You are on page 1of 38

26

GROUND ANCHORS AND SOIL NAILS IN


RETAINING STRUCTURES
ILAN JURAN, D.Sc.
Professor and Head
Department of Civil and Environmental Engineering
Brooklyn Polytechnic University

26.1

INTRODUCTION

Ground anchor and soil nail retaining systems are designed to


stabilize and support natural and engineered structures and to
restrain their movement using tension-resisting elements. The
basic design concept consists of transferring the resisting tensile
forces generated in the inclusions into the ground through the
friction (or adhesion) mobilized at the interfaces. These systems
allow the engineer to efficiently use the in-situ ground in
providing vertical or lateral structural support. They present
significant technical advantages over conventional rigid gravity
retaining walls or external bracing systems that result in
substantial cost savings and reduced construction period.
Therefore, during the past few decades, ground anchors, and
more recently soil nails, have been increasingly used in civil
engineering projects.
The use of these systems in permanent structures requires
careful evaluation of the durability of the structural elements
and assessment of the long-term system performance. Avariety
of inclusions, corrosion-protection systems, and installation
techniques have been progressively developed by specialty
contractors. This chapter briefly describes the construction
process and the main structural elements. It presents the main
aspects of ground-inclusion interaction, illustrates the observed
behavior of instrumented str\lctures, and outlines durability
considerations, performance criteria, and design approaches
that have been developed to ensure the internal and external
stability of these composite retaining systems.
26.2

26.2.1

VICTOR ELIAS, P.E.


V. Elias & Associates, P.A.
Consulting Engineers

structural element. Temporary ground anchors are used for a


specified construction period and their service life is generally
less than 2 years. Permanent ground anchors are corrosionprotected to insure their long-term performance throughout the
design service life of the structure.
Figure 26.1 shows a schematic diagram of a permanent
ground anchor. The basic components of the ground anchor are:

The tendon is made of prestressing steel wires, strands, or


bars and includes :
a. The anchor bond length-where the tendon is fixed in the
primary grout bulb and transfers the tension force to the
surrounding ground. The anchor bond length is designed
to provide the required load pull-out capacity of the
anchor.
b. The unbonded length-where the tendon is free to elongate
elastically transferring the resisting force from the anchor
bond length to the structural element (i.e., wall face, slab,
etc.). It is designed to reach the underlying substratum or, in

PRINCIPLES, HISTORICAL DEVELOPMENT,


AND FIELDS OF APPLICATION
Permanent Ground Anchors

Permanent ground anchors are prestressed cement-grouted


tendons used in soils or rock to restrain and control the
displacements of structural elements such as walls or slabs.
They have been developed mainly by specialty contractors
involved in temporary excavation support systems and in some
cases are proprietary. The anchors are installed in drilled holes
and prestressed to the design load in order to mobilize and
transfer the required resisting force from the ground to the

868
H.-Y. Fang (ed.), Foundation Engineering Handbook
Springer Science+Business Media New York 1991

Fig . 26.1

Permanent ground anchors.

Ground Anchors and Soil Nails in Retaining Structures 869


homogeneous soils, to locate the anchor bond length
beyond the potentially unstable soil mass adjacent to the
structural element.
The anchor grout, also called primary grout, is generally a
portland cement-based mixture or a polymer resin and is
used to transfer the anchor force to the ground. Secondary
grout can be injected into the drilled hole after stressing to
provide corrosion protection for unsheeted tendons.
The anchorage is a device attached to the tendon that consists
of a plate and an anchor head (or threaded nut) and permits
stressing and lock -off of the prestressing steel.
During the past 50 years, permanent ground anchors have
been extensively used by contractors to provide vertical and
lateral support for natural and engineered structures. Typical
applications of ground anchors are illustrated in Figure 26.2.
They have found widespread acceptance in a variety of civil
engineering projects including cut slope retaining systems,
tied-back diaphragm or soldier pile walls, bridge abutments,
stabilization of natural slopes and cliffs, tunnel portals, underpinning, repair or reconstruction of quay walls, dam spillways,
loading ramps, hangars, etc. They have also been frequently
used as tied own supports for dams, transmission towers, and
waterfront structures, primarily to resist uplift water pressures
and rotational loadings.
Tiebacks were first used to anchor structures in rock. The
earliest permanent rock tied owns were installed by the French
engineer Coyne for anchoring the Jument lighthouse (1930)
and raising the Cheurfas Dam, Algeria (1934). By the late 1950s,
use of permanent rock tiedowns had become common practice
in renovation and construction of dams (Evans, 1955; Morris,
1956; Middleton, 1961) and towers (Weatherby, 1982). In the
1950s contractors began to use tiebacks for temporary supports

.D

of deep excavations. The first permanent soil tiebacks in the


United States were installed in 1961 in a very stiff silty clay for
the construction of retaining walls for the Michigan expressway
(Jones and Kerkhoff, 1961). However, in spite of long-term
European experience, permanent ground anchors had not been
in common use in the United States until the late 1970s, mainly
because of engineering concerns with regard to long-term
performance, potential time-dependent (creep) movement, corrosion protection of the tendon, and the need to establish
reliable quality control testing procedures to verify the shortand long-term holding capacity. Technological efforts have been
continuously invested by specialty contractors to overcome
these limitations, develop efficient corrosion-protection systems,
improve grouting methods and installation procedures, and
increase the tension capacity of the prestressed tendons.
The rapid acceptance and growing use of ground anchors
can be attributed mainly to significant technical advantages
resulting in substantial cost savings and reduced construction
period. Specifically, in urban areas the use of ground anchors
often allows significant reduCtion in right-of-way acquisition
and permits the elimination of temporary support systems,
external bracings, or the need for underpinning existing structures
near to excavation sites. The increasing confidence in ground
anchor use for permanent structures is primarily due to reliable
quality control procedures that involve routine performance
and proof testing of all production anchors under loads
exceeding the design load. Performance specifications and codes
of practice, based on experience and long-term observations of
permanent anchor installations, have been developed in European countries (French Recommendations, Bureau Securitas,
1977; FIP Rules, 1974; German Standards, DIN, 1972, 1976;
PTI Recommendations, 1980) and more recently in the United
States (FHWA; see Cheney, 1984) to specify design, construction,
and monitoring procedures.

Wall

Fractured
sandstone ........... /

Permanent

tiebacks

Permanent
tiebacks

(b)

(a)

Existing dam

',.

"

'-

... . F. ..

Permanent
tiedown
(c)

(d)

Fig.26.2 Typical applications of permanent ground anchors. (a) Concrete wall. (b) Landslide and tunnel portal. (c) Permanent tower
tiedown. (d) Dams. (After Weatherby, 1982.)

870

Foundation Engineering Handbook

26.2.2 Soil Nailing


Soil nailing is an in-situ soil reinforcement technique that has
been used during the past two decades, mainly in France and
Germany, in cut slope retaining systems and slope stabilization.
The fundamental concept of soil nailing consists of reinforcing
the ground by passive inclusions, closely spaced, to create in
situ a coherent gravity structure and thereby to increase the
overall shear strength of the in-situ soil and restrain its
displacements. This technique has emerged essentially as an
extension ofthe "New Austrian Tunneling Method" (Rabcewicz,
1964-65), which combines reinforced shotcrete and rockbolting
to provide a flexible support system for the construction of
underground excavations.
Although soil nailing technology is relatively new, it has
been used in a variety of civil engineering projects including
stabilization of railroad and highway cut slopes (Rabejac and
Toudic, 1974; Hovart and Rami, 1975; Stocker et aI., 1979;
Cartier and Gigan, 1983; Schlosser, 1983); construction of
excavation retaining structures in urban areas, for high-rise
buildings and underground facilities (Louis, 1981; GassIer and
Gudehus, 1981; Shen et aI., 1981); landslide stabilization
(Guilloux et aI., 1983; Blondeau et aI., 1984); tunnel portals in
steep and unstable stratified slopes (Louis, 1981); and other
civil and industrial projects. Typical applications of soil nailing
are illustrated in Figure 26.3. Several nailed soil-retaining
structures have been instrumented to establish a data base for
evaluation of structure performance and development of reliable
design methods (Stocker et aI., 1979; GassIer and Gudehus,
1981; Schlosser, 1983; Plumelle, 1986). In North America, the
system was initially used in Vancouver, B.C., in the late 1960s
in temporary excavation supports for industrial and residential
buildings (Shen et aI., 1981). Presently, soil nailing systems can

be considered for any temporary or permanent application


where conventional cut retaining systems, such as cast-in-place
reinforced-concrete walls or tied-back walls, are applicable. As
demonstrated by GassIer and Gudehus (1981), soil-nailed
retaining structures can withstand both static and dynamic
vertical loads at their upper surface without undergoing excessive
displacements. Therefore, they can be effectively used in the
construction of bridge abutments. Soil nailing also appears
to provide an efficient and economical technique for repair and
reconstruction of existing structures, particularly tie-back walls
and reinforced soil retaining structures.
In soil-nailed retaining structures, the inclusions are generally
steel bars or other metallic elements that can resist tensile
stresses, shear stresses, and bending moments. They are either
placed in drilled boreholes and grouted along their total length
or driven into the ground. The nails are not prestressed but
are closely spaced (e.g., one driven nail per 2.5 ft2, one grouted
nail per 10- 50 ft 2) to provide an anisotropic apparent cohesion
to the native ground. The facing of the soil-nailed structure is
not a major structural load-carrying element but rather ensures
local stability of the soil between reinforcement layers and
protects the ground from surface erosion and weathering effects.
It generally consists of a thin layer of reinforced shotcrete (4- to
6-in thick), constructed incrementally from the top down. The
facing and the nails are placed immediately after each excavation
stage to restrain ground decompression and therefore to prevent
deterioration of the original mechanical properties and shear
strength characteristics of the native ground. Prefabricated or
cast-in-place concrete panels have increasingly been used in the
construction of permanent structures to satisfy specific aesthetic
and durability design criteria.
As with ground anchors, soil nailing has been primarily used
for temporary retaining structures. This is mainly due to the

----(b)

(a)

(i) Conventional
(ii) Austrian tunneling
method
--method

I'"

///t t
Reinforced
concrete

Anchor pin

2f}[
~
TESTB

~t

..~ .. 1

20 ft

(c)
Fig. 26.3

(d)

Typical applications of soil nailing. (a) Landslide. (b) Retaining structures. (c) Tunnel portal. (d) Abutments.

Ground Anchors and Soil Nails in Retaining Structures 871


engineering concerns with regard to durability of metallic
inclusions in the ground and shortcomings offacing technology.
In recent years, technological developments have been mostly
focused on producing low-cost corrosion-protected nails and
prefabricated concrete or steel panels to overcome these limitations. Soil nailing has now become a common construction
method in European countries and over 100 temporary and
permanent retaining structures have been constructed. In the
United States, the engineering use of this technology for
permanent structures is currently growing with increasing local
experience followed by development of standard specifications
for design, construction, quality control, and monitoring of
soil-nailed structures.

26.3 TECHNOLOGY, CONSTRUCTION PROCESS,


AND STRUCTURAL ELEMENTS
The main components of an anchored (or nailed) ground
retaining system are the in-situ ground, the tension-resisting
inclusions (anchors or nails), and the facing or the structural
retaining element. The economy of the system is predominantly
dependent upon the technology used (i.e., structural elements
and installation process of inclusions) and the construction rate
achieved.

26.3.1
A.

Inclusions and Installation Techniques

Ground Anchors

A variety of anchors have been developed by specialty contractors


using different anchor tendons, drilling methods, grout control
procedures, and corrosion-protection systems. Selection of

anchors and installation process for a specific application will


generally depend upon the subsurface soil (or rock) type,
groundwater conditions, site restrictions, and availability of
equipment. Figure 26.4 illustrates schematically the four basic
types of commonly used ground anchors.
Low-pressure-grouted straight-shafted ground anchors are
used in most types of soils. They are commonly installed using
hollow-stem augers or tremie grouting under low (or no) grout
pressure.
Hollow-stem augers are extensively used in the United States
in both granular and cohesive soils. The anchor installation
process includes three basic stages: insertion of the tendon into
the auger, drilling, and grouting through the auger while it is
being extracted.
Low-pressure-grouted anchors are usually tremie-grouted
(grout pressures lower than 150 psi) in drillholes that are cased
in cohesionless soils but can be open in competent rocks or
clayey soils. In granular soils, drilling is generally performed
by driving a casing (4 or 6 inches in diameter) and using air
or water flushing to remove the soil.
Cohesive soils are mostly rotary drilled without casing.
Bentonite and cement slurries may be used to prevent caving
of the drill holes.
Pressure-injected anchors are installed in sandy or gravelly
soils under grout pressures exceeding 150 psi. Drilled or driven
casing (3.5 or 3 inches in diameter) with a solid closure point
may be used to seal the hole and to allow the bond zone to be
grouted under high pressure during the extraction of the casing.
Postgrouting technique is generally used in both granular
and cohesive soils to enlarge the primary grout bulb by multiple
stages of controlled high-pressure grouting. An inflatable bag
or packer is used to isolate the bond length and allow the grout
to be pressurized through single or multiple postgrouting
phases. A special grout tube (called tube-a-manchette) using a
double packer is designed to allow repetitive grouting through

Tendon

Tendon

Grouting pressure = p
Effective increase in
diameter: grout
permeati~n, ground
compactIon

"-~~

Lateral friction
+ low end bearing

Lateral friction

(b)

(a)

0, D'

Ground root
system

Mechanical interlocking
(lateral friction + end bearing)

(c)

(d)

Fig. 26.4 Interaction mechanisms in ground anchors. (a) Tremie-grouted straight-shafted (rocks, stiff clays). (b) Low-pressure-grouted
anchors. (c) High-pressure-grouted anchors. (d) Underreamed anchors (stiff to hard cohesive clays).

872

Foundation Engineering Handbook

valves located along the bond zone. Typical examples of


regroutable anchors are the Soletanche IRP anchor (Jorge,
1969), the TMD-Bachy anchor (Clement and Navarro, 1972),
and the TPT anchor (Mastrantuono and Tomiolo, 1977). The
Soletanche IRP anchor and the tube-a-manchette packer are
schematically illustrated in Figure 26.5.
Single and multi-underreamed anchors are used in stiff cohesive
soils and weak rock. Their installation process involves the use
of an underreaming device that is basically a cutting tool. First,
the cylindrical anchor shaft is drilled, usually using a continuous
flight auger. Then, the cutting tool is mechanically expanded
at a controlled rate to the design size of the underream. The
soil is removed by water flushing; neat cement grout is
tremie-grouted into the drill hole; and the tendon is inserted.
Depending on the underreaming device, several underreams

B.

