ERS Ref Manual June 2008-343-792
ERS Ref Manual June 2008-343-792
ERS Ref Manual June 2008-343-792
8.1 INTRODUCTION
Externally supported structural walls rely primarily on the bending resistance of a vertical
structural element to resist the applied lateral loads. The vertical wall elements may consist
of discrete elements (e.g., soldier piles) spanned by a structural facing, or may be a
continuous structure (e.g., sheet pile wall, tangent pile wall, slurry wall, jet-grouted wall, and
deep mixing method (DMM) wall). The primary types of externally supported walls are
illustrated in Figure 8-1. The wall configurations include nongravity cantilevered walls and
structures with single or multiple levels of support. Wall support may be provided by a
system of struts or rakers on the exposed side of the wall, or anchors installed through the
wall (Figure 8-2).
Externally supported structural walls can be used for both temporary and permanent wall
applications. Typical applications for such walls are listed below:
• roadway cuts;
• roadway fill containment;
• abutments;
• cut-and-cover tunnel walls;
• slide stabilization; and
• waterfront structures.
5 to 10 ft 2 to 3 ft (e)
1.5 to 3 ft
(g)
6 to 10 ft
(h)
(c)
Figure 8-1. Primary Types of Externally Supported Structural Walls: (a) Soldier Pile and
Lagging Wall; (b) Soldier Pile and Cast-In-Place Concrete Lagging Wall; (c) Master Pile
Wall; (d) Sheet Pile Wall; (e) Slurry (Diaphragm) Wall; (f) Secant Pile Wall; (g) Tangent
Pile Wall; and (h) Interlocking H-Pile Wall (after Dismuke, 1991)
The primary advantages of externally supported structural walls (as compared to other types
of walls) include:
• top-down method of construction is used (i.e., wall elements are installed before the
start of excavation) which generally results in reduced ground displacement;
• when used as permanent walls they require;
o reduced quantity of excavation;
o reduced quantity of backfill;
o reduced work area;
• faster construction time;
• can be used to provide a seepage barrier (e.g., for a depressed roadway); and
• can effectively support large vertical loads as well as lateral loads.
8.2.1 General
A sheet pile wall consists of a series of interlocking sheet piles driven side by side into the
ground, thus forming a continuous vertical wall. Sheet pile walls are commonly used for
waterfront structures and for temporary earth support applications, but can also be used as
permanent walls for highway structures. Sheet piling is also used for stabilizing ground
slopes and for cellular cofferdam construction. Figure 8-3 show several applications for
sheet pile walls.
Sheet piles can be timber, reinforced concrete, steel, vinyl, or composite material. Timber
sheet piling is used for short spans, light lateral loads, and for temporary support of shallow
excavations. Concrete sheet piles are precast members, possibly prestressed, usually with a
tongue and groove joint, designed to withstand the permanent service loads and handling
stresses during construction. They are heavy and bulky, and require heavier equipment to
drive and handle. Concrete sheet piles are generally used only for permanent wall
applications. Steel sheet piles are the most commonly used type.
(a)
(b)
(d)
(e)
Figure 8-3. (a) Sheet Pile Wall for Earth Support Behind a Cast-in-Place Wall; (b)
Cofferdam for Construction of Foundations in Water; (c) Cofferdam for Footing
Construction on Land; (d) Anchored Bulkhead; and (e) Bridge Abutment.
• lightweight;
• easier to drive and extract;
• higher bending resistance; and
• sheet piles reusable several times for temporary walls.
Steel sheeting is fabricated either using hot rolling or a cold formed manufacturing process.
Interlocking Z-shaped or U-shaped sections are typically used for retaining wall applications
since these sections provide a higher bending resistance and corresponding greater moment
of inertia. Cold formed sections are generally lighter (i.e., lower section modulus) than hot-
rolled sections and the connection formed by cold-rolled interlocks is looser than for hot-
rolled interlocks. Flat sheets are generally limited to use in cellular cofferdams. Some
common steel sheet pile sections used for sheet pile walls are shown in Figure 8-4.
Figure 8-4. Steel Sheet Pile Sections Commonly Used for Retaining Walls and Cofferdams.
(a) Z -Section, (b) U-Section, (c) Cold Formed Section.
The two general construction sequences for sheet pile walls include: (1) driving sheet piles
into the ground and then backfilling behind the wall (i.e., “backfilled structure”); and (2)
driving sheet piles into the ground and excavating the soil in front of the wall (i.e.,
“excavated structure”).
The sequence of construction for a backfilled structure is illustrated in Figure 8-6 and
generally proceeds as follows:
Step 1: Excavate the in-situ soil in front and back of the proposed structure, if
necessary.
Step 2: Drive the sheet piles.
Step 3: Backfill behind the wall to the level of the anchor and install the anchor system.
Step 4: Backfill to the top of the wall.
For cantilever type of backfilled sheet pile wall (i.e., without anchors), the construction
sequence is the same except anchors are not installed.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-7 June 2008
Figure 8-6. Sequence of Construction for a Backfilled Sheet Pile Structure (after Das, 1990).
The sequence of construction for an excavated structure is illustrated in Figure 8-7 and
generally proceeds as follows:
For cantilever type of excavated sheet pile wall (i.e., without the anchors) the construction
sequence is the same except anchors are not installed.
There are three different sheetpile driving methods: (1) pitch and drive; (2) panel driving; and
(3) staggered driving. The pitch and drive method is the simplest way of driving but can be
used only for loose soils or short piles (Figure 8-8). This method involves driving each sheet
pile to full depth before installing the next one. Panel driving is used for dense sands and
stiff cohesive soils or for cases where there are possible obstructions in the ground (Figure 8-
9). Panel driving enables greater control of verticality and alignment. Staggered driving is
Figure 8-8. Sheet Pile Pitch and Drive Method (after TESPA, 2001).
used for difficult soil conditions and is combined with panel driving. The piles are installed
between guide frames and then driven in short steps as piles 1, 3, and 5 first then piles 2 and
4 (Figure 8-10). If the soil is very dense, sheet piles 1, 3, and 5 are reinforced at the toe. In
this case, these sheet piles are always driven first and piles 2 and 4 in the second stage.
Depending on the hammer type sheet piles may be installed in variety of in-situ soils,
however; it should be noted that sheet piles cannot penetrate very hard layers. The most
favorable in-situ soil conditions for vibratory hammer driving include rounded sand and
gravel and soft soils. The least favorable soil conditions are dense, angular soils. Also, dry
The most favorable in-situ soil conditions for impact hammer driving include soft soils such
as silts and peats, loosely deposited medium and coarse sands and gravels without rock
inclusions. The least favorable soil conditions are dense sands and gravels and hard clays.
As is the case with vibratory hammer driving, it is harder to drive sheet piles in dry soils than
moist or fully saturated soils.
For walls up to 15 ft, the typical cost range of cantilevered sheet pile walls (i.e., without
anchors) is about $15 to $40/ft2 of exposed wall face with the lower end cost associated with
temporary applications in which sheet piles are rented.
Because of increased productivity, sheet pile walls installed using vibratory hammers may be
less expensive than for walls installed with impact hammers. Vibratory hammers are also
very effective in pulling sheet piles out of the ground.
The wall facing is not included in the cost range given above. Costs for wall facings are
described in Chapter 10.
Inspector responsibilities for sheet pile wall construction are summarized in Table 8-1.
Table 8-1. Inspector Responsibilities for a Typical Sheet Pile Wall Project.
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor’s schedule
Review in-situ and backfill property test results (i.e., grain size, Atterberg limits, unit
weight, and shear strength)
Discuss anticipated ground conditions and potential problems with Contractor
Confirm that Contractor’s pile installation equipment and its size and type is consistent
with Specifications
Confirm that the layout and dimensions of sheet pile consistent with Plans
Confirm Contractor stockpile area consistent with Plans
Review corrosion protection requirements of metallic units (i.e., epoxy coating, paint)
and confirm consistency with Specifications
EXCAVATION, STEEL SHEETING AND REINFORCING ELEMENTS
Confirm that excavation of slopes and/or structural excavation support is consistent with
the Plans
Confirm that disposal of excavated material is performed according to Specifications
Confirm that during installation sheet pile with ball end driven first
Confirm embedment depth of sheet pile
Confirm the quality of interlock between adjacent sheet piles by visual inspection
Confirm that sheet piles are driven according to the Specifications
Visually inspect the top of sheet piles to asses any possible damage after installation
Report to the Engineer cave-ins, excessive yielding of sheet pile, threatened flooding of
excavation, unanticipated subsurface conditions
Confirm the reinforcing elements used are consistent with Plans and Specifications(1)
8.3.1 General
Soldier pile and lagging wall systems (also known as post and panel walls) are commonly
used for temporary excavation support (Figure 8-11) in dense or stiff soils where sheet pile
walls may not be suitable. They are also used frequently for permanent earth retaining
structures (Figure 8-12).
Soldier pile and lagging walls consist of soldier piles usually set at 5- to 10-ft spacing, and
lagging which spans the distance between the soldier piles. The lagging is used to retain the
soil face from sloughing and transmit the lateral earth pressure to the soldier piles. Included
in this category of walls is the master pile (or king pile) wall system, shown in Figure 8-1(c),
which consists of discrete vertical H-pile sections interlocked and alternating with steel sheet
pile sections.
The most common soldier piles are rolled steel sections, normally H-pile or wide flange
sections. However, soldier piles can be almost any structural member such as pipe or
channel sections, cast-in-place concrete, or precast concrete. Figure 8-13 shows several
types of soldier piles.
For temporary walls, lagging is most commonly wood, but may also consist of light steel
sheeting or precast concrete. Cast-in-place concrete sheeting (or facing) is generally used for
permanent wall applications, but such construction typically requires the use of temporary
sheeting to support the soil face during excavation and concrete placement operations.
(a) (b)
Figure 8-12. Permanent Soldier Pile and Lagging Walls for (a) Roadway Embankment, and
(b) Roadway Cut.
(c) (d)
Figure 8-13. Types of Soldier Piles (a) Wide Flange Section, (b) Pipe Section without
Anchor, (c) Double Channel Section, and (d) Pipe Section with Anchor (after Xanthakos, et
al. 1994).
