

~----':=fi=Fi ==Fi=====r---_ _/

/'" STEP 1: Excavate 0.6 - 1.0 m for concrete cap.

~, "~f~,,~g PI.., steel snd ,.""gated /


polyethylene sleeves for the reinforcing units.
_

/0 "
~,lIUI,~----'=F=F==F===:::r-- /

STEP 3: Pour 2 m wide by 1 m deep concrete cap.

/_~"""'-"~i IF=F===::::r-- /

STEP 4: Drill and grout micro-pile into place,


pressure grouted at 200 to 280 kPa.

/--f}=T~=1:1-,-,--"=Fi=F=i=F=====----...J'/
! ! 1 1 '"
STEP:5: Regrade shoulder and repair roadway.

Figure 8-24: Typical Construction Steps for a Micro-Pile Wall. (Bruce, 1992)

8 - 68
The drilling rigs are similar to those used for installation of soil anchors. They are often track-mounted,
hydraulically powered by diesel or electricity, and capable of using power sources stationed away from the
drilling area which would be beneficial when installing the micro-piles in confined space or steep side slopes.

Water or air can also be used as the drilling fluid. Water has the advantage of cooling the drill bit while
displacing the soil cuttings. The water level inside the drilled hole must be kept above the groundwater
level at all times to prevent bottom instability. Bentonite slurry should not be used because of the potential
for reduction of the frictional resistance along the pile-ground contact area.

The grout mix is prepared in a high-speed colloidal mixer and continually injected through the drill head
as the casing, or auger, is withdrawn. Based on the existing ground conditions and design requirements,
the grout can be placed either by gravity head only, or pressure grouting techniques. When placed by
gravity head, sand-cement grout as well as neat cement grout is used. With pressure grouting, neat cement
grout is injected under pressures typically in the range of 0.3 to 1 MPa; the grout pressures are limited to
avoid hydrofracturing of the ground around the drill hole and prevent excessive grout losses.

Regrouting of the pile bond length is sometimes perfonned to increase pile/ground friction resistance. In this
method neat cement grout is placed in the drillhole by gravity. Some hours later, when this primary grout
has hardened, similar grout is injected through a preplaced, sleeved grout pipe. A packer is used inside the
sleeved pipe so that specific locations can be regrouted, one or more times, at pressures of 2 to 8 MPa.

CIP Cap, Excavation and Facing

After completion of pile installation a reinforced cast-in-place concrete cap is placed to tie the tops of the
piles together (Photo 8-5(b». This cap may also serve as the upper part of the earth retaining system.
Following the required curing time for the cap, excavation is performed, if necessary, on the cut side of
the wall using conventional earth excavation equipment. The excavation is performed in controlled lifts,
and a construction facing of reinforced shotcrete is applied for each lift (Photo 8-5(c». Wall drainage
elements are installed using procedures described in Section 8.2.7D for soil nail walls. After the
excavation reaches final grade, a fmal facing is installed. This fmal facing may consist of cast-in-place
concrete (Figure 8-25), or a second layer of shotcrete.

8.3.4 Materials

Regardless of the installation method, micro-piles consist of two basic materials: grout and reinforcement.
In addition, a cap beam made of conventional cast-in-place reinforced concrete is provided to interconnect
individual pile heads. Following is a brief description of the basic requirements of the grout and
reinforcement.

Cement grout may contain aggregate and chemical additives. Typically, the grout mix has a water-cement
ratio ofOA to 0.5, by weight. Recently, neat cement grout has become very common as the presence of
aggregate may be harsher on the life of the equipment. In post-grouting, the water content of the grout
is slightly higher.

The pumpability of the grout mixture becomes a major concern when long piles are installed. The grout
in this case should be sufficiently fluid to travel the relatively long pumping distance, yet stable enough
to resist permeation losses after injection.

8 - 69
(a)

(b)

(c)

Photo 8-5: Construction of a Reticulated Micro-Pile (RMP) Wall: (a) Drilling for Pile Installation; (b)
Placement of CIP Concrete Cap; and (c) Excavation and Shotcrete Facing.

8 -70
"

Existing Grade

,-..
~ 305mm
S CIP Facing L.j;;l':IH-~"""
'-'
S Grade Beam / Pile Cap
102mm
<""l Shotcrete '-_-="_ Micropiles
r-... Facing

Ambedment Line

3 m Min.

.,,-
.,,-
.,,-
.,,-
.,,- "
.,,-
.,,-

-~. -~. .,,-


.,,-
.,,-
.,,-
.,,-
.,,-
.,,- .,,-
.,,-
.,,- .,,-
"

Figure 8-25 Cross-Section of a Completed Micro-Pile Wall

Portland cement types I, II, or III may be used in the grout. Chemical admixtures can be added to modify
the grout properties and control set time. Grout compressive strengths of 20 to 40 MPa are usually
achieved.

Pile Reinforcement

Steel bars, cages, strands (tension piles only), pipes or casings can be used for reinforcement. When
deflections must be minimized, or when the pile must support high lateral loads, steel pipes or casings are
normally used.

Figure 8-26 illustrates the different types of reinforcement used in micro-piles. When strands or bars are
used, centralizers are usually provided to center the reinforcement and assure adequate grout cover. When
working in low headroom areas, the reinforcement is placed in sections and connected by the use of
mechanical couplers.

The steel reinforcement is normally made of Grade 420 steel, but higher grades can be used to provide
extra strength (Dash, 1987). Deformed steel reinforcing bars, 19 to 57 mm diameter, are normally used.

A typical reinforcing pipe used for micro-piles may have a diameter ranging from 100 mm to 180 mm.

8 -71
In potentially corrosive environments, the reinforcing bars can be provided with an epoxy coating.

8.3.5 1restb1g

A pile load test program is typically performed to verify the adequacy of the micro-pile design and the
proposed installation procedures. Both compression and tension tests should be performed for conditions
(e.g., loadings, construction procedures, and subsurface conditions) similar to those anticipated for the
production piles.

There are no established guidelines for determining the number of load tests to be performed. The actual
number used is usually dependent on previous experience and the level of confidence in the soil data and
the design parameters. Generally, 1 to 3 percent of production piles are tested (Gorneck, et ai., 1993).
In addition, load tests should be performed at every significant change in the subsurface conditions.

Load testing procedures used for foundation piles (Module 8 - Deep Foundations) can be used for testing
micro-piles.

Tie Bar
~iPe

Gro~.t ~~mfure~
einforcing

Grom
fj Bar

Grom
~ '.
,0
i~
. . . •••..•. . . •. . .•'0
O>C)
. . Strand
.•. •. . •.••.. . .
Bar
Grout

(a) (b) (c) (d)

Figure 8-26: Different Types of Reinforcement for Micro-Piles. (a) Reinforcing Bar; (b) Pipe;
(c) Strand; and (d) Rebar Cage.

8.4 EXAMPLE PROBLEM 8-1

Statement

A soil nail wall has been proposed for a road cut through a site underlain by medium dense silty sands.
Other conditions and assumptions that are relevant for the design of this cut slope soil nail wall are listed
below:

• There are no existing utilities or structures. Modest settlements that may result from the
construction of the wall are not of concern.
• Groundwater table is well below the base of the planned cut.
• The site is located in an area of very low seismic risk.
• For the static loading condition, AASHTO Load Group I governs the design.
• The wall is classified as a permanent, non-critical structure.
• Site investigation data provided the following design soil parameters:

8 -72
In situ density (y) = 18 KN/m3
Internal friction angle (<I>u) = 34°
Cohesion (cu) = 5 KN/m2
Pullout resistance (Qu) = 60 KN/m (for 200 mm diameter nails)

Solution

Step 1 - Set Up Critical Design Cross-Section and Select a Trial Design

The wall will have a maximum vertical height of9.5 meters, with a face batter of 10 degrees from vertical.
The critical design cross-section is shown on Figure 8-27 and will have a 20 degree slope at the top of the
wall. The trial nail spacing will be at 1.5 meters, vertically and horizontally, and the nails installed at 15
degrees below horizontal for constructability reasons.

To provide the initial trial design for the nails, the simplified design charts presented in Section 8.2.4 are
used. Following the procedure presented Section 8.2.4, the initial nail design is derived as follows:

1. The critical design section has a face batter of 10° and a backslope angle of 20°. Therefore,
design charts presented in Figures 8-15(a) and 8-15(c) should be used.

2. Compute the factored soil friction angle, <1>0' and the dimensionless factored soil cohesion, Co,
using Equations 8-7 and 8-8 as follows:

0
= 1
tan- [tan(<I>u)] = tan-I [tan(34 )] = 26.50
F cI> 1.35

tan(<!>D) = tan(26.S 0) = 0.5

cu (5 kN/m 2)
= = 0.022
(FcyH) [1.35(18 kN/m 3)(9.5 m)]

(Note: Fel> and F c are global factors of safety applied to soil frictional strength, tan<l>u, and
cohesive strength, cu' respectively. For AASHTO Group I loading condition, Fel> = F c = 1.35,
see Table 8-5.)

Determine the dimensionless nail strength capacity, To, using the chart in Figure 8-15(a).

To = 0.23 (see Figure 8-28(a»

3. Determine the required nominal nail tensile strength, TNN , using Equation 8-9.

yHsvSHTD 3
= (18 kN/m )(9.5 m)(1.5 m)(1.5 m)(0.23) = 161 kN
TNN = ---.;..--.;..~
~ (0.55)

8 -73
8.0

~ Critical Circular Slip Surface


ES. = 1.35
Bilinear Approximation
1.0 to Slip Surface

H-9.5

00

~
,, NOTES:

,,/e l-7.7
All dimensions In meters
All allowable nail loads In kN

el-7.7
' - - -......
I-
Horizontal and vertical
nail spacings =1.5m
o Nallnumber
3.83

1.0

Figure 8-27: Sample Problem - Trial Design Cross-Section


0.5
i
•, ·· i .1 I .,·· i I
I

··
I I
I
·· .1
... I

0.4
I ·· i'

0.3

To

0.2

0.1

- - - - Face Baner. 10'

I I! II
0.0 -1-~.....J..-!--t-!....-!.--l......L-f--I--:"-"-...:.....Ilf-l--l..."':'-~~-'--'---t""":"'''':'-''''''''--1
0.2 0.3 0.4 0.5 0.6 0.7 0.8
Ian to
(a)

1.6

1.4

1.2

UH
1.0

0.6

0.4 +-J....l-Jl-J-.f-l-..J.L-t-.!......!..-!....:......;-!-!.....!...4...:.---+-..I-..!.~'-+-..I-..l......l-l-+-'---'---t
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
TdOo

(b)

Figure 8-28: Preliminary Design Charts for Backslope 20° and Face Batter 10°. (a) Determination of
TD ; and (b) Determination of LlH.

8 -75
(Note: CXN is the nail tendon tensile strength factor. For AASHTO Group I loading condition,
CXN = 0.55, see Table 8-5.)

4. Compute the dimensionless nail pullout resistance, Qo, using Equation 8-10.

= _ _(.:..,0_.5..:...;)(:.-60_kN_/m....,:.)_ _ = 0.74
(18 kN/m 3 )(1.5 m)(1.5 m)

(Note: cxQ is the nail pullout resistance factor. For AASHTO Group I loading condition, cxQ =
0.50, see Table 8-5.)

Compute the ratio of To / Qo, and determine the required nail length from the chart in Figure
8-15(c).

To / Qo = 0.23/0.74 = 0.31
LIH = 0.87 (see Figure 8-28(b»
L = 0.87 x H = 0.87 x 9.5 m = 8.3 m

In summary, the design charts indicate a required bar yield strength of about 161 kN (use No. 25, Grade
420 bars - corresponding to standard bar size No.8), and a nail length of about 8.3 meters. This nail
length could be slightly conservative since the effective backslope angle of the design selection is less than
20 0 (Le., the design charts are prepared for constant backslope angles only, whereas the design example
backslope angle is variable). For comparison purposes, a backslope angle of 10 0 would result in a
required nail length of about 7.3 meters. For initial trial design purposes, an intermediate nail length of
7.7 m is selected. This length is shown in Figure 8-27 as the maximum nail length. If global stability is
not of concern, the length of the nails located in the lower half of the wall height may be reduced (see Step
5 below).

Step 2 - Compute the Allowable Nail Head Loads

Based on local practice and material availability, the trial design will assume the use of a "standard"
temporary shotcrete construction facing (28-day compressive strength of 28 MPa), having a nominal
thickness of 100 mrn, reinforced with a single layer of 152x152, MW 18.7 x MW 18.7 (6x6-W2.9 x W2.9)
welded-wire mesh. Two No. 13 (Soft Metric Designation - corresponding to standard bar size No.4),
Grade 420 continuous horizontal waler bars will be used along each row of nails, and two No. 13 vertical
bearing bars at each nail head (see Figure 8-29).

The nails will be connected to the shotcrete construction facing by a 225 mrn square, 25 rnrn thick bearing
plate. The nail pattern will be staggered such that the horizontal location of each nail is at the mid point
of the nails of the adjacent rows. The permanent facing will be a Cast-in-Place (CIP) concrete wall (28-day
compressive strength of 28 MPa), 200 mrn thick, reinforced with No. 13 Grade 420 deformed bars on 300
MM centers vertically and horizontally, and connected to the nail heads with a headed-stud connection
system. Figure 8-29 shows the details of the temporary shotcrete facing and the permanent CIP concrete
facing.

Based on the data presented above, the nominal nail head strengths for Load Group I are computed for both
the temporary shotcrete facing and the permanent concrete facing. The computations are presented below.

8 -76
Nail Head Strength for Temporary Facing

(a) Facing Flexure

Using Equation 8-1, the nominal nail head strength under the flexural failure mode can be calculated as
follows:

= 8SH]
T FN C F (m V NEG + m V POS)
. . ( Sv
--

= 2.0 (5.88 kN -m/m + 2.53 kN -m/m) 8 x 1.5 m] = 135 kN


[ 1.5 m

Note that from Table 8-3 the facing flexure pressure factor, CF , for a 100 mm thick temporary facing is
2.0. The vertical nominal unit moment resistances, mV'NEG (at nail end), and mv,POs (at mid-span location),
are computed using conventional reinforced concrete beam theory.

(b) Facing Punching Shear

The nominal internal punching shear strength of the facing is computed from Equation 8-2, where he and
D'e are as indicated below:

he = 100mm
D'e = bPL + he = 225 mm + 100 mm = 325 mm

The resulting nominal internal punching shear strength of the facing is computed to be:

VN = 0.33 J 28.0 MPa (,,;)(325 mm)(lOO mm) = 178 kN

From Table 8-3 the pressure factor for punching shear, Cs, for a 1oo-mm thick temporary facing is 2.5.
The punching cone bottom diameter, Dc, is equal to D'e + he = 425 mm. The diameter of the grQut
column is estimated to be about 200 mm. The corresponding areas are computed as follows:

Ac = 0.25(1t)(Def = 0.25(1t)(425 mm)2 = 1.42 x lOS mm2

Aue = 0.25(1t)(Doe)2 = 0.25(1t)(2oo mm)2 = 3.14 x 1()4 mm2

Substituting the above values into Equation 8-3, the nominal nail head strength for the punching shear
mode is computed as follows:

TFN = (178kN) [ 1 = 204 kN


1-2.5 (1.42 x 10smm 2 - 3.14 x 104mm 2)
[(1500mm)(l500mm) - 3.14 x 104mm 2]

8 -77
'.~ • OIl:'

· .-:-:
Do ..icI .. r . . ... 152 X 152 MW18.7 X MW 18.7
Grade 420 Welded-Wire Mesh
. ,
"
Permanent CIP Concrete
Facing (Minimum 28-Day .'
.\
4
Compressive Strength ~ . .> 2- No. 13 Vertical Bearing
=28 MPa) : ~.' :'., : Bars (Grade 420)
.'
'. Do'.·· "
· v.
. ,~ . Temporary Shotcrete
'.-: .r·.·;' .' Facing (Minimum
28-Day Compressive
Strength = 28 MPa)

31.8 X 9.5 Stud Head ----. 225 X 25 Steel Plate


(Grade 420) (Grade 250)

105

125 X 19 Stud
(Grade 420) · ';',' ....; .'.
'. .
~ '.
.. :.~'. ~ " v. , ~

· " . r' .. ;. "


2- No. 13 Waler Bars
. .",..::. ? ~
:
~ (Grade 420)
..
~' d:
.!
J
No. 13 @ 300 Each Way --~-r
~

(Grade 420)
.. -:
d
'4
J
· ~.' ..~. " .... ~ d

I...-- 200 - ...1.-100-.1

All dimensions in millimeters

Figure 8-29: Facing Details.

8 -78
Nail Head Strength of Permanent CIP Concrete Facing

(a) Facing Flexure

Following the same procedure presented above, the nominal head strength of the permanent facing under
the flexural failure mode can be calculated (using Equation 8-1) as follows:

T FN = 1.0 (17.4 kN-m/m + 17.4 kN-m/m) 8 x 1.5 m] = 278 kN


[ 1.5 m

Note that the facing flexure factor used in this case, CF, is 1.0 (for a 2oo-mm thick permanent facing, see
Table 8-3).

(b) Facing Punching Shear

The nominal internal punching shear strength of the facing is computed from Equation 8-2, where he and
0' e are as indicated below:

he = 25 mm + 125 mm = 150 mm
D'e = SHS + he = 105 mm + 150 mm = 255 mm

The resulting nominal internal punching shear strength of the facing is computed to be:

VN = 0.33 J 28 MPa (1t)(255 mm)(150 mm) = 210 kN

From Table 8-3 the pressure factor for punching shear for a 2oo-mm thick permanent facing is 1.0. The
punching cone bottom diameter, Dc, is equal to D'e + he = 405 mm. The diameter of the grout
column is estimated to be about 200 mm. The corresponding areas are computed as follows:

Ae = 0.25(1t)(Dd2 = 0.25(1t)(405 mm)2 = 1.29 x lOS mm2


AGe = 0.25(1t)(DGe )2 = 0.25(1t)(200 mm)2 = 3.14 x 104 mm2

Substituting the above values into Equation 8-3, the nominal nail head strength for the punching shear
mode is computed as follows:

1
= (210kN) 1 - - - - - - - - - - - - - - - = 219 kN
1-1.0 (1.29 x lOsmm 2 - 3.14 x 104mm 2)
[(1500mm)(1500mm) - 3.14 x 104mm 2]

(c) Headed Stud Tension

The nominal nail head strength associated with the headed-stud tension failure mode is computed by
Equation 8-4. Assuming the studs have a 22 mm diameter and an ultimate tensile stress of 420 MPa, the
nominal nail head strength is computed to be:

TFN = 4(0.25)(1t)(22.0 mm)2(420 MPa) = 639 kN

8 -79
TABLE 8-12
SHOTCRETE CONSTRUCTION FACING

Nominal Nail Allowable Nail Head


Failure Mode Head Strength Load (Group I)
TFN TF
(kN) (kN)

Facing Flexure 135 0.67(135) = 90


Facing Punching 204 0.67(204) = 137
C;:hp!lr

TABLE 8-13
PERMANENT FACING

Nominal Nail Allowable Nail Head


Failure Mode Head Strength Load (Group I)
TFN TF
(kN) (kN)

Facing Flexure 278 0.67(278) = 186


Facing Punching 219 0.67(219) = 147
Shear
Headed-Stud Tensile 639 0.67(639) = 320
.....

The allowable nail head loads for Load Group I are computed from the nominal strengths and summarized
in Tables 8-12 and 8-13. The results indicate that the allowable nail head load, TF , is governed by the
flexural failure mode of the temporary facing(i.e., TF = 90 kN).

Step 3 - Minimum Allowable Nail Head Service Load Check

For a friction angle of 34 degrees, a face slope angle of 10 degrees from vertical and a backslope angle
of 20 degrees, the active earth pressure coefficient, KA, corresponding to a horizontal, triangular earth
pressure distribution is equal to 0.257 (ignoring the cohesive component of soil strength).

Using Equation 8-5, the nail head service load can be determined as:

tF = 49 kN < 90 kN

(OK - the estimated nail head service load does not exceed the allowable nail head load.)

Step 4 - Define the Allowable Nail Load Support Diagrams

Develop the allowable nail load diagram for each nail by determining the allowable nail head load, the
allowable pullout resistance, and the allowable nail tendon tensile load.

8 - 80
Allowable Nail Head Load

From Step 2, the allowable nail head load is 90 kN.

Allowable Pullout Resistance (Ground-Grout Bond)

Q = aQ u
aQ = 0.50 (From Table 8-5 for AASHTO Group I loading condition)
Qu = 60 kN/m
Q = (0.50)(60 kN/m) = 30 kN/m

Allowable Nail Tendon Tensile Load

TN = aNTNN
aN = 0.55 (From Table 8-5 for AASHTO Group I loading condition)
TNN = ABFy = (510 mm2)(0.42 kN/mm2) = 214 kN
TN = (0.55)(214 kN) = 118 leN

Step 5 - Select Trial Nail Spacings and Lengths

In accordance with Figure 8-11, the nail length distribution for design purposes is shown in Figure 8-27.
The maximum nail length used is 7.7 meters (see Step 1 above). The specific nail length for each row is
listed in Table 8-14.

TABLE 8-14
NAIL LENGTH DISTRIBUTION

Nail Row No. Nail Length (m)


1 7.7
2 7.7
3 7.7
4 6.8
5 5.2
6 3.5

The above nail length distribution is obtained from Figure 8-11, as follows:

• The dimensionless nail pullout resistance, Qo, is calculated:

(0.50)(60 kN/m)
= --..:....----:...:....----=---- = 0.74
(18 kN/m 3 )(1.5 m)(1.5 m)

8 - 81
• The dimensionless nail length is:

LlH = (7.7 m)/(9.5 m) = 0.81

• The ratio of the dimensionless nail pullout resistance to the dimensionless nail length is:

= 0.74 = 0.91, giving an "R" factor of 0.31.