I Steel tendon
Sleeve-pipe

2 Plastic sheath
~~~SI

can be cut simultaneously. The spacing between the underreams


is selected to induce a shear failure along a cylinder passing
through the tips of the underreams.
The main structural element of each ground anchor is the
steel tendon, which may consist of bars, wires, or strands.
Strands and wires have advantages with respect to tensile
strength (ultimate tensile strength: 270ksi for strands and
240 ksi for wires), and ease of transportation, storage, and
fabrication. However, bars (ultimate tensile strength of 150 to
160 ksi) are more readily protected against corrosion and, in
the case of shallow, low-capacity anchors, are usually easier
and cheaper to install. Often, availability and cost will be the
determining factors. In the United States, bars and seven-wire
strands are the most commonly used tendons. High-capacity
tendons made of 18 strands with a diameter of 0.50 or 0.60
inch are also available for high-capacity tied own applications.

IT-""'--3 Annular space (epoxy- or

cement-filled)

Anchor

Strand
Rubber manchette

Grouting pressure distends rubber


manchelle and forces grout through
sealing grout
sleeve grout to secure tube
it manchette in hole
(b)

Fig.26.5 (a) Schematic section of an IRP anchor. (b) Detail of


tube-a-manchette for pressure grouting control. (After Pfister et al.,
1982.)

Soil Nailing

The steel reinforcing elements used for soil nailing can be


classified as (a) driven nails, (b) grouted nails, (c) jet-grouted
nails, and (d) corrosion-protected nails.
Driven nails, commonly used in France and Germany, are
small-diameter (15 to 46 mm) rods or bars, or metallic sections,
made of mild steel with a yield strength of 350 MPa (50 ksi).
They are closely spaced (2 to 4 bars per square meter) and
create a rather homogeneous composite reinforced soil mass.
The nails are driven into the ground at the designed
inclination using a vibropercussion pneumatic or hydraulic
hammer with no preliminary drilling. Special nails with an axial
channel can be used to allow for grout sealing of the nail to
the surrounding soil after its complete penetration. This installation technique is rapid and economical (4 to 6 per hour).
However, it is limited by the length of the bars (maximum
length about 20 m) and by the heterogeneity of the ground (e.g.,
presence of boulders).
Grouted nails are generally steel bars (15 to 46 mm in
diameter) with a yield strength of 60 ksi. They are placed in
boreholes (10 to 15 cm in diameter) with a vertical and
horizontal spacing varying typically from 1 to 3 m depending
on the type of the in-situ soil. The nails are usually cementgrouted by gravity or under low pressure. Ribbed bars can be
used to improve the nail-grout adherence, and special perforated
tubes have been developed to allow injection of the grout
through the inclusion.
let-grouted nails are composite inclusions made of a grouted
soil with a central steel rod, which can be as thick as 30 to 40 cm.
A technique that combines the vibropercussion driving and
high-pressure (greater than 20 MPa) jet grouting has been
developed recently by Louis (1986). The nails are installed (Fig.
26.6) using a high-frequency (up to 70 Hz) vibropercussion
hammer, and cement grouting is performed during installation.
The grout is injected through a small-diameter (few millimeters)
longitudinal channel in the reinforcing rod under a pressure
that is sufficiently high to cause hydraulic fracturing of the
surrounding ground. However, nailing with a significantly lower
grouting pressure (about 4 MPa) has been used successfully,
particularly in granular soils. The jet-grouting installation
technique provides recompaction and improvement of the
surrounding ground and increases significantly the pull-out
resistance of the composite inclusion.
Corrosion-protected nails generally use double protection
schemes similar to those commonly used in ground anchor
practice. Proprietary nails have recently been developed by
specialty contractors (Intrafor-Cofor; Solrenfor) to be used in
permanent structures. These corrosion-protection schemes are
described in a later section of this chapter.

Ground Anchors and Soil Nails in Retaining Structures 873

Fig. 26.6a Construction process of a soil-nailed wall illustrating


excavation, shotcreting, and nailing.

Fig. 26.6b Jet bolting: installation of reinforcing elements. 1.


Vibropercussion hammer. 2. Sliding support. 3. Reinforcement to
be inserted. 4. Sliding guide. 5. Fixed guide. 6. Soil to be treated.

26.3.2

Facing and Structural Retaining Elements

In an anchored retaining system the wall has a major structural


role. It has to resist the tensile forces transferred by the anchors,
the lateral pressure applied by the retained soil, and bending
moments. The wall has to be stiff enough to restrain the ground
displacement induced by the excavation process. The facing in
a nailed soil-retaining system has only a minor mechanical role.
The maximum tensile forces generated in the nails are significantly greater than those transferred to the facing. The main
function of the facing is to ensure local stability of the ground
between the nails and to limit its decompression. Hence, the
facing has to be continuous, fit the irregularities of the cut slope
surface, and be flexible enough to withstand ground displacement
during excavation. The structural elements used to build the
anchored wall are therefore basically different from those used
to construct the facing of a nailed soil-retaining structure.
A.

Structural Elements of an Anchored Wall

An anchored wall can be constructed with a wide variety


of structural elements, using different installation techniques.
Selection of the structural element for a specific application will
generally depend on the subsurface soil (or rock) type, groundwater conditions, local construction practice, availability of
material and equipment, and performance requirements. The

structural elements can be evaluated in terms of their stiffness,


ease of handling and installation, durability, water-tightness or
continuity, and ease of removal. The elements commonly used
can be broadly classified into four major categories: driven
sheet piles, soldier piles and lagging walls, cylinder walls, and
concrete diaphragm or slurry walls. Typical properties of each
system are indicated in Table 26.1.
Sheet-pile walls usually consist of interlocking steel sheets
driven into the ground prior to excavation. They are fairly
impervious and easy to handle and install in soft clays,
cohesionless silts, or loose sands. However, they are difficult to
use in compact granular soils containing cobbles or boulders.
As compared with other elements, they are relatively flexible
and the wall displacement will in general be larger. They
are commonly used for marine bulkhead construction (see
Chapter 12).
Soldier piles and lagging (Figure 26.7a) usually consist of
steel H-beams that are either driven into the ground or placed
in predrilled boreholes prior to excavation. Concrete bored piles
with reinforcement or permanent casing have also been used.
As excavation proceeds, the ground between these piles is
retained by lagging of wood planks, cast-in-place, or precast
concrete elements. H-beam soldier piles and lagging walls are
probably the excavation support system most widely used in
the United States for temporary supports. They are easy to
install in most types of soils, and present a significant advantage
specifically in compact or irregular strata that would obstruct
sheet piling. They can be readily adapted to different site
conditions and irregular wall alignments. The main disadvantage
of this retaining system is that the wall is rather pervious and
subsurface water flow may cause local instabilities. A properly
lagged wall should permit drainage, draw down, and fluctuation
of water level without flow of the retained soil.
Cylinder walls consist of an array of cylindrical caissons that
are usually constructed of reinforced concrete or mixed-in-place
soil-cement and are closely spaced to form a continuous wall.
They can be cast-in-place and installed using several techniques
such as hollow-stem augers, rotary drilling equipment, deep
mixing methods, or jet-grouting. Depending on the stiffness of
the individual cylinders, such a wall may be rigid enough to
support lateral loads with limited deflection. To achieve watertightness and properly retain the soil, shotcrete or lagging in
the space between the cylinders may be required. Alternatively,
the cylinders can overlap to produce a continuous, impervious
wall. In addition to their rigidity, cylinder walls offer the
advantage of adaptability to irregular site alignments and can
be used in a variety of ground conditions.
Slurry walls or concrete diaphragm walls are generally
formed in a trench supported by viscous mud slurry (see Chapter
20). Concrete is tremied into the trench, displacing the mud
slurry upward. Reinforcement of the wall is made by vertical
steel sections, precast reinforced-concrete members, or cages of
reinforcing steel. Recent developments include the use of precast
concrete panels. These walls can be designed to achieve a
specified degree of stiffness and water-tightness, and can be
integrated in the permanent structure. They are often used
where lowering of the water table would adversely affect
adjoining structures. Their main disadvantage is the relatively
high cost and the need for specialized construction equipment
and experienced contractors. They also may present environmental problems pertaining to slurry disposal.
Cast-in-place reinforced-concrete panels have been used in
the construction of multi tied-back walls (Kerisel et aI., 1981).
Figure 26.7b shows a schematic cross section of a 30-m deep
open excavation retained by ten layers of prestressed anchors.
This anchored wall was constructed from the top down with
successive stages of (1) excavation, (2) in-place casting of the
reinforced-concrete panel (2-ft thick, 9-ft high), and (3) anchoring.

874

Foundation Engineering Handbook

TABLE 26.1 PROPERTIES OF STRUCTURAL WALL ELEMENTS.


Properties

System

~.

EI-KSF/F
x 10 3

Moment
KF/F

Depth
Range, ft

3 to 50

10 to 125

15 to 70

Fair

Readily available
Effective in soft ground
Low cost

3 to 40

7 to 70

15 to 60

Controlled
by lagging

Ease of installation in
competent ground
Readily available
Low cost

70 to 800

100 to 400

20 to 60

Poor to
fair

Common technique
Can be stiffened by
adding core
Can be widely spaced
Water-tightness can be
improved by
overlapping

350 to 1600

30 to 400

20 to 100

Good

High strength
Durable
Can be permanent wall
High cost

15 to 260

20 to 100

Good

High strength
Durable
Can be permanent
Higher cost

Watertight

Technical Features

VERTICAL STEEL SHEETING


SWF to 14WF al 6ft to 8ft c-<;

.~... : ..

~'i'"

.. ik

.........

,:~

SOLDIER BEAM A D CONCRETEI


WOOD LAGGI G

d=

ISin to
3(iin
YLI DER PILES-TANGENT

:<: '. :.'.:' '":.;. '.: ... r. .:;'f.


tT'
I =

.,

24in 10 36in

....

SLURRY WALL - STE


I

.'

.,. .....

B AM REINFORCEME T

= 24in to 36in

f::.~.'::~
/. >'.<.":')-:.:.~:' .<::".>:'.:~.~. i
~L..:-=-'':';,"-~'':~''':''''L;'''';':':''::L'' j~

300 to 1000

SLURRY WALL- REINFORCING BARS

For deep excavations this multianchored wall offers some


technical advantages as compared with a diaphragm wall: (1)
precasting of concrete prior to excavation is not required; (2)
as each panel is anchored, wall thickness is significantly smaller;
(3) adaptability to variation in ground conditions; and (4)
effective control of the wall performance during excavation by
properly adjusting the individual panels.
B.

Facing in Nailed Retaining Systems

Depending on the application and soil (or rock) type, four types
of facing are presently used.
Shotcrete facing (10 to 25 cm thick) is currently used for
most temporary retaining structures in soils. This facing technology provides a continuous, flexible surface layer that can
fill voids and cracks in the surrounding ground. It is generally
reinforced with a welded wire mesh (Fig. 26.7c) and its required
thickness is obtained by successive layers of shotcrete (each 9
to 12 cm thick). This technique is relatively simple and inexpensive, but it may not provide the technical quality and aesthetics

required for permanent structures. In particular the durability


of the shotcrete facing can be affected by groundwater, seepage,
and environmental factors such as climatic changes, e.g. freezing,
which may induce cracking. In addition, with shotcrete facings,
provision of efficient drainage at the concrete-soil interface is
difficult.
Welded wire mesh is generally used to provide a facing in
fragmented rocks or intermediate soils (chalk, marl, shales).
Concrete and steel facings: Cast-in-place reinforced-concrete
is frequently used for permanent structures.
Prefabricated concrete or steel panels have also been adapted
for permanent structures. These panels can be designed to meet
a variety of aesthetic, environmental, and durability criteria.
They also provide efficient means of integrating continuous
drainage behind the facing. Figure 26.8 shows Solrenfor metallic
panels for inclined facing (Louis, 1986). The rectangular steel
panels are bolted together and soil nails are installed through
their common corners. Also shown in Figure 26.8 is a prefabricated concrete panel with continuous geotextile drainage.
Concrete panels have also been combined with prefabricated
steel panels and cast-in-place concrete.

26m

C = 0 <\>'

= 32

26m
2
3
E
~

a:N

24m
28m

C' =lO kP a
$' = 32

28m

28m

28m

30 m
C= 20 kP a
$' = 28

Schematic eroS
(b)

----

eetion

(b)
fig . 2.7 (,) So ldi "
pH " "d logging ,n, ho
" .,,11. (C ou "." Ni,ho
/'on con,,"uwon Co.) (b)
L" .,I m "" open "" ""
Mo nte Carlo. (Kerisel
io n in
et 81., 1981 .)
cont.

87 6

6"

S H O iC
R E iE
f/l.CI!-lG
_GRO

Ui

f'/l-O /l
-$ REQ
\.\IRE'.O
","OR,/lR f't>
cl( , 0
,N O S
COl'f('/l-I
1/l-C;E
N
GROIft.
CO~iEO
N/l-I\..

GO. G".
~e ?\..A1
E
E.I'O'J.'I
' CO/)
.1.0

SII.~

c lJ
& f lE i i O f li
IJ)

~I,H EPO~'plI.INi
I' PII.tJ\

lSi S
iA

" ,0 @ G E GRO\.)"r
>' <Ie
,O

."e

l c ) con
strUcti

f i g . 26 1
.

~a)

fig

d 1'
on e ta l
01

, (b) .
" " . ., "
', , .

26 8 I a) Pletabli
cated ste
. \
el pane
s.

R ."
. o r ,"
. ' so
O ' " ov
<o"r.<r
<'

s h o tc r e
a

,,0'"'' P 'o

"'~" " ,e

te tacin
g

'" , o d
0'"

'00 0""""'\

e,u,""

SO/,,,'O

,-)

Ground Anchors and Soil Nails in Retaining Structures 877


26.3.3

where Sa is the uniaxial compressive strength. In cohesive soils,

Drainage

Groundwater (see Chapters 1 and 7) is a major engineering


concern relative to construction of anchored and nailed soilretaining structures. An appropriate drainage system must be
provided to (a) prevent generation of excessive hydrostatic
pressures on the facing (or the structural wall element), (b)
protect the facing element and particularly shotcrete facing from
deterioration induced by water contact, (c) prevent saturation
of the nailed ground, which can significantly affect the structure
displacement and may cause instability during and after excavation. In anchored walls, prefabricated vertical drains, porous
engineering fabrics, or subhorizontal drains can be used for
drainage ofthe subsurface flow. In soil nailing, shallow drainage
(plastic pipes, 10 cm in diameter, 30 to 40 cm long) is usually
used to protect the facing, while subhorizontal slotted plastic
tubes are used for deep drainage of the nailed ground. In the
case of permanent structures with prefabricated panels, a
continuous drain such as a geotextile can be placed behind the
facing.
26.4

SOIL-INCLUSION INTERACTION:
PULL-OUT CAPACITY ESTIMATES

'ult

Figure 26.4 illustrates the basic soil-inclusion interaction


mechanisms for the main types of ground anchors.
Tremie-grouted straight-shafted anchors, which are more
commonly used in rock and very stiff to hard cohesive soils,
generate their pull-out resistance through the lateral shear
mobilized at the grout-ground interface. The pull-out capacity
of these anchors is often estimated by

= n;'D'L"ult

= lO%'Sa

for Sa < 600 psi

(26.1c)

25

.::

20

"<i!

0.

:i:.-..

i.E

15

"-II

"0.,~

o 0. 10
-Z;:.;;:

;.::.
E

(26.1)

where 'ult is the ultimate lateral shear stress at the ground-grout


interface (also called shaft friction), D and L are, respectively,
the effective diameter and bond length of the grouted anchor.
It should be noted that the effective anchor diameter D is
difficult to estimate, since it is highly dependent upon ground
porosity and grout permeability. It is commonly assumed that
in competent rocks (Littlejohn and Bruce, 1975)
'ult

cfJ

~'"
""-.,i

Load Transfer in Ground Anchors

= p' A . tan

where p is the effective grout pressure, cfJ is the internal friction


angle of the soil, and A is a dimensionless empirical coefficient
smaller than 1 (Hanna, 1982). For practical applications p is
generally limited to less than 50 psi or 2 psi per foot of
overburden (Littlejohn, 1970).
High-pressure grouted anchors are installed under effective
grout pressures exceeding 150 psi often using postgrouting
techniques or pressure injection (feasible only in cohesionless
soils). Figure 26.9, by Jorge (1969), illustrates the significant
effect of grouting pressure on the ultimate load-transfer rate
(or ultimate lateral shear stress) of multiphase postgrouted
anchors in different types of soils. The high-pressure grouting

"'z
~~

26.4.1

(26.1b)

where IX is an adhesion factor, and S. is the average undrained


shear strength of the soil.
The adhesion factor (IX) generally varies (Tomlinson, 1957;
Peck, 1958; Woodward et aI., 1961) within the range of 0.3 to
0.75, with the lower values obtained for stiffer and harder clays.
Low-pressure grouted anchors are installed under an effective
grout pressure lower than 150 psi (or, in cohesive soils, under
pressure that would not fracture the ground) most commonly
using hollow-stem augers or tremie grouting, an open hole in
cohesive soils or cored rotary-drilled holes in cohesion1ess soils.
The grouting pressure will induce an increase of the effective
diameter of the grout bulb by permeation or local compaction
ofthe ground. Therefore, the pull-out resistance ofthese anchors
is highly dependent upon the grout pressure. It is primarily
derived from the ultimate interface shear stress but an endbearing resistance can be mobilized owing to an effective
increase of the grout bulb diameter. The pull-out resistance is
commonly estimated using Equation 26.1.
For cohesionless soils,
'ult

The load-transfer mechanisms between a grouted anchor (or


nail) and the subsurface soil (or rock) as well as the ultimate
pull-out capacity depend upon several parameters, including
installation technique, drilling and grouting method, grouting
pressure, size and shape of the grouted inclusion, engineering
properties of the in-situ soil and specifically its relative density
(or overconsolidation ratio), permeability, and shear strength
characteristics (see Chapter 3).
The grain size and porosity of the in-situ soil govern the
grout conductivity. In sands, gravels and weathered rock, with
hydraulic conductivities of 10- 1 to 1O- 2 cm/sec, grout will
permeate through the pores or natural fractures of the ground.
In fine-grained cohesionless soils (silts and fine sands), with
hydraulic conductivity smaller than 10- 3 cm/sec, the grout
cannot penetrate the small pores but rather compacts locally,
under pressure, the surrounding ground. Increasing the grout
pressure will induce a greater grout permeation into the ground
and/or a more effective ground densification. Consequently,
under high-pressure grouting, high radial stresses are locked
into the soil surrounding the anchor, increasing its pull-out
capacity.

= IX' S.

(26.la)

o~----~~----~----~

Grout pressure, psi


(I psi = 6.9 kPa)
Fig. 26.9 Influence of grout injection pressure on the ultimate
bearing capacity of anchors. 0. Medium sand. D. Alluvium and
marl. Alluvium sand. L.. Soft chalk. A. Gypsum marl. (After
Jorge. 1969.)

878 Foundation Engineering Handbook

2000

.><

Gravelly
sand

- U = dr:Jd
I--

1800

...

1600

SPT

Density

Typeo/soil

lO

= 1.6/0.16

Sandy gravel

N~,

blows/30 em

Very dense

120

Dense

60

Medium dense

43

Loose

II

> 130

1400

'"
'"<.>

1200

... 11.3 = Diameter of grouted body D = 11.3 em

'u

Q.

U= 15/0.3

Very den e

01)

'>, 1000

.......

'"u
'"
..Q

-0

800

600

'5E

400

;;;

Sandy gravel
U=5to10
dense

===

Gravelly and
U=8tolO

=-=.,.,,;,.,

to coarse sand
(with gravel)
U = 3.5to 4.5