In some cases, it may be more economical and rapid to attach the lagging boards to the
outside of the steel beams (i.e., contact lagging) using small steel attachment devices. Once
the attachment device is welded to the front of the beam, lagging board placement can be
done rapidly (see Figure 8-13 b and Figure 8-14). Where soils are particularly stiff, this
option may be considered.
Lagging may be omitted in hard clays, soft shales and soils with natural cementation,
provided that the soldier piles are installed at relatively close spacing and with adequate steps
taken to protect against erosion and spalling of soil at the face.
Soldier piles can be either driven or drilled in to the soil. Driven soldier piles can be installed
in dense or stiff soils where sheet pile walls may not be suitable. Drilled in soldier piles can
be installed in any soil type and even into bedrock. For a soldier pile and lagging wall to be
viable, the soil in between soldier piles has to be free standing to allow for lagging
installation. Soft clay and loose, cohesionless sands and silts may not be free standing.
• they are generally free draining and not suitable for applications where it is necessary
to maintain the groundwater level behind the wall;
• steel elements in direct contact with soil are more susceptible to corrosion;
• greater ground displacement in comparison to stiffer wall systems (i.e., slurry walls,
tangent pile walls, etc.), especially if used with relatively soft ground; and
• excavation for placement of the lagging between soldier piles increase risk of ground
loss.
The construction procedures presented below follow those described by Goldberg et al.
(1976) for soldier pile and lagging walls which are braced (using struts or rakers). Anchored
soldier pile and lagging walls are discussed in Section 8.8 and 8.10.
The soldier piles are installed by driving, or by concreting them within pre-drilled holes
(Figure 8-15). After installation of the piles, the excavation is performed from the top down.
If braces are used then the excavation is performed to the first support level, placing lagging
as the excavation proceeds. Brackets are then attached to the soldier pile to support the wale
(if used), and the wale is then placed in position and connected to the soldier piles. The brace
is cut slightly short to facilitate placement. This extra space is closed by plates and wedges
when the final connection is made. The above sequence of excavation, and installation of
lagging, walls and braces is continued until the required bottom of excavation elevation is
reached.
8.3.2.2 Preloading
Preloading of supports is conducted for braced soldier pile and lagging walls. Preloading is
usually required for installation of bracing members since it results in more reliable load
determination and load distribution within the structural support system, and it effectively
reduces ground displacements adjacent to the excavation. Preloading is particularly
important when the excavation is located near structures or other facilities which may be
damaged by settlement or lateral ground movement. Cross-lot bracing members are
commonly preloaded when they are installed. Rakers and corner braces are not generally
preloaded due to the more complicated skew connections required, however, if necessary,
these connections can be made allowing for preloading.
Preloading of internal braces is accomplished by loading hydraulic jacks to the desired load
followed by securing the member with steel blocking, steel wedges and welding. One
procedure is to jack to the desired load, and then to drive steel wedges between the member
and the wale until the jack load is essentially zero. A second procedure is to weld the
connection tight while maintaining the jack load and then drop the pressure in the hydraulic
jack, thus transferring the load through the connection to the wale. The second procedure
may result in additional wall movement as the load is transferred, although the magnitude of
movement is generally small.
Figure 8-15. Soldier Pile and Lagging Wall: (a) Drilling for Pile Installation, and (b)
Excavation between Soldier Piles in Cohesive Soil for Installation of Lagging.
High preloads may cause over stressing of struts because of unforeseen job conditions or
temperature effects. Accordingly, the general practice is to preload bracing members to
about 25 to 50 percent of the bracing loads. This preload removes the slack from the support
system and at the same time reduces the risk of over stressing. Larger preloads, up to as
much as 80 percent of the bracing load, may be desirable to further reduce ground
movements and protect adjacent structures from settlement.
Temporary support bracing, or struts, are generally removed during or after construction of
the permanent structure elements (i.e., invert and/or roof slabs). Also, it may occasionally be
necessary to remove and reinstall temporary bracing (rebracing) to maintain lateral support
while installing the permanent structure elements. Strut removal (and rebracing) may be an
additional source of wall and ground displacement. Factors controlling the amount of
displacement are the wall stiffness, the properties of the retained soil, the span distance
between remaining braces, and the quality and the compaction of backfill between the final
structure and the excavation support wall.
Drilled-in soldier piles require the predrilled holes to be filled with concrete backfill.
General design recommendations for concrete backfill include the use of structural concrete
from the bottom of the hole to the excavation base and lean-mix concrete for the remainder
of the hole. The design concept is to provide maximum strength and load transfer in the
8.3.3 Cost
Cost for soldier pile and lagging walls up to 15 ft in height is typically about 15 to 20/ft2 of
exposed wall face. One of the factors affecting the cost of soldier pile and lagging walls is
the equipment that is used for soldier pile installation. As mentioned previously, soldier piles
can either be driven or drilled in to the soil. Drilling is typically more expensive than driving
the pile; however, wall alignment can be achieved easier with drilling than with driving.
Compared to driving, the use of predrilling can add $10 to $20 per ft of soldier pile.
The inspector responsibilities for a typical soldier pile and lagging wall are summarized in
Table 8-2.
Table 8-2. Inspector Responsibilities for a Typical Soldier Pile and Lagging Wall Project.
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor’s schedule
Review test results and certifications for materials (e.g., steel soldier piles, wood lagging,
structural concrete, lean-mix concrete)
Discuss anticipated ground conditions and potential problems with Contractor
Check overall condition of Contractor’s pile driving or drilling equipment
Confirm that the layouts and dimensions of the wall consistent with Plans
Confirm that Contractor stockpile and staging area consistent with locations on Plans
Review Contractor’s survey results against Plans
SOLDIER PILE
Review corrosion protection requirements of soldier piles and confirm that Contractor is
following these requirements
Confirm soldier pile driving operations or drilling consistent with Specifications
Inspect driven soldier piles to assess damage at the top of the piles
Inspect the integrity of soldier pile holes before they are filled with concrete backfill
Confirm that loose material is removed from the bottom of the hole (for drilled-in)
Confirm embedment depth of soldier piles consistent with Plans and can be achieved
without damage to soldier pile
Check that soldier pile is properly aligned before concrete placement and remains aligned
during concrete placement
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-19 June 2008
LAGGING
Confirm that lagging is placed at an appropriate time after excavation to ensure no local
soil failure occurs
Report observations of over cutting and/or significant soil sloughing to Engineer
Confirm placement of lagging from top-down in sufficiently small lifts based on
Specifications
Confirm that gap is left between vertically adjacent lagging boards
DRAINAGE SYSTEMS AND BACKFIILL
Confirm backfill materials used are approved by Engineer
Verify alignment and elevation of drainage systems
Confirm that delivery, storage, and handling of the prefabricated drainage composite (i.e.,
geocomposite) is performed consistent with Plans and Specifications.
Confirm that geocomposite is not damaged in any way, while it is being installed and is
not exposed to excessive dust that could potentially clog the system
Confirm that geocomposite strips are placed and secured tightly against the lagging with
the fabric facing the lagging
Confirm placement of drainage aggregate in front of the lagging in horizontal lifts
Confirm that construction of drainage aggregate closely follow the construction of
precast facing elements
Confirm that perforated collector pipe is placed within the permeable material to the flow
line elevations and at the location shown on Plans
POST INSTALLATION
Verify pay quantities
8.4.1 General
A slurry wall (or diaphragm wall) is a structural, cast-in-place concrete wall constructed by
tremie placement of concrete in a pre-excavated, slurry-filled trench, as shown in Figure 8-
16. The wall is constructed in a series of panels which interlock to form a stiff continuous
structure. Commonly, trenches 2 to 3 ft wide are excavated in lengths of 10 to 20 ft.
Slurry walls are used at sites where it is necessary to restrict ground displacements adjacent
to the excavation. This is a particular concern when the excavation is in close proximity to a
building or other structure which is founded above the bottom of excavation level. As a
relatively rigid excavation support system, a slurry wall typically results in considerably
smaller ground displacements than a sheet pile or soldier pile and lagging support system.
• boulders, cobbles, and other obstructions are removed as part of the trenching
operation;
• can be used solely for temporary support of excavations, or serve both as a temporary
excavation support and as the permanent structural wall;
• eliminates the need for a costly cast-in-place concrete interior wall;
• allows a reduction in the width of the excavation; constructed as top-down
construction which may be advantageous for projects that require a minimum
duration for surface disturbance (i.e., construction of a depressed underpass at a busy
intersection); and
• minimal vibrations during construction.
Although previously described as a general sequence of construction for a wall system, the
term “top-down construction” may also be used to describe a method of constructing
excavations (usually in urban environments) that allow the structure (e.g., tunnel) to be built
within the excavation as the excavation proceeds rather than starting construction after the
bottom elevation of the excavation is reached. Oftentimes, a slurry wall is used for this
application (although other diaphragm wall types can be used). The general sequence for
top-down construction includes:
2. Soil is excavated to just below the elevation of the roof slab of the underground
structure. Struts are installed to support the walls.
3. The roof slab is constructed with access openings on the slab to allow work to
proceed downwards.
4. The next slab level is constructed and this process proceeds downwards until the base
slab is completed.
6. After the underground structure is completed, soil backfill is placed to the top strut
level and then this strut is removed. Backfilling to the ground surface is then
completed.
The two primary types of slurry walls are the conventional reinforced concrete wall and the
Soldier-Pile-Tremie-Concrete (SPTC) wall. Other wall types include the Precast-Concrete-
Panel wall and the Post-Tensioned Concrete wall. These walls are briefly described below.
A conventional reinforced concrete slurry wall includes steel reinforcement cages that are
placed in the slurry trench before the concrete tremie pour. The reinforcing bars are sized to
resist bending principally in the vertical direction between bracing levels. Supplemental
horizontal bars may be designed to serve as “internal walers” or to distribute forces around
inserts or openings. External wales are then eliminated provided that lateral bracing supports
are provided at panel joints. This method of reinforcement is currently most popular and is
used extensively throughout the U.S., (Figure 8-17). Panel end joints are formed by stop end
pipes, shown in Figure 8-16, or other suitable forming devices.