0.81

• Relative nail lengths are calculated from Figure 8-10 for the nail head elevations shown
in Figure 8-27 and an "R" value of 0.31.

The allowable nail load support diagrams are shown graphically in Figure 8-27, prepared in accordance
with the procedure previously presented in Figure 8-8.

Step 6 - Derme the Ultimate Soil Strengths

Ultimate Friction Angle, <l>u = 34.0 0


Ultimate Cohesion, Cu = 5 kN/m2

Step 7 - Calculate the Factor of Safety

For the trial cross-section and nail pattern shown in Figure 8-27, the soil strengths (i.e., <l>u = 34 0 , and
Cu = 5 kN/m2) and allowable nail load diagrams, a calculated minimum global factor of safety of 1.35
is obtained. The computations are based on the limit equilibrium procedure presented in Figure 8-7. The
minimum global factor of safety is obtained by analyzing a large number of potential slip surfaces. Because
of the extensive computations required, the analysis is performed most efficiently by a computer program.

The calculated minimum factor of safety (i.e., 1.35) meets the requirement for Group I loading for a
permanent, non-critical structure (see Table 8-5). For presentation purposes, the initially chosen "trial"
design for this example is in fact the design that meets the minimum required global factor of safety. In
a real world production design, some design re-runs and iterations would be required to arrive at an
acceptable design.

The results of stability assessments during construction are shown in Figure 8-30. Calculated factors of
safety for each stage of construction, following excavation of each lift and prior to installation of the
associated row of nails, are shown together with the corresponding critical slip surface in each case. It can
be seen that following installation of the first row of nails, the factor of safety during construction decreases
progressively as the height of the wall increases. However, the minimum calculated factor of safety is
1.25, which exceeds the minimum required value of 1.2 for temporary construction condition (see Table
8-5). Figure 8-30 indicates that prior to the installation of the first row of nails, the calculated factor of
safety is less than 1.2. Therefore, a test trenching is recommended in the field to verify the stability of the
initial cut.

Step 8 - External Stability Check of Nailed Block

The methods of analysis for external stability problems are the same as those used for conventional gravity
retaining structure. Work examples addressing this issue have been presented in Chapter 4. Computations
to check the external stability of the nailed block are therefore not presented.

8 - 82
Step 9 - Check the Upper Cantilever

The height of the upper cantilever above the top nail is identical (1.0 m) for both the temporary shotcrete
and permanent CIP wall facings. Therefore, the loading is identical in the two cases. Because both the
facing thickness and steel content are increased in the permanent facing, the permanent facing is less
critical by inspection. Therefore, only the upper cantilever of the construction facing needs to be
evaluated.

Shear Check

Compute the one-way unit shear force for the facing at the top row of nails:

= 2.3 kN/m

Compute the one-way unit shear strength of the facing based on Equation 8-49 of the Standard
Specifications for Highway Bridge (1992).

V NS = 0.166J f'c(MPa) (d) = 0.166"; 28.0MPa (0.05 m) = 43.9 kN/m

The allowable one-way unit shear is:

v = Q;F VNS = (0.67)(43.9 kN/m) = 29.4 kN/m > v (O.K.)

Flexure Check

Compute the one-way unit moment at the top row of nails:

h 1.0 m
ms = 0.33 ( ) v = (0.33)( )(2.3 kN/m) = 0.77 kN -m/m
coslO° coslO°

From Step 2, the nominal unit moment resistance in the vertical direction over the nail locations, mV,NEG'
is 5.88 kN-m/m. The allowable one-way unit moment for the upper cantilever is:

M = Q;F mV'NEG = (0.67)(5.88 kN-m/m) = 3.94 kN-m/m > ms (O.K.)


Step 10 - Check the Facing Reinforcement Details

Shotcrete Construction Facing

(a) Waler Reinforcement

Two No. 13 (soft metric designation - corresponding to standard bar size No.4) horizontal waler bars,
attached beneath the bearing place, will be placed continuously between nail heads in each nail row.
(O.K.)

8 - 83
8.0

/
/ / II'
I
I
/11
I
I
/
FS.. 2.20~ I I I
(Nalls 1-2) I / I
II
/ /I
/ / II
1.0
f
/ ~
/ ~

Slage 1 (2.0)
./
./ ./\'
/
4/
/;./
/
//

I//.;'-
/IJ/
H-9.5
~bL=7.7
00 Stsge 2 (3.5) ././ .1
./

Slage 3 (5.0) NOTES:

All dimensions In meters


All allowable nail loads In kN
1.5
Stsge 4 (8.5)
Horizontal and vertical
nail spacings =1.5m

Stage 5 (8.0)
o Nail number

eL=5.2

Figure 8-30: Construction Stability.


(b) Minimum Cover Requirements

For the temporary shotcrete construction facing, corrosion is not a concern and placing the reinforcement
at the center of the facing is adequate.

(c) Development of No. 13 Vertical Bearing Bars

For a uniformly loaded interior span with fixed-end supports, the point of zero moment theoretically occurs
at 0.213Lc from the support. The vertical bearing bars are not needed all the way to the point of zero
moment. However, per section 8.24 of AASHTO, they must extend the maximum of Lc/20, 15dB, or d,
past the point at which they are no longer needed.

L c/20 = (1500 mm)/20 = 75mm


lSdB = 15(12.7 mm) = 191 mm
d = (100 mm)/2 = 50mm

Section 8.24.3.3 of AASHTO provides additional requirements for embedments beyond points of
inflection. The only requirement that is potentially more critical than the above is an embedment at least
equal to:

L c/16 = (1500 mm)/16 = 93.8mm

This is less critical than the 15dB = 191 mm, computed above. Therefore, use a minimum development
length of 200 mm for the vertical bars.

Pennanent facing

(a) Waler Reinforcement

There are no applicable requirements.

(b) Minimum Reinforcement Ratios

Per Section 8.20 of AASHTO, the minimum required amount of shrinkage and temperature reinforcement
near exposed surfaces of walls and slabs is 265 mm2 per lineal meter. The No. 13 bars on 3OD-mm centers
provides (129 mm2)/(0.3 m) = 430 mm2 per meter, which is adequate.

(c) Minimum Cover Requirements

Per Section 8.22 of AASHTO, the minimum cover on the front side of the facing is specified at 50 mm.
Based on the design illustrated on Figure 8-29, the cover to the headed studs is as indicated below:

te = 200 mm - tpL - 4ts = 200 mm - 25 mm - 125 mm = 50 mm (O.K.)

The minimum required cover between the permanent facing reinforcing steel and the CIP
concrete/temporary shotcrete interface is 38 mm. Based on Figure 8-29, this cover is:

te = 100 mm -12.7 mm = 87.3 mm > 38 mm (O.K.)

8 - 85
For corrosion protection purposes, there must also be a minimum 75 mrn of cover between the facing steel
and the soil. Based on Figure 8-29, the 100-mrn thick temporary shotcrete provides adequate cover for
the permanent facing steel.

Step 11 - Serviceability Checks

Shotcrete Construction Facing

Because of the temporary nature of the wall, the serviceability requirements are not applicable to the
construction facing.

Permanent Facing

(a) Upper Cantilever Serviceability Check - Reinforcement Distribution

The reinforcement must be distributed properly such that the steel stress does not exceed an allowable
value to control the service crack widths. Based on procedure outlined in Section 8.16.8 of AASHTO,
the service steel stress is computed to be about 18 MPa and the allowable stress is computed to be about
200 MPa. Therefore, the reinforcement distribution is adequate for this case. Detailed calculations are
not presented for this example problem. Interested readers should refer to the FHWA Manual for Design
and Construction Monitoring of Soil Nail Walls (Byrne, et aI., 1996).

(b) Overall Displacement of the Wall

The construction-induced vertical and horizontal permanent displacements at the top of the wall can be
expected to be on the order of 0.2 % of the height of the wall, or about 20 to 25 mrn, for the given site soil
conditions (medium dense silty sands). Displacement can be anticipated to decrease back from the wall
in general accordance with the recommendations given in Figure 8-12.

Final Design Section

To simplify for construction as well as to add additional safety margin for global stability, the fInal design
cross-section of the soil nail wall is shown in Figure 8-31. It should be noted that the fInal design nail
lengths are uniform, independent of the nail length distribution calculated in Step 5 above. The nails in
the bottom three rows could be shortened, if desired, by constructability consideration and/or for added
cost savings.

8 - 86
10------,6 m l - - - - - -

Geocomposite Strip Drain


1.0m
1 Level 1
/soilNaii
Soil Nail Wall

._---
Level 2

Level 3
9.5m

Level 4

LevelS
T
1.5 m
(typ.)
-.L Level6
t
1.0m

7.7 rn

~
Note: All dimensions in
meters.

Figure 8-31: Final Design Section.

8 - 87
CHAPTER 9
CHEMICALLY STABILIZED EARTH WALLS

9.1 INTRODUCTION

Chemically stabilized earth (CSE) wall systems are relatively new in the United States. They are
constructed by mixing in situ soils with chemical grouts at depth to form overlapping soil-grout elements
(usually columns) that constitute the wall. The chemicals usually used are cement or lime, with or without
additives, applied to the soil in a liquid state.

The CSE walls can be designed as gravity structures or as externally-supported structural walls.
Reinforcement can be inserted into the wall elements before the soil-grout mix hardens. The resulting
structural combination can resist lateral earth pressures, and when the wall is properly supported laterally
(e.g., by ground anchors) it can support excavations of considerable depth. The overlapped columns can
also be used as a cut-off wall for seepage control.

The jet-grouted wall and the deep-soil-mix wall are the two primary types of CSE walls. These walls
commonly use cement for grout. A variation of these walls is the lime column wall wherein lime or lime-
cement mix is used in a dry or liquid form to stabilize the soil.

It should be noted that the CSE walls represent an emerging technology that requires special considerations
uncommon to the average geotechnical engineer. The successful application of this technology is largely
dependent on the use of specialists experienced in the design and construction aspects of these systems.

The objectives of this chapter are to introduce the most common CSE wall systems, discuss their
advantages and limitations, identify typical applications, describe construction methods and outline design
procedures.

9.2 JET-GROUTED WALL

9.2.1 Introduction

Jet grouting consists of injecting high pressure fluids into the ground through horizontal nozzles to
segregate the soil and mix it with a cementing agent. Segregation of the soil is achieved by the high energy
of the jet fluid(s) which may consist of the cementing agent or another cutting fluid. Jet grouting is
basically an erosion/replacement process which removes a portion of the soil particles and replaces them
with a mixture of soil and grout that has high strength and low permeability when hardened.

The jet grouted elements, which make up the wall, are typically either overlapping cylindrical columns or,
in some applications, panel elements. Cylindrical column elements are the most common and are formed
by rotation of the high pressure fluid jets. Panel elements are formed by allowing the jet to remain
stationary, thus cutting and mixing planar jet grouted elements.

In this chapter, three jet-grouting systems are introduced with their advantages and limitations. The
construction equipment, materials and procedures used for each system are discussed in detail. Because
the jet-grouted wall's performance is heavily dependent on the materials used, and construction procedures
employed, only a few analytical issues are presented. Instead, the parameters which affect the wall's
performance are discussed in detail and a step-by-step design procedure is outlined. A detailed quality
control/quality assurance program is also identified.

9-1
9.2.2 Wall Configurations and Applications

Figure 9-1 shows various grout column layouts used in construction of jet-grouted walls. The column
layout is governed by the intended use of the wall as well as stability considerations. For example, in
cases where water tightness is the main concern, as in the case of cutoff walls, then multiple rows of
overlapping columns should be used. Multiple rows are also required if the wall is to be designed as a
gravity structure. Figure 9-2 illustrates some jet-grouted wall applications.

9.2.3 Advantages and Limitations

Advantages

• Jet grouting can be used with a wider range of soil types than any other grouting technique (Figure
9-3). The system, however, is not suitable for highly plastic clays.
• Since the soil is partially or completely removed, its in situ permeability does not dictate grout
acceptance. Accordingly, the quantities of grout and, therefore, the costs can be more closely
predicted.
• Since large diameter columns can be created from relatively small boreholes, local obstructions
such as timber piles or large boulders can be bypassed or encapsulated into the jet-grouted soil
mass.
• Jet grouting can be conducted from any suitable access point, and can be terminated at any
elevation, providing treatment only in the target zones.
• It can be performed vertically or (sub)horizontally, above or below the water table.
• It can provide high strength relative to other methods of ground treatment.

Limitations

• It is an operator-sensitive construction process, which requires specialized skills and experienced


labor force.
• It requires specialized mechanical equipment subject to high maintenance costs.
• Due to the high pressure used, there is a possibility of ground heave or lateral movements which
may damage adjacent utilities and underground structures.
• Spoil handling and removal may be particularly difficult if the jet-grouted soil is contaminated.
• There is a general lack of a cost-effective method for measuring the dimensions of a large number
of columns in a routine production project.
• Jet-grouted soil strengths tend to be much more variable than concrete strength as they are strongly
influenced by the silt and clay content of the native soils. In addition, it is difficult to predict the
fInal strength of the jet-grouted soil during the design stage. Furthermore, if the groundwater flow
velocities are high, the fluid soil-grout mass may experience local removal of the cement (bleeding)
prior to its stiffening, and hence unevenness in quality may be observed.
• It requires a higher level of quality control than normal.
• There is no commonly accepted method available to engineers to design jet grouted structures. So
far, the design has been dependent to a great extent on experience and intuition.

9-2
:; :
:: .' .

.. "

:.'.'.
.
............
.. :.:
, .. , , .. ,
.....
:': ::.:
.' . , ...

Jet Grouted Column ~

/
/
1:1::;:::::'1::::::/::::;:::::':::;::::::::.::::;::::::::::::::::::: .
Injection Pipe _.._._ _••/
/

Figu~~ 9-1: . 'G~o~t: c6i~~~~t~y~J~:~:

V'",loE:--+1 J
r. PE _--\I1:;jol'J~ Utilities
::' i:
;,

I I ~
r~_"c.~~I-><",~')ll;<..J~ 1C)(~i.>{:::::;X:."::::7 ~~
\ ~ Diaphragm ~
Jet Grout Columns Wall
~ ~ 00
Figure 9-2: Jet-grouted Wall Applications for Excavation Support, Underpinning, Settlement Control
and Water Control.
9-3
Slit Sand Gravel
100% .,/
Q) //
I
>
"~
en / Chemical Grouting
0) f.....I - - - - - - - - 1 I - - - - - - - -...J
c:
"00 {
en 50% :
ctS
i
-...
0..

u
c:
Q)

/
;

,I
i
Jet Grouting

Q) .'
0.. I
../
0
0.002 0.06 2.0 60
Grain Size (mm)

Figure 9-3: Range of Soil Types Treatable by Chemical and Jet Grouting. (After Welsh, 1992)

9.2.4 Wall Construction

Jet Grouting Procedure

The basic procedure for jet grouting is presented in Figure 9-4 and summarized below:

a. Drill and stabilize a hole to the required depth in the soil to be treated.

b: If separate drilling and grouting equipment are used, withdraw the drilling bit and rods and insert
the jet grouting monitor and special grouting rods to the bottom of the predrilled hole. In many jet
grouting systems and ground conditions, however, the same equipment is used to drill the hole and
then perform the jet grouting. In these cases, after the hole is drilled to the required depth, a steel
ball (check ball) is inserted into the drill rods to redirect the fluid jets to the horizontal nozzles and
allow jet grouting to begin (see Figure 9-7).

c. Inject jet grouting fluids while slowly rotating and withdrawing the grout monitor and rods. The
grouting fluids cut the in situ soil and mix it with the grout to form a jet grouted column. Excess
soil-grout spoil is expelled out of the top of the drilled hole by the circulating fluids.

Jet Grouting Systems

There are basically three jet grouting systems in general use: the single, the double and the triple fluid
systems. A schematic representation of each system is shown in Figure 9-5. The principal characteristics
of each system are briefly presented below (Kauschinger and Welsh, 1989; Bruce, 1994).

9-4
Figure 9-4: Jet Grouting Procedure. (Pacchiosi, 1985)

(a)

Air Air
Grout Water
Air Air

Grout

(b) (c)

Figure 9-5: Schematic of Jet Grouting Systems: (a) Single Fluid System, (b) Double Fluid System, and
(c) Triple Fluid System.

9-5
Single Fluid System

• It is the simplest of the three systems.

• It uses a grout jet to simultaneously erode the soil and mix it with the cement grout.

• It involves only partial replacement of the soil and therefore results in the lowest volume of spoils
as compared to other systems.

• It produces the smallest column diameter.

• It can achieve significant compaction outside of the perimeter of the column up to a distance of
one-half a column diameter.

• It is the only method used for horizontal column formation.

Double Fluid System

• It uses a grout jet engulfed in compressed air. The compressed air enhances the cutting ability of
the grout jet.

• The equipment is more complex and susceptible to clogging. The pathway for the air between the
inner rod (carrying the grout) and the outer rod must be kept open, or the process will revert to
a single fluid system.

• A higher degree of soil replacement and a larger column diameter than that of the single fluid
system (almost twice the size) can be created. This enhancement is due to the following factors:

a. The air acts as a buffer between the jet stream and any groundwater present, thus permitting
deeper penetration by the jet.

b. The soil cut by the jet is prevented from falling back onto the jet, thus reducing the energy
lost through the turbulent action of the cut soil.

c. The cut soil is more efficiently removed from the region of jetting by the bUbbling action
of the compressed air.

• A potential drawback of this system is that the soil-grout mix may have a higher air content, and
therefore may have a lower strength than those of the other systems.

Triple Fluid System

• Fluids are emitted from two levels. An upper water jet engulfed in compressed air is used to
excavate the soil which is then mixed with, or replaced by, a grout jet emitted from a lower port.

• It is the most complicated of the three systems.

• It permits virtually full replacement of the jetted soil, and provides the largest column diameter.
This also results in the largest amount of spoils/cuttings.

• In this system, unlike the double fluid system, the grout is not injected with air. Hence, there is
no problem with high air content in the [mal jet-grouted soil mass.

9-6
• Fluid circulation is more efficient because each medium is carried in a separate passageway within
the rod (the grout in the central core, the water in the middle annulus, and the air in the outer
annulus). Hence, this system is the least likely of the three systems to exert high pressures on the
surrounding soils which may cause ground heave or lateral movement.

Hydrofracture Potential

In general, a jet grouting system conducts incompressible fluids and its success is contingent on maintaining
free and efficient fluid circulation. In particular, the cuttings must flow freely from the point of injection
up to the ground surface. Otherwise, pressures, up to the jetting pressures (30-55 MPa), may build up in
the soil. These pressures could hydrofracture the soil and cause severe lateral soil movement and ground
heave (Kauschinger and Welsh, 1989).

Jet Grouting Equipment

A typical jet grouting set up consists of: (a) a drilling rig; (b) an automated grout mixing plant; and (c)
a grout injection plant, which consists of automatic batchers and high-pressure pumps. In multiple fluid
systems, additional pumps and an air compressor are used (Figure 9-6). Photos 9-1 through 9-4 show the
jet grouting equipment in operation.

The drill rig automatically regulates rotation and withdrawal rates of a string of special drill rods and a jet
grouting monitor which is mounted at the end of the drill string. The jet grouting monitor is a special tool
through which the jet grouting fluids are passed and directed out of the nozzles and into the ground. Figure
9-7 illustrates details of jet grouting monitors for the single and triple fluid systems. Photo 9-5 shows the
triple rods and the monitor for a triple fluid jet grouting system.

The mode of drilling is selected according to the soil conditions. Rotary drilling is most commonly used.
In coarse-grained soils, which may include cobbles and boulders, rotary percussion is more suitable. The
mast length of the drilling rig is an important consideration on construction sites with overhead
obstructions.

Controlled jet grouting creates a spoil material during the erosion and mixing process. The volume of soil-
cement spoil can be predicted from the injected volumes. The spoil usually contains a significant cement
content and gains strength over time. Within 12 hours, it can typically be handled as a firm to stiff clay
and can be used as a construction material or be carried away from the site for disposal. If jet grouting
is used in contaminated ground, special handling procedures may be required.

Jet Grouting Materials

Neat, rapid setting, cement grout is typically used for jet grouting, although chemical grouts can also be
used. Grout viscosity and rigidity should be low to allow maximum penetration. Portland Cement Types
I, II or III are used. Type I is the most economical and is, therefore, used when possible.

Water/cement ratios of 1: 1 to 2: 1 are commonly used. Where high strength is required, ratios as low as
0.6: 1 are applied. Potable water is normally used. Although not recommended, salt water is sometimes
allowed, provided no steel reinforcement is used.

Fly ash is sometimes added to the cement grout in ratios of cement:fly ash between 1: 1 and 1: 10 by weight.
Where low permeability is needed, bentonite additives are often used. The addition of 2 percent of
bentonite to the grout mix can reduce shrinkage during curing. No aggregates are added to the grout mix.

9-7
GROUT
WATER
AIR

2 3 4

s-jL ---"' -=c- -- ,.

1) Silo
2) Mixing
3) Molor pump
4) Drill
-
5) Air Compressor

Figure 9-6: Jet Grout Set Up

cement mix
rods

-compressed air

water
coaxial nozzles

nozzle

check

vallie seat -cement mix

nozzle

bit

L--.J
(a) (b)

Figure 9-7: Details of Jet Grouting Monitors: (a) Single Fluid System, and (b) Triple Fluid System

9-8
Photo 9-3 Photo 9-4

Photo 9-1: Grout Mixing and Injection Plant.


Photo 9-2: Drilling and Grouting Rig.
Photo 9-3: Fanning a Jet-Grouted Column.
Photo 9-4: Jet Grouting Monitor.

9-9
Photo 9-5: Rods for a triple fluid grouting system.