~~~~and
~~~ Medium

200

Diameter of grouted bodies


D = 10to IS em

Bond-to-ground length L , m
Fig.26.10

Ultimate load holding capacity of anchors in sandy-gravel and gravelly sand. (After Ostermayer and Scheele, 1977.)

results in a grout root (or fissure) system that mechanically


interlocks with the surrounding ground, increasing substantially
the pull-out capacity of the anchor. Particularly, in dense
granular soils this interlocking phenomena generates high
tendency for the soil to dilate, which in turn results in a normal
stress concentration at the grout-ground interface. The effect
of pressure injection on the soil-anchor interaction is difficult
to evaluate. Empirical relationships were provided by Ostermayer (1974) for estimating the ultimate lateral shear stress for
high-pres sure-grouted anchors, with and without postgrouting,
in fine-grained soils (sandy silts to highly plastic clays). Ostermayer and Sheele (1977) developed empirical curves, reproduced
in Figure 26.10, to estimate the ultimate pull-out capacity of
pressure-injected anchors in granular soils as a function of
anchor length, soil type, density, and uniformity. These curves,
derived from 30 pull-out tests on anchors installed under grout
pressures of about 70 psi, illustrate that the ultimate capacity
of the anchor is not proportional to its length.
Underreamed anchors, which are mainly used in stiff to hard
cohesive soils derive their pull-out capacity from adhesion along
their shaft above the underreams, end bearing of the first
underream, and lateral shear along a cylinder established by
the tips of the underreams. For the cylinder to be effectively
established, the spacing between the underreams should not
exceed 1.5 times their diameter (Bassett, 1977). Estimate of the
pull-out capacity of these anchors (Littlejohn, 1970) is based
on empirical formulas that are conventionally used for pile
design in cohesive soils.
The load transfer along pressure-injected and high pressure
postgrouted anchors has been investigated by several authors
(Bustamente, 1975, 1976; Ostermayer and Sheele, 1977; Shields
et aI., 1978; Bustamente, 1980; Davis and Plumelle, 1982). Figure
26.11 illustrates, for ultimate pull-out loads, the distributions
of the lateral interface shear stress along pressure-injected
anchors in gravelly sands of different densities (Ostermayer and
Sheele, 1977). Similar results were reported for postgrouted
anchors (Bustamente, 1972) in river sands (Fig. 26.12a) and for
straight-shafted anchors (Feddersen, 1974) in highly overconsolidated, stiff, plastic clays. Figure 26.12b shows the results

Bond/engtll
L,.,m

Soil density
0

Very dense

2.0m
4.5m

Dense

3.0m

Medium
dense

2.0m
4.5m

Loose

2.0 m
4.5m

'..."

1300
1200

\Loo
1000
ME

Z
.><

900

...3

700

.2

600

:E

SOO

u
c

:.;;;

Very dense
~ ----- -1 --

800

.~axl'

0,

-:

til

Medium

den e
_" . _ _ _~O--== -,;;;.:;.;-" -,-

100

""
max l, = mean "
~'D,~... Loose
~--~
------ ------_ ...

-----"'_~ ...

max t, "" mean "

Length. m

L,. = 2.0 m

L,. = 3 .0 m

L, = 4.5 m

Fig. 26.11
Distribution of the lateral interface shear stress along
pressure-injected anchors at the ultimate load. (After Ostermayer
and Scheele, 1977.)

Ground Anchors and Soil Nails in Retaining Structures 879

train
200

400

E.<>

Pull-o ut load distribution

8
9

10
Displacement V, mm
Mobilization of lateral friction

(a)

(b)

Fig.26.12 (a) Distribution of deformation along the length of an IRP anchor. (After Bustamante, 1972) . (b) Mobilization of the lateral
friction along an anchor in a plastic clay. (Winnezeele, Bustamente, 1980.)

of a pull-out test on an instrumented anchor in a plastic clay


(Bustamente, 1980). The slope of the tension force distribution
along the anchor corresponds to the lateral interface shear stress
mobilized at a specific depth under the applied pull-out force.
As shown in Figure 26.12, the shear stress-upward anchor
displacement curves obtained for different depths indicate
overconsolidation of the subsurface soil layer and illustrate that
the anchor displacement required to fully mobilize the ultimate
shear stress is about 5 to 10 mm. The results of these studies
demonstrate that in dense granular soils and highly overconsolidated clays the load-transfer rate along the anchor is not
constant and the pull-out capacity is therefore not proportional
to the anchor length.
The variation of the load-transfer rate along the anchor is
mainly the result of its extensibility during pull-out testing. It
is primarily dependent upon the relative rigidity (or elastic
modulus ratio) of the anchor and the grout-soil interface and
is particularly pronounced in highly dilatant stiff soils. Wernick
(1977), Schlosser and Elias (1978), and Plumelle and Gasnier
(1984) have shown that the restrained tendency of the soil to
dilate during shearing results in a normal stress concentration
at soil-inclusion interfaces that affects significantly the load
transfer rate. As shown in Figure 26.13, vertical stresses as high
as four times the overburden pressure were measured in a
medium dense river sand at the anchor interface. The higher
the (anchor-to-soil) elastic modulus ratio the more uniform is
the load-transfer rate. Figure 26.11 shows that in loose to
medium dense granular soils the interface lateral shear stress
is approximately constant along the anchor. It is of interest to

indicate that these results are consistent with those obtained


for ribbed metallic strips in Reinforced Earth structures
(Schlosser and Elias, 1978).
Several authors have attempted to analyze the load transfer
along the anchor using the "t-z" method (Coyle and Reese,
1966; Davis and Plumelle, 1982), which is commonly applied
in design of friction piles, or more complex interface soil models
(Zaman et al., 1984; Frank et al., 1982). However, rational
anlysis of the ground-anchor interaction requires appropriate
interface properties that are difficult to estimate.

100

200

Distance, d, cm

Fig. 26.13 Restrained dilatancy effect around a ground anchor.


(After Plumelle and Gasnier, 1984.)

880 Foundation Engineering Handbook

26.4.2 Soil-Nail Interaction

In soil nailing, similarly to ground anchors, the load transfer


mechanism and the ultimate pull-out resistance of the nails
depend primarily upon soil type and installation technique.
The pull-out resistance of driven nails in a dense granular
soil was correlated by Cartier and Gigan (1983) with design
recommendations for Reinforced Earth walls (Schlosser and
Segrestin, 1979). These recommendations use the concept of an
"apparent friction coefficient" that is derived from Equation
26.1 assuming
(26.1d)

.2 4
U
:E

'0 3
C
Q)

'u
it:..,
0

C
..,

:;;

<

y is the unit weight of the soil


h is the overburden height above the nail
/1* is the apparent friction coefficient

o
Depth of nail, cm

As shown in Figure 26.14, the apparent friction coefficient (/1*)


obtained from pull-out tests in a nailed granular wall corresponds
to the design value generally used for the ribbed metallic strips
in Reinforced Earth walls. At relatively low depth, owing to
the restrained dilatancy effect, the value of /1* is significantly
greater than 1 and it decreases with depth to tan (4) ).
Laboratory-scale pull-out tests were conducted by Elias and
Juran (1988) in a medium dense sand to evaluate the effect of
the nail installation process on the apparent friction coefficient.
Figure 26.15 shows that the construction process for Reinforced
Earth (i.e., placing the nail during the construction and compacting the sand around the nail) produces a substantially
higher apparent friction coefficient than nailing by driving the
nail into the compacted sand embankment. In the latter case,
nail driving will significantly reduce the restrained dilatancy
effect on the pull-out resistance. Therefore, design guidelines
for Reinforced Earth walls cannot be extrapolated to soil-nailed
structures.
Grouted nails are generally gravity-grouted. Their pull-out
resistance is therefore expected to be approximately the same
as that of an equivalent straight-shafted anchor installed under
low (or no) grout pressure. The drilling of the borehole for the
grouted nail produces an unloading of the disturbed surrounding
soil that can significantly affect its mechanical properties. The
soil-nail interaction is primarily dependent upon soil recompaction due to grouting. In cohesionless soils, grouting pressures
of 50 to 100 psi are commonly used to prevent caving as the
casing is withdrawn. This grouting pressure will induce ground
t/(kN/m)
10
20

-,- t/AVERAGE
17 kN/ml

_ _

;;

c-

Q.
Q.

where

J;: 5

Z,m
SOIL: Sand <\> = 33 c
NAILS: Driven profile

= 10 kPa

(a)

-F

--

_
10

recompaction associated with grout penetration into permeable


gravelly seams, thereby increasing substantially the pull-out
resistance of the nail. Apparent friction coefficient values as
high as 3 to 6 have been reported (Elias and Juran, 1988).
Figure 26.16a shows a cross-sectional view of an excavated

10

Fig. 26.15 Laboratory pull-out test results. (1) Nails placed


during backfilling. (2) Nails inserted during excavation. (After Elias
and Juran, 1988.)

Specifications for
reinforced earth

tan<\>

Z,m

Fig.26.14 Soil-reinforcement friction between a driven nail and


a granular soil. (After Cartier and Gigan, 1983.)

(b)

Fig.26.16 (a) Cross-sectional view of an excavated grouted nail


in a granular soil, illustrating effect of grout permeation. (b)
Sectional view of an excavated nail in silty clay soil.

Ground Anchors and Soil Nails in Retaining Structures 881


~L-I

!mum:

D = 2in

200

100 kN/rn Lille Chack

150

-'"

100

summary of pull-out test results obtained with low-pressuregrouted nails in different types of soils.
Jet-grouted nails are installed under a grout pressure that
can exceed 20 MPa and is sufficiently high to cause hydraulic
fracturing of the surrounding ground (Louis, 1981). Similarly
to high-pressure grouting of anchors, the jet-grouting installation technique produces a mechanical interlocking between the
penetrating grout and the surrounding ground that results in
a substantial increase of the effective nail diameter. It also
provides recompaction of the surrounding ground that significantly improves the pull-out resistance of the composite nailedsoil inclusion. Field pull-out tests on jet-grouted nails (Louis,
1986) yielded ultimate lateral shear stress values as high as
400 kPa in sands and 1000 kPa in sandy gravels.
26.4.3

50

o
L , rn

Fig.26.17 Variation of pull-out resistance of grouted nails with


embedment length. (After Louis, 1986.)

nailed soil, illustrating grout permeation into an alluvial soil


grouted under pressure ofless than 70 psi. Figure 26.16b shows
that in a fine-grained cohesive soil the tremie grouting results
in a rather smooth soil-inclusion interface. The presence of
water at the interfaces, specifically in plastic soils, will generate
a lubrification effect, decreasing substantially the pull-out
resistance of the nail. Figure 26.17 (Louis, 1986) shows a

Estimates of Pull-Out Cpacity from In-situ


Tests

To date, estimates of the pull-out resistance of anchors and


nails are mainly based upon empirical formulas (or ultimate
lateral interface shear stress values) derived from field experience.
These formulas are useful for feasibility evaluation and preliminary design. Table 26.2 provides a summary of estimated
ultimate interface lateral shear stress (or ultimate load-transfer
rate) values for soil nails and ground anchors as a function of
soil (or rock) type and installation technique. Recently, increasing
attempts have been made to develop field correlations between
the ultimate lateral shear stress ('t ult) and the engineering
properties of soils obtained from commonly used in-situ tests
such as the Standard Penetration Test (Fujita et aI., 1977) or
the self-boring pressuremeter test (Bustamente, 1975, 1976).
Recognizing apparent similitude between the soil response to
high-pressure anchor grouting and to the expansion of a

TABLE 26.2 ESTIMATED ULTIMATE INTERFACE LATERAL SHEAR


STRESS VALUES FOR GROUND ANCHORS AND SOIL NAILS.
Ultimate Lateral Shear Stress, kips/ ft
Grouted Nails
Construction
Method

Soil Type

Soil Nailing
(Elias and Juran, 1988)

Permanent Ground
Anchors
(Cheney, 1984

Rotary drilled

Silty sand
Silt
Piedmont residual

2 to 4
1.2 to 1.6
1.5 to 2.5

5 to 9

Driven casing

Sand
Dense sand/gravel
Dense moraine
Sandy colluvium
Clayey colluvium

6
8

7 to 13
10 to 20

Jet grouted

Augered

Fine sand
(medium dense)
Sand
Sand / gravel
Soft clay
Stiff to hard clay
Clayey silt
Calcareous sandy clay
Silty sand fill

8 to 12
2 to 4
1 to 2

3.5 to 4.5 b

20

4.5 to 8.5 b
8.5 to 11 .5b

0.4 to 0.6
0.8 to 1.2

2 to 4

1 to 2

1.5 c

4 to 6
0.4 to 0.6

Cheney recommends a safety factor of 2.5 With respect to the ultimate lateral shear stress values indicated
in thiS table.
b Values obtamed for pressure-injected anchors by Jorge (1969).
C Design value proposed by Weatherby (1982) for hollow-stem augered anchor (assuming a diameter of 6
mches) in both sandy and clayey soils.

882

Foundation Engineering Handbook


(a)
0.8

0.7

00

0.5

0.4

~o

. :
<;>

.~

~y:

0~8,Q"""""~

\-t9 .. Y,"'/o/
.... v ~
..... lOXar

y>y......
'od.,,<
.
S:>'6"/ ,

o.

0.2

.. "

0.3

0.1

0.6

:2

00

..

;;.-"000

Type IRS:

Bustamante et al.
Fujita et al.
Type IGU: ... Bustamante et al.

~-...
Loose

Dense

Very dense

40

20

o Ostermayer and Scheele


K Koreck
v Ostermayer

60

80

100

120

SPT(N)
(b)

IRS

0.3

.7.

c..
:2 0.2

..

o.

~-.'"

~ o

vType-IRS:v-.- Bustamante et al.


~
....... _ '1'" "'t:>.~v-:r
01";~~0"'..
0

0.1

.,. ...._____

~ v.

?-:..""... ... v

'" "'",

"'''''"

1.5

10

15

oOstermayer
Type IGU: ... Bustamante et al. v Ostermayer
'" Jones, Turner, Spencer

0.5

IGU

20
SPT (N)

25

2.5

30

Fig.26.18 Empirical relationships for the determination of the lateral interface shear stress. (a) Lateral interface shear stress for sand and
gravel. (b) Lateral interface shear stress for silty clay soils. (After Bustamante and Doix, 1985.)

pressuremeter cell, the French Central Laboratory of Bridges


and Roads (L.c.P.c.) has conducted an extensive research
program including 94 pull-out tests in 34 sites to provide a
data base for field correlations.
Figure 26.18 shows the empirical relationships derived by
Bustamente and Doix (1985) to estimate the ultimate lateral
shear stress values (Tul l) in different types of soils and rocks as
a function of the limit pressure PI obtained from the pressuremeter
test or the SPT N value. These guidelines take into account
the improvement of the soil surrounding the anchors by different
modes of injection, considering single-stage pressure-grouted
(lGU) anchors (grout pressure of O.5PI < P < PI) and multistage postgrouted (IRS) anchors (grout pressure> PI). Also
shown in Figure 26.18 is the wide scatter of the field data
obtained by the L.c.P.c. and other investigators (Ostermayer,
1974; Fujita et aI., 1977; Ostermayer and Sheele, 1977; Koreck,
1978; Jones and Turner, 1980; Jones and Spencer, 1984) that
have been compiled by Bustamente and Doix to establish these
empirical relationships. The pull-out capacity of the anchor is
estimated using Equation 26.1. The effective anchor diameter
is estimated using a correction factor (a) to allow for diametral
expansion due to highpressure grouting. The a values for IGU

type anchors range from 1.1 in weathered rocks, silty clays, and
fine sands, to 1.4 in highly dilatant granular soils, while the a
values for IRS type anchors range from 1.4 in granular soils
and weathered rock, to 1.8 in stiff clays and marls.
The available field data pertaining to the pull-out capacity
of nails is presently still too limited to substantiate development
of reliable correlations. An attempt has been made by Guilloux
and Schlosser (1984) and Louis (1986) to correlate the measured
pull-out capacity of both driven and grouted nails with the
French recommendations (L.c.P.c. and S.E.T.R.A., 1985) for
the determination of lateral shaft friction on bored and driven
concrete piles from pressuremeter test results. Figure 26.19
shows that in fine-grained soils (i.e., fine sands, silts, non plastic
clays) predicted Tult values correlate reasonably well with
pull-out test results, while in dilatant gravelly soils, compacted
moraine, or fissured rocks they may significantly underestimate
the measured ultimate lateral shear stress.
It appears that further research and field testing could
significantly improve the database for estimating the pull-out
capacity of ground anchors and soil nails. The pressuremeter
test appears also to provide valuable data for grouting procedure, such as the maximum injection pressure that can be used

Ground Anchors and Soil Nails in Retaining Structures


1000

or lateral structural support. They present significant technical


advantages over bracing systems that provide external supports
or conventional gravity retaining structures (e.g., reinforcedconcrete walls, uplift slabs, etc.), which often require substantial
weight to maintain stability. For cut slope retaining structures,
which currently constitute the major application of these
systems, the main technical advantages can be summarized as
follows.

== Range of measured values

Range of calculated values

'"

Po.
~

1. Incorporation of the temporary excavation support system

-r::i
CLl

"3
u

100

'".,

....

.1

100
'ult

1000

Measured, kPa

Fig.26.19 Comparison between measured and estimated values


for ultimate lateral shear stress (Tests results reported by Guilloux

and Schlosser, 1984; Louis 1986.)


1.
2.
3.
4.
5.
6.
7.
8.
9.
10.
11.
12.
13.
14.
15.
16.
17.

883

Driven bars in fine-grained soil.


Grouted bars in fine-grained soil.
Driven bars in granular soil.
Grouted bars in weathered rock.
Grouted bars in soft clay.
Grouted bars in stiff clay.
Grouted bars in marl.
Grouted bars in still marl.
Grouted bars in clayey silt.
Drilled and grouted bars in silt.
Drilled and grouted bars in silty sand.
Driven casing grouted bars in sand.
Driven casing grouted bars in moraine.
Driven casing grouted bars in colluvium clay.
Drilled and grouted bars in marl-limestone.
Drilled and grouted bars in soft rock.
Drilled and grouted bars in fissured rock.

without fracturing the ground. However, in light of the large


variability of parameters affecting the load-transfer mechanism,
specifically in heterogeneous soils, empirical correlations can
only be used for preliminary design. The load-carrying capacity
of each production anchor should be tested according to
established standard testing procedures to assess its performance over the anticipated service life of the structure. In soil
nailing, pull-out tests are required to provide reliable data for
final design and to verify on site the design assumptions by
testing to failure of nonservice witness nails.

26.5 APPLICATION CRITERIA: ADVANTAGES


AND LIMITATIONS

Ground anchor and. soil-nail systems are designed to stabilize


and support natural and engineered structures and to restrain
their movement using tension-resisting inclusions. The basic
design concept consists of transferring the resisting tensile forces
generated in the inclusions into the ground through the friction
mobilized at their interfaces. These systems allow the engineer
to make use efficiently of the in-situ ground in providing vertical

in the permanent facility.


2. Reduction in the amount of excavation and the concrete
work required for footing.
3. Elimination of backfilling behind the wall.
4. Elimination of foundation piles to support the structure in
mountainous areas with unstable slopes or sites underlain
by compressible soils.
5. Reduction in quantities of the reinforced concrete required
for the construction of the retaining wall.
6. Reduction in construction disturbance and right-of-way
acquisition, which in urban sites may eliminate the need for
underpinning nearby structures.
7. Adaptability to different site conditions and soil profiles,
allowing for cost-effective use in repair and reconstruction
of existing structures.
In addition, soil nailing appears to offer some unique
advantages over tie-back systems that significantly affect the
cost and the construction rate of the structure, including:
1. Relatively rapid and flexible installation process of the

shotcrete facing and the unstressed nails.