SPTC walls are becoming increasingly popular, particularly in deep excavations requiring
high bending resistance. These walls are also referred to as “Soldier Pile and Concrete
Lagging Walls” owing to the similarity with soldier pile and lagging walls. SPTC walls
commonly include vertical wide flange sections set in a slurry-stabilized trench. In some
cases, a reinforcing bar cage is placed between the soldier piles to transfer earth and water
loads laterally to the soldier piles. Alternatively, the concrete may be designed to span (arch)
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-23 June 2008
3 in
2 ft
3 ft
6.5 to 26 ft
between the soldier piles which eliminates the need for reinforcing bar cages between soldier
piles. In this case, soldier piles should be spaced at least 4.5 ft or more on center to facilitate
concrete placement. This slurry wall type is relatively watertight, has significant strength in
the vertical direction, and allows for relatively easy connection for temporary cross lot
bracing and wales (Figure 8-18).
A variation of the above wall types is the Precast-Concrete-Panel wall in which precast
concrete wall elements are placed in a slurry-stabilized excavated trench. This method
produces the best quality finished wall but is limited in use by cost, transportation and
handling length limitations, and other specific site constraints. Wall panels are best cast on
site adjacent to the work and installed directly into an oversized trench excavation. This type
of wall requires the use of cement-bentonite slurry which will eventually harden in the void
between the wall and soil. Rubber water stops are usually required between panels since the
cement-bentonite joints are subject to drying and shrinkage cracking (Figure 8-19). The
precast panels are temporarily suspended within the excavation until the cement-bentonite
hardens. Wall installation usually proceeds in a linear fashion with precast panel installation
following closely behind trench excavation. The use of a Precast-Concrete-Panel wall
becomes less practical for deeper excavations and at locations where utility crossings or
obstructions are expected.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-24 June 2008
2 ft
3 ft
6.5 to 16 ft
3 to 6 in
1.5 ft
2.5 ft
5 to 10 ft
Another wall type used successfully in Europe but which has had limited use in the U.S. is
the Post-Tensioned-Concrete wall (Figure 8-20). For this type of wall, post-tensioning
provides increased bending resistance. However, major disadvantages of this type of wall
are its higher cost and the need for specialized construction techniques.
3 ft
6.5 to 26 ft
8.4.3.1 Slurry
The slurry used to stabilize the excavated trench consists of bentonite and water. When
mixed with water, the bentonite (which comes as a powder, chips or pellets) forms a colloidal
suspension (or slurry). To date, the use of additives in the bentonite slurry has been limited.
Polymer slurries are generally not used for slurry wall construction since they are not as
effective as mineral slurries for supporting large size excavations in coarse granular soils.
To maintain a fluid slurry until concrete is completed, the slurry must be circulated and
agitated. Desanding devices are typically used to remove a sufficient amount of suspended
soil so that the slurry can be used two or more times prior to disposal. The contractor should
verify that a heavily contaminated slurry suspension, which could impair the free flow of
concrete, has not accumulated in the bottom of the trench.
Tamaro and Poletto (1992) recommend the following specifications for fresh bentonite slurry
(at the beginning of excavation):
• Sand content not more than 5 percent, measured at about 5 ft above the bottom of the
trench;
• Specific gravity not more than 1.10; and
• Viscosity not more than 50 seconds.
Additional information on material requirements and the use of slurry is provided in O’Neill
and Reese (1999).
8.4.3.2 Concrete
The concrete must be a free-flowing mix capable of displacing the bentonite slurry and
bonding to the reinforcement. Tamaro and Poletto (1992) recommend the following for
concrete:
In addition, Goldberg, et al. (1976) recommend a water/cement ratio less than 0.6, a sand
content of 35 to 40 percent of the total weight of aggregate, and a cement content of at least
25 pcf of the tremie concrete.
Premature stiffening of the cement may negatively affect the tremie operation. Retarders are
sometimes added to the mix to keep the concrete workable during the entire pour. Some of
the retarders, however, may reduce the strength. The retarders most commonly used are
discussed in Xanthakos (1994).
8.4.3.3 Reinforcement
The slurry wall reinforcing can be in the form of a rebar cage, or a combination of a cage and
vertical wide flange sections. The rebar cage can be prefabricated and assembled either in a
shop or at the site.
The minimum bar spacing is generally 6 in. for vertical bars and 12 in. for horizontal bars (6
in. for horizontal spacing with internal wales). Selection of minimum bar spacing should
also consider reduced spacing caused by splices, vertical picking bars used to lift the cage,
and other items (e.g., inclinometer, pipes for non-destructive testing) tied to the cage. There
must be sufficient space between the bars to enable the free flow of tremie concrete and
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-27 June 2008
permit scouring of the reinforcing bars to remove slurry and provide good bond between
concrete and the reinforcing bars. The rebar cages, as well as inserts (e.g., sleeves for soil or
rock anchors, casing for instrumentation, etc.), should be secured with tie wire.
Spacer devices should be used on the outside of the rebar cage to provide a minimum
concrete cover of 3 in.
8.4.3.4 Equipment
Another type of slurry wall excavation equipment is the “hydromill” which was developed
by French and Italian equipment manufacturers (Figure 8-21b). This excavator is basically a
grinding device consisting of two milling heads rotating in opposite directions about axes
perpendicular to the trench. The rotating heads excavate soil and soft rock from the bottom
of the trench, and the excavated material is lifted by suction to the surface where the soil is
removed from the bentonite by sand separators.
For excavation in rock, heavy drop chisels, chisel drills, or large diameter roller bits can be
used. Pre-drilling can also be used to facilitate advancement of the clamshell through hard
layers.
8.4.3.5 Procedures
Construction of a slurry wall includes five primary elements: placement of guide walls,
trench excavation, placement of reinforcement, concreting, and the formation of joints.
Guide Walls
The construction of a slurry wall usually begins with the installation of guide walls. The
purpose of the guide walls is to: (1) prevent caving of the trench wall in the uppermost part of
the excavation; (2) align the trench; (3) contain the slurry; and (4) support suspended precast
elements in Precast-Concrete-Panel walls, or reinforcing steel in Cast-In-Place walls. Figure
8-22 shows a cross-section of a guide wall.
(b)
Figure 8-21. Slurry Trench Excavation Equipment: (a) Hydraulically Operated Clamshell
Bucket, and (b) Hydromill.
4 – 5 ± ft
Figure 8-22. Cross-Section of a Guide Wall (a) Compact Cohesive Soil, (b) Loose
Cohesionless Soil (Goldberg, et al., 1976).
The guide walls are usually 6 to 12 in. thick and 3 to 6 ft deep. To provide the necessary
support for suspended rebar cages or pre-cast panels, the guide walls should be cast on a
stable subgrade.
Trench Excavation
In conventional bucket excavation, the bucket brings the material to the surface, discharges
its load, and then is lowered back into the trench. With direct or reverse circulation
equipment, the material is broken up into smaller particles so that it may be suspended in the
bentonite slurry, which is circulated to the surface, screened and desanded. The cuttings are
brought to the surface by suction and/or air lift through suction pipes, or the excavation tool
itself.
The stability of the trench is maintained by the slurry pressure on the trench wall, and soil
arching. Also, local penetration of the bentonite into pervious soils will provide some
cohesion that helps to prevent spalling.
The slurry level in the trench is maintained at an elevation at least 4 ft higher than that of the
groundwater table. A membrane or a “mudcake” is formed against the walls of the trench by
a combination of hydrostatic pressure, osmotic pressure, and electrolytic properties of the
colloid. This mudcake will maintain the pressure against the trench walls and prevent fluid
losses through pervious materials.
The verticality of the wall is checked as the excavation advances. If the excavation is found
to be out of vertical tolerance, the trench can be backfilled with lean concrete and the
excavation operation repeated. This technique can also be used to fill cavities formed in the
side wall of the trench due to caving.
Reinforcement
Prior to placement of the rebar cage in the slurry filled trench, the trench bottom should be
sounded by a weighted tape, or other means, to verify the depth of the trench and assess the
cleanliness of the trench bottom. Also, the slurry in the trench should be sampled and tested
to assure that it meets specification requirements. If necessary, slurry circulation and trench
bottom cleaning operations should be continued to meet the specified requirements.
The rebar cage and end stops should be placed in the trench as soon as possible after the
trench is cleaned and inspected (Figure 8-23). Figure 8-24 shows a reinforced panel in cast-
in-place slurry wall.
Figure 8-23. Slurry Wall Reinforcement Cage (a) On Fabrication Bed, Showing Styrofoam
Knock-Out Panel, and (b) During Lifting for Installation into Slurry Filled Trench.
Figure 8-24. Reinforced Panel in Cast-In-Place Slurry Wall (Tamaro and Poletto, 1992).
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-32 June 2008
In the Soldier-Pile-Tremie-Concrete wall, the soldier piles, together with the reinforcing
cage, are set within the excavated trench prior to concreting. An alternative approach is to
set the soldier piles in pre-augered holes and then excavate and place tremie concrete
between consecutive piles.
Concreting
The concrete should be placed as soon as practical after installation of the reinforcement
cage, and concrete placement should proceed continuously until completion of the slurry wall
panel. Concrete placement is performed through one or more tremie pipes lowered to the
bottom of each panel (Figure 8-25). Two or more tremie pipes may be used for long panel
lengths and for SPTC walls. The tremie pipe must remain embedded in fresh concrete a
minimum of 6 ft and a maximum of 15 ft. The tremie concrete displaces the bentonite slurry
progressively as it rises uniformly to the surface. The concrete should be sampled and tested
at the start of placement and at defined intervals during placement, to verify that the
delivered concrete meets the specified requirements.
Figure 8-25. Slurry Filled Trench with Tremie Pipes Just Prior to Concrete Placement.
A construction joint is provided between two adjacent panels. This joint should allow
excavation of the new panel without significant disturbance to the previously poured panel.
It should be watertight and capable of transferring shear and compressive stresses. The joint
can be formed using different configurations and details, but generally consists of a steel
tube, steel plate, or steel beam (Xanthakos, 1979).
Unless permanent steel piles are used, the end stops are removed after the initial set of the
concrete. For deep walls, it may be necessary to partially extract the end stops while
concrete is placed in the upper part of the wall panel. The end stops should be removed in a
smooth and continuous manner.
Slurry walls can be supported using either internal bracing or soil or rock anchors. Similar to
soldier pile and lagging walls, the support elements are generally preloaded when they are
installed to reduce ground displacements behind the wall, and to obtain more reliable and
more uniform loading in the support elements.