This component is provided by the treated ground. Steel reinforcement can be used with jet grouting,
where the grouted columns are used as a shear key, or if they require high bending resistance. A high
tensile threaded bar, a Dwidag anchor bar or a steel tube can be used for reinforcement. Grade 420 steel
is usually used.

9.2.5 Wall Design


Introduction

The jet-grouted earth wall is designed based on conventional wall design procedures. The width of the
wall is usually defmed in terms of column diameter, configuration of overlapped columns, and minimum
overlapping length. Jetting parameters, which achieve the required width, strength and permeability of
the wall, are then established based on existing soil and groundwater conditions. The characteristics of the
jet-grouted soil (strength, permeability, etc.) are influenced by the properties of the in situ material, the
composition of the grout mix, and a number of operating parameters, such as injection pressure, flow rate,
withdrawal rate (lifting speed), rotation rate, etc.

Factors Affecting Design

The design of a jet-grouted column is influenced by a number of interdependent variables related to in situ
soil conditions, materials used, and operating parameters. Table 9-1 presents a summary of the principal
variables of the jet grouting system and their potential impact on the three basic design aspects of the jet-
grouted wall: column diameter, strength and permeability. Table 9-2 gives typical ranges of operating
parameters and results achieved by the three basic injection systems of jet grouting. It should be noted,
however, that the grout pressures indicated in this table are based on certain equipment and can vary. This
table can be used in feasibility studies and preliminary design of jet-grouted wall systems. The actual
operating parameters used in production are usually determined from initial field trials performed at the
beginning of construction.

Design Procedure

Due to the large number of variables involved in the jet grouting process, it is very difficult to establish
a rigorous design process. A typical design procedure is dependent to a great extent on the field and
laboratory test results, the specific equipment and method used, and the experience of the equipment
operator. The following steps are followed in the design of a jet-grouted wall:

9 - 10
1. Establish earth pressure diagrams using procedures described in Chapter 2. For determining the
lateral earth pressures, the jet-grouted wall may be considered as a gravity structure due to its large
mass. For thinner, externally supported and reinforced jet-grouted walls subject to bending, the
earth pressures acting on the wall may be similar to those of a slurry diaphragm wall (Chapter 7).

2. Perform stability and wall seepage analyses and determine wall thickness based on assumed
strength, permeability, column diameter and overlapping length. Use Table 9-2 in establishing
these assumptions.

3. Conduct a field trial, coupled with laboratory tests, to confirm the design assumptions and select
the parameters to use for production.

Field Trial

A field trial is considered an essential element of any jet grouting project. Its basic steps are as follows:

• Select a test site with ground conditions similar to those of the constructed project.

• Design a field trial to model actual construction. In general, it is necessary to (a) perform the field
trial with the same equipment to be used in the actual construction, and (b) provide adequate
instrumentation to study pore pressures, heave and lateral deformation of the surrounding ground.

• Construct a number of jet-grouted test columns to determine optimum combinations of the various
operational parameters and determine the column diameter and jet-grouted soil strength that can
be achieved in the field.

• Evaluate the quality of the jet grouted test section. This is usually done by: (a) excavating and
exposing columns to study the wall continuity; (b) drilling inclined cores to estimate column
diameter; (c) testing the strength of the cores obtained along the centerline and near the column
perimeters; and (d) performing field permeability tests on completed columns or a pumping test
on a completed cut-off enclosure.

9.2.6 Quality Control/Quality Assurance

Basic QC/QA Issues

To ensure the quality of the completed wall, it is critical to exercise strict control and monitoring during
the construction stage. Such control and monitoring will also provide data to verify the design criteria and
to modify, if necessary, the jet grouting parameters.

Depending on the scale of the project, the following controls can be exercised (Bruce, 1994):

• Alignment survey of the drilled holes.


• Assessment of the column diameter and verticality by drilling, inclinometers, or a special
verticality-measuring device introduced into the drilling rod.

• Monitoring the vertical movements on the ground surface or adjacent structures.

• Monitoring the horizontal displacements (e.g., by inclinometers) of the surrounding ground.

• Quantitative and qualitative evaluation of the ejected materials. This is particularly important
for estimating the quantity of cement refuse and checking the actual composition of the treated
ground.

9 - 11
TABLE 9-1
SUMMARY OF SYSTEM VARIABLES AND THEIR IMPACT
ON BASIC DESIGN ELEMENTS

Principal Variables General Effect of the Variable on Basic Design Elements


(Strength, Permeability and Column Diameter)

(a) Jet-Grouted Soil Stren2tb


Degree of mixing of soil and Strength is higher and less variable for higher degree of mixing.
grout
Soil type and gradation Sands and gravels tend to produce stronger material while clays and silts
tend to produce weaker material.
Cement Factor Strength increases with an increase in cement factor (weight of cement
per volume of iet-.e;routed mass.)
Water/cement ratio of grouted Strength of the jet-grouted soil mass decreases with increase in in situ
mass water/cement ratio.
Jet grouting system The strength of the double fluid system may be reduced due to air
entrapment in the soil-.e;rout mix.
Grout composition Richer .e;routs produce hi.e;her strengths than leaner grouts.
Age of grouted mass As the jet-grouted soil mass cures, the strength increases but usually at a
slower rate than that of concrete.
(b) Wall Permeability
Wall continuity Overall permeability of a jet grout wall is almost entirely contingent on the
continuity of the wall between adjacent columns or panels. Plumb,
overlapping multiple rows of columns would produce lower overall
permeability. In case of obstructions (boulders, utilities, etc.) if complete
encapsulations is not achieved then overall permeability may be reduced
due to possible leaka.e;e alon.e; the obstruction-.e;rout interfaces.
Grout composition Assuming complete wall continuity and complete replacement of in situ
soil, the lowest permeability which can be obtained is that of the grout
(typically 10-6 to 10-7 cm/sec). Lower permeabilites may be possible if
bentonite or similar wateroroofm.e; additive is used.
Soil composition If complete replacement is obtained (as may be possible with a triple
fluid system) then soil composition does not matter. Otherwise, if
uniform mixing is achieved then finer grained soils would produce lower
permeabilities as compared to granular soils.
(c) Column Diameter
Jet grouting system The diameter of the completed column increases in size as the number of
fluids is increased from the sin2le to the triple fluid systems.
Soil density and gradation As density increases, column diameter reduces. For granular soils, the
diameter increases with reducin.e; uniformity coefficient (DdD,n).
Degree of mixing of soil and Larger and more uniform diameters are possible with higher degree of
.e;rout mixin.e; .

9 - 12
TABLE 9-2
TYPICAL RANGE OF JET GROUTING PARAMETERS AND
JET-GROUTED SOIL PROPERTIES
(Adapted from Kauschinger and Welsh, 1989)

Parameter Units Single Fluid Double Fluid Triple Fluid


~lecf Ion Pressure
(a ) In·
Water jet MPa Not used Not used 30-55
Grout jet MPa 30-55 30-55 1-4*
Compressed air MPa Not used 0.7-1.7 0.7-1.7
(b) Flow Rates
Water jet liters/min Not used Not used 70-100
Grout Jet liters/min 60-150 100-150 150-250
Compressed air m3/min Not used 1-3 1-3
(c) Nozzles Sizes and Number
Water jet mm Not used Not used 1.8-2.6
Grout jet mm 1.8-3.0 2.4-3.4 3.5-6.0
Number of water jets - Not used Not used 1-2
Number of 2routiets - 2-6 1-2 1
(d) Cement
Grout w/c ratio - 0.8 - 2.1 0.8 - 2.1 0.8 -2.1
3
Cement Consumption k2/m 400-1000 150-550 150-650
()DriI1RdR
e 0 otation and Lifi·
tine S~Deed
Rotation speed rpm 10-30 10-30 3-8
Liftin2 Speed minim 3-8 3-10 10-25
(1) Column Diameter
Coarse-grained soil m 0.5-1 1-2 1.5-3
Fine-2rained soil m 0.4-0.8 1-1.5 1-2
C2) Jet-Grouted SoH Streneth
Sandy soil MPa 10-30 7.5-15 10-20
Clayey soil MPa 1.5-10 1.5-5 1.5-7.5

* Higher grout pressures of 10-18 MPa are used with certain equipment where large separation exists
between the water/air nozzle and the grout nozzle.

9 - 13
In addition to the above controls during construction, the following options may be considered where
applicable (Gallavresi, 1992):

• Coring to provide laboratory specimens to verify physical properties such as strength and
permeability. This is highly recommended.
• Direct examination by exposing completed columns.
• Penetration tests (such as standard penetration test or cone penetration test) and pressuremeter tests
to evaluate the bearing capacities of intercolumn soils.
• Sonic testing to evaluate the improvement of the mechanical properties, uniformity of the treatment
and the possible overlapping of adjacent columns. In practice, vertical pipes are placed at
preestablished spacings through which the transmitter and receiver are lowered.

Electronic Data Control

Several approaches based on field and laboratory experimentation, mass balance equations and energy
considerations have been proposed to monitor construction quality (Kauschinger, 1992). A monitoring
system, PAPERJET, reported by DePaoli et ai., (1991) appears to be the most sophisticated. This system
allows electronic measurement of the various jet grouting parameters; continuously evaluates various
energies expended in the jet grouting process; and uses a real-time electronic data control system capable
of operational decisions.

The electronic data control system enables the site engineers to continuously monitor the various
operational parameters both in graphical and numerical form and immediately identify and rectify any
malfunctions such as drop in pressure or flow rate, or clogging of the nozzles. The site engineer can
identify any change in the soil condition and can accordingly modify the jet grouting process.
Furthermore, the owner can monitor the execution of the work, confirming that it has been carried out
correctly and within design limits. Finally, such a system permits the use of the data stored magnetically
for preparation of graphical outputs showing the variation of the measured quantities with depth.

Responsibility of Field Personnel

Since there are no standard design procedures for jet-grouted works, quality control testing during the field
trial and production grouting is critical to the success of the project. One of the key elements of quality
control is to have experienced, qualified individuals from both the contractor and owner on site. In
addition, the equipment and procedures should be suitable for the site and the requirements of the project.

The first step in quality control involves the education of the field personnel in the details of the project,
the jet grouting process to be used, the jet grouting parameters and the design requirements. All field staff
should understand exactly what jet grouting can do, how the system works, and what data must be
recorded. They shoul~ also be familiar with the project's specifications and approved shop drawings.

The next phase is the actual implementation of the jet grouting work. At this phase, critical information
and observations must be made by the field personnel to ensure that the jet grouting procedures and
parameters are correctly applied for the project. Also, data collected by the quality control staff can be
used to identify potential problems or deficiencies in the jet-grouted wall, which may require remediation
prior to any excavation. Typical items which should be observed and monitored are usually included in a
check list prepared specifically for the project.

9 - 14
9.3 DEEP SOIL MIXING WALL

9.3.1 Introduction

The deep-soil-mixing (DSM) wall is a system of overlapped soil-cement columns fonned by mixing the in
situ soil with cement grout at depth. This wall system is also referred to as the Soil-Mix-Wall (SMW) and
the Soil-Cement-Column (SCC) wall.

The DSM system differs from jet grouting in two ways. First, the equipment does not use high pressure
grout. Second, the DSM columns are built to defined dimensions, while the jet-grouted columns are
fonned, under high pressure, to uncertain dimensions.

This section discusses the advantages and limitations of the DSM system and its use in civil engineering
applications. The construction equipment, material, installation procedures and operational parameters of
the system are discussed. A design approach is presented with recommended procedures for detennining
the factors that affect the design such as strength, penneability and modulus of elasticity. Other design
considerations and QC/QA issues are also discussed.

9.3.2 Wall Configuration and Applications

DSM columns can be arranged in a variety of patterns as shown in Figure 9-8. Wall applications include
containment/cutoff walls and structural retaining walls, as illustrated in Figure 9-9.

9.3.3 Advantages and Limitations

Advantages

• The DSM wall can be built in a broad range of soils including soft to very stiff and low to highly
plastic clay and silt, loose to dense sand, gravel and cobble, and soft rock.
• The process generates a smaller volume of spoils because most of the in situ soil is used as
construction material and remains underground.
• It has minimal environmental concerns since the soil is treated in place and at depth.
• A relatively short construction time and high production rate can be anticipated because the method
involves in situ materials and mechanized operations.
• The wall configuration is well defined, and the continuity between the columns is assured.
• Since the construction involves a drilling process, it will not create significant noise and vibration,
and the impact on adjacent facilities is minimal.
Limitations
• Since the wall relies on use of in situ soil as a construction material, all obstructions such as rubble,
pieces of concrete, abandoned pipes and boulders must be completely removed and replaced with
suitable soil. Otherwise, pre-boring with special drill bits may have to be used to partially loosen
and/or break up these obstructions prior to installation of the wall.
• The equipment and procedures used may not be easily amenable to variation in column geometry
with depth (e.g., the column cannot be wrapped around utilities). Therefore, relocation of utilities
may be required.

9 - 15
(a)

(b)
reinforc:ina _ _r

(c)

(d)

Figure 9-8: Plan View of Typical DSM Wall Layouts: (a) Cut-off Wall, (b) and (c) Excavation-Support
Wall, and (d) Lattice Pattern for Liquefaction Control.

Highway Highway

II

DSMWalls

Highway

Stress
ReinforcementS-.
Soil cement

(a) (b)

Figure 9-9: Various Wall Applications: (a) as Containment/Cutoff Wall, and (b) as Structural
Retaining Wall.

9 - 16
• Until the soil-cement is hardened, the constructed column may constitute a weak spot that may
trigger movement of adjacent structures or utilities. This may require underpinning or temporary
protection works.

9.3.4 Wall Construction


Basic Construction Methodology
Multiple hollow-stem augers equipped with mixing paddles penetrate the ground to the required depth and
back. During penetration and withdrawal, cement grout is pumped through the auger stem. The auger
flights and mixing paddles mix the soil and the grout in place to form continuous overlapping soil-cement
columns that constitute the wall (Figure 9-10). If needed, steel reinforcement is inserted in the DSM
column to provide bending resistance (Figure 9-8).
DSM Equipment
Photos 9-6 and 9-7 show typical DSM equipment. A typical DSM system consists of a mixing plant and
a drilling/mixing unit. The mixing plant consists of a grout mixer, a grout agitator, automatic batching
scales to control grout composition, and a computer for mixing and grout flow control. The drilling/mixing
unit consists of multiple axis augers guided by a vertical steel lead on a track-mounted base machine as
shown in Figure 9-10. Two to five augers can be used, with auger diameters varying from 550 to 1016
mm. The base machine, together with the lead, are supported at three points during operation for
maintaining accurate vertical alignment which is critical for eliminating unmixed zones between columns
and maintaining wall continuity.

The typical auger has an auger head, discontinuous auger flights and mixing paddles. The auger flights
are positioned so that they overlap with each other to form overlapped soil-cement columns. The
discontinuous auger flight is designed to provide some vertical displacement of the soil for mixing, but also
to prevent transporting the soil to the surface. Thus, the auger mixes the grout with the soil at its original
depth, uniformly and continuously.

In recently developed equipment, jetting or spreadable mixing tools are used at the tip of the auger to
enhance the grout penetration and increase the column diameter at specific depths.

DSM Installation Procedures

DSM walls are installed by constructing a series of sets or elements. Typically, a set consists of three
overlapping columns as shown in Figure 9-11. The stepped installation procedure, which is commonly
used, drills two primary column sets followed by drilling a secondary column set using two boundary
columns of the primary column sets as guide holes to construct a continuous wall as shown in Figure 9-8.
The redrilling in this procedure increases the uniformity of the soil-grout mix. The redrilling ratio can be
adjusted according to the required level of soil-mixing.

During the drilling process, grout is injected into the soil through the tip of the hollow-stem auger. A
separate positive displacement pump supplies the grout to each of the injecting augers for accurate control
of grout flow. The auger flights penetrate and loosen the soil, and lift it to the mixing paddles which blend
the grout with the soil. As the auger continues to advance, the soil and the grout are remixed by additional
paddles attached to the shaft. About 60 to 80 percent of the slurry is injected as the augers penetrate
downward and the remainder is injected as they are withdrawn so that the mixing process is repeated on
the way up. If reinforcing elements are to be used, they are inserted into the soil-cement column
immediately after the auger is withdrawn.

9 - 17
N
N
ClO

(a) (b) (c)

Figure 9-10: DSM Wall Equipment. (a) Mixing Shaft for General Use, (b) Mixing Shaft for Soil with
Boulders, and (c) Mixing Shaft for Cohesive Soil. (Taki and Yang, 1991)

9 - 18
Photo 9-6

Photos 9-6 and 9-7: DSM Equipment


Photo 9-6: Drilling and Mixing Unit.
Photo 9-7: Mixing Plant.

9 - 19
Figure 9-11: Installation Procedure. (Taki and Yang, 1991)

DSM Operational Parameters

There are three principal DSM operational parameters as follows:

1. Drilling Speed: The drilling speed is governed by the properties of the soil to be treated, and the
soil mixing effort required to obtain the design soil-cement properties.

2. Mix Design: The DSM wall is usually composed of three basic materials: the soil, the grout and
the reinforcement, if any. The most suitable soil and grout mixes are similar to those discussed
under jet grouting. Portland cement is commonly used. A small amount of bentonite is sometimes
added to increase the workability of the soil-mixing work. Cement-based additives such as silicate,
slag and gypsum are used for gaining strength in saline or organic soils, or for stabilizing
contaminated soils.

The mix design is governed by the required engineering properties of the DSM wall such as
strength and permeability. The type of the soil essentially determines the extent of improvement
in engineering properties. Laboratory strength and permeability tests are normally performed to
identify the cement proportion that could provide the required properties.

9 - 20
The final selection of a mix design is influenced by the selection of equipment and installation
procedures used, efficiency and economy. In the field, an automated batching system measures the
water, cement and other additives by weight to produce a more reliable grout than that produced
by a volumetric batch system. The desired weight of each grout component can be entered in the
computer at a control panel, and changes to the mix design can be made by simply adjusting the
component weights at the control panel.

3. Grout flow rates: The grout flow rate is usually adjusted constantly to accommodate varying drill
speeds in different soil strata, so that the design volume of grout per unit volume of in situ soil is
maintained. The grout flow is electronically controlled.

9.3.5 Wall Design

General

The design of a DSM wall is influenced, to a certain extent, by its planned function. When used as a cut-
off wall, for instance, the wall is designed to provide a required impermeability. When used for excavation
support, on the other hand, vertical reinforcing members are usually provided to resist bending moments
as well as shear stresses along the longitudinal direction of the wall. Between the reinforcements, the soil-
cement structure is designed to resist and redistribute the horizontal stresses to neighboring reinforcing
members. When used for liquefaction control, the DSM wall is installed in block or lattice pattern (Figure
9-8) to resist the stresses from embankment or structure loadings, should the embankment or foundation
soil liquefy during seismic ground shaking. The lattice structure of the wall is designed so that the excess
'pore pressure build up in the liquefiable soil is reduced to an acceptable level.

Factors Affecting Design

Strength, permeability and modulus of elasticity are the primary engineering properties of concern for the
DSM wall. These properties are affected by the soil type, cement and other admixture proportions,
water/cement ratio of the grout, degree of soil-cement mixing, curing environment and age. Based on
actual wall projects in various soil types, Taki and Yang (1989, 1991) presented design guidelines, many
of which are included in the following sections.

Strength

The strength is commonly expressed in terms of unconfmed compressive strength, quo It is primarily
influenced by the soil type and cement content. In general, an increase in cement content increases the
unconfined compressive strength. Higher strength increases are observed for sandy or gravelly soils as
compared to cohesive soils where the strength gain with increasing cement content may be minimal.

For a given cement proportion, the strength increases with age. The 28-day strength is approximately
twice the 7-day strength for both sandy and clayey soils, i.e.,

~(28) = 2 qU(7) (9-1)

For design purposes, one third of the unconfined compressive strength can be considered as the shear
strength, 'tf ,of the soil-cement, i.e.:

1
'tf = '3 QU(28)
(9-2)

9 - 21
Permeability

For a well mixed soil-cement, the coefficient of permeability ranges from 10-5 to 10-7 cm/sec. In general,
the coefficient of permeability and the porosity of the soil-cement matrix decrease with decreasing the sand
content and the water-cement ratio, and increasing the curing age.

It should be noted, however, that the above indicated coefficients of permeability are those of the soil-
cement mix and do not necessarily represent the overall permeability of the DSM wall. This permeability
is influenced by the tightness of the wall and the potential presence of "windows" which may be caused
by boulders or column deviations. The overall permeability of the DSM wall is best determined by
performing in situ permeability tests in the completed wall.

Modulus of Elasticity

For design purposes, the following formula can be used:

E50 = oc QU(28) (9-3)

where: E50 = Modulus of elasticity measured at 50 percent ultimate strength


qu(28) = 28-day unconfmed compressive strength
ex: = Coefficient of elasticity of the soil-cement ranging between 500 and 1350
Design Considerations

The structural design of DSM walls includes: (a) the design of reinforcing members to resist bending
moments, shear stresses and deflections along the height of the wall; (b) the design of soil-cement elements
to resist and redistribute the lateral pressures to the reinforcing members; and (c) the determination of the
minimum depth of embedment required for base stability or vertical load support.

The design of reinforcing members and the required depth of embedment are carried out as in the case of
conventional soldier pile and lagging walls (Chapter 7). The lateral earth pressures are similar to those
discussed in jet-grouted walls and externally supported structured walls.