2. Use of light and relatively inexpensive structural elements
(i.e., nails and relatively thin shotcrete or concrete facing)
that may be substantially cheaper than ground anchors and
structural elements (i.e., cast-in-place slurry wall, sheet piles,
soldier piles, etc.) used in permanent tie-back walls.
3. Greater flexibility in adapting on site the technology (i.e.,
structural elements and installation technique) and structure
geometry to site conditions and soil profile.
4. Light construction equipment can be used for drilling or
vibropercussion and gravity grouting, which is of particular
interest on sites with difficult access.
5. Soil nailing uses a large number of nails; therefore, failure
of any nail may not detrimentally affect the stability of the
system as would be the case for a conventional tied-back
system.
6. In heterogeneous ground with boulders, dense gravelly soils,
or weathered or hard rock, small-diameter drilling for soil
nailing is more feasible and cost-effective than installation
of soldier piles.
7. Structural flexibility: Nailed-soil structures are more flexible
than conventional reinforced-concrete retaining walls and
can therefore withstand larger total and differential settlements.
8. Resistance to seismic loading: As a coherent yet flexible mass,
nailed-soil structures provide a relatively high degree of
structural damping and therefore appear to be well adapted
for construction in seismically active regions.
The main limitations of ground anchors and soil nailing
technology are:
1. Permanent underground easements are required.
2. In fine-grained soils, effective groundwater drainage systems
may be difficult to construct and to maintain.
3. In plastic clayey soils, creep can significantly affect long-term
performance and structure displacements.
4. In soft cohesive soils, pull-out capacity of inclusions cannot
be economically mobilized.

884 Foundation Engineering Handbook

5. Durability considerations may impose severe limitations on


the use of metallic inclusions in aggressive environments.
. In addition, it should be pointed out that in a nai~ed-s~il
retaining structure a certain soil-to-reinforcement relatIve dIsplacement is required to mobilize their interaction and generate
the required resisting nail forces. Therefore, in urban areas the
potential use ofthis technique can be limited by the requirement
to prevent movement of structures in the imm~diate vicinity of
excavation sites. Monitoring of the structure dIsplacements can
be implemented to verify that they are compatible with the
required performance.

26.6

FEASIBILITY EVALUATION

26.7

To evaluate the feasibility and engineering use of permanent


ground anchors or soil-nailing systems, soi.l conditions and
existing physical constraints have to be consIdered.
The presence of utilities such as subways or other underground facilities and the need to obtain und~rground ease~e~ts
may preclude installation of anchors (or naIls) and can slgmficantly affect the project cost.
Durability considerations require an evaluation of the aggressiveness of the ground and the pore water, particularly when
field observations indicate corrosion of existing structures.
Ground anchors and nailing should not be used for permanent
structures in corrosive soils (e.g., soils with high contents of
cinder, ash or slag fills, rubble fills, industrial or acid mine
wastes, etc.). The soil tests most commonly used to evaluate
ground aggressiveness are electrical resistivity, pH, and .sulfate
concentration. The critical values for ground aggressIveness
commonly associated with ASTM standards are outlined in
FHW A DP-68-IR, Permanent Ground Anchors, and are summarized in Table 26.3.
Assessment of the suitability of the subsurface soil (or rock)
to provide short- and long-term pull-out capacity of the anc~or
(or nail) requires a determination of its engineering propertIes,
..
specifically, shear strength and creep characteristics.
In rock, the overall strength is controlled by the eXlstmg
joints or discontinuity system. Highly fractured rocks with open
joints or cavernous limestone are difficult to grout and therefore
potential use of ground anchors or soil nails should be carefully
assessed.
Permanent ground anchors and soil nails generally cannot
be cost-effectively installed in loose granular soils with SPT
blow count number (N) lower than 10 or relative density lower
than 0.30. Nailing becomes practically unfeasible in poorly
graded cohesionless soils with a uniformity coefficient of less
than 2. In such soils, nailing would require stabilization of the
cut face prior to excavation by grouting or slurry wall construction.
In fine-grained cohesive soils, long-term pull-out performance
of the anchors (or nails) is a critical design criterion. Permanent
ground anchors and soil nails are, generally, not feasible in soft

TABLE 26.3 FEASIBILITY CRITERIA WITH REGARD


TO GROUND AGGRESSIVENESS.
Test
Resistivity
pH
Sulfate
Chlorides

ASTM Standard
G-57-78 (ASTM)
G-51-77 (ASTM)
California DOT test 407
California DOT test 422

cohesive soils with undrained shear strength smaller than 0.5 tsf,
or soils susceptible to creep. A number of national codes
(German Standards and French Recommendations) index the
creep susceptibility to the Atterberg limits and natural moisture
content of the soil. They preclude the use of permanent ground
anchors in organic soils, and plastic clayey soils with liquid
limit (LL) greater than 50 and liquidity index (LJ) greater than
0.2 (or consistency index (Ic) less than 0.9). Soils with a plasticity
index (PI) greater than 20 must also be carefully assessed for
creep. In light of the limited experience with soil nailing in
clayey soils, the applicability criteria developed for ground
anchors are recommended for feasibility evaluation of soilnailed structures.

Critical Values
Below
Below
Above
Above

2000ohm/cm
4.5
500 ppm
100 ppm

SHORT- AND LONG-TERM PERFORMANCE


OF ANCHORS AND NAILS

The effective load transfer from the anchor to the surrounding


ground requires a relative displacement between these tW?
components of the retaining system. For ground anchors, thIs
relative displacement is generated by prestressing the anchor
immediately after installation. In the passive soil nails, resisting
forces are generated owing to ground displacement during the
construction. Evaluation of the short- and long-term performance
of ground anchors and nails requires determination of their
load-displacement-time behavior for the specific application
and site conditions. Short-term performance is defined by a
time-independent load-displacement relationship, while an
assessment of the long-term performance should account for
the effect of time-dependent phenomena such as creep and
relaxation.
26.7.1

Short-term Performance

A static loading of anchors or nails can cause several "shortterm" failure mechanisms:
a. Failure of the steel tendon or nails.
b. Shear failure of the soil mass owing to insufficient depth of
anchor embedment.
c. Failure of the grout-tendon or nail bond.
d. Failure of the soil-grout bond.
The engineering design of the anchored (or nailed) retaining
system for specific application and site conditions should
provide a proper selection of the inclusion (i.e., mechanical
properties, length, inclination, spacings, and corrosion protection) to prevent any of these failure modes.
(a) Selection of tendon or nail section should insure that the
working stress in the inclusion does not exceed its ultimate
tensile strength with an acceptable factor of safety. The Post
Tensioning Institute (PTI) recommends limiting the working
tensile stress in prestressed steel to 60 percent of the ultimate
tensile strength for permanent structures and 80 percent for
temporary applications.
(b) To prevent a shear failure of the shallow soil mass
overlying the upper anchors, the bond zone should be located
at a minimum depth of embedment that is generally of the
order of 15 ft (4.6 m). This embedment length should also permit
high-pressure grouting without damage to existing facilities.
(c) To insure that the strength of the ground is fully mobilized
the grout-tendon (or nail) bond should not be exceeded. The
mechanism of grout-tendon bond involves three components:
adhesion, friction, and mechanical interlock. The neat cement

Ground Anchors and Soil Nails in Retaining Structures 885


grout generally used provides steel-cement bonding values of
the order of 1 to 2 MPa, mainly owing to the mechanical grout
interlocking against irregularities in the tendon surface (i.e.,
ribbing of the bar, threading of rebars, or stranding of wires).
(d) Failure of the ground-grout bond will result in sliding
of the anchor or nail. The bonded anchor or nail length should
be designed to ensure that the force mobilized in the inclusion
does not exceed its pull-out capacity with an acceptable factor
of safety. The empirical relations currently employed to estimate
the pull-out capacity (or the ultimate lateral shear stress) of
anchors or nails can only be used for a preliminary design.
Pull-out tests on soil nails are required to provide reliable data
for final design, and for anchored walls each production anchor
should be tested to ensure its load-carrying capacity throughout
the anticipated service life of the structure. For practical reasons,
a minimum bond length of 15 ft is generally required for ground
anchors in soils. Experience has shown that bond lengths
exceeding 40 ft do not efficiently increase the anchor capacity.
Anchor inclination should be as small as possible. However,
steep inclinations may be dictated by practical considerations,
such as right-of way constraints, buried utilities, and soil profile.
A minimum inclination of about 10 to 15 is generally required
to facilitate and insure effective grouting, particularly under
low pressure. Higher inclination of tieback anchors will result
in a transfer of significant vertical forces to the structural element
(i.e., concrete wall or soldier pile) and is generally used only to
reach deep bearing strata or to avoid existing structures.

26.7.2

long-Term Performance

Long-term performance of anchors or nails depends primarily


upon the potential of the ground-inclusion system to creep.
Theoretically, creep can develop in the three basic components
of the system: the ground surrounding the bond zone, the grout,
and the steel (i.e., tendon and/or connections). However, in
practice, creep deformations of the cement grout and the steel
are found to be insignificant, while fine-grained clayey soils may
undergo large creep deformation that will result in timedependent anchor displacement. Large creep displacements
have been reported for multi-underreamed anchors (Ostermayer,
1974) and pressure-injected anchors (Bustamente et aI., 1978;
Bustamente, 1980) in plastic clayey soils. Relaxation of the steel
tendon (i.e., stress decrease under constant strain) can also affect
the long-term performance. However, for a stress level lower
than the elastic limit of the steel the stress loss will generally
not exceed 5 percent of the lock-off stress and its effect on the
displacement will be negligible.
Creep is a time-dependent deformation of the soil structure
under a sustained loading owing to a continuous fabric reorientation. The creep potential of a clayey soil is highly dependent
upon the composition and structure of its minerals, its depositional (preconsolidation) history, and its natural moisture
content (or consistency index). Several investigators (M urayama
and Shibata, 1958; Bishop, 1966; Singh and Mitchell, 1968;
Edgers et aI., 1973) have shown that, as illustrated in Figure
26.20a, for most soils, under a sustained deviatoric stress, the
log of strain rate is linearly decreasing with the log of time.
Singh and Mitchell (1968) reported that the slope m of this
linear relationship appears to be a soil property and is independent of the deviatoric stress level. The m parameter, which can
be obtained from laboratory creep tests, can be used to assess
the creep potential of the soil. Values ofm smaller than 1 indicate
a relatively high potential for accelerated creep associated with
a strength loss that will induce a creep rupture. Bustamente
(1980) showed that Singh and Mitchell's creep theory appears

to consistently describe the observed time-dependent anchor


displacement under a constant load. He therefore suggested
that the creep displacement under a sustained load can be
esitmated using Singh and Mitchell's type equation
AeT

L'll =!1lo + - - (t 1 1- m

m -

1)

(26.2)

where T and L'llo are respectively the applied pull-out force and
the initial displacement prior to creep; A, IX, and m are interface
creep parameters that are obtained from the experimental
log ~l-log t and log f,.l - T curves, and f,./ is the displacement
rate.
Figures 26.20b and c (Bustamente, 1980) illustrate the creep
behavior of an anchor in a plastic clay and the determination
of the relevant interface creep parameters. The test results
indicate a steady increase of the creep displacement almost up
to failure, which is consistent with the m = 1 value derived from
the experimental log f,.l-t curves.
In spite of the apparent similarity between the laboratory
creep test results and the soil-anchor interface creep behavior
observed in situ, more fundamental studies are required in order
to develop a rational creep model for anchors in plastic
fine-grained soils.
In practice, the critical creep load of an anchor or nail is
obtained from a load-controlled pull-out test following a
standard testing and interpretation procedure (DIN 4125,1972,
1974; Bureau Securitas, 1977; Cheney, 1984). The French
standard testing procedure is schematically illustrated in Figure
26.21a. Figure 26.21b shows actual results of a load-controlled
pull-out test on an anchor in a plastic clay (Bustamente, 1980).
It consists of I-hour sustained load increments of O.lF9 (where
Fg is the elastic limit strength of the steel tendon at which
permanent elongation is 0.1 percent). For each load increment
the anchor displacement (s) is plotted versus log time (T). An
upward concavity of the creep curve indicates an accelerated
creep inducing failure. The slope of the s vs.log T line is plotted
against the applied pull-out load to determine the critical creep
load Fc. The allowable anchor working load F w is the smaller
of either 0.9Fc or 0.6F g The loading increment period can
significantly affect the test result. Therefore, a second test is
conducted that includes a 72-hour sustained loading stage at
0.9F w to verify the long-term anchor performance.

26.7.3

Repetitive loading

Anchored structures are often subjected to repetitive (or fluctuating) live loads such as tidal variations, wind or sea wave
loadings, etc. Permanent ground anchors must be designed to
withstand such repetitive loadings throughout the service period
of the structure, which may include millions of cycles. Documented technical data on the long-term performance of anchors
under repetitive loadings are still very limited. Repetitive
loading tests on anchors for a seawall in France showed (Pfister
et a\., 1982) that for peak cyclic load levels smaller than 63
percent of Po (where Po is the ultimate static pull-out capacity)
anchor displacement became negligible after five cycles. However,
for larger cyclic loads anchor displacement continued to increase
at a constant or increasing rate. Begemann (1973) reported that
repetitive uplift loads on steel H-piles in sand under cyclic load
amplitude as low as 35 percent of Po generated progressive
pull-out of the piles. Laboratory model studies of repetitive
loading on plate anchors and friction piles have been conducted
by several investigators (Hanna et a\., 1978; Andreadis et a\.,
1978; Hanna, 1982) and suggest some trends in the anticipated
anchor response to cyclic loading. Specifically, Al-Mosawe
(1979) and Hanna (1982) showed that displacement rate (per

886 Foundation Engineering Handbook


D = deviatoric stress

sc::
~....
c::

"
rJi

0.0002
0.0001
0.00004
0.00002

Time, min
(a)

m= 1

~:::::::~~~:;rcreeiDTm > 1

Typical creep curves

logt
c::

~e

c::

] 10- 1

e
e

10- 3

10-3~~~~__~~~__~~~~__~_

8 .9 10

Tensile force, 102 kN

Time, min

Rate of displacement - load relationship

Rate of displacement - time relationship


(b)

Fig. 26.20 (a) Strain rate vs. time relationship during undrained creep of alluvial clay. (After Murayama and Shibata, 1958.) (b) Modelling
creep of anchors in clays. (Winnezeele, Bustamante, 1980.)

cycle) of plate anchors (Fig. 26.22a) and friction piles under


repeated tensile loads gradually decreased with increasing
number of cycles but did not cease. For a large number of
cycles (Fig. 26.22b), large strains occurred under a cyclic stress
amplitude as low as 25 percent of p . The higher the cyclic
load amplitude, the smaller is the number of cycles required to

generate large strains. Alternating cyclic loading (tension to


compression) accelerates the degradation of the anchor resistance. Prestressing the anchor increases its resistance to repeated
loading.
In granular soils the effect of repeated loading on the anchor
capacity appears to be mainly related to soil densification due

Ground Anchors and Soil Nails in Retaining Structures 887

/; 'f..

/; 'f..

I~.

10

/; 'f..

~il}

71]

(i) Creep curves

6()
~il]

log T

0.1
(ii) Critical creep load
(a)

- - lOOkN
- - - -e- - 200kN
-_e-e-e-e300
ekN""e_-e-e-e ___
E --e-e_-e-e_e-e-e-._e_e_e
E
- e - e-.---e-e--_-____
~kN
e_e
e
e
_
e
.
-e-e
500kN
"u -.-e-.-e_e 600 kN_ _ e-e-e-._e
E
.....
- e _-e-e-e_e_e_
7iXJekN'".-e-e-__-.
<f - - - e _ . 800 kN --.-e_.

C
<Ll

E
<Ll

OJ

i5..

is'"

--e___............. -----------e-e_......---.-.

"

bn
0

e\

::;

"

<:l

.'.900 kN

1:j

c..
<Ll
<Ll
....

2
1
0.5

Displacement - time relations


for each increment of load

Critical creep load Tc

200

400
Pull-out load, kN

(b)
Fig. 26.21 (a) Anchor tension test for determination of critical creep load. (Bureau Securitas, 1977.) (b) Load-controlled pull-out test
on an anchor in plastic clay. (After Bustamente, 1980.)

to particle reorientation, which results in a decrease of the


normal stress at the soil-inclusion interface. In fine-grained
soils of low permeability, the cyclic shear stress may result in
a gradual increase of pore water pressure, decreasing the
effective normal stress at the interfaces. However, further
research and field testing are required in order to develop a
database for a rational evaluation of anchor performance in
clayey soils under low-frequency repetitive loading.
26.7.4

Anchor Testing and Acceptance Criteria

Anchor load tests are conducted following standard testing


procedures (e.g., German DIN 4125, British Standard on
Ground Anchors, French TA 77, or FHWA recommendations

for Permanent Ground Anchors by Cheney, 1984) and are


conventionally classified in three categories:
1. Preproduction tests to evaluate pull-out capacity or critical
creep load and to establish the design load. The testing
procedure is illustrated in Figure 26.21.
2. Performance tests to verify on a limited number of production anchors that: (a) design load may be safely carried, (b)
effective free length corresponds to design requirements, and
(c) residual movement is within tolerable range.
3. Proof tests to verify that the load-deflection behavior of
each production anchor is consistent with the specified
acceptance criteria.
The performance test consists of incrementally applying
cycles of anchor loading and unloading until the reference test

888 Foundation Engineering Handbook


1.0

O.OOOOOll:-_ _+
1

__

__~~----,;;;-l;:=---:;~.

Number of cycles, N
(a)

Number of cycles, N
(b)

Fig. 26.22 (a) Effect of number of cycles on the rate of anchor displacement. (b) Effect of number of load cycles on anchor displacement.
Pu = ultimate pull-out load. (After AI-Mosawe. 1979.)

load is attained. In order to determine long-term creep potential,


each load increment is maintained until measured deflection is
negligible (Le., displacement rate is smaller than a specified
displacement increment per log cycle of time) and a I-hour
creep test is conducted under the reference test load. Cheney
(1984) recommends that the reference test load should be 1.5
times the design working load in cohesionless soils and 1.25
times the design working load in cohesive soils. The performance