8.4.4 Cost
The cost of a slurry wall is typically between $60 and $100/ft2 of exposed wall face. The use
of braces or ground anchors would add to the cost of the wall. Typical mobilization costs for
slurry walls are around $50,000.
Slurry walls may be cost competitive compared to other wall systems when at least two of
the following conditions can be met (Godfrey, 1987):
Slurry (diaphragm) wall construction requires relatively detailed construction inspection due
to the use of specialized equipment and materials. These walls are usually constructed by
specialty contractors.
On-site quality control of the slurry and concrete is also necessary. The inspection of slurry
should include: (1) documentation of bentonite source and quality; (2) records indicating age
of slurry (i.e., time for bentonite hydration); and (3) results of on-site tests for slurry
viscosity, specific gravity, pH, and sand content. During placement of concrete, plots of
concrete volume placed versus the rise of the concrete within the excavated trench should be
recorded. This information can be used to evaluate whether a cave-in has occurred.
After the wall has been exposed due to excavation, the wall should be checked to verify that
excessive seepage is not occurring through vertical panel joints or through any openings in
the wall face. Inspection responsibilities for slurry walls are summarized in Table 8-3.
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor schedule
Discuss anticipated ground conditions and potential problems with Contractor
Check overall condition of Contractor equipment
Verify that the layouts and dimensions of the wall consistent with Plans
Review Contractor stockpile area
GUIDE WALL
Verify that no caving occurs in the uppermost part of the trench excavation
Confirm alignment of the guide wall trench
Confirm consistency of Contractor’s guide wall details, width, and height with Plans
Observe placement of concrete to form the Guide Wall
Confirm that inside face of the guide wall does not have any ridges or abrupt changes and
the guide wall face and top does not vary horizontally or vertically from a straight line or
specified profile in the Plans
Confirm that the distance between inside faces of the guide wall is within the width of
slurry wall plus the distance specified in Specifications
Confirm consistency of placement and compaction of backfill soil behind the guide wall
with Specifications
Confirm that guide wall is effectively containing the slurry without losses
Confirm that guide walls are removed at the end of construction
8.5.1 General
A tangent or secant pile wall consists of a line of drilled shafts (also referred to as bored
piles) (Figure 8-26). If the bored piles are contiguous, or tangent, to each other, the wall is
called a “tangent pile” wall. In an alternate case, referred to as a “secant pile” wall, the pile
elements overlap so as to form an interlocking wall. Another variation of this wall type is
called an “intermittent wall” in which the piles are installed at a spacing exceeding the pile
diameter; this type of wall can be considered only if the ground is stable or secondary
elements, such as shotcrete or a cast-in-place facing, is used to provide a continuous wall.
Various configurations of bored pile walls are shown in Figure 8-27.
(a) (b)
Figure 8-26. Tangent Pile Wall a) With Structural Steel Section as Reinforcement, and b)
Face of Completed Wall.
Tangent pile walls and secant pile walls are stiff continuous walls that are constructed by the
top-down method. Similar to slurry (diaphragm) walls, tangent pile and secant pile walls can
be used when it is necessary to minimize groundwater lowering outside the excavation or to
reduce ground displacements. Also, similar to slurry walls, tangent pile and secant pile walls
can be used for either temporary or permanent ground support.
• they can be constructed using conventional drilled shaft excavation equipment and
procedures; and
(c) (d)
(e)
Figure 8-27. Various Configurations of Bored Pile Walls: (a) Tangent Pile Wall; (b)
Staggered Tangent Pile Wall; (c) Secant Pile Wall; (d) Intermittent Pile Wall with Grouted
Openings; and (e) Intermittent Pile Wall with Lagging.
• tangent / secant pile walls may be more suitable than slurry walls at work sites with
limited space since less area is needed for slurry containment and treatment, and for
fabrication and handling of rebar cages.
Limitations include:
Bored piles for a tangent or secant pile walls are usually installed using auger type drill rigs.
Use of temporary casing may be necessary when penetrating into unstable soil strata, which
may be susceptible to squeezing or caving. A faster and more economical construction
method, however, involves rotary drilling and the use of bentonite (or polymer) slurry to
keep the drilled-hole stable and remove the excavated material to the surface by reverse
circulation (FHWA, 1988).
Figure 8-28 illustrates the typical construction sequence for tangent pile and secant pile
walls. The piles should be installed in a staggered pattern to avoid disturbing the concrete in
an adjacent pile that has not completely cured.
Figure 8-28. Construction Sequence (a) Tangent Pile and (b) Secant Pile Walls
(after Xanthakos, 1994).
For a tangent pile wall, shown in Figure 8-28a, the direction of pile installation is from the
edges of a section towards its center. This sequence prevents interference between adjacent
piles during concreting, and allows all the piles in the section to be installed along the same
alignment except possibly the center pile which may have to be displaced slightly to fit in the
remaining space and still be tangent with the two adjacent piles.
In secant pile walls (Figure 8-28b), alternate piles (numbered 1, 3, 5, etc.) are drilled and
concreted first with or without reinforcement. Reinforced piles (numbered 2, 4, 6, etc.) are
cut into these piles about one day later after the concrete in the first group has achieved its
initial set but before it becomes too hard. Sometimes, reinforcement is provided in every
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-39 June 2008
pile; however, this is generally practical only when the piles are reinforced with steel
sections. Due to the difficulty in cutting reinforcing bars, it is usually necessary to place
rebar cages only in alternate piles.
Favorable soil conditions for tangent and secant piles include soft clay and granular silty and
sandy soils. Maintaining vertical tolerances for a wall constructed in hard soils is difficult.
The wall facing may consist of reinforced shotcrete, pre-cast concrete panels, or cast-in-place
concrete. A drainage gallery may be provided behind the concrete facing to intercept and
channel any seepage that may penetrate the wall. Grouting is sometimes performed behind
the joints between piles to reduce water seepage.
8.5.3 Cost
Tangent pile walls are typically less expensive than secant pile walls simply because less
shafts are required for a given wall length. The cost of tangent pile walls up to 30 ft in height
is about $25 to $40/ft2 of wall face and the cost of secant pile walls up to 30 ft is about $30 to
$45/ft2 of exposed wall face.
One of the factors affecting the cost of tangent/secant pile walls is the drilling. Equipment
used for drilling should have adequate capacity to excavate a hole through soft and hard soils,
as well as, obstructions and in secant pile walls, through previously installed concrete piles.
When hard soils, or other material encountered cannot be drilled using conventional earth
augers, more costly alternative equipment has to be used.
For very competent hard soils, piles can be installed apart from each other (i.e., intermittent
pile wall). This wall type requires fewer piles for a given length than tangent or secant piles,
resulting in a smaller cost as compared to tangent and secant pile walls. The cost of an
intermittent wall (up to 30 ft in height) is about $20 to $35/ft2 of exposed wall face.
The cost of ground anchors and a permanent wall face are additional costs that should be
included. These costs are all described in Chapter 10.
Construction inspection for tangent pile walls and secant pile walls is similar to that for
drilled shafts. Inspection includes verifying that drilling techniques are consistent with the
soil type and ground-water conditions at the site and that all construction tolerances are
maintained. For tangent pile walls, it is important that vertical tolerances are maintained if a
relatively watertight wall is required. The inspection responsibilities of tangent pile and
secant pile walls are summarized in Table 8-4.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-40 June 2008
Table 8-4. Inspector Responsibilities for a Typical Tangent/Secant Pile Wall Project
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor schedule
Discuss anticipated ground conditions and potential problems with Contractor
Check overall condition of Contractor’s equipment
Review test results and certificates for materials (e.g., steel reinforcing, concrete)
Confirm consistency of location and positioning of each pile with Plans
PILE EXCAVATION
Observe Contractor’s test pile installation and confirm that drilling technique is
consistent with ground conditions
Confirm that drilling fluid is tested for density, fluid loss, viscosity, shear strength, sand
content, and pH according to Specifications
When penetrating unstable soil strata, confirm that Contractor is using temporary casing
(if required).
Confirm that upon completion of pile excavation that all loose soil and sediment is
removed to expose a firm base
Inspect pile excavation to its full length where practical and confirm no squeezing or
caving exist on the sides
Confirm that all drill holes are protected from surface water
REINFORCING
Confirm that reinforcing is free from rust and mud
Confirm that reinforcement is maintained in its correct position during concreting
For secant pile walls, confirm that overcutting of previously installed piles occurs
approximately one day after installation. Note, this will be site-specific and depends on
time required for concrete to reach minimum strengths and type of overcutting drilling
equipment used.
CONCRETING
Confirm that before placing concrete, no accumulation of drilling fluid or other
deleterious material exist at the base of the drill hole
Confirm that concrete for each pile is from the same source
Confirm that method of concrete pouring (e.g., free fall, tremie pipe, or concrete pumps)
is consistent with Specifications
Confirm that vertical tolerances are maintained as concrete is poured
To check for possible necking or loss of ground, verify the actual volume of concrete
pored for each pile consistent with calculated volume required
If required in the Specifications, observe the pile integrity test on each pile to locate
potential defects, necking, soil inclusions and to verify pile length. Report the results to
Engineer
Confirm that Engineer-rejected piles are replaced by the Contractor
WALL FACING
Confirm that wall facing is consistent with Plans and Specifications
If grouting is required, confirm consistency of grout and grouting operations behind the
joints between piles with Specifications and Plans.
8.6.1 General
Jet-grouted walls commonly use cement for grout. A variation of these walls is the lime
column wall wherein lime or lime-cement mix is used in a dry or liquid form to stabilize the
soil.
Jet grouting consists of injecting high pressure fluids into the ground through horizontal
nozzles to segregate the soil and mix it with a cementing agent. Segregation of the soil is
achieved by the high energy of the jet fluid(s) which may consist of the cementing agent or
another cutting fluid. Jet grouting is basically an erosion/replacement process which
removes a portion of the soil particles and replaces them with a mixture of soil and grout that
has high strength and low permeability when hardened.
The jet grouted elements, which make up the wall, are typically either overlapping
cylindrical columns or, in some applications, panel elements. Cylindrical column elements
are the most common and are formed by rotation of the high pressure fluid jets. Panel
elements are formed by allowing the jet to remain stationary, thus cutting and mixing planar
jet grouted elements.