The design of the soil-cement lagging proceeds by considering it as a horizontal member supported on two
adjacent reinforcing members. If a reinforcing member is installed in every column, as shown in Figure
9-12(a), it is only necessary to consider the punch-through shear force Q in calculating the shear stresses.
Where a reinforcement is not installed in every soil-cement column, the soil-cement element may be
analyzed using the hypothetical model shown in Figure 9-12(b). In this case, in addition to the punch-
through shear stresses, the analysis must consider compressive stresses along a hypothetical parabolic arch
with a configuration as shown. In both cases, although the resistance to the punching shear (Q) is
calculated at the point of maximum shear along the edge of the reinforcing member (H-pile), it should also
be checked at the necking point between columns. Although this point has lesser shear, it also has the
minimum section which may result in the lowest safety factor.

An empirical design guideline for spacing reinforcing members to avoid bending failure in the soil-cement
element is presented in Figure 9-12(c). This guideline was based on [mite element simulation and full-scale
model tests to study various failure modes of soil-cement between reinforcing members (Taki and Yang,
1989,1991).

9 - 22
HIOJ"I"'J'J; w
I

I'Q I I
~~I _'_ _

L.J
(a)

(b)

o h

If L2 ~ 0 + h -2e. no bendi.ng failure

(c)

Figure 9-12: DSM Wall Design: (a) Analysis for Punch-Through Shear; (b) Analysis for Compressive
Action of Arching Effects; (c) Empirical Guideline for Avoiding Bending Failure. (Taki
and Yang, 1989, 1991)

9 - 23
9.3.6 Quality Control/Quality Assurance

A laboratory testing program is performed to provide parameters required in the DSM semiempirical
design. Unconfmed compression tests are usually performed. Direct shear and triaxial compression tests
may also be performed to provide a direct evaluation of the shear strength of the soil-cement mix. In
addition to the structural design requirements, permeability tests are also performed if the wall is designed
to serve as a cutoff barrier. Prior to construction, these tests are performed on samples prepared in the
laboratory using in situ soil mixed with cement.

In order to verify the design parameters, field sampling is mandatory during the construction of the wall.
Wet soil-cement samples should be obtained routinely and cured in the laboratory for testing and quality
control. After construction, core samples should be obtained from the exposed wall according to a testing
schedule to be determined based on the site conditions observed during excavation. Usually, the core
samples are tested after 7 and 28 days to establish the strength increase with time.

It is recommended that for the fmished DSM wall, direct shear tests and/or triaxial compression tests be
conducted for strength assessment, in addition to the unconfined compression tests. For quality control
during construction, however, the unconfined compression test is adequate, and the results may be used
as standard values.

Monitoring of equipment and operational procedures is a key element of the QC/QA program. The
number of revolutions of the mixing paddles per unit volume of in situ soil is usually used in monitoring
the soil mixing effort. The drilling time spent in installation of each DSM wall panel is recorded and used
as part of the QC/QA program. The grout injection rate is also monitored and adjusted constantly during
production to accommodate varying drilling speeds in different soil strata, so that the design volume of
grout per unit volume of column is maintained.

As with the jet-grouted wall, the successful completion of a DSM project is heavily dependent on the
qualifications, experience and behavior of the contractor's personnel during all phases of construction.

9 - 24
CHAPTER 10.0
WALL SELECTION

10.1 INTRODUCTION

The earth-retaining structure used for highway construction is usually a part of a comprehensive highway
project. Depending on the contracting policy adopted for the project, the responsibility for selecting the
wall type may be with the owner, the consulting engineer representing the owner, or the contractor.

Because of the many systems available on the market and in the interest of economy, alternative systems
are sometimes selected and designed for the project. These alternative designs are usually made in one
of three ways, (a) as a design task performed by the design engineer, (b) as the result of a value
engineering study performed during design or construction, or (c) as an alternative design proposed by
the contractor.

The objective of this chapter is to develop a systematic approach for wall selection considering many
factors that may affect the implementation and performance of the wall. Ten selection factors are studied
in detail and a selection process is proposed. A wall selection example, which includes a detailed selection
matrix, is presented to demonstrate the selection process. Ten case studies are discussed in detail, with
a particular emphasis placed on wall selection.

10.2 FACTORS AFFECTING SELECTION

With the variety of systems available on the market, building an earth-retaining structure should not be a
difficult task. With the rapid influx of new techniques, however, and the many restrictions imposed on
construction, it is becoming more difficult for the design engineer to determine which of the available
systems is the most suitable for the project. To do so, a thorough evaluation of many factors involving the
design, construction, use and maintenance of each system is required. This evaluation also allows the
engineer to avoid a system that may appear to be acceptable at first,. but proves to be unsuitable after all
the evaluation factors are considered.

The ten most important factors affecting wall selection are:


• The ground
• Groundwater
• Construction considerations
• Geometry and right-of-way requirements
• Aesthetics
• Environmental concerns
• Durability and maintenance requirements
• Cost
• Tradition
• Contracting Practice

Following is a brief summary of each.

10.2.1 The Ground

An earth-retaining structure is obviously influenced by the earth it is designed to retain, and the one on
which it rests. The influence of the earth is particularly important in "earth walls" where the retained earth

10 - 1
itself has a major load-carrying function. In mechanically stabilized earth (MSE) walls, for instance, which
usually involve some sort of reinforcement, the pull-out force in the reinforcement is resisted by (1) the
friction along the soil-reinforcement interface and (2) the passive resistance along the transverse members
of the reinforcement, if any (grid reinforcement). Therefore, these systems are best suited for soils with
high internal friction such as sands and gravels.

Strain compatibility is another element affecting the design of an earth wall. In systems using inextensible
soil-reinforcing elements, such as steel strips, the strains required to mobilize the full strength of the
reinforcing elements, are much smaller than those needed to mobilize the full strength of the soil. For
extensible reinforcing materials such as geotextiles, the required strains are much larger. Therefore,
relatively large internal deformations usually occur in these systems and the soil properties are measured
at large strains (residual strength). Obviously, these systems are less compatible with soils of relatively
low residual strength.

When in situ reinforcement is used to support excavations, such as in soil-nail walls, the possible saturation
and creep of the in situ soil can have a large negative impact on the long-term performance of the system.
Therefore, these systems are less suited for clayey soils than for granular materials. For the same reason,
the anchored wall also is not frequently used in these soils.

Gravity-type structures are less influenced by the type of soil than the systems involving soil reinforcement.
For soils with large vertical and horizontal deformations, a very flexible system such as a gabion wall may
be chosen in lieu of a more rigid system that attempts to resist such deformations.

In chemically-stabilized earth walls, the compatibility of the ground and the structure depends largely on
the type of chemicals used, and hence on the specific system employed. While deep soil mixing walls, for
instance, are normally used in sandy and liquefiable soils, the lime columns walls are mostly suitable for
use in deep clay deposits rich in pozzolans. In jet-grouted walls, the in situ properties and structure of the
soil are not as important since the concept of jet grouting is to break down the soil structure and replace
it with a self supporting composite mass of soil and grout sometimes called soilcrete. The strength of the
soilcrete, however, and the permeability of the wall may be influenced by certain soil elements such as peat
or boulders.

10.2.2 Groundwater

Generally, the groundwater table behind an earth-retaining structure is lowered to: (1) reduce the
hydrostatic pressure acting on the structure, (2) reduce the likelihood of corrosion of any metal reinforcing
and facing elements used in the system, and (3) prevent saturation of the soil which may significantly
increase displacements and cause instability during or after excavation, particularly in systems involving
staged excavation and support such as soil nailing. To reduce the negative impact of groundwater, a free-
draining system such as a MSE wall can be used. Alternatively, for relatively watertight structures (soil
nailing, element wall, etc.), an appropriate groundwater drainage system is provided.

Sometimes, it is desirable to keep the water table high to prevent settlement of adjacent stru~tures or
protect existing untreated timber pile foundations from fungus decay due to exposure to oxygen. In these
cases, a relatively rigid watertight structure is used (slurry wall, tangent piles, jet-grouted wall, etc.).
These structures usually are designed to support the full hydrostatic pressure.

10 - 2
10.2.3 Construction Considerations

Construction schedule, availability of material, site accessibility, equipment availability and labor
considerations are important factors affecting the selection of an earth-retaining structure. Construction of
a Doublewal or similar prefabricated modular system, for instance, is fast and simple. With a crane and
a crew of four, a construction rate of about 200 square meters per day may be accomplished. When the
site is inaccessible to heavy equipment, such as in rough mountainous terrain, a system that can be installed
with a minimum of equipment (geotextile wall) is preferred. On the other hand, labor-intensive systems
usually are not cost-effective in areas with strong labor union requirements.

Material availability is an important factor affecting the selection of an earth-retaining structure. Where
rock is abundant, a gabion wall may be suitable and economical. This may not be the case, however,
where suitable rock backfill has to be hauled a long distance to the project site. Where suitable aggregate
is not available, structures built of concrete usually are avoided.

A major construction-related advantage of earth walls, as compared to conventional cast-in-place concrete


structures, is that no form work is required on the job, and construction can be accomplished regardless
of temperature. The behavior of some walls (chemically stabilized walls, soil-nail walls, bin walls),
however, may be influenced by the soil temperature during and after construction. The temperature at the
site, for instance, may affect the hydration of the cement injet grouting. Also, repeated freeze-thaw cycles
may soften up frost-susceptible in situ or backfill soils and cause ground deformations that may have
negative impacts on the structure. Hunt, et al. (1994) discusses the failure of a bin wall and attributes it
to the freeze-thaw behavior of the backfill material.

10.2.4 Right-of-Way and Space Requirements

In MSE walls, a relatively large space is required behind the structure face as compared to that needed for
construction of conventional walls (the length of the reinforcing elements is typically 0.7 times the wall
height).

To support an excavation in a very tight space, a top-down staged excavation and support system such as
soil nailing may be the most suitable. The feasibility of such a structure, however, is influenced by the
presence of utilities and buried structures nearby and the additional cost of permanent underground
easement for placement of the reinforcing elements.

Site congestion may be a drawback for some systems such as slurry diaphragm walls. When low headroom
does not allow the operation of conventional construction equipment, walls which can be implemented in
a limited operating space (soil nails) or from a remote operating area Get grouting) are preferred.

10.2.5 Aesthetics

In addition to being functional and economical, the permanent earth-retaining structures, in most cases,
have to be aesthetically pleasing. Different types, shapes and color facings are used in construction of
earth walls. The types of facings range from built on-site continuous facings (shotcrete, welded wire mesh,
cast-in-place concrete) to prefabricated concrete or steel panels. Cast-in-place facing and precast panels
usually are more attractive than shotcrete or soldier pile and lagging walls. In permanent bored-pile walls,
an architectural wall facing usually is provided, but at an additional cost to the structure.

10 - 3
The aesthetic factor is extremely important when building a retaining structure in parks, forests and natural
habitat. A number of attractive wall systems (Criblock, Evergreen, etc.) are usually considered for those
areas because of their aesthetic, acoustic and anti-graffiti advantages. The Evergreen wall, for instance,
consists of precast concrete units with open spaces at the face into which are planted shrubs, vines, etc.
With adequate water supply for the foliage, the concrete facing will no longer be visible a few years after
construction.

10.2.6 Environmental Concerns

Like most structures, the selection of an earth-retaining system is influenced by its potential environmental
impact during and after construction. Excavation and disposal of contaminated material at the project site,
and discharge of large quantities of water or slurry fluids (jet grouting, slurry diaphragm walls) are of
primary concern. MSE walls which allow construction of roadway embankments with vertical sides to
minimize encroachment on wetlands have positive environmental and ecological benefits. The
environmental advantages of these walls can be further enhanced by the use of wood chips, and other waste
materials in construction.

To reduce noise and vibration impacts, the systems which use pile driving or heavy construction machinery
may be rejected. On a recent excavation support project adjacent to the Johns Hopkins University hospital
in Baltimore, Maryland, for instance, an extremely tight vibration control policy was adopted due to the
presence of sensitive laboratory optical equipment nearby.

To reduce traffic noise in environmentally-sensitive areas, the gravity-type gabion and Evergreen walls
offer specific advantages. The open nature of the face and the presence of foliage covering are effective
in absorbing the noise hitting their facings, making these walls acoustically superior as compared to other
earth-retaining structures where the traffic noise is reflected on hard or smooth continuous surfaces.

10.2.7 Durability and Maintenance

An earth-retaining structure built of concrete has a higher durability against corrosion and weathering
effects than a structure constructed of metal, or which uses metal or synthetics for reinforcement and/or
facing. Corrosion of the reinforcement, for instance, is one of the major design issues of MSE walls that
use metal reinforcement. Although, these walls are sometimes avoided in higher than normal corrosive
environments, corrosion-protection measures can be provided if necessary, but at an added cost to the
structure. Gabion walls have durability concerns similar to those of MSE walls with metal reinforcement.

When geosynthetics are used for reinforcement, their long-term creep behavior and resistance to
deterioration due to chemical attack and exposure to ultraviolet light are major considerations that have
to be addressed in the design. Selection of the types of geosynthetic is sometimes dictated by durability.

The durability of a concrete structure (gravity wall, slurry diaphragm wall, etc.) is influenced by the
quality of the aggregates and water used in the mix, and by the casting procedures. As indicated before,
concrete walls are not recommended in areas where quality aggregates are not economically available.

The durability factor is extremely important when selecting a maintenance-free earth-retaining structure
in highly corrosive surroundings, or when the structure is subjected to attack by non-conventional elements
such as waves, chemicals or marine borers.

10 - 4
10.2.8 Cost

The total cost of an earth-retaining system has many components including the structure, right-of-way,
temporary or permanent easement, excavation and disposal of unsuitable material, drainage, etc.

The construction costs of specific types of retaining structures have been discussed in previous chapters
and are summarized in Tables 10-1 and 10-2. It should be noted, however, that a structure with the least
construction cost does not necessarily mean an economical alternative, as the ultimate cost of the system
is influenced by many indirect cost factors such as those listed above as well as schedule, permitting and
maintenance requirements.

10.2.9 Tradition

Tradition may dictate, or prevent, the use of a certain-type of structure irrespective of its technical rating.
Although earth walls are very popular in certain states, for instance, they are rarely built in others. While
still considered novelty in certain parts of the U.S.A., slurry diaphragm walls are heavily used in
construction of underground facilities in other areas such as Boston. Where drilled caissons are
traditionally used as deep foundations, bored-pile walls are popular and generally economical since the
local contractors are equipped for, and experienced with, that type of construction. Tradition plays a
greater role in construction in underdeveloped countries.

10.2.10 Contracting Practice

The contracting policies and procedures followed in the United States may discourage or even preclude
the use of certain types of walls, particularly those involving patented equipment, materials or procedures.
Contracting issues will be discussed in more detail in Chapter 11.

10.3 THE SELECTION PROCESS

The selection process consists of three steps:

• Identification of project requirements and site constraints


• Identification of appropriate wall alternatives
• Qualitative rating of suitable alternatives

10.3.1 Identification of Project Requirements and Site Constraints

The first step in the selection process is to identify the project requirements such as permanency of the
wall, construction time table, available budget, "buy-America" guidelines, aesthetic requirements,
maintenance policy, etc., and site constraints such as right-of-way limitations, site accessibility, presence
of utilities and obstructions, presence of wetlands, environmental requirements, restrictive weather periods,
etc.

It should be noted that based on the initial site evaluation, a decision should be made regarding the need
for an earth retaining system at all. For instance, if sufficient right-of-way is available, it may be feasible
and economical to construct stable soil slopes and, thus, eliminate the need for a retaining structure.

10 - 5
TABLE 10-1 SYSTEM SELECTION CHART FOR FILL WALLS. (after Sabatini, et. al., 1997)
Wall Type Perm. Temp. Cost Cost in $ Required Differential Advantages Disadvantages
Effective per m2 ROWe) Settlement
Height ··ofwall Tolerance e)
Ran2e face (I)
Concrete .I 1- 3 m 270 - 370 0.5 - 0.7H'"' 1/500 • durable • deep foundation support may be necessary
gravity wall • requires smaller quantity of select backfill as compared • relatively long construction time
to MSEwalls
• concrete can meet aesthetic reQuirements
Concrete .I 2-9 m 270 - 650 0.4 - 0.7H(4) 1/500 • durable • deep foundation support may be necessary
cantilever wall • requires smaller quantity of select backfill as compared • relatively long construction time
to MSEwalls
• concrete can meet aesthetic reQuirements
Concrete .I 9 -18 m 270 - 650 0.4 - 0.7H(4) l/5oo • durable • deep foundation support may be necessary
counterfort wall • requires smaller quantity of select backfill as compared • relatively long construction time
to MSE walls
• concrete can meet aesthetic reQuirements
Conrete crib .I .I 2 - Il m 270 - 380 0.5 - 0.7H lI300 • does not require skilled labor or specialized equipment • difficult to make height adjustments in field
wall • rapid construction
Metal bin wall .I .I 2 -Il m 270 - 380 0.5 - 0.7H l/3oo • does not require skilled labor or specialized equipment • difficult to make height adjustments in field
• rapid construction • subiect to corrosion in ae:e:ressive environment
Precast .I .I 2 -7 m 270 - 380 0.5 - 0.7H 1/300 • does not require skilled labor or specialized equipment • difficult to make height adjustments in field
...... Concrete ·blocks are easily handled
o Module Wall
0'1 Gabion wall .I .I 2-8m 270 - 540 0.5 - 0.7H l/50 • does not require skilled labor or specialized equipment • need adequate source of stone
• construction of wall requires sie:nificant labor
MSEwall .I .I 3 -20 m 240 - 380 0.7 - LOH l/Ioo • does not require skilled labor or specialized equipment • requires use of select backfill
(precast facing) • flexibility in choice of facing • subject to corrosion in aggressive environment
(metallic reinforcement)
MSEwall .I .I 2 -7 m 175 - 275 0.7 - LOH 1/200 • does not require skilled labor or specialized equipment • requires use of select backfill
(modular block • flexibility in choice of facing • subject to corrosion in aggressive environment
facing) • blocks are easily handled (metallic reinforcement)
• positive reinforcement connection to blocks is
difficult to achieve
MSE wall (geo- .I .I 2 - 15 m 165 - 380 0.7 - 1.0H 1/60 • does not require skilled labor or specialized equipment • facing may not be aesthetically pleasing
textilel geogridl • flexibility in choice of facing • geosynthetic reinforcement is subject to
welded wire degradation in some environments
facing)
Reinforced Soil .I .I 3 - 30 m 80 - 260 0.5 - 1.0H l/6O • does not require skilled labor or specialized equipment • facing may not be aesthetically pleasing
Slopes (RSS) • flexible in choice of facing • geosynthetic reinforcement is subject to
• vegetation provides ultraviolet light protection to degradation in some environments
geosynthetic reinforcement • vegetated soil face requires significant
maintenance

Notes: (1) Total installed costs in 1995 U.S. dollars.


(2) ROW requirements expressed as the distance (as a fraction of wall height, H) behind the wall face where fill placement is generally
required for flat backfill conditions, except where noted.
TABLE 10-2 SYSTEM SELECTION CHART FOR CUT WALLS. (after Sabatini, et. al., 1997)
Wall Type Perm. Temp Cost Effective Cost in $ per Required Lateral Water Advantages Disadvantages
Height m2 0fwall ROW (5) Movements Tightness
Range face CI )
Sheet-pile ~ ~ up t05 m 160 - 430 None (6) large fair • rapid construction • difficult to construct in hard ground or through
wall 6 to 11 m(2) • readilv available obstructions
Soldier pilei ~ ~ upt05m 110 - 380 None") medium poor • rapid construction • difficult to maintain vertical tolerances in hard
lagging wall 6 to 24 m (2) • soldier beams can be drilled or ground
driven • Dotential for I!round loss at excavated face
Slurry ~ ~ 6-24m(2) 650 - 930 None(6) small good • can be constructed in all soil types or • requires specialty contractor
(diaphragm) weathered rock • significant spoil for disposal
wall • watertight • requires specialized equipment
• wide range of wall stiffness
Tangent pile ~ ~ 3-9m 430 - 810 None(6) small fair • adaptable to irregular layout • difficult to maintain vertical tolerances in hard
wall 6 - 24 m(2) • can control wall stiffness ground
• requires specialized equipment
• silmificant sDoil for disDosal
Secant pile ~ ~ 3 -9 m 430 - 810 None") small fair • adaptable to irregular layout • requires specialized equipment
wall 6 - 24 m(2) • can control wall stiffness • sil!nificant sDoil for disDOsal
Soil mixed ~ ~ 6 - 24 m(2) 430 - 590 None(6) small fair • adaptable to irregular layout • requires specialized equipment
wall • relativelv small bendinl! caDacitv

-
o
-...J
Anchored wall

Soil-nailed
wall
~

~
~

~
5-20m(J)

3 - 20 m
160 - 810

160 - 600
0.6H +
anchor bond
lenl!th
0.6 - I.OH
small-
medium

small-
medium
N/A

N/A
• can resist large horizontal pressures
• adaptable to varying site conditions

• rapid construction
• requires skilled labor and specialized
equipment
• anchors mav reouire Dermanent easements.
• nails may require permanent easements
• adaptable to irregular wall alignment • difficult to construct and design below water
table
Micro-pile ~ ~ N/A 3,200 - Varies medium N/A • may not require excavation • requires specialty contractor
wall 9800(4)

Notes: (1) Total installed costs in 1995 U.S. dollars.


(2) Height range given is for wall with bracing or anchors.
(3) For soldier pile and lagging wall only.
(4) Cost per linear meter of wall.
(5) ROW requirements expressed as the distance (as a fraction of wall height, H) behind the wall face
where wall anchorage components (Le., ground anchors and soil nails) are installed.
(6) ROW required if wall includes tieback anchors.

* Fill walls can also be considered in cut situations. However, temporary excavation support and additional earthwork costs need to be added to the fill
wall costs as shown on Table 10-1
10.3.2 Identification of Appropriate Wall Alternatives

The second step in the selection process is to perform an initial screening of wall alternatives to identify
those that are not feasible nor practicable considering the project requirements and site constraints
identified above. Tables 10-1 and 10-2, which summarize general characteristics of fill and cut walls, can
assist in the initial screening. The potentially suitable alternatives are then identified for further evaluation
and qualitative rating.