tests are conducted on the first anchors (minimum of two


anchors) to verify the selected installation procedure and
provide reference data for the proof tests.
The prooftest consists of a single cycle of incremental loading
to the reference test loads specified above followed by unloading.
Each load increment is maintained until measured deflection
is negligible. The test results are compared with performance
test results on an adjacent anchor.

Ground Anchors and Soil Nails in Retaining Structures

Three acceptance criteria have been established (Cheney,


1984):
1. To ensure that the load transfer reaches the anchor bond

length, the deflection of the anchor head should exceed 80


percent of the calculated elastic elongation of the unbonded
tendon length.
2. The total anchor deflection measured at the maximum test
load should not exceed the calculated elastic elongation of
the tendon length measured from the anchor head to the
center of the bond length. This criterion (not valid for anchors
in layered soils or for underreamed anchors) ensures that
the center of gravity of the bond stress distribution has not
been transferred beyond the midpoint of the bond length.
3. Creep displacement should not exceed 0.08 inches during
the final log cycle of time.
26.8

DURABILITY CONSIDERATIONS

The long-term resistance of permanent ground anchors (or


nails) depends primarily upon their resistivity to corrosion,
ground aggressiveness, and groundwater compositions. Underground steel corrosion is induced when a difference in potential
exists between two points that are electrically connected and
immersed in an electrolyte. Electrochemical cells may develop
between the steel tendon and an external metal element or in
local regions of inhomogeneities within the metal surface of the
same tendon. In either case the chemical reaction between the
groundwater and anodic regions in the exposed steel tendon
results in time-dependent metal loss. For the corrosion process
to occur, oxygen has to be supplied to the metal and therefore
air-moisture solutions, specifically in industrial areas, and soil
layers containing high oxygen content are highly corrosive. The
major variables that affect the corrosion rate are:
1. Ground aggressiveness: organic soils, and acidic or highly
alkaline soils that contain large concentrations of soluble
salts such as sulfates, chlorides or bicarbonates, are highly
corrosive.
2. Groundwater composition: acidic, alkaline, or salt solutions
have high electrical conductivity, inducing high corrosion
rate.
3. Differential aeration: high oxygen concentration (e.g., in fill
or near the ground surface) results in a cathodic environment,
its local variation in the ground generates electrochemical
cells and thereby accelerates the corrosion rate.
4. High stresses or cyclic stresses in the steel tendon accelerate
corrosion and may generate, in anodic environments, brittle
stress-corrosion cracking.
5. Environmental hazards including bimetallic action, large
temperature changes, anerobic bacteria, and stray currents
in the ground (i.e., currents caused by a mass transit facility,
electrical transmission, or transport systems) will generate
a highly corrosive environment.
The corrosion process can develop through different mechanisms, such as uniform surface corrosion,localized pit corrosion,
stress corrosion, corrosion fatigue, and hydrogen embrittlement.
The type of corrosion will significantly affect the degradation
rate of the steel tendon and the efficiency of potential protection
systems.
Documented technical data on the long-term corrosion
performance of ground anchors are very limited since only few
permanent installations have been in service more than 25 years
(Weatherby, 1982). However, performance trends can be anticipated on the basis of extensive research that has been conducted
by the National Bureau of Standards (Romanoff, 1957) on
underground corrosion.

889

A detailed review of corrosion-induced anchor failures by


Weatherby (1982) yielded pertinent conclusions, specifically in
demonstrating that:
1. Quenched and tempered prestressing steels have been involved in a significant number of tieback failures.
2. The unprotected portion of the tendon just behind the anchor
head is highly susceptible to corrosion.
3. All reported failures occurred in the unbonded length of the
tendon, mostly near the anchor head, where poor corrosion
protection (or none) was provided.
Based on these conclusions, FHWA recommendations for
permanent ground anchors (Cheney, 1984) require that all
anchors used for permanent applications be corrosion-protected
in the unbonded length and at the anchor head. For routine
applications, only a single degree of corrosion protection is
required, which may consist of a grease-filled sheath along the
free stressing length and grout cover (minimum 0.5-inch thick)
in the bond zone.
A variety of corrosion-protection systems have been developed. They mostly rely on the following basic principles
(Weatherby, 1982; Hanna, 1982).
Simple corrosion protection relies upon cement grout to
generate a noncorrosive high-pH environment and protect the
tendon in the bond zone. Plastic sheaths filled with anticorrosion
grease, special epoxy pitch, or cement mix, and heat-shrinkable
sleeves are commonly used for corrosion protection along the
free stressing length.
Coating with electrostatically applied resin-bonded epoxy
can be applied to increase the corrosion protection in the bond
zone. Intact resin-bonded coatings, being dielectric, will preclude
the formation of galvanic cells in areas affected by microcracking.
Complete encapsulation of the steel tendon is accomplished
by grouting it into a uniformly corrugated plastic or steel tube
to provide double protection. The annular space between the
tube and the tendon is usually filled with neat cement grout
containing admixtures to control bleeding of water from the
grout.
Compression steel tubes are used by European contractors
to protect the tendon in the bond zone. The tube, which is
high-pressure grouted into the ground, maintains the pressureinjected grout under compression, preventing microcracking.
The unbonded length is generally protected using a grease-filled
PVC tube.
Secondary grouting is commonly used to protect the unbonded
length of the tendon. First, the anchor (primary) grout is tremied
into the bond zone and the tendon is tested and locked-off.
Then, the secondary (antibleed) grout is tremied around the
unbonded length of the anchor, bonding it to the surrounding
ground. Cheney (1984) recommends that secondary grouting
be used only for semipermanent or low-risk applications.
Postgrouting technique can be effectively used to provide
repeated high-pressure grouting in the bond zone with corrosion
protection of the tendon.

For permanent applications of soil nailing, based on current


experience, it is recommended (Elias and Juran, 1988) that a
minimum grout cover of 1.5 inches be achieved along the total
length of the nail. Secondary protection should be provided by
electrostatically applied resin-bonded epoxy on the bars with
a minimum thickness of about 14 mils. In aggressive environments, full encapsulation is recommended. It may be achieved,
as for anchors, by encapsulating the nail in corrugated plastic
or steel tube grouted into the ground. For driven nails, a
preassembled encapsulated nail, shown in Figure 26.23, has
been developed by the French contractor Solrenfor (Louis,
1986).

890

Foundation Engineering Handbook


Protective tube
with injection holes

Fig. 26.23

Welding

"TBHA" nail patented and developed by Solrenfor for

permanent structu res.

Anti-corrosion
grease

Bearing plate
Trumpet

Anchorage

Anchor
head

Anti-corrosion
grease or grout

Fig. 26.24 Anchor head corrosion protection. (After Weatherby,


1982.)

The anchor head is highly susceptible to corrosion, particularly below the bearing plate. It is usually protected by
encapsulation within a plastic or steel cap filled with anticorrosion grease, or cement grout. This encapsulation should
permit prestressing of the tendon and accommodate load
changes in the anchor during its service life. Figure 26.24 shows
anchor head details for multistrand and bar tendons.

26.9
26.9.1

DESIGN OFANCHOREDWALLSAND NAILED


SOIL-RETAINING STRUCTURES
Basic Behavior and Design Concepts

The basic design concept of an anchored or soil-nailed retaining


structure relies upon the transfer of resisting tensile forces
generated in the inclusions (anchors or nails) into the ground
through friction (or adhesion) mobilized at the interfaces. As
illustrated in Figure 26.25, the ground exerts the driving forces
(i.e., lateral earth pressure on the wall or weight of a potentially
sliding soil mass) while providing the anchor bond resistance.
The frictional interaction between the ground and the quasi
"nonextensible" steel inclusions restrain the ground movement
during and after excavation. The resisting tensile forces mobilized
in the inclusions induce an apparent increase of normal stresses
along potential sliding surfaces (or rock joints), increasing the

overall shear resistance of the native ground. The main engineering concern in the design of these retaining systems is to ensure
that ground-inclusion interaction is effectively mobilized to
restrain ground displacements and can secure the structure
stability with an appropriate factor of safety.
In an anchored wall, the resisting tension force is mobilized
by prestressing the anchor to the design working load immediately
after its installation. As excavation proceeds, wall deflection is
mainly controlled by the bending stiffness of the wall, the
prestress load in the anchors, and the anchor longitudinal
stiffness (or elastic modulus).
The effect of prestress anchor loads on wall movement
measured in tied-back walls constructed both in sands and in
clayey soils is illustrated in Figure 26.26 (Clough, 1975). In
spite of the large scatter in the field data, which is mainly due
to the differences in the construction process, structural wall
components, and subsurface soil type, these results demonstrate
that an increase in the prestress level results in a significant
decrease of ground movement. For the sake of comparison,
prestress loads based on earth pressure design values proposed
by Terzaghi and Peck (1967) for braced excavations, as
discussed later in this chapter, are reported.
Figure 26.27 (Goldberg et ai., 1976) illustrates for braced
excavations the effect of the wall bending stiffness, defined as
EI I S~ (where E and I are, respectively, the elastic modulus and
moment of inertia of the wall, Sv is the vertical spacing between
braces), on its lateral deflection in clayey soils characterized by
the stability numberyH ISu (where H is the total structure height
and Su is the undrained shear strength of the soil). Clough and
Tsui (1974) have reported that wall movement and ground
settlement in anchored walls are generally smaller than those
observed in braced excavations.
The design procedure of anchored walls should include the
following steps.
1. Select structural wall element, and for the specified wall type,
estimate the design working prestress loads in the anchors
required to limit ground movement to allowable displacement values.
2. Select anchor type, corrosion protection system, length, and
spacings, and verify that the anchor resistance (tensile
strength and pull-out capacity) is sufficient to withstand the
design working load and testing overloads to ensure longterm performance.
3. Verify that the anchor bond length is located beyond the
potential sliding surface.
4. Verify that the global stability of the retaining system
(ground-anchors-wall) and the surrounding ground with
respect to general sliding along a potential failure is maintained with an acceptable factor of safety.
5. Structural design of the wall element with respect to the
applied system of forces and bending moments.
6. Evaluation of basal stability of the wall elements (i.e.,
required soldier pile penetration, bearing capacity of the
foundation soil below a diaphragm wall, required embedment
depth of a sheet pile, etc.)
7. Select drainage system.

In soil-nailed retaining structures, the reinforcement by


closely spaced passive inclusions results in a composite coherent
material and, as schematically illustrated in Figure 26.25b, the
maximum tensile forces generated in the nails are significantly
greater than those transferred to the facing. The locus of
maximum tensile forces separates the soil-nailed mass into two
zones: an active zone (or potential sliding soil or rock wedge)
where lateral shear stresses are mobilized at the interfaces to
restrain the outward ground movement, and a resistant (or
stable) zone where the generated nail forces are transferred into

Ground Anchors and Soil Nails in Retaining Structures 891

-.. ... .,. Anchor

(a)
Fig. 26.25

:x:

..

1.6

.E

1.4

'0

..c

1.2

.g

1.0

.e -8
c

> c
>(

E 0.
.,"E>
<> "

E
::J

Load-transfer mechanism in ground anchors and soil nails. (a) Ground anchor. (b) Soil nailing.

...

0.6

0.4

0.2

'x

0.65 KA for cj> = 35

yH

~.

--,

:x:

..

<l

1.8

.c

1.6

'0

..c
c

1.4

1.2

5<
>

13
>< -c:

"c: ""

~ ~

0.

Movements generally
<3cm

Flexible -

1fS,., k Im 21m ------- Stiff

Fig. 26.27 Effect of wall stiffness on lateral wall movement. (After


Goldberg et al., 1976.)

Increasing
prestress

1.0

OAy H
'-

0.6
0.4

0.2
0

'"
:::E

Lateral wall denection

0.8

E
E

'x

o SCll lement

(a)

::J

S"

~~
rCdicted by F.E.M.

(Maximum design pressure)/(unit weight x excavation height).


Pm./yH

'"

:::E

Movements generally
>8cm

O~
o __~~~==~~==~~~
0.1
0.2
0.3
0.4
0.5

"E
"c>

10

o Selliement
Lateral wall dcnection

Increasi ng
prestress

0.8

(b)

o
- E
~ E

0. 1

0.2

0.3

0.4

E....:,,"o.r.l
'" .'"
:a-=
c;o

0.5

(Maximum design pressurc)/(unit weight x excavation height),


Pm"lyH

-0.

Co
0_
u

.~

~..c

0_

:r:

(b)

Fig. 26.26 Effect of prestress pressure on wall movement. (a)


Sands. (b) Clays. (After Clough, 1975.)

the ground. Laboratory model tests (Juran et aI., 1984) have


demonstrated that this maximum tensile force line coincides
with the potential sliding surface in the soil.
The soil-nail interaction is mobilized during construction
and structure displacement occurs as the resisting forces are
progressively generated in the nails. Therefore, it has been
essential to monitor actual structures, to measure the facing
displacements in different types of soils and to verify that they
are compatible with performance criteria. Figure 26.28 shows
field measurements of horizontal facing displacements in several
soil-nailed structures (Gassier and Gudehus, 1981; Shen et aI.,
1981; Cartier and Gigan, 1983; Juran and Elias, 1987; Plumelle,
1986). In spite of the differences in the types of inclusions,

Wall height . m
Soil

+ Medium sand

*
0

Silty sand (SM)


Fine sand (SP) 10
clayey sand (SC)
Residual clayey silt
wealhered shale, sandstone
Fontainbleau Sand (SP)

Nail

Driven

Reference

Grouted
Driven

Gassleretal. (1981)
Shcnelal. (1981)
Carlier and Gigan (1983)

Grouted

Juran and Elias (19 6)

Grouted

Plumelle (1986)

Fig. 26.28 Horizontal displacement ot sotl-nailed walls. (After


Juran and Elias, 1987.)

892

Foundation Engineering Handbook

installation techniques, and soil profiles, these data illustrate


that in non plastic soils maximum facing displacement does not
exceed 0.3 percent of the structure height. This ground movement
is comparable to that observed in braced and anchored retaining
systems.
The design procedure of a soil-nailed retaining structure
should include the following steps.
1. For the specified structure geometry (depth and cut slope
inclination), ground profile, and boundary (surcharge) loadings, estimate working nail forces and location of the
potential sliding surface.
2. Select the reinforcement (type, cross-sectional area, length,
inclination, and spacing) and verify local stability at each
reinforcement level, that is, verify that nail resistance
(strength and pull-out capacity) is sufficient to withstand the
estimated working forces with an acceptable factor of safety.
3. Verify that the global stability of the nailed-soil structure
and the surrounding ground is maintained during and after
excavation with an acceptable factor of safety.
4. Estimate the system offorces acting on the facing (i.e., lateral
earth pressure and nail forces at the connections) and design
the facing for specified architectural and durability criteria.
5. For permanent structures, select corrosion protection relevant
to site conditions.
6. Select the drainage system for groundwater piezometric
levels.
The following section outlines the design procedures currently
used for development of earth pressure diagrams, selection of
inclusions, and evaluation of the stability of multi anchored and
soil-nailed retaining structures.
26.9.2

Estimate Working Loads and Structure


Displacements

Several approaches have been developed to estimate the prestress


anchor loads required to limit wall displacements to a tolerable
range. Similar approaches have been proposed for the design
of nailed soil-retaining structures. They can be broadly classified
into four main categories:
a. Empirical design earth pressure diagrams
b. "p-y" lateral soil reaction method
c. Finite-element analyses
d. Kinematical limit analysis method
Notes:
- Vertical cut slope
- Horizontal upper surface

A.

Empirical Earth Pressure Diagrams

Selection of an appropriate earth pressure diagram for the


determination of anchor prestress or nail loads depends upon
the tolerable level of wall and ground movements. Generally,
the allowable displacements of anchored walls are limited to
the range of displacements observed in braced excavations.
Therefore, it is common practice to estimate the prestress loads
using the design diagrams proposed by Terzaghi and Peck
(1948,1967) and Tschebotarioff( 1951) to estimate bracing loads
(Hanna, 1982; Pfister et aI., 1982; Cheney, 1984). These
diagrams are schematically illustrated in Figure 26.29 (also see
Chapter 12). For a multianchor system with uniform anchor
spacings, the design anchor load F w can be expressed as a
normalized, nondimensional parameter:
(26.3 )
at the relative depth of z / H. For sands,
(c/yH < 0.05):

TN

= 0.65K.

For a cohesive soil with both cohesion (c) and friction angle (cfJ):
TN =

K.( 1 - ~. ~)

CJh =

Yh =

lateral earth pressure


overburden pressure

:Terzaghi and
Peck (1967)

..n-ATI'~oC- _ _ _ _ _ J

= 0.65Ka
)0

TN

SAND: _c_ .;; 0.05


yH
Ka = tan 2 (n14 - <1>/2)
Fig. 26.29

(26.3b)

where K. = tan 2 (n/4 - cfJ/2) is the Rankine active earth pressure coefficient, H is the total excavation depth, Sh and Sv are,
respectively, the horizontal and vertical anchor spacings.
Juran and Elias (1987) have shown through analysis of
field measurements obtained on soil-nailed retaining structures
that these earth pressure design diagrams provide a rational
estimate of working tensile forces generated in the nails. On
the basis of the reviewed field data, Terzaghi and Peck's design
diagram for sands has been slightly modified (Fig. 26.29) in
order to calculate nail forces. Figure 26.30 shows nail forces
and structure displacements measured in two instrumented
soil-nailed structures (field data reported by Shen et aI., 1981;
Plumelle, 1986). Table 26.4 summarizes the main characteristics
of these structures. The measured nail forces were found to
agree fairly well with the assumed earth pressure design diagram.
These results illustrate that the observed behavior of nailed cut
slopes is similar to that of braced excavations.
The use of the empirical earth pressure diagrams in the
design of anchored and soil-nailed retaining walls presents some

T
TN
10(

(26.3a)

I 0(

)0

CLAYEY
SAND:
TN = Ka
CLAY: TN

(4C
1- -

1)
,-;- .;; 0.65 Ka
yH vKa

= 0.2 yH --> 0.4 yH

Empirical earth pressure design diagram. (After Juran and Elias, 1987.)

Ground Anchors and Soil Nails in Retaining Structures 893


Horizontal displacement, mm

"'"

0.25

"I

0.50

0.75

1.00

e Inclinometer at
1.5 m from the
facing

;'

yH

I
I
leo
I I i =0.124
I Y
..- <I>=W
o
)

= 0.114

<I> = 33

o Inclinometer at 4.5 m
from the facing
e End of construction
Experimental results
ZIH

H=9.2m

(a)

Horizontal displacement, mm
0

20

TN=

10

0.05

y. H SH Sv
0.10

0.15