Figure 8-29 shows various grout column layouts used in construction of jet-grouted walls.
The column layout is governed by the intended use of the wall as well as stability
considerations. For example, in cases where water tightness is the main concern, as in the
case of cutoff walls, then multiple rows of overlapping columns should be used. Multiple
rows are also required if the wall is to be designed as a gravity structure. Figure 8-30
illustrates jet-grouted wall applications for excavation, underpinning, settlement control, and
water control.
• jet grouting can be used with a wider range of soil types than any other grouting
technique, however, it is not suitable for highly plastic clays (see Figure 8-31).
• since large diameter columns can be created from relatively small boreholes, local
obstructions such as timber piles or large boulders can be bypassed or encapsulated
into jet-grouted soil mass;
• jet grouting can be conducted from any suitable access point, and can be terminated at
any elevation, providing treatment only in the target zones; and
• due to the high pressure used, there is a possibility of ground heave or lateral
movements which may damage adjacent utilities and underground structures;
• spoil handling and removal may be particularly difficult if the jet-grouted soil is
contaminated;
• jet-grouted soil strengths tend to be much more variable than concrete strength as
they are strongly influenced by the silt and clay content of the native soils and it is
difficult to predict the final strength of the jet-grouted soil during the design stage;
furthermore, if groundwater flow velocities are high, the fluid soil-grout mass may
experience local removal of the cement (bleeding) prior to its stiffening, and hence
variability in quality may be observed;
• it requires a higher level of quality control than for other wall systems; and
The basic procedure for jet grouting is shown in Figure 8-32 and summarized below:
• Drill and stabilize a hole to the required depth in the soil to be treated.
• If separate drilling and grouting equipment are used, withdraw the drilling bit and
rods and insert the jet grouting monitor and special grouting rods to the bottom of the
predrilled hole. In many jet grouting systems and ground conditions, however, the
same equipment is used to drill the hole and then perform the jet grouting. In these
cases, after the hole is drilled to the required depth, a steel ball (check ball) is inserted
into the drill rods to redirect the fluid jets to the horizontal nozzles and allow jet
grouting to begin (Figure 8-33).
There are basically three jet grouting systems in general use: the single, double, and triple
fluid systems. A schematic representation of each system is shown in Figure 8-34. The
principal characteristics of each system are described below (Kauschinger and Welsh, 1989
and Bruce, 1994).
• It uses a grout jet to simultaneously erode the soil and mix it with the cement grout.
• It involves only partial replacement of the soil and therefore results in the lowest
volume of spoils as compared to other systems.
• It uses a grout jet engulfed in compressed air. The compressed air enhances the
cutting ability of the grout jet.
• The equipment is more complex and susceptible to clogging. The pathway for the air
between the inner rod (carrying the grout) and the outer rod must be kept open, or the
process will revert to a single fluid system.
o The air acts as a buffer between the jet stream and any groundwater present,
thus permitting deeper penetration by the jet;
o The soil cut by the jet is prevented from falling back onto the jet, thus
reducing the energy lost through the turbulent action of the cut soil; and
o The cut soil is more efficiently removed from the region of jetting by the
bubbling action of the compressed air.
• A potential drawback of this system is that the soil-grout mix may have a higher air
content, and therefore may have a lower strength than those of the other systems.
• Fluids are emitted from two levels. An upper water jet engulfed in compressed air is
used to excavate the soil which is then mixed with, or replaced by, a grout jet emitted
from a lower port.
• It permits virtually full replacement of the jetted soil, and provides the largest column
diameter. This also results in the largest amount of spoils/cuttings.
• In this system, unlike the double fluid system, the grout is not injected with air.
Hence, there is no problem with high air content in the final jet-grouted soil mass.
In general, a jet grouting system conducts incompressible fluids and its success is contingent
on maintaining free and efficient fluid circulation. In particular, the cuttings must flow freely
from the point of injection up to the ground surface. Otherwise, pressures, up to the jetting
pressures may build up in the soil. These pressures could hydrofracture the soil and cause
severe lateral soil movement and ground heave (Kauschinger and Welsh, 1989).
A typical jet grouting set up consists of: (1) a drilling rig; (2) an automated grout mixing
plant; and (3) a grout injection plant, which consists of automatic batchers and high-pressure
pumps. In multiple fluid systems, additional pumps and an air compressor are used (Figure
8-35). Figures 8-36 through 8-39 show the jet grouting equipment in operation.
The drill rig automatically regulates rotation and withdrawal rates of a string of special drill
rods and a jet grouting monitor which is mounted at the end of the drill string. The jet
grouting monitor is a special tool through which the jet grouting fluids are passed and
Rotary drilling is most commonly used. In coarse-grained soils, which may include cobbles
and boulders, rotary percussion is more suitable. The mast length of the drilling rig is an
important consideration on construction sites with overhead obstructions.
Controlled jet grouting creates a spoil material during the erosion and mixing process. The
volume of soil-cement spoil can be predicted from the injected volumes. The spoil usually
contains significant cement content and gains strength over time. Within 12 hours, it can
typically be handled as a firm to stiff clay and can be used as a construction material or be
carried away from the site for disposal. If jet grouting is used in contaminated ground,
special handling procedures may be required.
Neat, rapid setting, cement grout is typically used for jet grouting, although chemical grouts
can also be used. Grout viscosity and rigidity should be low to allow maximum penetration.
Portland Cement Types I, II or III are used. Type I is the most economical and is, therefore,
used when possible.
Water/cement ratios of 1:1 to 2:1 are commonly used. Where high strength is required, ratios
as low as 0.6:1 may be used. Potable water is normally used. Although not recommended,
salt water is sometimes allowed, provided no steel reinforcement is used.
Fly ash is sometimes added to the cement grout in ratios of cement:fly ash between 1:1 and
1:10 by weight. Where low permeability is needed, bentonite additives are often used. The
addition of 2 percent of bentonite to the grout mix can reduce shrinkage during curing. No
aggregates are added to the grout mix.
Figure 8-38. Forming a Jet-Grouted Column. Figure 8-39. Jet Grouting Monitor.
A field trial is considered an essential element of any jet grouting project. Its basic steps are
as follows:
• Select a test site with ground conditions similar to those of the constructed project.
• Design a field trial to model actual construction. In general, it is necessary to: (1)
perform the field trial with the same equipment to be used in the actual construction;
and (2) provide adequate instrumentation to study pore pressures, heave and lateral
deformation of the surrounding ground.
• Evaluate the quality of the jet grouted test section. This is usually done by: (1)
excavating and exposing columns to study the wall continuity; (2) drilling inclined
cores to estimate column diameter; (3) testing the strength of the cores obtained along
the centerline and near the column perimeters; and (4) performing field permeability
tests on completed columns or a pumping test on a completed cut-off enclosure.
The costs of jet grouted walls typically range from $60 to $90/ft2 of exposed wall face.
Additional costs will be incurred if ground anchors are used.
The mobilization/demobilization cost for jet grouted walls is similar for slurry walls, which is
typically around $50,000. The cost of cement has a significant impact on the overall cost of
the jet grouted walls. Typically, cement is 40 to 45 percent of the entire jet grouted wall cost.
Also, jet grouted walls require a specialty contractor with special equipment. This
requirement makes jet grouted walls expensive compared to sheet pile, soldier pile, or
tangent/secant pile walls.
8.6.4.1 General
To ensure the quality of the completed wall, it is critical to exercise strict control and
monitoring during the construction stage. Such control and monitoring will also provide data
to verify the design criteria and to modify, if necessary, the jet grouting parameters.
Depending on the scale of the project, the following controls can be exercised (Bruce, 1994):
In addition to the above controls during construction, the following options may be
considered where applicable (Gallavresi, 1992):
• Penetration tests (such as standard penetration test or cone penetration test) and
pressuremeter tests to evaluate the bearing capacities of intercolumn soils.
An electronic data control system enables the site engineers to continuously monitor the
various operational parameters both in graphical and numerical form and immediately
identify and rectify any malfunctions such as drop in pressure or flow rate, or clogging of the
nozzles. The site engineer can identify any change in the soil condition and can accordingly
modify the jet grouting process. Furthermore, the owner can monitor the execution of the
work, confirming that it has been carried out correctly and within design limits. Finally, such
a system permits the use of the data stored magnetically for preparation of graphical outputs
showing the variation of the measured quantities with depth.
Since there are no standard design procedures for jet-grouted works, quality control testing
during the field trial and production grouting is critical to the success of the project. One of
the key elements of quality control is to have experienced, qualified individuals from both
the contractor and owner on site. In addition, the equipment and procedures should be
suitable for the site and the requirements of the project.
The first step in quality control involves the education of the field personnel in the details of
the project, the jet grouting process to be used, the jet grouting parameters and the design
requirements. All field staff should understand exactly what jet grouting can do, how the
system works, and what data must be recorded. They should also be familiar with the
project’s specifications and approved shop drawings.
The next phase is the actual implementation of the jet grouting work. At this phase, critical
information and observations must be made by the field personnel to ensure that the jet
grouting procedures and parameters are correctly applied for the project. Also, data collected
by the quality control staff can be used to identify potential problems or deficiencies in the
jet-grouted wall, which may require remediation prior to any excavation. Typical items
An example Guide Specification for Jet Grouting is provided in Elias et al. (2006) including
additional information on QA/QC for jet grouting projects.