10.3.3 Qualitative Rating of Alternatives

Once the feasible alternatives are identified, they can be rated with respect to some or all of the selection
factors discussed in Section 10.2. The more suitable alternatives are then evaluated based on cost, and the
one with the lowest cost is usually selected for the project.

A selection matrix can be used for a qualitative evaluation of the wall alternatives. Based on each
evaluation factor, a qualitative rating between 1 and 4 can be given for each alternative with the highest
number representing the best rating. Each qualitative rating is usually multiplied by a weighting factor
reflecting the importance of that particular evaluation factor for the project at hand. For instance, a wall
to be built in a national park for the purpose of future leisure activities may require a higher weighting
factor for environmental impact than for construction schedule. The trend would be reversed, however,
if the wall is to be constructed in a remote barren land with little environmental concerns, but with a very
tight completion schedule dictated by severe weather conditions.

The weighting factors, which differ from one project to another, are normally established jointly by the
owner and the designer of the facility. Usually, cost and durability are given higher weights than the rest.
After the evaluation matrix is completed, the top rated alternative(s) are selected for detailed analysis and
cost estimate. A sample evaluation matrix is shown in the following wall selection example.

Figure 10-1 is a flow chart illustrating the wall selection process discussed above.

10.4 WALL SELECTION EXAMPLE

To build a road embankment adjacent to a creek, an earth-retaining wall is needed to allow construction
of the embankment with a near vertical side slope and avoid encroachment on the wetland. The function
of the road is to provide temporary access to remote areas while a permanent highway is built nearby. The
embankment and its retaining wall will be designed for a temporary life of 5 years, but with minimum
maintenance requirement.

The site lies in a U.S. national forest, thus, the constructed project has to meet strict environmental and
aesthetic requirements. Based on the roadway profile and the site topography, the required wall height
varies along the alignment, reaching 10 m in certain sections. The soil consists of medium dense silty
sand, with zones of soft compressible silty clays that may cause long-term differential settlement problems.

Initial screening of all possible wall alternatives is performed using Table 10-1. Because of the required
wall height and potential differential settlement, the concrete walls identified in the table can be quickly
eliminated from further consideration. Two concrete-type walls, however, are still evaluated: a cast-in-
place concrete cantilever wall and a concrete module wall -- the first to satisfy traditional construction
practice, and the second to meet the short construction period allowed for the project.

10 - 8
Identify Project Requirements

I
Identify Appropriate Wall Alternatives
If FILL, use Table 10-1
If CUT, use Table 10-2

I
Evaluate 10 Selection Factors

I
Prepare Selection Matrix Using
Qualitative Rating

I
Identify Most Suitable Wall Type or
Alternative Wall Types

Figure 10-1 Wall Selection Flow Chart

The qualitative rating matrix prepared for the project is illustrated in Table 10-3. Weighting factors
between 1 and 3 are established for the project with the highest weight of 3 given to the ground, the cost
and the environmental impact factors. Although minimum maintenance requirement is specified for the
project, a weighting factor of 2 is used for durability because of the temporary nature of the wall.

Based on each selection factor, a qualitative rating of 1 to 4 is given to each wall alternative. Based on
that, the initial rating is multiplied by the appropriate weighting factor to determine the comparative rating
of that alternative. Both, the initial and weighted ratings are shown in the matrix (Table 10-3). The total
score ranges from as low as 47 for a metal bin wall to as high as 63 for a geotextile MSE wall. After the
qualitative rating is completed, preliminary cost estimates are performed for the three top ranked
alternatives: the geotextile wall, the reinforced soil wall, and the gabion wall. The geotextile wall is
selected for the project.

10 - 9
TABLE 10-3
SELECTION MATRIX

Project Selection Factor Total


Score
Weighting Factors 3 1 2 2 2 3 2 3 1 1

Wall Alternative Ground Ground- Construction Right- Aesthetics Environment Durability Cost Contracting Tradition
water of- Wav
Reinforced Soil Wall 2*/6** 3/3 4/8 2/4 3/6 3/9 3/6 3/9 4/4 3/3 58

Geotextile Wall 3/9 2/2 4/8 2/4 3/6 3/9 3/6 4/12 4/4 3/3 63

..... Gabion Wall 2/6 3/3 3/6 4/8 3/6 3/9 3/6 3/9 3/3 2/2 58
o
.....
o
I

Bin Wall 2/6 2/2 3/6 3/6 2/4 2/6 3/6 2/6 3/3 2/2 47

Concrete Module Wall 3/9 2/2 3/6 3/6 2/4 2/6 4/8 3/9 3/3 4/4 57

Cast-in-Place Concrete 3/9 2/2 3/6 3/6 2/4 2/6 4/8 1/3 4/4 4/4 52

* Initial Rating (1 to 4)
** Weighted Rating
10.5 CASE STUDIES

Following are 10 case studies of non-conventional earth retaining structures designed by Parsons
Brinckerhoff Quade & Douglas, Inc. In all of those projects, conventional and non-conventional
alternatives were considered. Although the design and construction aspects of these structures are briefly
discussed, a greater emphasis is placed on their selection process which is the basic theme of this chapter.
These case histories were previously published in a comprehensive paper on the selection, design and
performance of earth retaining structures published by ASCE (Munfakh, 1990).

10.5.1 Case 1 • Aviation Corridor Highway, Tucson, Arizona

The Aviation Corridor Highway (SR 210) in Tucson, Arizona had an extensive requirement for earth-
retaining structures, mostly in the downtown area where the roadway is in a depressed section. Retaining
walls were also required in the fill areas, but to a much lesser extent.

Following a qualitative evaluation of a number of alternatives, a preliminary design and cost study was
performed on the wall systems believed to be the most appropriate for the soil conditions and construction
practice in the Tucson area.

Two wall configurations were considered: the single wall system and the double wall system (stepped-back
wall). The double wall system was recommended by the project's landscape architect to mitigate the
canyon effect of high parallel walls on each side of the roadway.

The single walls studied were: (a) cast-in-place retaining walls with spread footings, (b) cantilever drilled
shaft retaining walls, (c) circular drilled shafts with ground anchors and (d) soil-nail walls (Figure 10-2).
The double wall alternatives consisted of combinations of the single wall systems (Figure 10-3).

The typical soil profile used in the design consisted of cohesionless medium dense to dense sand,
characterized by an angle of internal friction of 36 and a total unit weight of 2185 kg/m3 to a depth below
0

the base of the wall equal to the wall height. This soil profile is found at the locations of the majority of
cut walls within the project limits. "Non-typical" soil conditions were also analyzed at a few wall locations
where cohesive clayey soils were present beneath the wall, or cemented sands behind it. The cemented
sands were treated as a cohesive material with an undrained shear strength of 100 kPa. The clay was
assigned an undrained shear strength of 75 kPa.

The cast-in-place concrete walls were designed using classical soil mechanics theory. The cantilever
drilled shaft walls were designed as soldier pile and lagging but using drilled piers instead of the
conventional steel H-piles. Movements in the drilled shafts and wall deflections were calculated by treating
the drilled shafts as laterally-loaded piles (using the program COM 624 developed by Reese and Sullivan
at the University of Texas). The anchored drilled shaft walls were analyzed using the U.S. Army Corps
of Engineers computer program CSHTWAL and the FHWA handbook on permanent anchors (Cheney,
1984). A reinforced drilled shaft, 600 mm in diameter, was sufficient for all cases. However, for walls
9 m or higher, a 900 mm diameter was used to reduce the required embedment and the wall deformation.

The soil-nail walls were analyzed for internal and external stability according to procedures presented in
Mitchell and Villet (1987). A typical spacing of 1.5 m was considered for the 25 mm steel bar nails. The
nail length was governed by the external stability of the composite soil-nail structure.

10-11
Q)i
=,
CI

~
c.

"
; .. '0-

c~al
WI , , Drilled Shaft

--
I
I ( -,
457 mm -- ~ ,.
' .• /.-.

Cast-in-Place Wall Cantilever Drilled Shaft Wall


ROW ROW
610 mm -:-0.. • Permanent Easement 762mm
I

E .. ----===----:.
.-IH-- 305 mm
-'
"1

e --- ----------
:x:
-
"1
",
1.5 m (Typ.)

---------
I E
"1
-- - ------------
---- ---
----- ---- ---~----------;.::.
e
"1

-----
-----::.-:,-----
-----:·::.::::::3
762mm

x
Drilled Shaft Wall with Anchors Soil Nail Wall
Figure 10-2: Single Wall Alternatives Considered for Aviation Corridor Highway. (Munfakh, 1990)

10 - 12
Permanent Permanent Construction
ROW Easement Easement Easement
ROW

2.1 m

::J:

...
::J:

." .

Two Cast-in-Place Walls Drilled Shaft & Cast-in-Place Walls

Construction Easement l ·
Permanent Easement . ROW
ROW 914 mm """ 610 mm

:--
C\I - ==. -:. ---- -----------..: _... .:. -- :: :)
::J: C\I
::J:
::J: 1--_ _2_,.:..1.::.;m~_-l< ::J:

I-.----:. :=.-.--~ :. : :.: :: : :- .",=-~


... - - -- - -•-- .....J

C\I ::J:
::J:
-: ==:=.'.'.'.~".' -'.'. )1
I
30Smm
L.
Drilled Shaft
I. x+H,

Two Drilled Shaft Walls Two Soil Nail Walls

Figure 10-3: Double Wall Alternatives Considered for Aviation Corridor Highway, (Munfakh, 1990)

10-13
Double cantilevered drilled shaft walls were selected for all areas where the cut depth is greater than 3 m.
These structures were the least costly among the double wall configurations dictated by the aesthetic
requirements of the landscape architect. Their estimated construction costs were from $305 to $338 per
square meter (1988 cost) for wall heights of 4.5 m to 9 m. Where the depth of cut was 3 m or less, a
single cantilevered drilled shaft was selected.

Although soil-nail walls appeared to be the most economical for heights less than 3 m, the tight right-of-
way limitations in the downtown area, and the difficulty in obtaining permanent easements to the tips of
the nails, prohibited their use. For the few areas where soil nailing was feasible, it was not believed to be
cost effective to require the contractor to mobilize specialized equipment needed for this type of
construction. The equipment required for construction of drilled shaft walls, on the other hand, was
readily available at the site for construction of bridge foundations. Unfortunately, due to funding reasons,
the construction of the project was put on hold indefinitely.

Selection of the earth-retaining structures for this project was based on the following factors: cost,
aesthetics, right-of-way, construction requirements and tradition.

10.5.2 Case 2 - H-3 Tunnel Access Roads, Hawaii

Building of Interstate Route H-3 through the Koolau Mountain Range of the Hawaiian Island of Oahu
involved construction of a major tunnel. To reach the remote tunnel portal location during construction,
access roads having extensive retaining walls were required on both sides of the mountain. Since the
retaining walls were to be constructed in mountainous terrain with difficult accessibility, alternatives
requiring heavy machinery were ruled out, and the wall selection concentrated on three most promising
alternatives -- an inextensible MSE wall, a gabion wall and a geotextile wall. To minimize cost, all three
alternatives were designed for the North Halawa Valley side and the prospective bidders were requested,
but not required, to bid on all three. All walls were required to have a minimum service life of 10 years
and to be resistant to the moderately to highly acidic in situ soils.

Figure 104 presents typical cross-sections of the three alternatives. The gabion wall, which was designed
as a gravity structure, consisted of steel wire baskets filled with good quality gravel or crushed rock
resistant to degradation and chemical deterioration. This type of rock, however, was difficult to fmd in
Oahu. Due to the temporary nature of the walls, cold rolled galvanized steel face panels were specified for
the inextensible MSE wall alternative instead of the conventional but more expensive precast concrete
panels. For the geotextile wall alternative, woven or non-woven sheets of polypropylene were required to
be placed between layers of compacted granular backfill. The fabric sheets were wrapped upward then
overlapped for anchorage at the wall face.

Design of the geotextile walls is described in detail by Castelli and Munfakh (1986). The fabric spacing
was based on the ultimate tensile strength of the fabric assuming a safety factor of 1.5. Both internal and
external stabilities were considered in determining the required fabric width. The external stability of the
reinforced soil system governed the design, requiring a minimum width to height ratio of 0.6. For depths
less than 1.5 m below the top of the wall, an overlap length of 1.4 m was calculated. For greater depths,
a minimum length of 0.9 m was used. When the required earth-retaining structure was greater than 6 m
in height, a two-level stepped wall was used. The bench between the two walls and the faces of the walls
were protected by a wire-mesh reinforced gunite layer (Figure 10-4(c».

The average bidding price for the geotextile wall was approximately 32 percent less than that for the
inextensible MSE wall, and 42 percent less than that for the gabion wall. The unit cost of the geotextile

10 - 14
Approx. Existing - .
Ground Y
1
7.3m < /<~:J
Granular c:--
Backfill l>~ I *'
Reinforced
1g~~~2~/§I~
:'-! -=!

Earth Wall Steel Facing I
. Z;,
Reinforcing) < !--I

Stri.PS (1'.yp.) I'..J


p _ ; ; Additional Wall Width
Dram Ipe / for Wall Height
/\ I ! I Greater than 4.9 m
;;
AlII
,.f.,!,

(a)

Approx. Existing _ / '


Ground,/ IL-

7.3 m /
./ >< .. \
II

Gabion Wall

(b)

Approx. Existing Ground - ,


/1"
/1 I

I-
7.3m
7
<
/
"'I
I
'

Granular
Backfill
Geotextile

Wall 5, ==
Geotextile rDrains-
Wall \ /

(c)

Figure 10-4: Retaining Structure Alternatives for North Halawa Access Road; (a) Inextensible MSE
Wal; (b) Gabion Wall; and (c) Geotextile Wall. (Castelli and Munfakh, 1986)

10 - 15
wall from the overall low bid was $155 per square meter of wall face (1983 price). Immediately after the
bids were received, the construction was halted due to a legal action relating to the environmental aspects
of the project. The issue was later resolved and the geotextile wall contract was awarded in 1987 at a low
bid price of $175 per square meter for both the North Halawa Valley and the Haiku (Figure 10-5) access
roads. Several wall sections were constructed since and are functioning properly. The same design was
also applied for protection of a landslide area nearby.

Selection of the earth retaining structure was based on the following factors: cost, material availability,
right-of-way. construction requirements and durability.

10.5.3 Case 3 - Newark Transit Rehabilitation, New Jersey

To accommodate widening of the track beds at a depressed section of the Newark Rapid Transit System,
steepening of the side slopes was required. To maintain an adequate safety factor against deep-seated
failures, retaining walls 2 to 3 m high were required for partial support of the excavated slopes.

The subsurface proflle along the slopes consisted of 2 to 6 m of loose fill, underlain by a 0.6 to 1 m layer
of soft to firm silty clay, over glacial till. Angles of internal friction of 29 and 38 degrees were assumed
for the loose fill and the glacial till, respectively. The silty clay layer had an undrained shear strength of
21 to 37.5 kPa.

The earth retaining structures considered included cantilevered concrete or steel sheeting, an inextensible
MSE wall, a geotextile wall and a concrete modular block wall. The sheet pile wall required steel sections
ofPS28 and PDA27, or concrete panels 180 mm and 360 mm thick. One disadvantage of the sheet pile
wall was the presence of boulders in the glacial till which could damage the concrete panels or split the
interlocks of the steel sheeting.

The inextensible MSE wall and the geotextile wall were analyzed for internal and external stability. A
major concern of the inextensible MSE wall was the possible accelerated corrosion of the metal strips due
to stray currents from existing electric lines located above the transit right-of-way.

A cost comparison of the five alternatives showed the geotextile wall to be the least expensive. Its
preliminary engineering cost was about 80 percent that of the concrete modular block, 66 percent that of
the inextensible MSE wall, 65 percent that of the conventional steel sheet pile wall, and 55 percent that
of the cantilevered concrete sheet pile wall.

Based on the above cost analysis and considering durability and constructability, the geotextile wall shown
in Figure 10-6 was selected for the project.

10.5.4 Case 4 - Mt. McDonald Tunnel Ventilation, Canada

As part of the ventilation system for the 15-km railroad tunnel passing through Mt. McDonald (Rogers
Pass) in British Colombia, Canada, a 300 m deep ventilation shaft connects the tunnel to a building housing
the ventilation equipment at the ground surface. To construct the ventilation building, a 7.8 m deep
excavation was made at a remote and rugged hillside location.

The overburden soil at the site generally consisted of dense sand and gravel containing some cobbles and
boulders. The angle of internal friction of the soil was estimated to vary between 35 and 40 degrees. No
permanent groundwater table was present above the excavation level.

10 - 16
Ground /
SUrface-y./
./
",
/

-
.02
./
",

(Sta. 39 + 95)

--
~-

1-'--4.3 m ---t.'
Figure 10-5: Haiku Access Road Typical Wall Cross-Section. (Munfakh, 1990)

Catenary Pole

1.5 m
Clearance

Gunite Cover Layer


Level A
Wall Drain h~~~..,.~__ Anchor Pin
- - - - -......·~~:-i!r,.....--...IM

Figure 10-6: Cross-Section of Geotextile Wall for Newark Transit. (Munfakh, 1990)

10 - 17
Several earth-retaining structure alternatives were considered. The presence of cobbles and boulders in
the soil made the constructibility of a conventional soldier pile and lagging structure questionable. Due
to difficult topography and potential access problems at the site, a top-down staged excavation and support
method that requires a minimum right-of-way was desirable. Because of the severe cold climate, the use
of cast-in-place concrete was avoided as much as possible. Based on these and other factors, and after
pricing the most feasible alternatives, a precast concrete "element wall" was selected.

The design, construction and performance of the element wall were discussed in detail by Abramson and
Hansmire (1988). This earth support system used prestressed rod reinforcements installed in the ground
in a manner similar to that of construction of soil anchors. Each reinforcement had an independent facing
in the form of a precast concrete panel (Figure 10-7). The difference between this system and that of an
anchored wall is that, in the element wall, the soil and the reinforcement form a coherent material that act
as the retaining wall and, therefore, the conventional structural elements (soldier piles, wales, lagging) are
not necessary. As an added advantage, the excavation support system became part of the permanent
building structure, thus reducing the space needed for construction of the permanent wall at the back of
the building.

A long-term earth pressure diagram based on the average of the active and at-rest earth pressure envelopes
was used to design the element wall. Water pressure was not considered in the design since the permanent
groundwater table is below the excavation level.

The performance of the element wall was monitored during and, for a long period of time, after
construction. The wall deflections, measured by inclinometers placed behind the wall facing, were
between 0.02 and 0.07 percent of the wall height as compared to the 0.25 percent of the wall height
expected from a conventional soldier pile and lagging system. This relatively low deflection can be
attributed to the top-down level-by-Ievel excavation and support method of construction. Shotcreting the
wall face and placing the wall element immediately after excavation, and the high ground anchor prestress
force apparently prevented decompression of the soil and reduced the associated horizontal deformation.
Another important observation was the appearance of wet spots on the wall face during snow melt.
Although horizontal drains were used, water continued to seep between concrete panels and along the
ground anchors. For two snow-melt seasons after construction, water seepage was relatively low and was
handled by sumping. In the long-run, however, water seepage may be eliminated by placing additional
horizontal drains behind the wall.

Selection of the element wall was based on cost, right-of-way, climate and construction considerations.

10.5.5 Case 5 - Sterling Mountain Tunnel Portal, North Carolina

The Sterling Mountain Tunnel in North Carolina carries Interstate Highway 1-40 through the Blue Ridge
Mountains. In March, 1985, a large rock slide destroyed the tunnel portal. After removal of the debris,
an emergency action was implemented to stabilize the side of the mountain and to reopen the tunnel to
traffic.

A combination of permanent rock reinforcement and horizontal drainage was used for the large-scale slope
protection. To provide protection from small rock falls, a 10.6 m-high retaining wall supporting an
embankment was constructed parallel to the roadway with a 90 degree end wall abutting the rock face. The
wall and its embankment were designed to serve as a buffer zone between the roadway and the rock slope,
absorbing the kinetic energy of any falling rock and accumulating loose rock behind a fence erected above
the wall.

10 - 18
Cast·ln-Place
Concrete Joint Ground Anchor
Precast / rTopwall
Panel
/ rParapet wall
,
\

·• -·• ...,"• • • •" ·• ....,::...


\
I J I
, ~ Precast Panel
- " -~ Aliva Drain

· · . . . . · .-:i
Porous Backfill
Cone. Footing
Struct. Backfill
"
"- Reinforced Concrete

Front View Cross Section

Figure 10-7: Element Wall for Mt. McDonald Ventilation Building. (Munfakh, 1990)

To meet the requirements of the emergency situation, three types of earth retaining structures were
analyzed: an inextensible MSE wall structure which incorporated the embankment in its design, a gravity-
type concrete modular block structure, and a concrete-faced anchored wall. All three were required to be
constructed in a very limited time and to be aesthetically pleasing. The cost differential between the MSE
wall and the concrete modular block wall was within the degree of precision of a preliminary cost estimate.
The anchored wall was 60 percent higher in construction cost, mainly because it required difficult drilling
through very hard quartzite for installation of the soldier piles and anchors. To speed up construction, the
material had to be procured in a separate advance contract and be stored at the site. The space required
for material storage was greatest for the concrete modular block wall and least for the anchored wall.

The anchored wall alternative was eliminated because of its relatively high cost. Because of the tight
construction schedule and the limited space available for operation and storage of construction material,
the MSE wall was judged to be the most appropriate. Construction of the wall involved placement of steel
reinforcing strips within layers of compacted granular material, supported by a wall facing of prefabricated
concrete panels. The total construction time, including material procurement, was about two months for
the 63 m long wall. The wall and its embankment are functioning properly.