of construction
Experimental results:

0
\

oH=3m
eH=5m
.H=7m

\
\

0.25

\
\

,,
,

o.
0.75

1.00

e Facing
displacement
(H = 7 m)
o Facing
displacement
(H = 3 m)

Kinematical
approach

ZIH
(b)

Fig. 26.30 (a) Davis wall-Experimental data and theoretical predictions of tension forces. (b) Full-scale experiment CEBTP-Experimental
data and theoretical predictions of tension forces.

TABLE 26.4 CHARACTERISTICS OF ANALYZED SOIL-NAILED STRUCTURESa .

Structure

Hom

Soil
Classification

Davis wall
(Shen et al..
1981 )

9.2

Heterogeneous
SM

CEBTP wall
(Plumelle.
1986)

7.5

SP

<p.

degrees

c.
kPa

I'.
kN/m3

36.5

18.5

16.3

38

o to

After Juran and Elias (1987).


Facing. reinforced shotcrete.
C Installation technique. A-grouted nails in 10-cm boreholes
B-grouted nails in 6.3-cm boreholes.

15

Type of Nail
#8 rebars

AI. tubes
40 x 1 mm or
30 x 2 mm

L.m

Sh/Sv.

Inclination.b
degrees

Installation
Technique C

1.85
1.85

20

6 to 8

1.15
1.00

10

F,.
kN/m

4.5t05.5

894 Foundation Engineering Handbook

limitations. In particular, these diagrams have been developed


for the conventional cases of bracing supports with the simple
geometry of a vertical wall, horizontal ground surface, and
lateral braces. Therefore, they cannot be used to assess the effect
of varying design parameters such as inclination of the facing
or the wall slope surcharge, wall stiffness, inclination, and
rigidity of the inclusions, etc., on the working forces in the
inclusions and structure displacements. In addition, as shown
in Figure 26.30a, in cohesive soils the empirical earth pressure
diagram is highly sensitive to small variations in soil properties
and is, therefore, difficult to use confidently in design.
B.

"p-y" Lateral Soil Reaction Model

This model is commonly used in design of laterally loaded


structures (piles, foundations, or walls) to assess the lateral
displacement and earth pressure distribution generated by a
given system of external loads (or boundary displacements). It
is based on Winkler's (1867) solution for the elastic bending of
beams assuming that the soil can be represented by a series of
independent elastic springs. At any point of the soil-structure
interface the lateral soil stress p is assumed to be proportional
to the wall deflection y, that is,
p=Kh'Y

(26.4)

where Kh is the modulus of horizontal subgrade reaction.


Nonlinear, elastoplastic "p-y" relationships are generally
used to more adequately represent the soil response. The
elastoplastic reaction model proposed by Terzaghi (1948) for
laterally loaded retaining walls has been adapted (Pfister et a!.,
1982). This model, which relates the earth pressure coefficient
K to the relative wall displacement y/ H, is schematically
illustrated in Figure 26.31. For an assumed "p-y" relationship,
the differential equation of the elastic bending of the wall is
analytically or numerically integrated, yielding wall deflection,
bending moments, and earth pressure distribution.
This design approach permits an evaluation of the effect of
actual wall stiffness and anchor elasticity on the wall movement.
As reported by Pfister et al. (1982), it generally predicts fairly
well earth pressure distributions and wall deflections measured
in actual structures. An iterative procedure is commonly used
to assess the effect of the main parameters and search for design
optimization.
The main drawback of this design approach lies in the
difficulty of determining an appropriate, characteristic "p-y"
relationship (or lateral reaction modulus) for the soil. Several
Idealized
curve
Actual curve

Kp

\./--~-=-~
y

HrW

dd~

K=L
yH

p=ky
Ko
-0.1 -0.2

Compression
Extension
Relative wall displacement
(y/H)

Fig. 26.31

Terzaghi"s idealized relationship used inp-yanalysis

investigators (Baguelin et aI., 1978; Briaud et aI., 1983) have


proposed semiempirical methods for deriving Kh values from
pressuremeter test results. Pfister et al. (1982) provided useful
charts, based on field experience, relating the Kh value to the
shear strength parameters of the soil.
C.

Finite-Element Analysis

The finite-element method has been used by several investigators


to analyze the behavior of anchored walls (Clough et aI., 1974;
Clough and Tsui, 1974; Simpson et aI., 1979; Barla and
Mascardi, 1974; Egger, 1972) and soil-nailed retaining structures
(Shen et aI., 1981; Juran et aI., 1985; Shafiee, 1986). These
analyses involve different constitutive equations for the soil and
interface elements to simulate soil-wall and soil-inclusion
interaction. Attempts have been made by several investigators
(Clough et aI., 1974; Simpson et aI., 1979; Shen et aI., 1981) to
compare finite-element predictions with observed behavior of
instrumented structures. However, the use of finite-element
methods in design is currently limited by the relatively high
cost and raises significant difficulties with regard to the following.
1. The actual construction stages and installation process of

the inclusions are difficult, if not practically impossible, to


simulate.
2. The complex soil-inclusion and soil-wall interaction is
difficult to model. Several interface models have been developed (Zaman et aI., 1984; Frank et aI., 1982), but their
implementation in design requires relevant interface properties that are difficult to determine properly.
3. Various elastoplastic soil models can presently be used to
predict soil behavior during excavation. However, determination of soil model parameters generally requires specific
and rather elaborated testing procedures, which limits the
practical use of these models.

The finite-element method has therefore been used primarily


as a research tool to evaluate the effect of the main design
parameters on the engineering behavior of the structure, ground
movement, and working forces in the inclusions.
Clough and Tsui (1974) have shown through finite-element
simulations of anchored sheet pile and slurry walls in a medium
clayey soil, illustrated in Figure 26.32a, that an increase in
flexibility of the wall and/or tiebacks results in larger structure
displacements. With respect to soil-nail walls, Figure 26.32b
shows the results of a parametric finited-e1ement study (Juran
et aI., 1985; Shafiee, 1986) to evaluate the effect of bending
stiffness and nail inclination on facing displacement in vertical
nailed cut slopes. These results illustrate that, for nail inclinations
used in practice (10 to 15), the greater the nail bending stiffness
is, the smaller is the facing displacement. As shown in Figure
26.33a, for inclined nails an increase in the bending stiffness
results in a decrease of the maximum tensile forces. The behavior
of the inclined nails is substantially different from that of
horizontally placed nails. During construction, the relatively
flexible inclined nails tend to undergo a local deformation,
approaching the horizontal direction of maximum extension
strain in the soil. This local deformation, which is controlled
by the bending stiffness of the nails, results in an increase of
the structure/facing displacements. For horizontal nails, as
illustrated by both reduced-scale model tests and numerical test
simulations (Figs. 26.32b and 26.33b), the bending stiffness has
practically no effect on the mobilized nail forces and structure
displacements. Although the finite-element results are rather
qualitative, they provide a significant insight into the fundamental understanding ofthe system behavior and relevant input
into the selection of the main design parameters.

Ground Anchors and Soil Nails in Retaining Structures 895

e
-5p..

Distance from excavation, m


18
27
37

r-----

10

D.

45

I
I
I
I
I
I
I
I

15

Slurry wall
'" Sheet pile wall

20

Medium clay

C1)

Cl 25

W Deflection scale

30mm

(a)

Distance from excavation, m


12
18
24

30

o
5

Medium clay

10

15
'" Stiff tiebacks
Flexible tiebacks
L.......J Deflection scale
30mm

20
25
(b)

Kinematical Limit Analysis Design Method

This limit analysis approach was developed (Juran and Beech,


1984; Juran et aI., 1988) for the design of nailed soil-retaining
structures. It allows for the evaluation of the effect of the main
design parameters (i.e., structure geometry, inclination, spacing,
and bending stiffness of nails) on the tension and shear forces
generated in the nails during construction. The main design
assumptions, shown in Figure 26.34, are:
a. Failure occurs by a quasi-rigid body rotation of the active
zone, which is limited by a log-spiral failure surface.
b. The locus of the maximum tension and shear forces at failure
coincides with the failure surface developed in the soil.
c. The shearing resistance of the soil, defined by Coulomb's
failure criterion, is entirely mobilized along the sliding
surface.
d. The shearing resistance of stiff inclusions, defined by Tresca's
failure criterion, is mobilized in the direction of the sliding
surface in the soil.
e. The horizontal components of the interslice forces Eh (Fig.
26.33) are equal.
f. The effect of a slope (or horizontal surcharge F h), at the
upper surface of the nailed soil mass, on the tension forces
in the inclusions is linearly decreasing along the failure
surface.
The effect of the bending stiffness is analyzed using a
conventional "p-y" analysis procedure, assimilating the relatively flexible nail to a laterally loaded infinitely long pile. This
solution implies that at the failure surface the bending moment
in the nail is zero, whereas the tension and shear forces are
maximum. It involves a nondimensional bending stiffness
parameter, defined as
N=

12

10

Bending stiffness of reinforcement

E%

1OEolo

I
I
I
I

I
I

Inclination 30

I
I
I
I

:~!
I

I
I

6~Y ____
:
I
I

I
I
I
I

Inclination 20<>----~

9
I

l00E%

I
I

I
I
I
I
I
I

I
I
I
I
I
I

I
I
I
I
I
I

I
I
I
I

I
I
I
I

: Inclination 0 :

I
I

I
I

I
I
I
I

I
I

O~--~--------~----------~----

(c)
Fig. 26.32 Effect of (a) wall and (b) anchor stiffness on the
movements of an anchored structure in clay; finite-element simulations. (After Clough and Tsui, 1974.) (c) Effect of the bending
stiffness and the inclination of reinforcement on the facing displacements. (After Shafiee, 1986.)

L6
( KhD)
yH Sh'Sv

(26.5)

where Lo = [( 4El)/(K h D)]1/4 is the transfer length, which


characterizes the relative stiffness of the inclusion to the soil,
E, I and D are the elastic modulus moment of inertia and
diameter of the nail, respectively.
Generally, the length of the inclusion L is substantially
greater than three times the transfer length Lo and it can
therefore be considered as infinitely long.
A unique, kinematically admissible, failure surface that
verifies all the equilibrium conditions of the active zone can be
defined. In order to establish the geometry of this failure surface,
it is necessary to determine its inclination Ao with respect .to
the upper ground surface. Observations on both full-scale
structures (Schlosser, 1983; Juran and Elias, 1987) and laboratory model walls (Juran et aI., 1984) show that for relatively
flexible nails the failure surface is practically vertical at the
upper part of the structure (Ao = 0).
The normal soil stress along this failure surface is calculated
using Kotter's equation. The maximum tension force (Tmax) in
each inclusion is calculated from the horizontal force equilibrium
of the slice comprising the inclusion. Following assumption (d),
analysis of the state of stress in the inclusion yields the ratio
of the mobilized shear (7;) to tension (Tmax) forces as a function
of the inclination of the inclusion with respect to the failure
surface.
In order to implement the kinematical analysis approach in
a detailed design of soil-nailed structures, a computer code has
been developed (Juran et aI., 1988). It provides for each
reinforcement level at the relative depth of Z / H (where H is
the total excavation depth) the nondimensional design parameters corresponding to the normalized maximum tension force

896 Foundation Engineering Handbook

(a)

~ __
T",ma:;;:cx_ _
Tmax '

f
0.2
0.4
ZIH

0.6
0.8
1.0

"A ... '0

<II

0.2.1

i 0~~Iq,50
"

\ I=

-~\
'max
f~Tc '(Y.H.SH.SV
"'... /
\

Tmax

\\

t!
'b.
Tc k'/ y .
0.4 i:/Tmax

jO,

y. H SH Sv
0.1
0.2

ZIH

i
i /

....".".

:----~

0.8 i!..

Tmax
H
SH Sv

I
I

Rigid
/J--inclusions

l---

Flexible
;----- inclusions

1.0

Tc = shear force
Tmax = tensile force

/3=0

(i) Finite element analysis

(ii) Reduced-scale laboratory models


(b)

Fig.26.33 (a) Effect of bending stiffness of the inclusions on the maximum tensile forces in inclined nails. (After Shafiee, 1986.)
(b) Effect of the bending stiffness of the inclusions on the maximum tensile forces in horizontal nails. (After Juran et al., 1985.)

= T.m/Y H S~ So),

the normalized shear force (TS =


and the normalized distance of the locus of the
maximum tension force from the facing (SH = S/ H). Figure
26.35a illustrates these design parameters for a typical soilnailed vertical cut slope, 12 m deep, in a silty soil (q, = 35,
cfyH = 0.05) using #8 rebar nails (N = 0.33) at an inclination
of P= 15. For preliminary design in homogeneous soils,
simplified yet conservative, design charts have been prepared
(Juran et aI., 1988). Figure 26.35b provides design charts
established for the common geometry of vertical facing and
horizontal ground surface considering perfectly flexible nails
with 15 inclination.
Figure 26.30 shows a comparison between predicted and
measured values of maximum tension forces in nailed soilretaining structures. It illustrates that the kinematical design

(TN

~/Y H . S~ So),

approach provides a reasonable estimate of tension forces


mobilized in the inclusions. Specifically, the results of the
full-scale experiment conducted in France on a 7-m deep
granular soil-nailed wall (field data reported by Plumelle, 1986),
which are reported in Figure 26.3Oc for several excavation
depths, illustrate that the total excavation depth has only a
negligible effect on both the normalized tension forces in the
nails and the geometry of the active zone. Therefore, at any
relative depth (Z / H) the maximum nail tension force is
approximately proportional to the total excavation depth. The
predicted distribution of the maximum tension forces agrees
fairly well with the earth pressure design diagrams proposed
for braced excavations. It can be concluded that the kinematical
design approach can also be used to estimate working forces
in bracing supports and anchor prestress loads.

Ground Anchors and Soil Nails in Retaining Structures 897

o
p

Tmax = UN' As' Tc = TN' As


As = Section area of nail
State of stress in the inclusion

K(a)FH

Loading effect
on nail forces

(a)

/ Failure surface

(b)

Fig. 26.34 Kinematical limit analysis approach. (a) Mechanics of failure and design assumptions. (b) Theoretical solution for infinitely long
bar adopted for design purposes. (After Juran et al., 1988.)

26.9.3 Stability Analysis of Anchored and

where

Soil-Nailed Retaining Structures

TN

')'RS.S.

The design of anchored or soil-nailed retaining structures should


verify:
a. The local stability at the level of each inclusion.
b. The global stability of the structure and the surrounding
ground with respect to a rotational or translational failure
along potential sliding surfaces.
c. For anchored walls, the basal stability must be ensured by
sufficient penetration of the wall element (i.e., soldier pile,
diaphragm wall, or sheet pile).
A.

Local Stability Analysis (soil-nailed and multianchored

walls)

= __T.-"m::;.:.=-._
fult'D

J.L=--

')'S.S.

and S is the inclusion length in the active ( or unbonded) zone.


(2) Breakage failure of the inclusion: For flexible nails and
anchors that withstand only tension forces,

fall A ~ TN
')'RS"S.

where fall and A. are the allowable tension stress and crosssectional area of the inclusion, respectively.
For rigid nails that can withstand both tension and shear
forces,

At the level of each inclusion the design should satisfy the


following internal failure criteria.
(1) Pull-out failure of the inclusion:

(26.8)

(26.9)
where

K.q = [(TN)2
(26.6)
TS

where fult is the limit interface lateral shear stress, Tm .. is the


maximum tensile force in the nail or the design prestress load
Fw in the anchor, LII is the adherence (or bond) length, and F,
is the safety factor with respect to pull-out. This design criterion
implies that for a multianchored wall or a soil-nailed cut slope,
the structure geometry defined by the LIR ratio (where L is
the total inclusion length) should satisfy
(26.7)

+ 4'(TS)2r /2

1'c
')'RS.S.

and 1'c is the maximum shear force in the inclusion.

(3) Failure by excessive bending of a stiff inclusion:

Mp>FrnMm..

(26.10)

where M p is the plastic bending moment of the inclusion and


F rn is a factor of safety with respect to plastic bending (usually,
allowable tension stress is used to calculate M p with F rn = 1).
The bending moment Mm is derived from the "p-y" analysis:

Mm

= 0.321'c Lo

898 Foundation Engineering Handbook


SIH

o 0 1 020304
1

0.2 0.1

0.3

0.7 -

0.9
1.0

-1
I

I'-'

hence,

Mp/Lo > O.32F",' TS


Y'H'Sv'S h

B.

0"

(26.11)

Global Stability Analysis

ITS

,\TN
0

0.6 -

\O~

0.5

0.8

1 ~/

0.4
21H

.1

TN and TS
004 006 008 0 10 0 12 0 14

\
0

I
I

P
(a)

This analysis consists of evaluating a global safety factor of the


anchored or soil-nailed retaining structure and the surrounding
ground with respect to a rotational or translational failure along
potential sliding surfaces. It requires determination of the
critical sliding surface, which may be dictated by the stratification
of the subsurface soil or, in rock, by an existing system of joints
and discontinuities. The potential sliding surface can be located
inside or outside the anchored or soil-nailed mass. Evaluation
of the global safety factor is generally based on limit equilibrium
approaches. Slope stability analysis procedures have been
developed to account for the available limit pull-out, tension,
and shearing resistance of the inclusions crossing the potential
sliding surfaces.
Limit equilibrium methods commonly used in the design of
anchored walls (Kranz, 1953; Broms, 1968; Bureau Securitas,
1977) and soil-nailed retaining structures (Stocker et aI., 1979;
Shen et aI., 1981; Schlosser, 1983) involve different definitions
of the safety factors, and a variety of assumptions with regard
to the shape of the failure surface, the type of soil-inclusion
interaction, and the resisting forces in the inclusions.

0.40

1. Anchored Walls For anchored walls it is common practice


to assume a planar failure surface passing through the toe of
the wall at an inclination of c/J or (n/4 - c/J/2). Rankine's failure
surface has been recommended by Cheney ( 1984). The shearing
resistance of the soil, as defined by Mohr-Coulomb's failure
criterion, is assumed to be entirely mobilized along the potential
failure surface. The global safety factor is defined as the ratio
of the sum of the available resisting limit forces (Rd in the
inclusions to the total force (Rm) required to maintain limit
equilibrium, that is,

TN

0.05 L-_---''--_----L_--'--'-_ _--'-_----'


0.08
0.12
0.16
0.20
C

yH

RL

(26.12)
Rm
As shown in Figure 26.36 the total force Rm required to maintain
limit equilibrium is readily obtained using the polygon of forces
acting on the rigid soil wedge. The resisting forces (R L ) are
provided by the pull-out capacity of the anchors. Cheney
recommends that the anchor bond length be located at a
distance of at least H /5 beyond the assumed failure surface or
a minimum distance of 15 ft from the facing, whichever is greater.
FS=-

SH

-50..

Failure sllrface

Sc

Q)

"0

.9

~T

'OJ
>

'u"

><
u.l

---..... T

p
/
p

~T

(b)
Fig.26.35 (a) Typical example of design output provided by the
kinematical limit analysis approach. (b) Design charts for perfectly
flexible nails. Notes: 1, Nail inclination 15. 2, Vertical force.
3, Horizontal backfill. 4, No surcharge. (After Juran et a/., 1988.)

FS = RL = I:Fw
Rm
I:T
Fig. 26.36 Global stability analysis of an anchored wall using force
equilibrium method with a plane failure surface.

Ground Anchors and Soil Nails in Retaining Structures 899


The early work of Terzaghi (1948) demonstrated that the
shape ofthe failure surface behind rigid retaining walls is highly
dependent upon the displacement mode and associated earth
pressure distribution. In braced excavations or anchored walls,