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor’s schedule
Discuss anticipated ground conditions and potential problems with Contractor
Review quality assurance/quality control program for the project
Review Contractor’s survey for the location of the jet grout holes and confirm their
consistency against Plans
TEST SECTION
Confirm that location of the test section is consistent with Plans
Confirm consistency of Contractor’s drilling, batching, and jetting technique with
Specifications and report it to the Engineer
Retrieve wet-grab and core samples from the test section and confirm that satisfactory
unconfined compressive strengths is achieved
DRILLING
Confirm that angle and depth of drilling consistent with Plans
Confirm that technique used for drilling consistent with approved method after test
section results
Confirm drilling parameters (i.e., lift, speed, and rotation rate) consistent with what is
approved during test section operations
Review the soil obtained from drilled holes and report to the Engineer for his comparison
against the assumed soil type for design
BATCHING
Confirm the preparation of grout slurry for consistency in material content
Obtain sample from grout slurry and confirm its physical and chemical properties with
Specifications
JETTING
Confirm that volume of spoil created from jetting is consistent with what is being
predicted in Design
Test the water-cement ratio and confirm the results with Specifications
In the case spoil is considered as construction fill, confirm that it is stockpiled and left in
place according to Specifications
If monitoring system is required to be used, confirm that real-time electronic data is
obtained continuously during jet grouting
Monitor cement-grout injection pressures and rates, rotational speeds, penetration and
withdrawal rates, horizontal and vertical alignments during jet grouting
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-56 June 2008
Write daily reports explaining the production and submit it to the Engineer
Within 60 minutes of the withdrawal of jet grouting equipment, obtain wet grab soil-
cement samples at pre-determined depths according to the Specifications
Ensure that one wet grab soil-cement cylinder from each sampling depth, selected by the
Engineer, is tested to determine 7 and 28 days unconfined compressive strength. The
remaining samples should be tested at 56 days
As directed by the Engineer, obtain core samples to evaluate compressive strength, unit
weight, and composition of the soil-cement. Ensure that coring is not performed until
wet grab samples are tested
Ensure that the Contractor obtains vertical alignment profiles over the length of one soil-
cement element per day as directed by the Engineer. The Contractor shall advise the
Engineer within one hour after measuring the vertical alignment of any non-compliance
with tolerance requirements
POST INSTALLATION
Verify pay quantities
8.7.1 General
The DMM wall is a system of overlapped soil-cement columns formed by mixing the in-situ
soil with cement grout at depth. The overall process is termed “deep mixing method”
(DMM).
DMM differs from jet grouting in two ways. First, the equipment does not use high pressure
grout. Second, the DMM columns are built to defined dimensions, while the jet-grouted
columns are formed, under high pressure, to uncertain dimensions.
DMM columns can be arranged in a variety of patterns as shown in Figure 8-41. Wall
applications include containment/cutoff walls and structural retaining walls, as illustrated in
Figure 8-42.
• the DMM wall can be built in a broad range of soils including soft to very stiff and
low to highly plastic clay and silt, loose to dense sand, gravel and cobble, and soft
rock;
Figure 8-42. Various DMM Wall Applications: (a) as containment/cutoff wall and
(b) as structural retaining wall.
• a relatively short construction time and high production rate can be anticipated (as
compared to slurry walls, for example) because the method involves in situ materials
and multi-axis augers;
• the wall configuration is well defined, and the continuity between the columns is
maintained; and
• construction does not create significant noise and vibration, and the impact on
adjacent facilities is minimal.
• since the wall relies on use of in-situ soil as a construction material, all obstructions
such as rubble, pieces of concrete, abandoned pipes and boulders must be completely
removed and replaced with suitable soil, otherwise, pre-boring may have to be used to
partially loosen and/or break up these obstructions prior to installation of the wall;
• the equipment and procedures used may not be easily amenable to variation in
column geometry with depth (e.g., the column cannot be wrapped around utilities),
therefore, relocation of utilities may be required;
• until the soil-cement is hardened, the constructed column may constitute a weak spot
that may trigger movement of adjacent structures or utilities, thus making
underpinning or temporary protection works necessary; and
• although the quantity of waste is generally less than for other methods, soil mixing
may produce a spoil volume equivalent to between 30 and 100 percent of the in-situ
soil volume, depending on in-situ soil moisture content thus making disposal of the
waste a significant cost item.
Multiple hollow-stem augers equipped with mixing paddles penetrate the ground to the
required depth and back. During penetration and withdrawal, cement grout is pumped
through the auger stem. The auger flights and mixing paddles mix the soil and the grout in
place to form continuous overlapping soil-cement columns that constitute the wall (Figure 8-
43). If needed, steel reinforcement is inserted in the DMM column (Figure 8-41). Additional
details on deep mixing methods are provided in Elias et al. (2006).
Figures 8-44 and 8-45 show typical DMM equipment. A typical DMM system consists of a
mixing plant and a drilling/mixing unit. The mixing plant consists of a grout mixer, a grout
agitator, automatic batching scales to control grout composition, and a computer for mixing
and grout flow control. The drilling/mixing unit consists of multiple axis augers guided by a
vertical steel lead on a track-mounted base machine as shown in Figure 8-43. Two to five
augers can be used, with typical auger diameters varying from 22 to 42 in. The base
machine, together with the lead, are supported at three points during operation for
maintaining accurate vertical alignment which is critical for eliminating unmixed zones
between columns and maintaining wall continuity.
The typical auger has an auger head, discontinuous auger flights and mixing paddles. The
auger flights are positioned so that they overlap with each other to form overlapped soil-
cement columns. The discontinuous auger flight is designed to provide some vertical
displacement of the soil for mixing, but also to prevent transporting the soil to the surface.
Thus, the auger mixes the grout with the soil at its original depth, uniformly and
continuously.
In recently developed equipment, jetting or spreadable mixing tools are used at the tip of the
auger to enhance the grout penetration and increase the column diameter at specific depths.
The interested reader is referred to Table 3 of Chapter 6 of Elias et al. (2006) for detailed
operational information for installation equipment currently available.
DMM walls are installed by constructing a series of sets or elements. Typically, a set
consists of three overlapping columns as shown in Figure 8-46. The stepped installation
procedure, which is commonly used, drills two primary column sets followed by drilling a
secondary column set using two boundary columns of the primary column sets as guide holes
to construct a continuous wall as shown in Figure 8-41. The redrilling in this procedure
increases the uniformity of the soil-grout mix. The redrilling ratio can be adjusted according
to the required level of soil-mixing.
During the drilling process, grout is injected into the soil through the tip of the hollow-stem
auger. A separate positive displacement pump supplies the grout to each of the injecting
augers for accurate control of grout flow. The auger flights penetrate and loosen the soil, and
lift it to the mixing paddles which blend the grout with the soil. As the auger continues to
advance, the soil and the grout are remixed by additional paddles attached to the shaft.
About 60 to 80 percent of the slurry is injected as the augers penetrate downward and the
remainder is injected as they are withdrawn so that the mixing process is repeated on the way
up. If reinforcing elements are to be used, they are inserted into the soil-cement column
immediately after the auger is withdrawn.
1. Drilling Speed: The drilling speed is governed by the properties of the soil to be
treated, and the soil mixing effort required to obtain the design soil-cement properties.
2. Mix Design: The DMM wall usually comprises three basic materials: the soil, the
grout and the reinforcement, if any. The most suitable soil and grout mixes are
similar to those discussed under jet grouting. Portland cement is commonly used. A
small amount of bentonite is sometimes added to increase the workability of the soil-
mixing work. Cement-based additives such as silicate, slag and gypsum are used for
gaining strength in saline or organic soils, or for stabilizing contaminated soils.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-63 June 2008
The mix design is governed by the required engineering properties of the DMM wall
such as strength and permeability. The type of the soil essentially determines the
extent of improvement in engineering properties. Laboratory strength and
permeability tests are normally performed to identify the cement proportion that
could provide the required properties.
The final selection of a mix design is influenced by the selection of equipment and
installation procedures used, efficiency and economy. In the field, an automated
batching system measures the water, cement and other additives by weight to produce
a more reliable grout than that produced by a volumetric batch system. The desired
weight of each grout component can be entered in the computer at a control panel,
and changes to the mix design can be made by simply adjusting the component
weights at the control panel.
3. Grout flow rates: The grout flow rate is usually adjusted constantly to accommodate
varying drill speeds in different soil strata, so that the design volume of grout per unit
volume of in situ soil is maintained. The grout flow is electronically controlled.
8.7.3 Cost
Given the large cranes required to support the multiple mixing augers, coupled with the
specially designed on site batcher, the current minimum cost for mobilization/demobilization
for DMM is in the vicinity of $100,000. In addition, the cost per unit volume of mixed soil
may vary depending on the normal project variables, e.g. labor, size of the project, depth and
type of in-situ soil being treated. A typical cost for DMM walls is $40 to $55/ft2 of exposed
wall face.
During construction of the wall, field sampling of the soil-cement is performed regularly and
unconfined compressive strength and total unit weight is performed to confirm the design
parameters. Wet soil-cement samples should be obtained routinely and cured in the
laboratory for testing. After construction of a soil-cement column, core samples should be
obtained from the exposed wall according to a testing schedule to be determined based on the
site conditions observed during excavation. Usually, the core samples are tested after 7, 28
and 56 days to establish the strength increase with time. Based on the results of unconfined
compression testing for several projects as part of the Central Artery Project in Boston,
average strengths from wet soil-cement samples were approximately 50 percent greater than
for core samples from hardened columns. This was attributed to wet soil-cement sampling
It is recommended that for the finished DMM wall, direct shear tests and/or triaxial
compression tests be conducted for strength assessment, in addition to the unconfined
compression tests. For quality control during construction, however, the unconfined
compression test is adequate, and the results may be used as standard values.
The inspection responsibilities for DMM walls are summarized in Table 8-6.
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor’s schedule
Discuss anticipated ground conditions and potential problems with Contractor
Review test results of in-situ soils for consistency with design assumptions
Review Contractor’s surface water control methods and confirm its consistency with
Specifications
Review Contractor’s survey results to confirm column locations are consistent with Plans
TEST COLUMN
Confirm that location of the test column is consistent with Plans
Confirm that test column is constructed full scale as specified in Specifications
Review the test results for consistency with Specifications and report to the Engineer
DRILLING
Confirm that location of drilling consistent with Plans
Confirm that angle and depth of drilling consistent with Plans
Confirm that drilling speed consistent with Specifications
Review the soil from drilled holes and report to the Engineer for his comparison against
the assumed soil type for design
If needed, confirm that redrilling is performed according to Specifications
8.8.1 General
Ground anchors, sometimes referred to as tiebacks, are structural tension elements that
receive their support in soil or rock. The basic anchor components include: (1) anchorage;
(2) unbonded length; and (3) bond length. These and other components of a ground anchor
are shown schematically in Figure 8-47. The anchorage is the combined system of anchor
head, bearing plate, and trumpet that is capable of transmitting the prestressing force from the
prestressing steel (bar or strand) to the ground surface or the supported structure. The
unbonded length is that portion of the prestressing steel that is free to elongate elastically and
transfer the resisting force from the anchor bond length to the structure. A bondbreaker is a
smooth plastic sleeve that is placed over the tendon in the unbonded length to prevent the
prestressing steel from bonding to the surrounding grout. It enables the prestressing steel in
the unbonded length to elongate without obstruction during testing and stressing and leaves
the prestressing steel unbonded after lock-off. The tendon bond length is that length of the
prestressing steel that is bonded to the grout and is capable of transmitting the applied tensile
load into the ground. The anchor bond length should be located behind the critical failure
surface.