Selection of this earth-retaining structure was based on the following factors: construction schedule, right-
of-way, aesthetics and cost.

10.5.6 Case 6 - New Jersey Turnpike Widening, New Jersey

At Interchange 15W of Section D of the New Jersey Turnpike, a ramp was required to be added under an
existing bridge to accommodate traffic volume increases. To allow construction of the new lane while
maintaining traffic on the bridge deck above, a top-down construction method was selected for the earth
retaining structure using soil nailing for support. Conventional earth retaining structures, such as sheet
piles, soldier piles and lagging or reinforced concrete walls, were more expensive and required removal
of the bridge deck and interruption of the bridge traffic. Maintaining that traffic was the highest priority
placed on the selection process by the Turnpike Authority.

Figure 10-8 illustrates a cross-section of the soil nail wall. Three rows of nails were required with vertical
spacings of 0.75 to 1 m and a horizontal spacing of 1.5 m. Each nail consists of a 3.5 m long, 25 mm
diameter steel reinforcing bar, installed in a 150 mm diameter drilled hole. The entire length of the nails
were within the compacted granular fill material of the bridge abutment.

10 - 19
7.5 on

~
Plate
!J -.I Shotcrete
Thread-Bar
Grade 60
Grout

Cross Section Detail

Figure 10-8: Soil Nail Retaining Structure for New Jersey Turnpike Widening. (Munfakh, 1990)

The internal stability of the system was checked for tensile rupture of the nail and adhesion failure of the
soil-nail interface. The external stability of the coherent soil nail structure, however, governed its design.
A conservative approach using ~ was followed for determining the horizontal pressure acting on the wall
face in order to reduce any horizontal soil movement that may have a negative impact on the existing
bridge piles behind the wall.

The earth retaining structure is constructed in three repetitive stages. Each involves excavation of a soil
layer, installation of a row of nails and shotcreting of the exposed face. A cast-in-place concrete facing
provides the final wall finish. No internal drainage was required since the permanent groundwater table
was below the base of the soil nail mass.

Selection of the soil nail wall was based on the following parameters: right-of-way, ease of construction
and cost.

10.5.7 Case 7 - Shot Tower Metro Station, Baltimore, Maryland

To construct the Shot Tower Station of the Northeast Line of the Baltimore Metro, three types of retaining
structures were used: a soldier pile tremie concrete (SPTC) slurry wall, a steel soldier pile and timber
lagging wall, and a jet-grouted wall (Figure 10-9). These walls were necessitated by the ground conditions
at the project site, a high groundwater level, the presence of three high voltage electric transmission lines
across the site, and the proximity of existing structures to the excavation. Construction of the three types
of retaining structures at the Shot Tower Station has been completed and the facility is currently in
operation.

In general, the soil profile at the station consists of 4.5 to 7.5 m of fill, underlain by 6 to 10.5 m of
Cretaceous soils, followed by residual material and rock. The station excavation is mainly in the
Cretaceous soils which consist of dense to very dense sands and gravels with up to 49 percent fmes
(primarily homogeneous silt binder with varying percentages of clay). The design groundwater table is
about 16.5 to 21 m above the bottom of the excavation.

10 - 20
1.4 mcast-In-Place \ 115 kV Lines
Cone. Wall I
~ 914mm

I~_--....---
• • -- -t--

Overtapping Jet
_ _..l--203m 2.8m 2.8m 1.7m Grout Columns (Typ.)
1.4m

Soldier Piles in Augered


Holes and Timber Lagging

Figure 10-9: Shot Tower Station Retaining Structures. (patel and Castelli, 1992)

Due to the high density of the soil (SPT N-values of more than 100 blows/ft were common), the use of
sheet piling was not feasible. To use conventional soldier pile and lagging walls, the ground would have
to be dewatered ahead of excavation. The effectiveness of dewatering, however, was questionable and its
impact on nearby structures was of concern. Based on that, the SPTC slurry wall was selected for
excavation support. This excavation support structure was designed also to be the permanent station wall.

The design of the SPTC wall is discussed in detail by Patel and Castelli (1992). Lateral earth pressures
were calculated for short-tenn, long-tenn and unbalanced loading conditions. The theoretical active earth
pressures were converted to rectangular apparent earth pressure diagrams for the design of the temporary
bracing system. Seepage analyses were perfonned to detenninethe design water pressure distribution
behind the wall. Water seepage under the wall was· a factor in reducing the full hydrostatic pressure
applied to the wall during construction. The maximum horizontal and vertical ground movements behind
the SPTC wall are expected to be less than 19 mm.

Four 115 Kv electric transmission lines cross the station excavation (Figure 10-9). Due to the difficulty
in constructing the SPTC wall around and below these live lines (which could not be relocated), an
alternative excavation support system using soldier piles and lagging was selected for that location. The
soldier piles consist of steel H-piles placed in pre-drilled holes to ensure damage-free installation and
proper alignment in the very dense soils, and to minimize noise and vibration during construction. To
achieve contact between the back flange of each pile and the soil, the pre-drilled hole is filled with concrete
after the pile is seated plumb in its center. The soldier pile and lagging system was designed for the
temporary loading condition using a rectangular apparent earth pressure diagram appropriate for design
of braced excavations, but assuming full water pressure. The final· station wall at that location consisted
of cast-in-place reinforced concrete.

Alternative methods of preventing seepage into the excavation through the lagging were considered.
Chemical grouting was eliminated because of the high fines content in the Cretaceous soils at that location.
Ground freezing was rejected due to the limited space available adjacent to the excavation for placement
of freezing equipment and pipes. A seepage cut-off wall comprised of overlapping jet-grouted columns,
on the other hand, was detennined to be feasible for the subsurface conditions at the site.

10 - 21
Jet grouting is a relatively new method of ground improvement that is less dependent on the type of ground
than conventional grouting methods. It involves applying high speed water jets vertically and horizontally
through a special drill bit to excavate the in situ soil, then mixing it with grout pumped at high pressure
through the nozzles of the drill bit as it is withdrawn. By rotating the drill bit, a high-strength jet-grouted
column is created (Munfakh, et aI., 1987). Substantial reductions in the permeabilities of sandy soils (up
to 0.0001 of their original values) were reported due to jet grouting. The specified jet-grouted columns
were 0.75 m in diameter, placed about 1 m behind the soldier pile and lagging wall.

Although the jet grouted wall could have been designed to provide structural support (by placing a
reinforcing element in the grouted column), thus eliminating the need for the soldier pile and lagging wall,
a decision was made to use it as a cut-off wall only due to the sensitivity of the proposed excavation, and
the lack of prior experience in this type of construction in the Baltimore area.

Selection of the retaining structures at this project was dictated by ground and groundwater conditions,
right-of-way limitations, and the presence of existing structures at, and adjacent to, the excavation.

10.5.8 Case 8 - Kismayo Port, Somalia

The Port of Kismayo lies 45 kIn below the Equator in the East African country of Somalia. As part of a
major reconstruction of the port, funded by the United States Agency for International Development
(USAID), a bulkhead was constructed outboard of an existing deteriorated wharf, with granular backfill
placed underwater behind it.

The subsurface conditions at the site consist generally of a surface layer of soft silt and clay about 1 m
thick, underlain by approximately 20 m of loose to very dense silty sand followed by a hard coral surface.
No suitable sources of aggregate material were present locally or at an economical hauling distance from
the project site. The low quality aggregates used in the original construction were major contributors to
the rapid deterioration of the concrete deck and pile caps.

Three types of bulkhead structure were judged to be capable of meeting the design criteria established for
the project: (1) a cellular structure consisting of a series of large diameter steel sheet pile cells filled with
granular backfill and connected by sheet pile arcs, (2) a gravity-type unreinforced concrete block wall and
(3) an anchored steel sheet pile wall supported by an A-frame pile anchorage or a continuous sheet pile
deadman.

The cellular structure was rejected due to its relatively high cost. The concrete block wall was the most
desirable since it utilizes local material, requires unskilled labor, and requires minimum maintenance. This
structure, however, was also rejected due to the unavailability of the large quantities of suitable aggregates
required to cast the concrete blocks. The anchored steel sheet pile wall, on the other hand, required a
minimum quantity of structural concrete, had a relatively low cost and provided a low maintenance
structure. This scheme was selected for construction of the bulkhead (Figure 10-10).

To build the new bulkhead, the existing wharf platform was demolished and the tops of the existing
concrete piles were cut off. A steel sheet pile structure was then constructed 10.7 m outboard of the
existing platform, supported by continuous steel sheet pile deadman located near the back of the platform.
Behind the bulkhead, a sand backfill was placed underwater then compacted from the surface using deep
vibratory compaction. The deep compaction was specified to minimize future settlement of the backfill and
provide a stable foundation for a maintenance-free rigid concrete pavement above it.

10 - 22
The bulkhead design is discussed in detail by Castelli and Secker (1987). Briefly, the sheet pile structure
was analyzed using the free earth support method and Rowe s moment reduction factor. Angles of internal
I

friction of 30 and 35 degrees were used for the granular soils. A factor of safety of 1.5 was applied to the
passive soil resistance on the outboard side of the bulkhead. A water level differential of 0.5 m was used
in the design.

The required section modulus of the sheet pile exceeded available standard American sections. The use
of a European master pile system of interlocking Hand Z sections, though well suited for the bulkhead,
was excluded by the "Buy America" requirement of the USAID funded project. A master pile system
fabricated from American sections on the other hand, was uneconomical. The bulkhead structure fInally
selected utilized standard Z sections (PZ-40) reinforced with cover plates on the outer faces of both
flanges. A multiple corrosion protection system using coal-tar epoxy coating and cathodic protection was
provided.

The spacings, equipment and procedures used in the deep vibratory compaction of the hydraulic fIll were
selected on the basis of fIeld trials performed at the beginning of construction. Terra probe, vibro wing
and two types of vibroflots were tested, and compaction was investigated with or without the use of water
jet and with or without the addition of backfIll during compaction. The best results were accomplished
using two vibroflots rigidly connected together and applying water jets during compaction (Castelli, 1989).
Relative densities of more than 80 percent were measured following compaction. Construction of the
Kismayo Port was completed in early 1989. The new bulkhead is functioning properly.

In addition to cost, selection of the earth retaining structure for this project was influenced by durability,
availability of material, maintenance requirements and contracting considerations.

Edge of Existing Plattorm

Ii n n
Existing 'I I , , •

/Aock --. II
___ I' ... ~ _~
"'\ -,'I'!"- -:.~,,
I ·Exis!ing Surface
~New Filter Layer Sheet Piles With
DikeI'll I,
, I, ,. '
II
~I II I ' " ~ Cover Plates
II
II i I I; 1 ~ I .....' t t i l ~ Sand
II
II
I
':
I
::
I
':
I
;-i-.....
I '" l'......
Fill .r==-
L-...
9.5 m
II ,I I ~:::""""_----it'--'----:::
__~"~$~ws~'-
II
"
"
,I
I~
"
I, ,:
II
" " ..... 1--
t..
~E:ist:ngL:Conc~;~e-Pile~ - E~;~ation &: Backfill
of Soft Silt & Clav in
Vicinity of Berths',2 & 3

Figure 10-10: Cross-Section of Bulkhead for Kismayo Port. (Castelli and Secker, 1987)

10 - 23
10.5.9 Case 9 - Canton-Seagirt Facility, Baltimore, Maryland

In conjunction with construction of the Fort McHenry Tunnel in Baltimore, Maryland, an offshore disposal
and containment facility was built in the Baltimore Harbor to receive approximately 900,000 cubic meters
of dredged material generated by construction of the immersed tube tunnel. Due to stability and
environmental considerations, construction of a dike was not allowed and an earth retaining structure was
needed to contain the hydraulically placed dredged material.

Borings drilled along the alignment of the retaining structure, where the harbor bottom was about 3.7 to
4.6 m below water, revealed very soft to medium stiff organic silts or silty clays varying in thickness from
7.6 to 12 m, underlain by dense sands and hard Cretaceous clays.

The design criteria (loading, elevations, etc.) were set with the objective of turning the disposal facility into
a container port operated by the Maryland Port Administration. The containment structure, therefore, had
to provide complete separation of the dredged material from the harbor water and to support the horizontal
loads from the hydraulic fill and an additional surcharge load of 50 kPa.

The least costly retaining structure suitable for the severe loading criteria was an H-Z master pile wall
system, which uses interlocking H-pile and sheet pile sections to provide the required high section
modulus. The structure would be built at a batter to reduce the applied pressure and, in tum, the required
section modulus. However, the use of foreign steel was precluded by the Maryland State Specifications
and a similar system had to be fabricated using domestic steel H-pile and sheet pile sections (Figure 10-11).
Since the inclined master pile-sheet pile structure was unfamiliar to U.S. contractors and required building
a special template for driving both the H and Z sections at a batter, and to meet the extremely tight
construction schedule (the facility had to be completed prior to the start-up of tunnel dredging), an alternate
containment structure consisting of 76 19 m diameter steel sheet pile cells was built by the contractor.

Normal construction practice for cellular structures would require removal of all soft soils within the cell
and replacement with granular backfill. The presence of a soft cohesive layer sandwiched between the
hard bottom and the granular fill would reduce both the vertical and horizontal shear resistances. In
addition, this layer would settle under the superimposed load resulting in downdrag forces on the steel
sheet piles. Since the vertical stress applied on the outboard side of the cell is larger than that applied on
the inboard side due to the unbalanced loading condition, settlement of the clay and the resulting downdrag
force would be larger at the outboard side causing tilting of the cell. Therefore, removal of all cohesive
soil from within the cell was desirable.

The environmental restrictions on construction in the Baltimore Harbor and the unavailability of a disposal
site precluded predredging of the soft soils. Excavation of all clays from within the cells was not believed
feasible due to the potential collapse of the cell. Therefore, a decision was made to leave some of the clay
inside the cells. It was also decided to monitor the cellular structure and to implement remedial measures
once the lateral movement of any cell reached 0.6 m.

As the contractor fIlled the contained site with dredged material, several cells moved. Once the allowable
limits were exceeded, different remedial measures were undertaken to stabilize the structure. The
contractor installed deep wells, excavated behind the cells, installed underdrains behind the cells and placed
a berm in front of the cells. A few months later, the outboard lateral movement had reached 1.2 m and
the settlement had reached 0.9 m in some cells. The remedial measures applied to that date, therefore,
were only partially successful. As a result, it was decided to install stone columns inside the cells. The

10 - 24
presence of stone columns within the clay layer would (1) increase its resistance to horizontal shear, (2)
reduce the total settlement and resulting downdrag forces and (3) accelerate consolidation and strength gain
in the cohesive soil.

Figure 10-12 illustrates a layout of stone columns inside a cell. The column spacing varied from 1.7 to 3.1
m across the cell to accommodate the distribution of vertical stresses throughout the cell. A column
diameter of 1.1 m, a friction angle of 42° and a stress ratio of 3 were used in the design (Castelli, et aI,
1984). The resistance of the clay layer to horizontal shear was increased by 67 percent and its anticipated
additional settlement was reduced by approximately 300 percent due to reinforcement with stone columns.
Movements of the cells stopped within 30 days after installation of the stone columns.

Selection of the cellular earth retaining structure was seriously influenced by the following factors:
construction schedule, environmental restrictions, contracting issues and tradition.

~ Combined Pile
(W36 + PSA23)
I
I
.
254mm .

;
,PSA23
Cut &
254mm J
1
406mm
PDA27
Welded~' (A-36 Steel)
I
i

.~~~ ::::::::!:=5 ~ -'~


~/ ,f ~ \ I~
11
I '/1
~ /
W36->l'
x 135 ·.

Coverplate I
b

Detail of Master.Pile-Sheet Pile System


Cross-Section of Retaining Structure

Figure 10-11: Master Pile Retaining Structure Proposed for Canton-Seagirt Facility. (Munfakh, 1990)

Inboard

Figure 10-12: Cellular Structure Reinforced with Stone Columns Used for the Canton-Seagirt Facility.
(Castelli, et aI, 1984)

10 - 25
10.5.10 Case 10 - Jourdan Road Terminal, New Orleans, Louisiana

Construction of Berth 4 of the Jourdan Road Terminal of the Port of New Orleans presented an innovative
use of a combination of earth retaining structures. The generalized soil profile at the site consisted of
approximately 18 m of soft clays and silty clays with interbeds of silt and fine sand, underlain by medium
to very dense sands which act as the bearing layer. The undrained shear strength of the clay ranged from
about 10 kPa at the top to approximately 22.5 kPa immediately above the sand stratum.

Several alternative wharf structures were considered for the project, many of which involved the use of
earth retaining structures including a steel sheet pile bulkhead, a cellular structure, a concrete caisson wall,
a slurry wall relieving bulkhead and others. A selection matrix was prepared for a qualitative evaluation
of the feasible alternatives. Twelve selection factors were considered representing construction, operation,
maintenance and relative cost. The factors related to operation and cost were given higher weights than
the rest. As a result of the preliminary engineering study, an innovative, cost-effective scheme was
selected using a combination of earth retaining structures and ground improvement methods.

Figure 10-13 illustrates the selected wharf structure. Basically, it involves construction of a narrow pile-
supported deck with a steel sheet pile wall connected to its outboard edge. The sheet pile wall allows
dredging in front of the wharf while retaining the material behind it. To improve the overall stability of
the system, the soft cohesive soil behind the deck is reinforced with stone columns extending from the
ground surface to the top of the bearing stratum. On top of the stone columns, an MSE wall is constructed.
The wall and its reinforcing embankment have three functions: (1) to provide a semi-rigid mat above the
stone column zone, thus forcing potential failure surfaces to pass beyond the mat and intersect the
maximum number of stone columns, (2) to transfer the superimposed loads to the stone columns by arching
over the in situ soils and (3) to allow vertical filling behind the deck without a bearing capacity or shallow
type failure. The combined use of in situ reinforcement and earth retaining structures improved the overall
stability of the wharf structure, eliminating the need for a wide pile-supported deck.

1. 81 m .,I
J
i

Transit Shed--ll::lf; ~=======F===:;:::~;;;:;:::::::::J::::J


I'
EI. + 3.1 m !! RR Tracks
I /1

Sheet Pile

L..-.J
Steel Pipe Piles
104. St. .:e . :. . . .:Pi. .:·pe~Pi. . .:·: .:le:.: s II. Stone COlumns..l.
... Wick Drains

Figure 10-13: Composite Wharf Structure for Jourdan Road Terminal. (Castelli, et aI, 1983)

10 - 26
The outboard sheet pile wall was designed to support the horizontal loads from the soil behind it (the deck
and operating loads were carried by the piles) using conventional methodology. The MSE wall was
designed to meet internal stability requirements. Except for a possible deep-seated failure, the external
stability of the system was not a concern since the width to height (LIH) ratio was quite high (the length
of the reinforcement was dictated by the width of the stone column reinforced area). The overall stability
of the stone columnlMSE system was analyzed using a composite shear strength value for the reinforced
soil, determined from weighted average material properties. The stone columns had a 2.1 m spacing, an
average diameter of 1.11 m, a friction angle of 42 degrees and a stress ratio of 3.

The vertical and horizontal deformations of the wharf structure and adjoining soils were monitored during
and after construction (Castelli, et al., 1983). No stability or settlement problems were experienced.

The presence of stone columns below the reinforced embankment reduced the anticipated settlements by
about 70 percent. Although the horizontal deformations of the subsoil were relatively high - a maximum
value of 127 mm -- they did not cause instability. The use of this composite structure realized an estimated
saving of $1.25 million (1980 price) per berth as compared to the cost of a conventional pile-supported
platform.

Selection of the Jourdan Road Terminal Wharf structure was based on the following factors: cost, stability.
constructability, impact on an existing levee. and operation and maintenance requirements.

10.6 SUMMARY

Selection of an earth retaining structure is influenced by many factors involving technical, practical,
economical and contractual considerations. Alternate bids are often allowed to secure a cost-effective
solution.

The ten cases discussed in this chapter demonstrate clear advantages of the innovative solutions used. Up
to 45 percent reductions in construction cost were realized on these projects. Table 10-4 summarizes the
factors that governed the wall selection on these projects.

The following approach to selecting earth retaining structures is recommended:

• Select a structure compatible with the soil it supports and the one it rests on - Strain compatibility
with the soil is particularly important when the soil is incorporated within the structure.

• Beware of groundwater - It is the geotechnical engineer's enemy.

• Consider the contractor's capabilities, constraints and experience - Remember, he will implement
your design. Perform a constructibility review including assessment of construction schedule and
staging.

• Consider contracting issues and tradition when making your selection.

• Be sensitive to the environment.

• Place a special emphasis on aesthetics.

10 - 27
• Consider the long-term behavior and function of the structure -- A structure with the least
construction cost does not necessarily mean an economical solution if it has to be replaced within
a few years after construction, or if it requires continuous maintenance and rehabilitation.

• Provide the contractor with sufficient information and enough flexibility - It avoids contingencies
and future claims.

• Allow alternate bidding on your project -- It encourages competition and assures minimum cost.

• Monitor the behavior of your structure and the soil it supports. Document your instrumentation
results and make them available to the profession.