restraining the lateral soil displacement results in a nonplanar
failure surface that is likely to be vertical in the upper part of
the wall. Therefore, the assumption of a Rankine failure surface
is inconsistent with the actual behavior of these structures.
Cheney's recommendations provide an initial assessment for
the location of the anchor bond zone that has to be verified
by global stability analysis.
Kranz's (1953) limit equilibrium method generalized by
Broms (1968) considers (Fig. 26.37) a bilinear sliding surface
passing through the midpoint of the anchor shear force (or
hinge) in the wall. The soil resistance to shearing (i.e., cohesion
component Sc and frictional component SiP) along the lower
segment ofthe potential failure surface is assumed to be entirely
mobilized. The retained soil is assumed to apply an active lateral
earth pressure p. on the vertical segment of the potential failure
surface calculated using Rankine's theory (i.e., p. = t K. Y . Z~,
where Zo is the depth of the midpoint of the anchor bond
length).
The polygon of forces acting on the soil wedge at limit
equilibrium involves the active earth pressure of the retained
soil p., the shearing resistance of the soil (i.e., Sc and SiP)' and
the wall reaction P A' The polygon of forces acting on the
structural wall element involves the soil pressure P A, the
resisting prestress anchor load required to maintain equilibrium,
the passive resistance of the foundation soil Pp and the basal
soil reaction F b
Kranz (1953) considered the limit equilibrium of the soil
wedge neglecting the passive (P p) and basal (F b) reaction forces
applied by the foundation soil. Equation 26.12 is used to define
the factor of safety, that is, FS = RL/ Rm , with RL = Fwand
Rm= T.
Broms (1968) extended Kranz's method, taking into account
the passive soil resistance Pp and the basal reaction Fb in the
equilibrium of forces acting on the anchored wall-soil wedge
system. The factor of safety was defined as FS = p~m / P p' where
p~m is the limit passive resistance of the foundation soil that
can be estimated using Rankine's theory and P p is the passive
soil reaction required to maintain the limit equilibrium of the
anchored wall-soil wedge system.
In its generalized form (Fig. 26.37), Kranz's (1953) and
Brom's (1968) limit equilibrium method requires evaluation of
the passive soil resistance P p and of the basal reaction F b' The
passive soil resistance on the wall element is estimated using
Rankine's theory or Coulomb's limit equilibrium method,
which are commonly used in design of retaining walls. The
basal reaction F b depends upon the interaction between the
wall element and the foundation soil.

W,

Internal force
soil- wall
interaction
p_____
/
_

Fig. 26.37 Generalized Kranz's force equilibrium method for


stability analysis of an anchored wall. (After Broms. 1968.)

For soldier piles and sheet piles, the basal reaction is


mobilized through lateral shaft friction and can be estimated
using bearing-capacity formulas that are commonly used in
design of friction piles. For cast-in-place diaphragm walls, the
basal reaction can be estimated using Meyerhof's (1953)
bearing-capacity formula for shallow foundations under inclined
eccentric loads.
In the global stability analysis of the anchored wall-soil
wedge system, the safety factors with respect to the passive soil
resistance P p and basal reaction F b should be consistent with
current design procedures commonly used for retaining walls,
friction piles, or shallow foundations. The prestress anchor load
required to maintain the system equilibrium can be estimated
from the polygon of forces. The global safety factor should be
evaluated with respect to pull-out failure ofthe anchor following
Kranz's original definition given by Equation 26.12.
The major advantage of Brom's method is its simplicity.
However, it is noted that this method is based on a restrictive
assumption concerning the shape of the potential sliding surface
and under its extended form requires an evaluation of the
soil-wall element interaction, which is difficult to assess.
Furthermore, the assumed definition of the failure surface is
difficult to generalize for the case of multianchored walls. A
generalized Kranz's method is used in the French code of
practice (Bureau Securitas, 1977) to allow for the design of
multianchored walls; however, it leads to significant overestimates of the required prestress anchor loads. Kranz's and
Brom's methods provide a safety factor with respect to failure
along a potential sliding surface passing through the anchors.
However, the global stability of the anchored wall system and
the surrounding ground with respect to general sliding along
potential failure surfaces that do not cross the anchors should
be investigated using conventional slope stability analysis
methods.
Slope stability analysis procedures are increasingly used in
design of anchored wall systems for the evaluation of the global
safety factor with respect to general sliding along a potential
failure surface that may pass within or outside the anchored
wall system. These procedures (Pfister et aI., 1982; Blondeau
et aI., 1984) mostly use conventional limit equilibrium slice
methods (see Chapter 10) that have been modified to account
for the prestress load and pull-out capacity of the anchors
crossing the potential sliding surface.
2. Soil-Nailed Retaining Structures The three limit equilibrium
methods currently most used for nailed soil-retaining structures
are the so-called German method (Stocker et aI., 1979), Davis
method (Shen et aI., 1981), and French method (Schlosser,
1983).
(a) THE GERMAN METHOD. Stocker and coworkers (1979)
proposed a limit force equilibrium method (Fig. 26.38) that,
following the principles of Kranz's method, assumes a bilinear
sliding surface and adapts the same definition for the global
factor of safety (Eq. 26.12). The inclination (0 1 ) of the failure
surface is iteratively determined to yield the minimum factor
of safety. Gassier and Gudehus (1981) have shown, through
stability analyses, that the minimum safety factor value is
usually obtained for O2 = (n/4 - 4>/2) and Os = 90. In this
analysis, the resisting limit nail force is provided by the pull-out
capacity of the nails (i.e., the pull-out capacity of the portion
of the nail located beyond the potential failure surface).
The assumed bilinear failure surface is mainly based on a
limited number of model tests where failure was generated by
substantial surcharge loading. However, it does not appear to
be consistent with the observed behavior of nailed soil-retaining
structures that are primarily subjected to their self-weight. In
particular, stability analyses show that this bilinear failure

900 Foundation Engineering Handbook


surface is generally not entirely contained within the nailed-soil
mass and therefore yields an active zone (or potential failure
wedge) that is substantially larger than that observed on actual
structures.
(b )THE DAVIS METHOD. Shen et al. (1981) developed a similar
force equilibrium method assuming a parabolic failure surface,
passing either entirely or partially within the nailed-soil mass.
Their assumption is based on the contours of factor of safety
derived from finite-element simulations, as shown in Figure
26.39a. Nails are assumed to withstand only tension forces and
their failure can therefore be generated by either breakage or
pull-out. In .this analysis it is implicitly assumed that the safety
factors with respect to the shear strength of the soil (i.e.,
Fe = clcmand Fq, = tan /tan m, where Cmand m are, respectively, the soil cohesion and internal friction angle mobilized
along the potential sliding surface) and the ultimate interface
lateral shear stress (i.e., F L = TUlt/Tm, where Tm is the lateral
shear stress mobilized at the soil-inclusion interfaces) are equal
to the global safety factor FS, that is,
(26.13)
A slope stability analysis procedure, using a modified
method-of-slices, has been implemented to iteratively determine
the critical sliding surface and the minimum factor of safety.

Fig. 26.38 Force equilibrium method of global stability analysis


of nailed soil-retaining structure. (After Stocker et al., 1979.)

~::-'=-+--4..-1

_ _- - - Potential failure
surface
3.0

H=25ft
L = 15ft

(a)
~-----aH----~~

Element 1

H
Element 2

N2
Element 1
(b)

Tp

= 'ull . 1t D . La; Fe = F~ = FL = FS

Fig. 26.39 (a) Contours of factor of safety derived from finite-element analysis. (b) Limit force equilibrium method for stability analysis
of nailed soil-retaining structures. (After Shen et al., 1981.)

Ground Anchors and Soil Nails in Retaining Structures

To calculate the interslice forces, a stress ratio parameter K


(i.e., ratio of the lateral to vertical stresses at the interslice) is
input with K values of 0.4 for frictional soils and 0.5 for cohesive
soils.
Shen et al. (1981) have evaluated their design procedure
through analysis of observed failure surfaces and failure heights
of centrifugal soil-nailed model walls. As shown in Figure 26.39b
the method's predictions agree fairly well with the experimental
results.
(c) THE FRENCH METHOD. Common to all the limit
equilibrium methods specified above is the assumption that the
inclusions withstand only ,tension forces. A more general
approach for the stability analysis of soil-nailed retaining
structures has been developed by Schlosser ( 1983), considering
the two fundamental mechanisms of soil-inclusion interaction
(i.e., lateral friction and passive normal soil reaction). This
method takes into account both the tension and shearing
resistance of the inclusions as well as the effect of their bending
stiffness. For an inclusion that withstands both tension (Tmax)
and shear (7;) forces, the mobilized limit forces are calculated
according to the principle of maximum plastic work considering
Tresca's failure criterium. The 7;/ Tmax ratio is a function of the
inclination of the inclusion with respect to the potential failure
surface. The tensile strength of the inclusion is defined by the
elastic limitf.l1 = fy, and the shear resistance by Re = fy/2. The
limit shear force that can be generated in the nail depends upon
the mobilized passive lateral soil pressure on the inclusion. In
order to prevent plastic flow (or creep) of the soil between the
inclusions, the maximum lateral soil pressure Plim should not
exceed half of the ultimate lateral pressure of the characteristic
"p-y" curve. In French practice, this lateral soil pressure is
limited to the creep pressure obtained from a pressuremeter
test. The shear force in the inclusion should therefore not exceed

901

of the effect of soil stratification, groundwater flow, and seismic


loading on the global structure stability. It can also be used
for the design of mixed structures associating ground anchors
and soil nailing.
Postfailure analyses of several nailed soil-retaining structures
reported by Blondeau et al. (1984) have illustrated that with
an appropriate input design value of the ultimate lateral shear
stress this design procedure could predict fairly well the pull-out
failure of the structures.
3. Evaluation of Global Stability Analysis Procedures The
Davis and the French design procedures have been evaluated
(Juran and Elias, 1987) through the analysis of field data from
the full-scale structures. These analyses suggest that:
1. In soil-nailed cut slopes the factors of safety with respect to
soil strength are fairly close to 1.
2. The Davis method generally yields safety factor values that
are about 15 percent lower than those predicted using the
French method.
3. Observed loci of the maximum tension forces in these
structures agree fairly well with predicted location of the
potential failure surface (Fig. 26.41).
These results further suggest that in soil-nailed retaining
structures, owing to the staged construction process the soil
resistance to shearing along the potential failure surface is

(26.14 )
A multicriteria analysis, illustrated in Figure 26.40, is
conducted to evaluate the global stability of the nailed-soil
system with respect to four potential failure modes: shear failure
of the soil along the critical sliding surface, pull-out failure of
the nail, nail breakage by either excessive bending or combined
effect of tension and shear forces, and creep or plastic flow of
the soil between the nails. The global factor of safety is defined
by Equation 26.13 (Le., Fe = F4> = F, = F.) and a minimum
safety factor of 1.5 is generally required. This multicriteria
analysis procedure uses a slices method (see Chapter 10) that
is modified to take into account the resisting nail forces in the
equilibrium of each slice. This procedure permits an evaluation

----- Davis method


- - French method
- - - Kinematical approach
(a)

(b)

Fig. 26.41 Predicted and observed locus of maximum tension


forces in nails. (a) Full-scale experiment CEBTP. (Experimental
results. Plumelle. 1986.) (b) Parisian wall. (Experimental results.
Cartier and Gigan. 1983.) (After Juran and Elias. 1987.)

Forces in the bar

~-"""''"fE

~~__ Tmax

fmax

1/ ...O(E----;L,....-I:~

--I,';..........-~
I

I"---Slip surface
Failure criteria:
Tensile strength and shear resistance of the bar: Tmax
Soil bar friction: Tmax ~ ltDtu l t L.
Normal lateral earth thrust on the bar: P ~ P max
Shear resistance of the soil: t < c + (1 tan '"
(After Schlosser. 1983.)

Fig. 26.40

= As fy;

Tc ~ Rc

Multicriteria slope stability analysis method.

= As fy/2

902 Foundation Engineering Handbook


practically mobilized at the early stages of excavation. As the
excavation proceeds, the load increments are entirely transferred
to the inclusions. Therefore, for internal stability analysis of
these structures, the definition of the factor of safety adapted
by Kranz for anchored walls and Stocker et al. for soil-nailed
structures appears to be more consistent with the observed
structure behavior. It implies that the global factor of safety
should be evaluated with respect to the pull-out resistance of
the inclusions considering a safety factor of 1 with respect to
shear resistance of the soil. For permanent anchored walls
Cheney (1984) recommends a minimum safety factor of 1.5,
whereas for soil-nailed walls Elias and Juran (1988) recommend
a minimum safety factor of 2. GassIer and Gudehus (1981)
recommend for soil-nailed structures the use of residual soilstrength parameters (c and <p) factored by 1.25 to comply with
statistical evaluation criteria concerning the probability of
failure.
The major limitation of the slope stability analysis procedures
currently used in design of soil-nailed retaining structures lies
in the basic definition of a global factor of safety. Observations
on both full-scale structures and reduced-scale laboratory
models have illustrated that pull-out failure is a progressive
phenomenon that is generally induced by the sliding of the
upper inclusions. Therefore, this internal failure mechanism
cannot be adequately defined using a "global" value of a unique
safety factor for all the inclusions. The local stability (or safety
factor) at the level of the sliding inclusion can be significantly
more critical than the "global" stability with respect to general
sliding in the retaining system or the surrounding ground. For
a reliable design of these composite structures, it is therefore
essential that the engineer should attempt to rationally evaluate
both the local internal stability at the level of each inclusion
and the" global" structure stability.
C.

Structural Design of Anchored Walls or Facing


Elements and Stability Analysis of the Wall Base

The design of the structural elements of an anchored wall or


of the facing of a soil-nailed retaining structure involves three
main steps:
1. Estimation of the system of applied forces and bending

moments.
2. Structural analysis (i.e., calculation of internal stresses and
moments) and selection of appropriate wall or facing elements.
3. Stability analysis of the structural element base.

The basic differences in construction process of a soil-nailed


facing and anchored wall elements result in a substantially
different interaction of these structural elements with the
foundation soil. As illustrated in Figure 26.25, the flexible facing
of a soil-nailed structure does not transfer any load to the
foundation soil while the anchored wall is designed to mobilize
the passive soil resistance and transfer the vertical component
of the anchor load to the foundation soil. Therefore, in the case
of inclined anchors, the basal stability of the structural wall
element has to be carefully evaluated.
The system of forces acting on the facing or wall elements
includes the resisting forces in the inclusions (anchors or nails),
the lateral soil pressure and the reaction of the foundation soil
on an anchored wall. The forces in the inclusions are the
prestress anchor loads or the forces in nail connections, which
are assumed to be equal to the maximum tension nail forces.
The lateral soil reaction on the anchored wall is governed by
the passive resistance of the foundation soil, which can be
estimated using conventional limit equilibrium approaches.
The design of the structural element (i.e., soldier pile, concrete
wall, sheet pile, reinforced shotcrete, etc.) is generally done by

assimilating this element to a beam or raft (width equal to the


vertical or horizontal spacing between the inclusions) on simple
supports formed by the inclusions. For the design loads in the
supports and the estimated lateral passive soil reaction, the
moments in the structural element can be calculated and the
selection of the element follows usual structural design procedures. "p-y" type analyses are commonly used in design of
diaphragm walls (Pfister et aI., 1982) to assess the effect of wall
stiffness and anchor longitudinal rigidity on the structure
displacements.
Evaluation of the basal stability of anchored walls requires
a rational estimate of the vertical load transferred to the
foundation soil. This vertical load is highly dependent upon
the frictional interaction between the structural element, the
foundation soil, and the retained ground, which is difficult to
assess. A conservative working assumption neglecting the effect
of soil-structure interaction above the foundation level has
been recommended by Cheney (1984). This design assumption
governs the required wall penetration and may significantly
affect construction cost. Therefore, additional research and site
monitoring are needed to provide a relevant database for the
development of a more rational design assumption.
For the estimated vertical load, the basal stability of the
structural wall element can be analyzed using conventional
bearing-capacity limit equilibrium methods. Required penetration depth of soldier sheet piles can be estimated using
bearing-capacity formulas that are commonly used in design
of friction piles. The basal stability of a diaphragm wall can be
evaluated using Meyerhof s (1963) formula for the bearing
capacity of shallow foundations under inclined eccentric loads.

26.10 TECHNOLOGICAL DEVELOPMENT AND


RESEARCH NEEDS

The increasing use of ground anchors and soil nails in permanent


structures is a key parameter in current technological developments. Durability of the inclusions, long-term performance in
fine-grained soil, and environmental/ architectural requirements
for the anchored wall or soil-nailed facing have become major
design considerations.
The potential application of the technology in more aggressive
environments has stimulated specialty contractors to continuously invest in the development of more reliable and
cost-effective corrosion-protection schemes. In particular it is
noted that driven soil nails, which are commonly used in Europe,
are not corrosion-protected. Their implementation in the construction of permanent structures requires innovative improvements in the manufacturing and/or installation process to
provide adequate protective coatings that may resist construction damage. Attempts have been made by French contractors to provide corrosion protection of driven nails using
a variety of techniques such as coupling nail driving with jet
grouting, driving encapsulated nails, or driving prefabricated
nails that consist of prestressed bars in compression tubes.
For the construction of permanent soil-nailed retaining