Un
bo
nd
Trumpet ed
Le
ng
th
Anchor Head
Sheath
Bearing Plate
An
ch
Te or
nd Bo
on nd
Bo Le
Wall nd ng
Le th
ng
th
Unbonded Tendon
Anchor Grout
An
Bonded Tendon Di cho
am r
ete
r
Anchors may be used for either temporary or permanent support applications. The design
and construction of both are similar, except that permanent anchors generally are designed
for lower stresses in the tendon, and are provided with added protection against corrosion.
The method of anchor construction is greatly influenced by the type of ground at the site.
Ground anchors can be used essentially in all cohesionless soil deposits. Good performance
has also been observed for anchors in stiff to hard cohesive soils with N-value greater than 9
(Weatherby, 1982). Soil deposits not generally suitable for the anchor bond length (Cheney,
1988) include:
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-68 June 2008
• organic soils;
• cohesive soils with an average liquidity index greater than 0.2; and
• cohesive soils with an average liquid limit greater than 50.
In addition to these criteria, caution should be exercised with soils that have a plasticity index
greater than 20 percent as they may exhibit excessive creep.
There are three main ground anchor types that are currently used in U.S. practice: (1) straight
shaft gravity-grouted ground anchors (Type A); (2) straight shaft pressure-grouted ground
anchors (Type B); and (3) post-grouted ground anchors (Type C). Although not commonly
used today in U.S. practice, another type of anchor is the underreamed anchor (Type D).
These ground anchor types are illustrated schematically in Figure 8-48.
Straight shaft gravity-grouted ground anchors (Type A) are typically installed in rock and
very stiff to hard cohesive soil deposits using either rotary drilling or hollow-stem auger
methods. Tremie (gravity displacement) methods are used to grout the anchor in a straight
shaft borehole. The borehole may be cased or uncased depending on the stability of the
borehole. Anchor resistance to pullout of the grouted anchor depends on the shear resistance
that is mobilized at the grout/ground interface.
Straight shaft pressure-grouted ground anchors (Type B) are most suitable for coarse granular
soils and weak fissured rock. This anchor type is also used in fine-grained cohesionless soils.
With this type of anchor, grout is injected into the bond zone under pressures greater than 50
psi. The borehole is typically drilled using a hollow stem auger or using rotary techniques
with drill casings. As the auger or casing is withdrawn, the grout is injected into the hole
under pressure until the entire anchor bond length is grouted. This grouting procedure
increases resistance to pullout relative to tremie grouting methods by: (1) increasing the
normal stress (i.e., confining pressure) on the grout bulb resulting from compaction of the
surrounding material locally around the grout bulb; and (2) increasing the effective diameter
of the grout bulb.
Post-grouted ground anchors (Type C) use delayed multiple grout injections to enlarge the
grout body of straight-shafted gravity grouted ground anchors. Each injection is separated by
one or two days. Postgrouting is accomplished through a sealed grout tube installed with the
tendon. The tube is equipped with check valves in the bond zone. The check valves allow
additional grout to be injected under high pressure into the initial grout which has set. The
high-pressure grout fractures the initial grout and wedges it outward into the soil enlarging
the grout body. Two fundamental types of post-grouted anchors are used. One system uses a
packer to isolate each valve. The other system pumps the grout down the post-grout tube
without controlling which valves are opened.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-69 June 2008
Figure 8-48. Main Types of Grouted Ground Anchors (after Littlejohn, 1990).
8.8.2.1 Materials
Cement Grout
The cement grout consists of sand, cement, and water. Under normal conditions, most
specialty contractors prefer not to use chemical additives in the grout (Munfakh, et al., 1987).
Portland cement with low sulfate content is normally used. The grout mix should attain a
minimum cube strength (AASHTO T 106) of 3,500 psi at 7 days (Cheney, 1988). The
water/cement ratio of tendon bonding grouts is usually in the range of 0.35 to 0.60.
Both bar and strand tendons are commonly used for soil and rock anchors for highway
applications in the U.S. Material specifications for bar and strand tendons are codified in
American Society for Testing and Materials (ASTM) A722 and ASTM A416, respectively.
Indented strand is codified in ASTM A886. Bar tendons are commonly available in 1 in.,
1.25 in., 1.375 in., 1.75 in., and 2.5 in. diameters in uncoupled lengths up to approximately
60 ft. For lengths greater than 60 ft and where space constraints limit bar tendon lengths,
couplers may be used to extend the tendon length. As compared to strand tendons, bars are
easier to stress and their load can be adjusted after lock-off.
Strand tendons comprise multiple seven-wire strands. The common strand in U.S. practice is
0.6 in. in diameter. Anchors using multiple strands have no practical load or anchor length
limitations. Tendon steels have sufficiently low relaxation properties to minimize long-term
anchor load losses. Couplers are available for individual seven-wire strands but are rarely
used since strand tendons can be manufactured in any length. Strand couplers are not
recommended for routine anchor projects as the diameter of the coupler is much larger than
the strand diameter, but strand couplers may be used to repair damaged tendons. Where
couplers are used, corrosion protection of the tendon at the location of the coupler must be
verified.
Spacer/centralizer units are placed at regular intervals (e.g., typically 10 ft) along the anchor
bond zone and within 3 ft of either end of the anchor tendon. For strand tendons, spacers
usually provide a minimum interstrand spacing of 0.25 to 0.50 in. and a minimum outer grout
cover of 0.5 in. Both spacers and centralizers should be made of non-corrosive materials and
be designed to permit free flow of grout. Figures 8-49 and 8-50 show a cut away section of a
bar and a strand tendon, respectively.
Epoxy-coated bar (AASHTO M284) and epoxy-coated strand (supplement to ASTM A882),
are becoming more widely used. The epoxy coating provides an additional layer of corrosion
protection in the unbonded and bond length as compared to bare prestressing steel.
Additional details on epoxy-coated bar and epoxy-coated strand can be found in PTI (2004).
CORRUGATED
PLASTIC
SHEATHING
CEMENT
GROUT
SPACER
BAR
Figure 8-49. Cut away view of bar tendon (after Sabatini et al., 1999).
CENTRALIZER
STRAND
SHEATH
GROUT
TUBE
SPACER
Figure 8-50. Cut away view of strand tendon (after Sabatini et al., 1999).
Ground anchors can be used with both flexible and stiff nongravity cantilevered walls,
however; soldier pile and lagging walls are the most commonly used type of anchored wall
system in U.S. The construction sequence for a permanent soldier pile and lagging wall is
illustrated in Figure 8-51 and the components of ground anchor construction is described
below.
Drilling
After drilling is completed and the hole is thoroughly flushed out, the drilled hole should be
probed to verify that no collapse of material has occurred before installation of the
prestressing element. This installation and the subsequent grouting should be carried out on
the same day as drilling to avoid potential ground deterioration.
Tendon Installation
Steel tendons should be stored indoors, if possible. If left outdoors, however, they should be
stacked off the ground under a waterproof cover that allows air circulation. Neither bare nor
coated tendons should be dragged across abrasive surfaces.
The tendon is lowered in the borehole manually or using mechanical equipment (Figure 8-
53). Immediately before its installation, the tendon should be carefully inspected for damage
and corrosion.
Grouting
The grouting operation involves injecting cement grout at the lowest point of the borehole so
the hole will fill evenly without air voids. A grout pipe is often tied to the tendon before
inserting it in the borehole.
The cement grout is prepared by batching the dry materials by mass and mixing them
mechanically, with water added, for at least two minutes in order to obtain a homogeneous
mix. High-speed colloidal mixers (1000 rpm minimum) or paddle mixers (150 rpm
minimum) are used for grout preparation. The colloidal mixers are preferred when grouting
water-bearing ground since grout dilution is minimal. After mixing, the grout is kept in
continuous motion until it is pumped to its final position. Grout should not be used after a
period equivalent to its initial setting time. Prior to grouting, all air in the pump and the line
should be removed. An airtight system should be maintained at all times during grouting.
Figure 8-52. Drilling Equipment for Installation of Ground Anchors: (a) Hollow Stem Auger
and (b) Down-the-Hole Hammer.
The stressing operation involves fitting the jack assembly on the anchor head, loading and/or
unloading of the anchors, locking off the load by anchor nuts or wedges, then removal of the
assembly from the anchor head. The equipment used for stressing operations need to be
calibrated. Stressing should not begin until the grout strength has reached a crushing strength
of approximately 3,500 to 4,500 psi. Cheney (1988) recommends a waiting period of 7 days
before stressing can take place.
The load testing of ground anchors should be considered an integral part of the design. The
typical types of load tests include:
• Preproduction Tests: These tests are used to verify the available resistance in the
anchor bond length and establish anchor loads for designs.
• Performance Tests: These tests are conducted at the beginning of construction and
periodically during construction to verify short and long-term performance of the
anchor under the design load.
• Proof Tests: These tests are used to determine the behavior of each production anchor
after installation.
• Lift-Off Tests: These tests are used to confirm the load in the tendon after completion
of installation and lock-off of the applied load.
Figure 8-54 shows a typical set-up for strand and bar anchor testing.
Preproduction tests may be specified when unusual conditions are identified during the
design stage. Situations prompting such tests include: (1) soil deposits in which no previous
experience exists; (2) very long anchors; (3) difficult drilling conditions; or (4) creep
susceptible soils in the bond length.
The preproduction tests can be performed under a separate test contract awarded during the
design stage, or at the beginning of the construction contract. The anchors used for
preproduction testing are usually not incorporated in the final structure because of the
potential damage that may be induced by the high test loads. Also, the bond length for
preproduction test anchors is typically shorter than that of the production anchors.
JACK
FIXED
BASE
(a)
(b)
Figure 8-54. Typical Equipment For Load Testing Of (A) Strand Ground Anchor And (B)
Bar Ground Anchor.