TABLE 104
SUMMARY OF SELECTION FACTORS

Selection Factor
»~ ....
""' Q Q
~ .Sl
.... ~ lI) 0.0
~ 0 <+!..
0 ....""
.~ S
Q
g .5
....
0
Q
.Sl
"'0
Q
"'0
Q ....""'=' ....
I
..c::
lI)
-5
0
""'
:c ....
~
....""' ....
=B
0='
~
0=' ""0Q
~
Q
""'
0.0
""
lI) ""'=' ""0 0
~
""'
c:J c:J u ~ <: Q U U ""'
Eo-<
Project
Aviation Corridor ./ ./ ./ ,/ ,/ ./
H-3 Access Roads ./ ./ ,/ ./ ./ ./
Newark Transit ./ ./ ./ ./ ./
Mt. McDonald ./ ./ ,/ ./ ,/
Sterling Mountain ./ ./ ,/ ,/ ./
N.J. Turnpike ./ ,/ ,/ ./
Shot Tower Station ./ ,/ ./ ,/ ./ ./
Kismayo Port ./ ./ ./ ,/ ,/
Canton-Seagirt ./ ,/ ,/ ./ ,/ ./
Jourdan Road ./ ,/ ,/ ,/

10 - 28
CHAPTER 11.0
CONTRACTING METHODS AND BIDDING DOCUMENTS

11.1 INTRODUCTION

Since the early 1980s, the non-conventional earth-retaining systems have gradually displaced conventional
systems through bid alternates, experimental applications, and value engineering proposals. It is estimated
that hundreds of millions of dollars have been saved through the use of these systems (FHWA, 1988).
However, an increase in frequency of wall design and construction problems associated with these
alternative systems has accompanied their escalated use. Although the actual causes of each particular
problem may be unique, most problems can generally be attributed to four main causes:

• Selection of inappropriate wall alternatives


• Use of unsuitable materials
• Inadequate bidding documents
• Lack of specification enforcement

The purpose of this chapter is to provide technical guidance for issues relating to contracting methods and
bidding documents for different wall systems.

There are two other FHWA publications on this subject: Geotechnical Guideline No.2 (FHWA, 1988)
and Geotechnical Engineering Circular No. 2 (Sabatini, et. ai., 1997). Some material from these
publications is included in this chapter. In addition to this chapter, the reader is referred to the reference
manual of Module 2 on the topics of Contracting and QA/QC.

11.2 CONTRACTING APPROACHES

Permanent earth retaining systems are typically contracted using either a method approach or a
performance approach. These approaches are briefly described below. A detailed discussion is presented
in subsequent sections.

The method approach is used for projects designed by the owner or the consulting engineer acting on
behalf of the owner. In the contract documents, wall component details, construction materials, and
methods of execution of construction are explicitly specified. This contracting approach is discussed
further in Section 11.3.

In the performance approach, pre-approved or generic wall systems or components are used. Included in
the contract documents are lines and grades, as well as specific geometric and design criteria. The
contractor selects a specific type of wall that meets the contract's guidelines, and submits detailed project-
specific plans, specifications and design calculations for the owner's review, in addition to normal working
drawing submittals. This contracting approach is discussed further in Section 11.4.

Both contracting approaches are valid for most earth retaining systems, if properly implemented. Often,
the approach will be selected based on the experience of the owner agency and its engineering consultants
with various wall systems, the complexity of the project, the availability of specialty contractors or material
suppliers, and the agency's philosophy with respect to contracting methods. In addition to the method and
performance approaches, a mixed approach employing both methods can be used. Mixed approaches are
not discussed further in this chapter because their appropriateness depends on project-specific conditions.

11 - 1
Regardless of which contracting approach is chosen for a specific project, it is highly desirable that each
owner agency develop a formal policy with respect to contracting for earth retaining systems. The general
objectives of such a policy would be to:

• obtain uniformity in selection of earth retaining system alternatives;

• establish guidelines for using either the method or the performance contracting approach;

• establish standard policies and procedures for technical review and acceptance of proprietary and
generic earth retaining systems;

• establish responsibility for the acceptance of new earth retaining systems and/or components, and
for the design, preparation of bid documents, and construction control of such systems;

• develop uniform design, performance criteria, and standard specifications for different earth
retaining systems.

11.3 METHOD CONTRACTING APPROACH

11.3.1 Introduction

The method contracting approach involves the development of a detailed set of plans and construction
specifications for inclusion in the bidding documents. The advantage of this approach is that it enables the
design engineer to examine various earth retaining system options during design, with impartiality that
cannot be expected from the contractor. The design engineer also has more time to optimize the design
and develop technical details that would minimize uncertainties and disputes during construction.

A disadvantage of the method approach is that for alternate bids, more systems must be evaluated, and
more sets of design must be developed. Therefore, the owner's resources may be expended, even though
only one wall system will be constructed. Another disadvantage is that the designer may be unfamiliar
with newer and potentially more cost-effective systems and thus may not consider them during the design
stage. Similarly, the proprietary wall systems, may have technical details known only to the proprietors,
thus, the owner or the consulting engineer may not feel comfortable enough to use them.

The method contracting approach is best suited for walls supporting fill where the available technology is
either traditional or widely disseminated and reasonably well-established. Knowledgeable contractors and
material suppliers of fill-type wall systems are widespread throughout the United States. Detailed plans
and special technical provisions are often furnished to the design engineer, at no expense, by specialty
contractors and proprietary material suppliers, especially those involved in construction of MSE wall
systems.

For wall applications in cuts, the method contracting approach is generally appropriate for the more
widely used and well developed wall systems such as cantilevered sheet-pile walls, anchored walls, and
soil nail walls. The approach is mostly suitable for cut wall systems with which the owner or consulting
engineer has developed significant experience in design and construction. The method contracting
approach is commonly used for permanent wall construction.

11 - 2
11.3.2 Bidding Documents

The bidding documents in the method approach consist of drawings, specifications, and bidding quantities.
Depending on agency preference, the contract can be bid on a lump-sum basis or following a detailed unit
price list. Because the wall is designed and estimated prior to bidding, the contract drawings and
specifications should be very detailed to ensure that all the assumptions made by the designer and the
estimator are followed by the contractor.

Typically, the bidding documents (plans and specifications) should include the following items as a
minimum:

• the horizontal alignment of the wall identified by stations and offsets from a horizontal control line
to the wall face, the beginning and end stations for wall construction, and all appurtenances that
affect construction of the wall;

• the elevations of the top and the base of the wall along its aligmnent, and the final ground surface
behind the wall;

• the site plan cross sections showing limits of construction, backfIll requirements, excavation limits,
groundwater levels, etc;

• typical wall sections and special details;

• construction constraints such as staged construction, vertical clearance, right-of-way limits,


construction easements, noise and air quality requirements, dimensional and aligmnent tolerances
during construction, etc.;

• required construction procedures including foundation preparation, and external and internal
drainage details;

• details of ancillary structures such as traffic barriers, copings, parapets, noise walls, and attached
lighting;

• a boring location plan and boring logs;

• general notes required for construction and contractor submissions;

• payment limits and quantities.

Inclusion of detailed technical specifications in the bidding documents is essential, particularly for the
method contracting approach. Proper implementation of these specifications will ensure that the
assumptions made by the designer with regard to the construction and performance of the wall are satisfied.
Clear, concise and detailed specifications will also assist in assessment of pay requests, evaluation of
contractor's claims, and dispute resolution.

Following is a list of references where key specification elements used for earth retaining structures can
be found.

11 - 3
• Federal Highway Administration

• Corps of Engineers

• Veterans Administration

Guide specifications for specialized types of walls are also available from specialty contractors and
propriety material suppliers.

11.3.3 Key Plans and Specifications Items

In addition to the generally required bidding items discussed above, the following items are required for
specific wall types:

Cast-in-Place Gravity and Semi-Gravity Walls (Chapter 4)

• footing size, depth, and reinforcing steel details;

• location and details for deep foundations, if required, including pile elevations and pile cap details;

• location of construction and expansion joints and applicable details;

• typical sections of the wall showing concrete dimensions and, for cantilever and counterfort walls,
reinforcing steel details;

• special technical specifications such as pile driving, testing, etc.

Modular Gravity Walls (Chapter 5)

• length, size, and type of the gravity unit (concrete module, gabion basket, bin cell, etc.), and
positions for which unit dimensions change;

• wall cross-section showing unit arrangements and positions for which different unit sizes are used;

• footing location, depth, dimensions, and details;

• properties and methods of placement of fill material;

• corrosion protection requirements of metallic units;

• planting or seeding requirements in the facing blocks.

Mechanically Stabilized Earth Walls (Chapter 6)

• length, size, and type of soil reinforcement, and positions for which the reinforcing elements
change in length or size;

11 - 4
• layouts, dimensions and elevations of the footings and/or leveling pads;

• aligmnent and elevation of internal drainage systems, and method of passing reinforcing elements
around the drainage systems;

• details of facing panels and panel connections with reinforcing elements;

• details of wire-mesh reinforcement for shotcrete facing and steel reinforcement for cast-in-place
facing;

• details of architectural treatment or surface fInish of the facing;

• details for construction along curved alignments, and around drainage facilities, overhead sign
footings or other structures;

• corrosion protection requirements/details for reinforcing elements.

Externally Supported Stroctural Walls (Chapter 7)

• size, location, and minimum embedment depth of all vertical structural elements;

• size of lagging and details for temporary and fInal facing connections to the soldier piles and/or
walers;

• location, inclination, minimum unbonded length, anchor bond length, required anchor capacity,
and lock off load for each anchor;

• corrosion protection requirements for the anchor head, the unbonded length, and the anchor length;

• requirements, details, frequency and acceptance criteria of anchor testing;

• dimensions of the slurry wall panel, top and bottom wall elevations, panel joints and depth of
embedment below the bottom of the excavation;

• details of reinforcing steel, soldier piles, or precast concrete panels in slurry walls;

• bentonite slurry and concrete mix and concreting details in slurry walls;

• diameter, tip elevation, reinforcement arrangement and construction sequence of secant and tangent
piles in bored-pile walls;

• details for facing treatment or permanent facing installation including drainage requirement and
water proofIng.

In Situ Reinforced Walls (Chapter 8)

• nail size, spacing, length, and inclination;

11 - 5
• nail installation procedure, including minimum drill hole diameter
nail pullout capacity;

• corrosion protection requirements;

• minimum thickness of the temporary and/or permanent facing with details for facing reinforcement
and nail cover plates;

• details for facing treatment or permanent facing installation;

• nail testing requirements, including methods and frequency of testing;

• sequence and dimensions of the step-by-step top-down construction of the soil-nailed wall

• micro-pile installation·procedures;

• grout mix and grouting pressure for micro piles;

• size and details of the pile reinforcement;

• details of the capping beam;

• pull-out capacity and testing requirements of the micro-pile.

Chemically Stabilized Earth Walls (Chapter 9)

• jet grouting system, e.g., single, double or triple fluid, etc.;

• minimum required design parameters, e.g., grout pressure, w/c ratio, strength, permeability, etc.

• minimum column diameter;

• layout of stabilized columns;

• installation sequence of deep soil mixing columns;

• vertical and horizontal tolerances for stabilized columns including minimum overlap requirements;

• field and laboratory testing requirements and frequency;

• reinforcing details, if any;

• special considerations where peat, boulders, utilities, etc. may be encountered;

• quality control procedures and field acceptance check list.

11 - 6
11.4 PERFORMANCE CONTRACTING APPROACH

11.4.1 Introduction

For the performance contracting approach, often referred to as "line and grade" or "two line drawing"
approach, the agency prepares drawings showing the geometric requirements of the wall system, material
specifications, and wall performance requirements. The wall systems or components which are permitted
for use are either: (a) specified by the agency, or (b) on a pre-approved list maintained by the agency.

Design of the structure is the responsibility of the contractor and is usually performed by a trained and
experienced contractor, wall supplier, or engineering consultant. Prequalified material components have
been successfully and routinely used, which is a departure from the method approach where generic
material specifications are used. The technical basis for detailed designs to be submitted after the contract
award are included in special contract provisions. Performance criteria, which usually include permissible
horizontal/vertical deformations at completion of construction and/or maintenance of groundwater levels,
must be defmed and made part of the contract documents.

The advantage of the performance approach is that it enables the contracting agency to reduce its
engineering cost and manpower requirement since the responsibility of the design is shifted to the
contractor. Its disadvantage is that the agency engineers may not be experienced with the wall technology
proposed by the contractor and, therefore, may not be fully qualified to review and approve the wall design
and any construction modifications. Newer and potentially more cost-effective systems may be rejected
by the agency due to the unfamiliarity of the agency personnel with such systems, which may lead to
disputes and controversies.

In the performance contracting approach, different wall designs may have unique pay quantities reflecting
the materials and methods required, or the wall may be bid as a lump sum with unit prices for added or
deleted items based on changed conditions. A lump sum payment for the completed wall system is often
used, since the actual quantities of materials are dependent on the wall system to be designed, and thus,
cannot be included in the bidding documents.

11.4.2 Bidding Documents

The bidding documents prepared by the contracting agency consist of three main sections: (a) geometric
and site data, (b) design guidelines, and (c) performance requirements.

Geometric and Site Data

All the geometric and site data collected for the project or available from previous construction should be
included in the bidding documents. Specifically, the bidding documents should include:

• a plan showing the horizontal alignment and offset from the horizontal control line to the face of
the wall, the beginning and end stations, and all appurtenances that affect construction of the wall;

• profiles indicating the elevations at the top and bottom of the wall along its alignment, and locations
and elevations of the final ground line;

11 - 7
• cross sections showing limits of construction, excavation and backfilling limits, as well as mean
high water level, design water level, and drawdown conditions, if applicable;

• construction site constraints such as vertical clearance, right-of-way limits, construction easements,
etc.;

• locations of utilities, signs, etc., and any loads that may be imposed by these appurtenances;

• geotechnical data including previous construction records;

• environmental data including type and degree of groundwater contamination.

Design Guidelines

The design guidelines should include the following:

• reference to specific governing sections of an agency design manual (materials, structural,


hydraulic, and geotechnical), construction specifications, and special provisions; if none is
available, reference to current AASHTO Standard Specifications (both Division I - Design and
Division II - Construction, should be provided)

• geotechnical design parameters such as friction angle, cohesion, unit weight, coefficient of
permeability, etc.;

• magnitude, location, and direction of external loads from bridges, overhead signs and lights, and
traffic surcharges;

• minimum wall footing embedment depth and foundation preparation requirements;

• minimum factors of safety for various failure modes such as overturning, sliding, stability of
temporary construction slopes, overall rotational stability, bearing capacity, uplift resistance,
tensional capacity, pull-out resistance, etc.;

• seismic design requirements and minimum factor of safety under seismic conditions;

• in fill and backfill requirements, including material specifications and placement procedures;

• drainage requirements beneath, behind, above, or through the structure;

• type, size and architectural treatment of the permanent wall facing;

• erosion protection requirements, if applicable, and methods of shore protection design.

Performance Requirements

An instrumentation monitoring program is usually prepared as part of the design. The bidding documents,
therefore, establish guidelines for the minimum level of instrumentation to be used by the contractor, and
the threshold against which the monitoring data will be evaluated. At a minimum, the following

11 - 8
information should be included in the bidding documents:

• design life for the structure and, if applicable, required corrosion protection;

• tolerable horizontal and vertical movements of the structure and acceptable methods of measuring
these movements;

• pile design capacity, and static and dynamic pile testing requirements for walls supported on deep
foundations;

• pull-out design capacity of micro-piles, soil nails, etc., and methods of testing for pull-out
resistance;

• acceptance criteria, and performance and proof testing requirements for ground anchors;

• anticipated creep behavior of anchors, soil nails, etc., and methods of measuring creep movements;

• durability requirement of chemically stabilized earth walls, and methods of testing long-term
behavior of soilcrete;

• permissible range of variation in groundwater levels, and methods of groundwater level


measurement.

11.4.3 Review and Approval

If a performance contracting approach is used, the contractor's submittals are reviewed and approved by
the agency, or its consultant, before construction can commence. The evaluation by the agency's structural
and geotechnical engineers must be rigorous and must consider, as a minimum, the following items:

• conformance to the project line and grade;

• conformance of the design calculations to the design guidelines and applicable standards or codes;

• adequacy of the safety factors used in the design;

• conformance of the bidding documents to the performance requirements of the system;

• adequacy of the QA/QC program developed by the contractor, including monitoring methods and
testing details.

11 - 9
CHAPTER 12.0
REFERENCES

AASHTO. (1997). American Association of State Highway and Transportation Officials, Washington,
D.C.

AASHTO. (1992). Standard Specifications for Highway Bridges, 15th Edition, American Association
of State Highway and Transportation Officials, Washington, D.C.

AASHTO. (1993). "Epoxy coated reinforcing bars, in standard specifications for transportation materials
and methods of sampling and testing, part I: specifications. " M-284, American Association of State
Highway and Transportation Officials, Washington, D.C.

Abramson, and Hansmire. (1988). "Three examples of innovative retaining wall construction." Proc. 2nd
Int. Conj. On Case Histories in Geotechnical Engineering, 6.11, June, St. Louis, 1191-1200.

ACI. (1995). ACI building code requirements for reinforced concrete, 318-95, American Concrete
Institute, Detroit, Mi.

ASTM. (1996). "Concrete and aggregates." Annual Book of ASTM Standards, Vol. 04.02, ASTM, West
Conshohocken, Pa.

Bell, F G (1992). Engineering in Rock Masses, Butterworth Heinemann Ltd., Boston, Mass.

Bowles, J. E. (1988). Foundation analysis and design, Fourth Edition, McGraw-Hill Book Co., New
York, NY.

Brandl, H (1992). "Retaining structures for rock masses." Engineering in Rock Masses, Chp. 26,
pp 530-572, Butterworth-Heinemann Ltd, Boston, Mass.

Bruce, D. A. (1994). "Jet grouting - Chapter 8." Ground control and improvement, John Wiley & Sons,
Inc., New York, NY.

Bruce, D.A. (1994). "Small diameter cast-in-place elements for load bearing and in situ earth
reinforcement - Chapter 6." Ground Control and Improvement, P.P. Xanthakos, L.W. Abramson,
and D.A. Bruce, eds., John Wiley and Sons, New York, NY.

Bruce, D. A. (1992). "Two new specialty geotechnical processes for slope stabilization." ASCE Specialty
Conf on Stability and Performance of Slopes and Embankments - II, Geotechnical Special
Publication No. 31, Vol. 2, June 29- July 1, Berkeley, Calif., Seed, R. B. and Boulanger, R. W.,
eds., ASCE, 1505-1519.

Bruce, D. A. (1992b). "Two new speciality geotechnical processes for slope stabilization." ASCE
Speciality Conf. On Stability and Performance of Slopes and Embankments - II, June 29-July 1,
Berkeley, Ca., 15.

Bruce, D. A. (1989b). "Methods of overburden drilling in geotechnical construction: a generic


classification." Ground Engineering, 22 (7), 25-32.

12 - 1
Bruce, D. A. (1988). "Developments in geotechnical construction processes for urban engineering."
Civil Engineering Practice, 3 (1).

Bruce, D. A., and Jewell, R. A. (1987). "Soil nailing - application and practice." Ground Engineering,
20, (1), Emap Construct Ltd., Middlesex, United Kingdom, 21-32.

Bruce, D. A., and Jewell, R. A. (1986). "Soil nailing - application and practice." Ground Engineering,
19 (8), Emap Construct Ltd., Middlesex, United Kingdom, 10-15.

Bruce, D. A., and Juran, I. (1997). "Drilled and grouted micropiles - State of practice review, Volume
II: Design." FHWA-DTFH61-93-D0128, Federal Highway Administration, Washington, D.C.

Byrne, RJ., Cotton, D., Porterfield, J., Wolschlag, C., and Ueblacker, G. (1996). "Manual for design
and construction monitoring of soil nail walls." Federal Highway Administration, FHWA-SA-96-
069.

Canadian Foundation Engineering Manual. (1992). Third Edition, Canadian Geotechnical Society c/o
Biotech Publishers Ltd., Richmond, British Columbia.

Caquot, A., and Kerisel, F. (1948). "Tables for the calculation of passive pressures." Active Pressure
and Bearing Capadty ofFoundations, Gautheir-Villars, Paris, France.

Castelli, R J., (1991). "Vibratory Deep Compaction of Underwater FilL", Deep Foundation
Improvements: Design, Construction and Testing, ASTM STP 1089, Philadelphia, PA.

Castelli, R. J., Hsu, S.R. and Tanal, V. (1984). "Stabilization of a Cellular Cofferdam Using Stone
Columns." International Conference on In Situ Soil and Rock Reinforcement, Paris.

Castelli, R. J., and Munfakh, G. A. (1986). "Geotextile walls in mountainous terrain." Third International
Conference on Geotextiles, Vienna, Austria.

Castelli, R. J., Sarkar, S. K., and Munfakh, G. A. (1983). "Ground treatment in the design and
construction of a wharf structure." Advances in Piling and Ground Treatment for Foundations,
Thomas Telford, London, England.

Castelli, R J., and Secker, N. (1986). "Port of Kismayo (Somalia) rehabilitation." ASCE Specialty
Conference, Ports 86, Oakland, Ca.

Cheney, R. S. (1988). "Permanent ground anchors." FHWA-DP-68-1R, Federal Highway


Administration, Washington, D. C.

Christopher, B. R., Gill, S. A., Giroud, J. P., Juran, I., Mitchell, J. K., Schlosser, F., and Dunnicliff,
J. (1990). "Design and construction guidelines for reinforced soil structure - Volume 1." FHWA-
RD-89..()43, Federal Highway Administration, Washington, D. C.

Clough, G. W., and Duncan, J. M. (1991). "Earth pressures - Chapter 6." Foundation Engineering
Handbook, Second Edition, H-Y Fang, ed., Van Nostrand Reinhold, New York, NY, 224-234.

12 - 2
Clough, G. W., and O'Rourke, T. D. (1990). "Construction induced movements of in situ walls." Proc.
of Con! on Design and Performance of Earth Retaining Structures, Geotechnical Special
Publication No. 25, Cornell University, Ithaca, P. C. Lambe and L. A. Hansen, eds., ASCE, New
York, NY, 439-470.

Clough, G. W., Smith, E. M., and Sweeney, B. P. (1989). "Movement control of excavation support
systems by iterative design." Proc., Foundation Engineering: Current Principles and Practices,
Vol. 2, ASCE, 869-884.