structures, shortcomings of the shotcrete facing technology have
generated development of prefabricated facing elements (e.g.,
concrete or steel panels) that may efficiently accommodate
adequate facing drainage and economically comply with durability and architectural requirements. In addition, technological
efforts have been invested by European contractors to develop
cost-effective driving-installation processes, such as jet nailing,
that may efficiently improve the pull-out capacity while substantially increasing construction rate and thereby decreasing the
project cost.

Ground Anchors and Soil Nails in Retaining Structures 903


As it is often the case in geotechnical engineering, technology
and construction practice with ground anchors and soil nailing
have always preceded any fundamental research on the behavior
or long-term performance of the anchored or nailed ground
system and any experimental or theoretical basis for development
of appropriate design methods. However, implementation of
newly developed corrosion-protection systems has generated a
substantial requirement for innovative quality control procedures
to properly assess the performance of the proposed protection
scheme. Furthermore, the growing use of the technology in
permanent structures and potential economical applications in
fine-grained soils have raised significant research needs pertaining to the following.
(1) Long-term performance in fine-grained soils. The creep
response of the soil is highly dependent upon the strain path
and the induced shear strain rate, which in turn are mainly
governed by the mode of soil-inclusion interaction (i.e., lateral
shaft friction or passive soil resistance on transversal underreams)
and the applied stress level. Therefore, basic research on the
effect of the installation process (i.e., regrouting procedure,
grouting pressure) and inclusion shape on the long-term
inclusion performance may provide a relevant database for
extending the range of fine-grained soils in which ground
anchors and soil nails can be confidently used.
(2) Estimate of the pull-out capacity of grouted nails. The
current design procedure relies upon the estimate of an ultimate
interface lateral shear stress. However, the load transfer along
a grouted inclusion is nonlinear and therefore the interface
ultimate lateral shear stress derived from pull-out test results
is length-dependent. This scale effect is difficult to assess and
the use of short testing nails may result in an overestimate of
the pull-out capacity of production nails. Development of a
more rational load-transfer model and relevant testing interpretation procedure may significantly improve the estimate of the
pull-out capacity of grouted nails.
(3) Evaluation of soil-structure interaction in the anchored
wall-retained ground-foundation soil system. This complex
soil-structure interaction may significantly affect the load
transferred to the foundation soil and thereby the required
penetration depth of the structural wall element.
(4) Performance of ground anchors and soil nails under
repetitive cyclic loadings.
It is anticipated that monitoring of the long-term performance
of structures, development of reliable quality control procedures,
testing of witness inclusions for corrosion studies, and basic
research on soil-inclusion and soil-structure interactions under
both monotonic and cyclic loadings, will significantly contribute
to improving the state of design and practice. This may substantially enhance technological innovations and cost-effective use
in more aggressive environments and in a wider range of
fine-grained soils.

REFERENCES
AI-Mosawe, M. M. (1979), The effect of repeated and alternating
loads on the behavior of dead and prestressed anchors in sand,
Thesis, University of Sheffield, England.
Andreadis, A., Harvey, R. c., and Burley, E. (1978), Embedment
anchors subjected to repeated and alternating loads, Ground Engineering, 11, No.3.
Baguelin, F., Jezequel, J. F., and Shields, D. H. (1978), The Pressuremeter
and Foundation Engineering, Trans Tech Publications, Clausthal,
Germany.

Barla, G. and Mascardi, C. (1974), High anchored wall in Genoa,


Conference on Diaphragm Walls and Anchorages, Institute of Civil
Engineers, London, pp. 173-128.
Bassett, R. H. (1977), Underreamed ground anchors, Specialty session
No.4, Proceeding of the 9th International Conference on Soil
Mechanics and Foundation Engineering, Tokyo, pp. 11-17.
Begemann (1973), Alternating loads and pulIing tests on steel I-beam
piles, Proceedings of the 8th International Conference on Soil
Mechanics and Foundation Engineering, 2.1, pp. 13-17.
Bishop, A. W. (1966), The strength of soils as engineering materials,
Geotechnique, 16, pp. 91-128.
Blondeau, F., Christiansen, M., Guilloux, A., and Schlosser, F. (1984),
TALREN: methode de calcul des ouvrages en tere renforcee,
Proceedings of the International Conference on In-situ Soil and Rock
Reinforcements, Paris, pp. 219-224.
Briaud, J. L., Smith, T., and Meyer, B. (1983), Pressuremeter gives
elementary model for laterally loaded piles, International Symposium
on In-situ Testing, Paris, 2, pp. 217-221.

British Standards Institute ( 1980), October 1980 Draft British Code of


Practice for Ground Anchors, DSB 22 Committee.
Broms, B. B. (1968), Swedish tieback system for sheet pile walls,

Proceedings of the 3d Budapest Conference on Soil Mechanics and


Foundation Engineering, pp. 391-403.
Bureau Securitas (1977), Recommendations Concerning the Concepts,
the Calculation, the Execution, and the Control of Ground Anchors,

T.A. 77, Editions Eyrolles, 61, Boulevard Saint-Germain 75005,


Paris.
Bustamante, M. (1972), Essais pre'alables de tirants precontraints
definitifs pour la rive gauche de la Seine, Pont de Saint-Cloud-Pont
de Sevres, Travaux, No. 450.
Bustamente, M. (1975), Mesure des elongations dans les pieux et tirants
a l'aide d'extensometres amovibles, Travaux, No. 489.
Bustamante, M. (1976), Essais de pieux de haute capacite scelles par
injection sous haute pression, Proceedings of the 6th European
Conference on Soil Mechanics and Foundation Engineering, Vienna.
Bustamante, M. (1980), Capacite d' ancrage et comportement des
tirants injectes, scelles dans une argile plastique, These docteuringenieur ENPC, Paris.
Bustamante, M., Delmas, F., and Lacour, J. (1977), Behavior of
prestressed anchors in plastic clay, Proceedings of the 9th International Conference on Soil Mechanics and Foundation Engineering,

Special Session No.4, Tokyo.


Bustamante, M., Delmas, F., and Lacour, J. (1978), Comportement de
tirants preconstraints dans une argile plastique, Revue Franraise de
Geotechnique, 3.

Bustamante, M. and Doix, B. (1985), Une methode pour Ie calcul des


tirants et des micropieux injectes, Bulletin de liaison des Laboratoires
des Ponts et Chaussees, No. 140.
Cartier, G. and Gigan, J. P. (1983), Experiments and observations on
soil nailing structures, Proceedings of the 7th European Conference
on Soil Mechanics and Foundation Engineering, Helsinki, Finland.
Cheney, R. S. (1984), Permanent Ground Anchors, Federal Highway
Administration Report No. FHWA-DP-68-1R.
Clement, P. and Navarro, M. (1972), Les tirants en terrain meuble type
TM, Travaux, No. 450.
Clough, G. W. (1975), Deep excavations and retaining structures,
Proceedings of the Conference on Foundations of Tall Buildings,

Lehigh University.
Clough, G. W. and Tsui, Y. (1974), Performance of tied-back walls in
clay, Journal of the Geotechnical Engineering Division, ASCE, 100,
No. GT-12, pp. 1259-1273.
Clough, G. W., Weber, P. R., and Lamont, J. (1974), Design and
observations of a tied back wall, Proceedings of the Specialty
Conference on Performance of Earth and Earth Supported Structures,

ASCE, Purdue, 2, pp. 1367-1389.


Coyle, M. M. and Reese, L. C. (1966), Load transfer for axially loaded
piles in clay, Journal of the Soil Mechanics and Foundations Division,
ASCE, 92, No. SM-2, pp. 1-26.
Darbin, M., Jailloux, 1. M., and Montuelle, J. (1978), Performance and
research on the durability of reinforced earth reinforcing strips,
ASCE Symposium on Earth Reinforcement, Pittsburgh, Pa., pp.
305-333.
Davis, A. and Plumelle, C. ( 1982), Identification et etude des parametres
controlant Ie comportement des tirants d'ancrage dans un sable fin,
Annales de ITBTP, No. 401.

904 Foundation Engineering Handbook


DIN (1972,1976), Deutsche Industrie Norm, Soil and Rock Anchors:
Temporary Soil Anchors, Analysis, Structural Design and Testing,
DIN 4125, Part I, pp. 1-9, 1972; Part 2, pp. 1-9, 1976.
Edgers, L., Ladd, C. c., and Christian, J. T. (1973), Undrained Creep of
Atchafalaya Levee Foundation Clays, Vol. I, Report R73-16, Soils
Publication 319, Department of Civil Engineering, MIT.
Egger, P. (1972), Influence of wall stiffness and anchor prestressing on
earth pressure distributions, Proceedings of the 5th European Conference on Soil Mechanics and Foundation Engineering, Madrid, I, pp.
259-264.
Elias, V. and Juran, I. (1988), Draft Manual of Practicefor Soil Nailing,
prepared for V.S. Department of Transportation, FHWA, Contract
DTFH-61-85-C-00142.
Evans, R. H. (1955), Application of prestressed concrete to water supply
and drainage, Public Health Engineering Division Meeting, Public
Health Paper No. 12, London.
Feddersen, I. (1974), Verpessanker in Lockergestein (Grouted Anchors
in Soils), Bauingenieur, 49, No.8, pp. 302-310.
FIP (1974), Federation Internationale de la Prt)contrainte, Ground
Anchors, Proceedings of the 7th Congress of the FIP, New York,
pp.33-42.
Frank, R., Guenot, A., and Humbert, P. (1982), Numerical analysis of
contact in geomechanics, Proceedings of the 4th International
Conference on Numerical Methods in Geomechanics, Edmonton.
Fujita, K., Veda, K., and Kusabuka, M. (1977), A method to predict
the load-displacement relationship of ground anchors, Specialty
Session No.4, Proceedings of the 9th International Conference on
Soil Mechanics and Foundation Engineering, Tokyo.
Gassier, G. and Gudehus, G. (1981), Soil nailing: Some mechanical
aspects of in situ reinforced earth, Proceedings of the 10th International Conference on Soil Mechanics and Foundation Engineering,
Stockholm, Sweden, 3, pp. 665-670.
Goldberg, D. T., Jawouski, W. E., and Gordon, M. D. (1976), Lateral
Support Systems and Underpinning, Construction Method, Final
Report No. FHWA-RD-75-130.
GuiIloux, Notte and Gonin (1983), Experience on a retaining structure
by nailing, Proceedings of the 7th European Coriference on Soil
Mechanics and Foundation Engineering, Helsinki, Finland.
GuiIloux, A. and Schlosser, F. ( 1984 ~ Soil nailing: Practical applications,
Proceedings of the Symposium on Soil and Rock Improvement
Techniques, AIT, Bangkok.
Hanna, T. H. (1982), Foundations in Tension, Trans Tech Publications,
Series on Rock and Soil Mechanics, Vol. 6.
Hanna, T. H., Sivapalon, E., and Senturk, A. (1978), The behavior of
dead anchors subjected to repeated and alternating loads, Ground
Engineering. 11, No.3.
Hovart, C. and Rami, R. (1975), Elargissement de I'emprise SNCF
pour la desserte de Saint-Quentin-en-Yvelines, Revue Travaux.
Jones, D. A. and Turner, M. J. (1980), Load tests on post-grouted
micropiles in London clay, Ground Engineering, 6, No. 13.
Jones, D. A. and Spencer, I. M. (1984), Clay anchors: A Caribbean
case history, Ground Engineering, 17, No.1.
Jones, N. C. and Kerkhoff, G. O. (1961), Beleld caissons anchor walls
as Michigan remolds an expressway, Engineering News Record, pp.
28-31,195-197.
Jorge, G. R. (1969), The regroutable IRP anchorage for soft soils, low
capacity or karstic rocks, Proceedings of the 7th International
Conference on Soil Mechanics and Foundation Engineering, Specialty
Session No. 14 and 15, pp. 159-163.
Jorge, J. (1970), Le tirant IRP reinjectable pour terrains meubles,
karstiques, ou a faibles caracteristiques geotechniques, Proceedings
of the 2d International Congress on Rock Mechanics, Mexico.
Juran, I., Beech, J., and Delaure, E. (1984), Experimental study of the
behavior of nailed soil retaining structures on reduced scale models,
Proceedings of the International Coriference on In-situ Soil and Rock
Reiriforcements, Paris.
Juran, I. and Beech, J. (1984), Analyse Theorique du Comportement
d'un Soutenement en Sol Cloue, Proceedings of the International
Coriference on In-Situ Reinforcement of Soils and Rock Reinforcements, Paris, pp. 301-307.
Juran, I., Shafiee, S., and Schlosser, F. (1985), Numerical study of nailed
soil retaining structures, Proceedings of the 11 th International
Conference on Soil Mechanics and Foundation Engineering, San
Francisco, 4, pp. 1713-1717.

Juran, I. and Elias, V. (1987), Soil nailed retaining structures: Analysis


of case histories, ASCE Special Geotechnical Publication No. 12,
pp. 232-245.
Juran, I., Baudrand, G., Farrag, K., and Elias, V. (1988), Kinematical
limit analysis approach for the design of nailed soil retaining
structures, Proceedings of the International Geotechnical Symposium
on Theory and Practice of Earth Reinforcement, Fukuoka Kyushu,
Japan.
Kerisel, J., Robert, J., Schlosser, F., Juran, I., Causse, G., and
Romon, C. ( 1981 ), Experimentation d'un M ur d' Ancrages Multiples
(Experiments on a multi-anchored wall), Proceedings of the
10th International Coriference on Soil Mechanics and Foundation
Engineering, Stockholm, 2, pp. 157-161.
Koreck, W. (1978), Small diameter bored injection piles, Ground
Engineering, May.
Kranz, E. (1953), (Jber die Verankerung von Spundwiinden, Wilh. Ernst
& Sohn, 2 Aufl., Berlin.
Littlejohn, G. S. (1970), Soil anchors, Proceedings of the Conference on
Ground Engineering, Institution of Civil Engineers, London, pp.
33-44.
Littlejohn, G. S. and Bruce, D. A. (1975), Rock anchors state-of-the-art,
Part I: Design and Part II: Construction, Ground Engineering, May.
Louis, C. ( 1981 ), Nouvelle methode de soutennement des sols en deblais,
Revue Travaux No. 533.
Louis, C. (1986), Theory and practice in soil nailing temporary or
permanent works, ASCE Annual Conference, Boston.
L.C.P.C.-S.E.T.R.A. (1985), Regles de justification des fondations sur
pieux.
Mastrantuono, C. and Tomiolo, A. (1977), First application ofa totally
protected anchorage, Proceedings of the 9th International Conference
on Soil Mechanics and Foundation Engineering, Specialty Session,
Tokyo, pp. 107-112.
McKittrick, D. P. (1979), Reinforced earth: Application of theory and
research practice, Ground Engineering, 12, No. I, pp. 19-31.
Meyerhof, G. G. (1953), The bearing capacity of foundations under
eccentric and inclined loads, Proceedings of the 3d International
Conference on Soil Mechanics and Foundation Engineering, Zurich,
I, p. 440.
Meyerhof, G. G. (1963), Some recent research on the bearing capacity
offoundations, Canadian Geotechnical Journal, I, pp. 16-26.
Middleton, H. (1961), Raising the Argal dam, The Consulting Engineer,
II.
Mitchell, J. K. et al. (1987), Reinforcement of earth slopes and
embankments, National Cooperative Highway Research Program
Report No. 290, Transportation Research Board, June.
Morris, S. S. (1956), Steenbras dam strengthened by post tensioning
cables, Civil Engineering, 2.
Murayama, S. and Shibata, T. (1958), On the Rheological Characteristics
of Clays, Part I, Bulletin No. 26, Disaster Prevention Research
Institute, Kyoto, Japan.
Ostermayer, M. (1974), Construction, carrying behavior and creep
characteristics of ground anchors, Coriference on Diagram Walls
and Anchorages, Institute of Civil Engineers, London.
Ostermayer, M. and Sheele, F. (1977), Research on ground anchors in
non-cohesive soils, Proceedings of the 9th International Conference
on Soil Mechanics and Foundation Engineering, Tokyo.
Peck, R. B. (1958), A study of the comparative behavior of friction
piles, Highway Research Board Special Report No. 36, p. 72.
Pfister, P., Evers, G., Guilland, M., and Davidson, R. (1982), Permanent
Ground Anchors: Soletanche Design Criteria, Federal Highway
Administration Report No. FHWA-RD-81-150.
PTI (1980), Post-Tensioning Institute, Recommendationsfor Prestressed
Rock and Soil Anchors, PT!, 301 W. Osborne, Suite 3500, Phoenix,
Ariz., 85013, p. 57.
Plumelle, C. (1986), Full scale experimental nailed-soil retaining
structures, Revue Franc:aise de Geotechnique, No. 40, pp. 45-50.
Plumelle, C. (1987), Experimentation en vraie grandeur d'une paroi
clouee, Revue Franc:aise de Geotechnique, No. 40, pp. 45-50.
PlumeIle, C. and Gasnier, R. (1984), Etude Experimentale en Vraie
Grandeur de Tirants d'Ancrage (Full scale tests on ground anchors),
Proceedings of the International Symposium on In-Situ Reinforcement
in Soils and Rocks, Paris.
Rabcewicz, L. V. (1964-65), The new Austrian tunnelling method, Parts
I to III, Water Power, London, Dec. 1964 and Jan. 1965.

Ground Anchors and Soil Nails in Retaining Structures


Rabejac, S. and Toudic, P. (1974), Construction d'un mur de soutennement entre Versailles-Matelos, Revue Generale des Chemins de Fer,
pp.232-237.
Romanoff, M. (1957), Underground corrosion, National Bureau of
Standards, Circular 579, p. 227.
Schlosser, F. (1983), Analogies et differences dans Ie comportement et
Ie calcul des ouvrages de soutennement en Terre Armee et par
clouge du sol, Annals de I'Institut Technique du Batiment et des
Travaux Publics, No. 418.
Schlosser, F. and Elias, V. (1978), Friction in reinforced earth, Symposium
on Earth Reinforcement, ASCE Annual Convention, Pittsburgh.
Schlosser, F. and Segrestin, P. (1979), Dimensionnement des Ouvrages
en Terre Armee par la Methode de I'Equilibre Local, International
Conference on Soil Reinforcement: Reinforced Earth and Other
Techniques, Paris.
Shafiee, S. (1986), Simulation Numerique de Comportement des Sols
Cloues; Interaction Sol-Renforcement et Comportement de L'ouvrage,
Ph.D. Dissertation, ENPC, Paris.
Shen, C. K., Herrmann, L. R., Romstand, K. M., Bang, S., Kim, Y. S.,
and Denatale, J. S. (1981a), In-situ Earth Reinforcement Lateral
Support System, Report No. 81-03, Department of Civil Engineering,
University of California, Davis.
Shen, C. K., Bang, S., Romstad, J. M., Kulchin, L., and Denatale, J. S.
(1981b), Field measurements of an earth support system, Journal
of the Geotechnical Engineering Division, ASCE, 107, No. GT-12.
Shen, C. K., Bang, S., and Hermann, L. R. (1981c), Ground movement
analysis of an earth support system, Journal of the Geotechnical
Engineering Division, ASCE, 107, No. GT-12.
Shields, D. R., Schnabel, Jr., H., and Weatherby, D. E. (1978), Load
transfer in pressure injected anchors, Journal of the Geotechnical
Engineering Division, ASCE, No. GT-9, pp. 1183-1196.
Simpson, B., O'Riordan, N. J., and Croft, D. D. (1979), A computer

905

model for the analysis of ground movements in London clay,


Geotechnique, 29, No.2, pp. 149-175.
Singh, A. and Mitchell, J. K. ( 1968), General stress-strain -time function
for soils, Journal of the Soil Mechanics and Foundations Division,
ASCE, 94, No. SM-l, pp. 21-46.
Welsh, J. P. et al. (1987), Soil improvement-A ten year update, ASCE
Geotechnical Special Publication No. 12.
Stocker, M. F., Korber, G. W., GassIer, G., and Gudehus, G. (1979),
Soil nailing, International Conference on Soil Reinforcement, Paris,
2, pp. 463-474.
Terzaghi, K. (1943), Theoretical Soil Mechanics, John Wiley and Sons,
Inc., New York, N.Y.
Terzaghi, K. and Peck, R. B., (1948 & 1967), Soil Mechanics in
Engineering Practice, John Wiley and Sons, Inc., New York, N.Y.
Tomlinson, M. J. (1957), The adhesion of piles driven in clay soils,
Proceedings of the 4th International Conference on Soil Mechanics
and Foundation Engineering, 2, pp. 66-71.
Tschebotarioff, G. P. (1951), Foundations, Retaining and Earth Structures,
McGraw-Hill Book Co., Inc., New York, N.Y.
Weatherby, D. E. (1982), Tiebacks, Federal Highway Administration,
Report No. FHWA/RD-82/047.
Wernick, R. (1977), Stresses and strains on the surfaces of anchors,
Specialty Session No.4, Proceedings of the 9th International
Conference on Soil Mechanics and Foundation Engineering, Tokyo,
pp. 113-119.
Winkler, E. (1867), Die Lehre von Elastizitat und Festigkeit, Prague.
Woodward, R. 1., Lundren, R., and Boitono, J. D. (1961), Pile loading
tests in stiff clays, Proceedings of the 5th International Conference
on Soil Mechanics and Foundation Engineering, 2, pp. 177-184.
Zaman, M. M., Desai, C. S., and Drumm, E. C. (1984), Interface model
for dynamic soil-structure interaction, Journal of the Geotechnical
Engineering Division, ASCE, 110, No. SM-9, pp. 1257-1273.

You might also like