Clouterre. (1991). "Recommendations clouterre - soil nailing recommendations." French National


Research Project, FHWA-SA-93-026, Federal Highway Administration, Washington, D.C. (in
English).

Contech. (1997). Trade literature.

Coulomb, C. A. (1776). "Essai sur une application des regles de maximis et minimis aquelques problemes
a
de statique, relatifs l'architecture." Mem. Acad. Roy. des Sciences, Paris, France, Vol. 3, 38.

Das, B. M. (1990). Principles offoundation engineering, Second Edition, PWS-KENT Publ., Boston,
Ma., 333.

Dash, U. (1987). "Pin-pile wall evaluation. " FHWA-PA-86-047-84-36, Federal Highway Administration,
Washington, D.C.

DePaoli, B., Stella, C., and Perelli, C. A. (1991). "A monitoring system for the quality assessment
of the jet grouting process through an energy approach." Proc., 4th Int. Con! on Piling and
Deep Foundations, April 7-12, Sttesa, Italy.

Dismuke, T. D. (1991). "Retaining structures and excavations - Chapter 12." Foundation Engineering
Handbook, Second Edition, H. Y. Fang, ed., Van Nostrand Reinhold, New York, NY, 447-510.

DKP. (1989). "Baltimore metro northeast line-shot tower line and station structure, phase 1."
Geotechnical Design Summary Report prepared for the Maryland DOT, Mass Transit
Administration, Baltimore, Maryland.

DoubleWal Corporation. (1993). Trade Literature.

DoubleWal Corporation. (1982). Trade Literature.

Duncan, J. M., Williams, G. W., Sehn, A. L., and Seed, R. B. (1993). "Estimated earth pressures due
to compaction." Journal of Geotech. Engrg. Div., 119 (7), ASCE, New York, NY, 1162-1176.

Duncan, J. M., Williams, G. W., Sehn, A. L., and Seed, R. B. (1991). "Estimated earth pressures due
to compaction." Journal ofGeotech. Engrg. Div., 117 (2) ,ASCE, New York, NY, 1833-1847.

Elias, V. (1997). "Corrosion/degradation of soil reinforcement for mechanically stabilized earth walls and
reinforced soil slopes." FHWA-SA-96-072, Federal Highway Administration, Washington, D.C.

12 - 3
Elias, V., and Christopher, B. R. (1997). "Mechanically stabilized earth walls and reinforced soil slopes
design and construction guidelines." FHWA-SA-96-071, Federal Highway Administration,
Washington, D. C.

Elias, V., and Juran, 1. (1991). "Soil nailing for stabilization of highway slopes and excavations."
FHWA-RD-89-198, Federal Highway Administration, Washington, D. C.

Elias, V., Welsh, J., Warren, J., and Lukas, R. (1996). "Ground improvement technology manual."
Demonstration Project No. 82 - Ground Improvement, FHWA-DP-3, Federal Highway
Administration, Washington, D.C.

Federal Highway Administration. (1988). "Earth retaining structures - review and acceptance procedures."
Geotechnical Guideline No.2, FHWA Geotechnical Engineering Notebook, Washington, D. C.

Fournier, P. (1975). "Poured pile pamper patients." New England Construction, December 1.

Frydman, S., and Keissar, 1. (1987). "Earth pressure on retaining walls near rock faces." Journal of
Geotech. Engrg. Div., 113 (6) , ASCE, New York, NY, 586-599.

Fukuoka, M. (1977). "The effects of horizontal loads on piles due to landslides." Proc., Specialty
Session 10, Ninth Int. Con! on Soil Mechanics and Foundation Engineering, Tokyo, Japan, 27-42.

Gallavresi, F. (1992). "Grouting improvement of foundation soils." Proc., ASCE Con! Grouting, Soil
Improvement and Geosynthetics, 2 Vols., Feb 25-28, New Orleans, La., 1-38.

Goldberg, D. T., Jaworski, W. E., and Gordon, M. D. (1976). "Lateral support systems and
underpinning." VoU Design and construction, April, FHWA-RD-75-128, Federal Highway
Administration, Washington, D. C.

Gorneck, P. B., Bruce, D. A., Greenman, J. H., and Bingham, G. (1993). "Pin piles save silos."
Civil Engineering, September, ASCE, New York, NY.

Gould, J. P. (1990). "Earth retaining structures - developments through 1970." Proc. ofConf on Design
and Peiformance ofEarth Retaining Structures, Geotechnical Special Publication No. 25, Cornell
University, Ithaca, P. C. Lambe and L. A. Hansen, eds., ASCE, New York, NY, 8-21.

Gray, H. (1958). "Contribution to the analysis of seepage effects in backfills." Geotechnique, The Institute
of Civil Engineers, London, England.

Herbst, T. F. (1982). "The GEWI pile - a solution for difficult foundation problems." Symp. on Soil
and Rock Reinforcement Techniques, Nov. 29 - Dec. 3, AlT, Bangkok, Thailand.

Hetenyi, M. (1946). Beams on elastic foundations, The Univ. Of Michigan Press, Ann Arbor, MI, 255.

HKGEO. (1993). Guide to retaining wall design - GEOGUIDE 1. Geotechnical Engineering Office,
Civil Engineering Department, Hong Kong, China.

Holtz, R. D., Christopher, B. R., and Berg, R. R. (1995). Geosynthetic design and construction
guidelines, May, FHWA- H1-95-038, Federal Highway Administration, McLean, Va.

12 - 4
Hunt, S.W, Frank, RL., Tarvin, P.A., and Blazek, J.J. (1994). "Racine's Water Street Retaining Wall."
ASCE Geotechnical Special Publication No. 42, Serviceability of Earth Retaining Structures,
American Society of Civil Engineers, New York, pp 137-154.

Ito, T., and Matsui, T. (1975). "Methods to estimate lateral force acting on stabilizing piles." Soils
and Foundations, 15, (4), 43-59.

Jaky, J. (1944). "The coefficient of earth pressure at rest." Journal of the Society of Hungarian
Architects and Engineers, Vol. 7, 355-358.

Jewell, R A., and Pedley M. J. (1990). "Soil nailing design: the role of bending stiffness." Ground
Engineering, Emap Construct Ltd., Middlesex, United Kingdom, 30-36.

Jones, W. C., Anderson, L. R, Bishop, J. A., Holtz, R. D., and Wu, T. H. (1987). "Reinforcement of
constructed earth." State-of-the-Art Report, Soil Improvement - A Ten Year Update, Geotechnical
Special Publication No. 12, ASCE, New York, NY.

Jumikis, A. R. (1983) Rock Mechanics, Trans Tech Publications, Clausthal-Zellerfeld, BRD

Juran, I., and Elias, V. (1991). "Ground anchors and soil nails in retaining structures - Chapter 26."
Foundation Engineering Handbook, Second Edition, H.Y. Fang, ed., Van Nostrand Reinhold,
New York, NY, 868-902.

Kauschinger, J. L., and Welsh, J. P. (1989). "Jet grouting for urban construction." Proc., Geotechnical
Lecture Series, Boston Society of Civil Engineering: Design Construction and Performance of
Earth Support Systems, MIT, Nov. 18, Boston Society of Civil Engineers Section (ASCE), Boston,
Ma., 8.

Kauschinger, J. L., Hankour, R, and Perry, E. B. (1992). "Methods to estimate composition of jet grout
bodies." Proc., ASCE Con! Grouting, Soil Improvement and Geosynthetics, 2, Feb. 25-28, New
Orleans, La., 194-205.

Kauschinger, J. L., Perry, E. B., and Hankour, R. (1992). "Jet grouting: state-of-the-practice." Proc.,
ASCE Con! Grouting, Soil Improvement and Geosynthetics, 2, Feb. 25- 28, New Orleans, La.,
169-181.

Kerisel, J. (1992). "History of retaining wall designs." Proc. of Con! on Retaining Structures, Institute
of Civil Engineers, C.R.!. Clayton, ed., Thomas Telford, Robinson College, Cambridge, England,
1-16.

Kerr, W., and Tamaro, G. (1990). "Diaphragm walls - update on design and retaining structures."
Cornell University, Ithaca, NY.

Korfiatis, G. P. (1984). "Field testing of short pile systems for floor support." Con! on In-Situ Soil and
Rock Reinforcement, Paris, France.

Lawson, C. R (1992). "Soil reinforcement with geosynthetics." Proc., Workshop on Applied Ground
Improvement Techniques, Southeast Asian Geotechnical Society, Asian Institute of Technology,
Bangkok, Thailand.

12 - 5
Littlejohn, G. S. (1990). "Ground anchorage practice." Proc. of Conf on Design and Performance
ofEarth Retaining Structures, Geotechnical Special Publication No. 25, Cornell University, Ithaca
P. C. Lambe and L. A. Hansen, eds., ASCE, New York, NY, 540-550.

Littlejohn, G. S. (1982). "Design of cement based grouts." ASCE Conf, Grouting in Geotechnical
Engineering, Feb. 10-12, New Orleans, La., 35-48.

Littlejohn, G. S., and Bruce, D. A. (1977). "Rock anchors - state of the art." Foundation Publications,
Ltd., Brentwood, Essex, England, 50.

Lizzi, F. (1982). "The pali radice (root piles)." Symp. On Soil and Rock Improvement Techniques
Including Geotextiles, Reinforced Earth and Modem Piling Methods, Paper Dl, Dec., Bangkok.

Lizzi, F. (1980). "The use of <pali radice> (root pattern piles) in the underpinning of monuments and
old buildings and in the consolidation of historic centres." L'Industria Delle Construction, No.
110, December, Fondedile Corp., Rome, Italy (in English).

Lizzi, F. (1978). "Reticulated Root Piles - To Correct Landslides." ASCE Convention & Exposition,
Chicago, II., October 16-20.

Maccaferri. (1987). Trade Literature.

Mascardi, C. A. (1982). "Design criteria and performance of micropiles." Symp. on Soil and Rock
Improvement Techniques Including Geotextiles, Reinforced Earth and Modem Piling Methods,
Paper D-3, Dec., Bangkok.

Mayne, P. W., and Kulhawy, F. H. (1982). "Ko-OCR relationships in soil." Journal of Geotech. Engrg.
Div., 113, (11), ASCE, New York, NY, 1406-1410.

Mitchell, J. K., and Villet, W. C. B. (1987). "Reinforcement of earth slopes and embankments," NCHRP
Report 290, Transportation Research Board, Washington, D.C.

Module 1: Arman, A., Samtani, N., Castelli, R., and Munfakh, G. (1997), "Training Course in
Geotechnical and Foundation Engineering, Module 1 - Subsurface Investigations," Principal
Investigator: George Munfakh, Report No. FHWA-HI-97-021, NHI Course No. 13231, U.S.
Department of Transportation, Federal Highway Administration, National Highway Institute,
Arlington, Virginia.

Module 2: (1999), "Training Course in Geotechnical and Foundation Engineering, Module 2 - Contracting
and QA/QC." Principal Investigator: George Munfakh, Under Preparation, NHI Course No.
13233, U.S. Department of Transportation, Federal Highway Administration, National Highway
Institute, Arlington, Virginia.

Module 3: (1999), "Training Course in Geotechnical and Foundation Engineering, Module 3 - Soil Slopes
and Embankments." Principal Investigator: George Munfakh, Under Preparation, NHI Course
No. 13233, U.S. Department of Transportation, Federal Highway Administration, National
Highway Institute, Arlington, Virginia.

12 - 6
Module 4: (2000), "Training Course in Geotechnical and Foundation Engineering, Module 4 - Ground
Improvement Techniques," Principal Investigator: George Munfakh, Under Preparation, NHI
Course No. 13234, U.S. Department of Transportation, Federal Highway Administration, National
Highway Institute, Arlington, Virginia.

Module 5: Wyllie, D. and Mah, C. (1998), "Training Course in Geotechnical and Foundation
Engineering, Module 5 - Rock Slopes," Principal Investigator: George Munfakh, Report No. HI-
99-007, NHI Course No. 13235, U.S. Department of Transportation, Federal Highway
Administration, National Highway Institute, Arlington, Virginia.

Module 7: (1999), "Training Course in Geotechnical and Foundation Engineering, Module 7 - Shallow
Foundations." Principal Investigator: George Munfakh, Under Preparation, NHI Course No.
13237, U.S. Department of Transportation, Federal Highway Administration, National Highway
Institute, Arlington, Virginia.

Module 8: (1999), "Training Course in Geotechnical and Foundation Engineering, Module 8 - Deep
Foundations." Principal Investigator: George Munfakh, Under Preparation, NHI Course No.
13238, U.S. Department of Transportation, Federal Highway Administration, National Highway
Institute, Arlington, Virginia.

Module 11: Dunnicliff, J. (1998), "Training Course in Geotechnical and Foundation Engineering, Module
11 - Geotechnical Instrumentation," Principal Investigator: George Munfakh, Report No. HI-98-
034, NHI Course No. 13241, U.S. Department of Transportation, Federal Highway
Administration, National Highway Institute, Arlington, Virginia.

Mononobe, N. (1929), "Earthquake-Proof Construction of Masonary Dams," Proc. World Engineering


Conference, Vol. 9, p.275.

Munfakh, G. A. (1990). "Innovative earth retaining structures: selection, design and performance."
Proc. of Con! on Design and Peiformance ofEarth Retaining Structures, Geotechnical Special
Publication No. 25, Cornell University, Ithaca, P. C. Lambe and L. A. Hansen, eds., ASCE,
New York, NY, 85-118.

Munfakh, G. A. (1989). "Soil stabilization at depth - practical applications." Seminar on "Foundations


in Difficult Soils, State of the Practice." Met. Section, ASCE.

Munfakh, G. A. (1984). "Soil reinforcement by stone columns - varied case applications." International
Conference on In-Situ Soil and Rock Reinforcement, Paris, France.

Munfakh, G. A., Abramson, L. W., Barksdale, R. D., and Juran, I. (1987). "In-Situ ground
reinforcement." ASCE Geotechnical Special Publication No. 12, New York, NY 1-66.

Myles, B., and Bridle, R. J. (1991). "Fired soil nail- the machine." Ground Engineering, July/August,
Emap Construct Ltd., Middlesex, United Kingdom.

NAVFAC. (1982). Design manual 7.2: foundations and earth structures, Dept. of the Navy, May,
U.S. Government Printing Office, Washington, D. C.

Okabe, S. (1926), "General Theory of Earth Pressures," Journal of the Japan Society of Civil Engineering,
Vol. 12, No.!.

12 -7
O'Rourke, T. D., and Jones, C.J.F.P. (1990). "Overview of earth retention systems: 1970-1990." Proc.
of Con! on Design and Performance of Earth Retaining Structures, Geotechnical Special
Publication No. 25, Cornell University, Ithaca, P. C. Lambe and L. A. Hansen, eds., ASCE,
New York, NY, 22-51.

Pacchiosi. (1985). Trade Literature.

Patel, A. And Castelli, RJ. (1992), "Permanent Slurry Walls at Baltimore Metro's Shot Tower Station."
Slurry Walls: Design Construction, and Quality Control, ASTM STP 1129, Philadelphia, PA.

Pearlman, S. L., Campbell, B. D., and Withiam, J. L. (1992). "Slope stabilization using in-situ earth
reinforcements." ASCE Specialty Con! on Stability and Performance of Slopes and Embankments-
II, Geotechnical Special Publication No. 31, Vol. 2, June 29- July 1, Berkeley, Ca., R. B. Seed
and R. W. Boulanger, eds., ASCE, New York, NY, 1333-1348.

Peck, R B. (1969). "Deep excavation and tunneling in soft ground." Proc., 7th Int. Con! on Soil Mech.
and Foundation Engrg., State-of-the-Art Volume, Mexico City, Mexico, 225-290.

Penner, E. (1970). "Frost heaving forces in leda clay." Canadian Geotechnical Journal, Vol. 7, 8.

Plumelle, C. (1984). "Amelioration de la portante d'un sol par inclusions de froupe et. reseaux de
micropieux." Int. Symp. on In Situ Soil and Rock Reinforcement, Paris, Oct., 9-11, 83-90.

Porterfield, J. A., Cotton, D. M. and Byrne, R J. (1994). "Soil nailing field inspectors manual." FHWA-
SA-93-068, Federal Highway Administration, Washington, D. C.

Post-Tensioning Institute (1996) "Recommendations for Prestressed Rock and Soil Anchors", Phoenix,
Arizona.

Rankine, W. M. J. (1857). "On stability of loose earth." Philosophic Transactions of Royal Society,
Part I, London, England, 9-27.

Reinforced Earth Company. (1988). Trade Literature.

Rowe, P. W. (1957). "Sheet pile walls in clay." Proc., Inst. of Civil Engrg., Part I, Vol. I., London,
England.

Rowe, P. W. (1955). "Sheet pile walls encastre at anchorage." Proc., Inst. of Civil Engrg., Part I,
Vol. I., London, England.

Sabatini, PJ., Elias, V., Schmertmann, G. R, and Bonaparte, R (1996). "Earth retaining systems."
Geotechnical Engineering Circular No.2, Federal Highway Administration, Washington, D.C.

Schlosser, F. (1982). "Behavior and Design of Soil-Nailing," Int. Symp. on Recent Developments in
Ground Improvement Techniques, Nov. 29 to Dec 3, Asian Institute of Technology, Bangkok, pp.
399-413.

Schmidt, B. (1966). Discussion of, "Earth pressures at-rest related to stress history." Canadian
Geotechnical Journal, 3 (4), 239-242.

12 - 8
Seismic Design ofHighway Bridges. (1986). (NHI Course No. 13048), FHWA/93/040, Federal Highway
Administration, Washington, D. C.

Shen, C. K., Hermann, L. R., Romstad, K. M., Bang, S., Kim, J. S., and De Natale, J. S. (1981). "An
in situ earth reinforcement lateral support system." Dept. of Civil Engrg., Univ. of California,
Davis, Rep. No. 81-03, US National Science Foundation Rep. No. NSF/CEE-81059, March.,
130.

Simac, M. R, Bathurst, R J., Berg, R. R, and Lothspeich, S. E.(1996). Design manual for segmental
retaining walls, Second Edition, Collin, J. G., ed., National Concrete Masonry Association, 7.

Skempton, A. W., and Hutchinson, J. (1969). "Stability of natural slopes and embankment foundations."
Proc., 7th Int. Conf Soil Mech. and Found. Engrg., 3, 291.

SNAIL PROGRAM. (1991). "A user's manual." California Department of Transportation, Division of
New Technology, Materials and Research, Office of Geotechnical Engineering.

Spangler, M. G., and Handy, R L. (1984). Soil engineering. Harper and Row, New York, NY.

Stocker, M.F., Korber, G.W., GassIer, G., and Gudehus, G. (1979). "Soil nailing." Int. Conf on Soil,
Reinforcement, Paris, France, 469-474.

Taki, 0., and Yang, D. S. (1991, 1989). "Excavation support and groundwater control using soil-cement
mixing wall for subway projects." Proc. RETC, June 11-14, Los Angeles, Calif., R A. Pond
and P. B. Kenny, eds., Society for Mining, Metallurgy, and Exploration, Inc (SME), Littleton,
Colorado, 164-165.

Tamaro, G. J. (1990). "Slurry wall design and construction." Proc. ofConf on Design and Peiformance
ofEarth Retaining Structures, Geotechnical Special Publication No. 25, Cornell University, Ithaca,
P. C. Lambe and L. A. Hansen, eds., ASCE, New York, NY, 540-550.

Tamaro, G. J., and Gould, J. P. (1993). Conf. on new developments in urban geotechnical engineering
processes, University of Wisconsin-Milwaukee, Nov. 15-16, Atlanta, Ga.

Tamaro, G. J., and Poletto, R J. (1992). "Slurry walls - construction quality control." Slurry Walls:
Design, Construction, and Quality Control, ASTM STP 1129, David B. Paul, Richard R Davidson
and Nicholas J. Cavalli, eds., American Society for Testing and Materials, Philadelphia,
Pennslyvania.

Task Force 27. (1990). "Design guidelines for mechanically stabilized earth (MSE) walls in permanent
applications." AASHTO-AGC-ARTBA Joint Committee, Washington, D. C.

Teng, W. (1962). Foundation design, Prentice-Hall, Inc., Englewood Cliffs, NJ.

Terzaghi, K., and Peck, R B. (1967). Soil mechanics in engineering practice, John Wiley and Sons, New
York, NY.

USS Steel Sheet-Pile Design Manual. (1975). July, ADUSS.

12 - 9
Vidal, H. (1966). "La Terre Armee." Annales de L'Institute Technique du Batiment et Travaux Publics,
Vol. 19, 223-224.

Watts, G. R. A., and Brady K. C. (1990). "Site damage trials on geotextiles." Proc., Fourth Int. Con!
on Geotextiles, Vol. 2, Balkema, 691-696.

Weatherby, D. E. (1982). "Tiebacks." FHWAIRD-82/047, Federal Highway Administration, Washington,


D.C.

Welsh, J. P., Rubright, R. M., and Coomber, D. B. (1986). "Jet grouting for support of structures."
Proc. ASCE Con! Groutingfor the Support of Structures, April 16, Seattle, Wa.

Weltman, A. (1981). "A review of micro pile types." Ground Engineering, Vol. 14, No.4, May, Emap
Construct Ltd., Middlesex, United Kingdom.

Wu, J. T. H. (1994). "Design and construction of low cost retaining walls, the next generation in
technology." Report No. CTI-UCD-I-94, US Forest Service, Portland, Oregon, Colorado
Department of Transportation and University of Colorado at Denver.

Xanthakos, P. P. (1994). Slurry walls as structural systems, Second Edition, McGraw-Hill, New York,
NY, 2.

Xanthakos, P. P., Abramson, L. W., and Bruce, D. A. (1994). Ground control and improvement, John
Wiley & Sons, Inc., New York, NY.

12 - 10

You might also like