Performance tests provide the necessary information to verify that the production anchors
will be able to hold the design load without excessive movement or creep. Cheney (1988)
recommends performance tests be conducted on the first two anchors installed, then on 2
percent of the remaining anchors in rock or cohesionless soils and 10 percent in cohesive
soils.
The performance test is a cyclic test made by incrementally loading and unloading the anchor
until the maximum test load is reached. The typical loading and unloading sequence (PTI,
2004) is as follows:
AL, 0.25P, AL, 0.25P, 0.50P, AL, 0.25P, 0.50P, 0.75P, AL, 0.25P, 0.50P,
0.75P, 1.00P, AL, 0.25 P, 0.50 P, 0.75 P, 1.00P, 1.20 P, AL, 0.25 P, 0.50 P,
0.75P, 1.00P, 1.20P, 1.33P, 1.00 P, LL
where AL is the alignment load, P is the design load (i.e., unfactored anchor load), and LL is
the lock-off load.
Each load or unload increment is held constant just long enough to obtain the movement
reading but no longer than one minute. The maximum load should generally be held for one
hour in cohesive soils to determine the long-term creep potential. Coarse granular soils and
rock do not generally exhibit creep; creep tests in such deposits may be terminated if
negligible creep, i.e., less than 0.04 in. movement, is observed between the 1 minute and 10
minute readings of the test. The deflection measurements at the maximum load level (i.e.,
1.33 P) are taken at the following intervals: 1, 2, 3, 4, 5, 6, 10, 20, 30, 40, 50 and 60 minutes.
At each load increment, the total movement of the pulling head should be recorded to the
nearest 0.001 in.
Every production anchor (that is not performance tested) is proof tested. The proof test is a
single cycle test in which the test load is applied in increments until the maximum load. For
granular soils and rock, the maximum test load should be 1.33 times the anchor design (i.e.,
unfactored) load; for cohesive soils, the maximum test load should be 1.50 times the anchor
design (i.e., unfactored) load. The maximum load is held constant for at least one minute or
until the measured deflection is negligible. The typical loading and unloading sequence is as
follows:
An extended creep test is a long duration test (e.g., approximately 8 hours) that is used to
evaluate creep deformations of anchors. These tests are required for anchors installed in
cohesive soil having a plasticity index (PI) greater than 20 or liquid limit (LL) greater than
50. For these ground conditions, a minimum of two ground anchors should be subjected to
extended creep testing. Where performance or proof tests require extended load holds,
extended creep tests should be performed on several production anchors.
The test arrangement for an extended creep test is similar to that used for performance or
proof tests. The increments of load for an extended creep test are the same as those for a
performance test. At each load cycle, the load is held for a specific period of time and the
movement is recorded. During this observation period, the load should be held constant.
The load is assumed to remain reasonably constant if the deviation from the test pressure
does not exceed 50 psi. The loading schedule and observation periods for each load cycle in
an extended creep test for a permanent anchor are provided in Table 8-8.
Table 8-8. Load Schedule And Observation Periods For Extended Creep Test For Permanent
Anchor.
Extended creep tests are not normally performed on rock anchors since they do not exhibit
time dependent movements. However, anchors installed in very decomposed or argillaceous
rocks may exhibit significant creep behavior.
Three criteria are commonly used to determine the acceptability of anchor tests (PTI, 1996).
These are as follows:
• Creep;
• Movement; and
• Lock-off Load.
Creep
The creep amount shall not exceed 0.04 in. at the maximum test load during the period of 1
to 10 minutes. If this value is exceeded, then the total creep movement within the period of 6
to 60 minutes shall not exceed 0.08 in.
The creep behavior of epoxy coated strand itself is significant and the measured anchor creep
movements must be adjusted to reflect the behavior of the material. At a maximum test load
of 80 percent of the guaranteed ultimate tensile strength of the tendon, creep movements of
epoxy coated strand can be estimated to be 0.017 percent of the apparent free stressing length
during the 6 to 60 minute log cycle, but may be significantly higher than this value (PTI,
2004). PTI (2004) provides an approach to evaluate creep test data and acceptability based
on creep test results for epoxy-coated strand. For epoxy coated bars, these considerations do
not apply.
Movement
a) Residual Movement
The minimum apparent free length is calculated to verify that the anchor load is being
transferred beyond any potential failure or slip plane in accordance with the overall stability
requirements of the anchor-structure system. The minimum apparent free length at the
maximum test load, as calculated on the basis of elastic movement, should be not less than 80
percent of the designed free tendon length plus the jack length. If this criterion is not met,
the anchor should be reloaded up to two times more from the alignment load to the maximum
test load, and the calculation repeated on these cycles. If the criterion is still not met, then (i)
the cause of this inefficiency in load transfer should be investigated; and (ii) the anchor may
be rejected or assigned a lower resistance (or capacity).
A limit higher than 80 percent of the designed free length should be set in cases where future
additional movements occurring as a result of redistribution of the free length friction would
cause unacceptable structural movement.
The acceptance criterion based on maximum apparent free length was used in the past when
load transfer along the bond length was assumed to propagate at a uniform rate as the applied
load was increased. For that assumption, the maximum value of apparent free length was
restricted to elastic movements of 100 percent of the free length plus 50 percent of the bond
length plus the jack length. However, the concept of uniform distribution of bond is not valid
for soil anchors and only approximates the behavior of most rock anchors. The primary use
of this criterion is as an alternate acceptance criterion for proof tests in sound rock where
creep tests are waived. Anchors that do not pass this preliminary criterion are subsequently
creep tested to determine acceptability before a decision is made to reject the anchor.
The criteria for the minimum and maximum apparent free length, as described above, are not
strictly relevant if only total movement data are available. However, it is conventional to
apply these criteria also to total movement data when, from past experience or previous tests
in the same conditions, the magnitude of the residual movements is well known, and elastic
movements can, therefore, be estimated. In such cases, the criteria listed above should be
Lock-Off Load
After the factored anchor load of a production anchor has been verified by testing, the anchor
load is immediately transferred to the structure. The magnitude of this initial transfer load
must be determined based on structural design assumptions and the mechanical losses
associated with the tendon type selected.
The mechanical losses associated with transferring load to various anchorage systems and
long-term relaxation or creep should also be considered in determining the final transfer load.
Seating losses may vary from 0.08 in. for bars to 0.25 in. for strand; however, these values
will be dependent on specific details of the anchorage system.
The magnitude of the transfer load is generally specified, and should not exceed 70 percent
of the guaranteed ultimate tensile strength of the tendon. In practice, transfer loads of 80
percent of the design load are commonly specified. Higher transfer loads are sometimes used
when it is required to minimize long-term wall deflection. In selecting the transfer load, an
evaluation should be made to verify that the transfer load, particularly at the top anchor level,
will not exceed the ultimate passive resistance of the soil behind the wall since this could
cause ground displacement and heave which might damage existing facilities.
Lift-off tests are performed either during construction to check the magnitude of seating and
transfer losses or after construction to determine if long-term load losses are taking place.
The test is performed by applying load gradually until the tendon begins to elongate. When a
sudden deflection is observed on the dial gauge, the jack extension should be immediately
terminated, and the load required for lift-off recorded. This load should be approximately
equal to the design load plus an allowance for long-term losses. If the lift-off load varies
more than 5 percent from this value, the tendon load is adjusted and the lift-off test is
repeated. When the load in a strand tendon is more than 5 percent above the desired lock-off
load, and where no shims have been prepositioned under the wedge plate for later extraction,
then it is preferable to accept this load and so avoid the danger of having wedge marks below
the wedge plate as a result of strand/wedge regripping.
If an anchor does not reach the maximum test load as a consequence of bond failure,
subsequent actions depend on whether the anchor can be postgrouted or not. Regroutable
anchors should be postgrouted and then subjected to all the original acceptance criteria.
If an anchor fails the creep test at the maximum test load, then the anchor should be
postgrouted and subjected to an enhanced creep criterion, assuming the other acceptance
criteria are met. This enhanced criterion requires a creep movement of not more than 0.04 in.
between 1 and 60 minutes at the maximum test load. Anchors, which cannot be postgrouted
may be rejected or should be locked off at 50 percent of the maximum test load. In this
event, no further acceptance criteria are applied.
8.8.4 Cost
The cost of ground anchors ranges from approximately $45 to $60 per ft of anchor. Higher
costs are associated with longer anchors, rock drilling, and the need for larger diameter holes.
Typically, strand tendons are less expensive than steel bars, however strand tendons are not
suitable for ground anchors less than 40 ft long.
Ground anchors require typically require 0.6 H (where H is the height of the wall) distance
behind the wall face for installation. As with most wall systems, unit costs will increase
where access is limited and where right-of-way costs are significant.
Ground anchors are contracted using either method approach, performance approach, or
design/build approach. Method approach typically includes the development of a detailed set
of plans and material and construction specifications for the bidding documents. However,
the selection and installation of the anchors should be the responsibility of the contractor.
The contract documents should only establish minimum dimensions for drill hole diameter,
unbonded length, and bond length. The contractor should select the necessary anchor
installation dimensions and techniques to successfully pass the acceptance tests. In no case
should the owner specify the installation details for the anchors.
For anchored walls, the spacing, inclination, method of drilling, drill hole size, and length of
each anchor must be consistent with the design assumptions. Inspection personnel must pay
close attention to these issues because the ground anchors are the principal load-carrying
components of the wall system. Significant deviations in any of these items should be
reported to the design engineer.
CONTRACTOR SET UP
Review project Plans and Specifications
Review Contractor’s schedule and work plan describing ground anchor number, anchor
design load, type and size of tendon, minimum anchor length, minimum bond length,
minimum tendon length, and minimum unbonded length for consistency with
Specifications
Review the Plans submitted by Contractor for the ground anchor tendon and corrosion
protection system and check their consistency with Specifications
Review Contractor’s grout mix design and procedures for placement of the grout for
consistency with Specifications
Discuss anticipated ground conditions and potential problems with Contractor
Check overall condition of Contractor’s equipment
Confirm that the layouts and dimensions of the ground anchors consistent with Plans
Review Contractor’s stockpile area and confirm its consistency with Plans
VERTICAL WALL ELEMENT
Follow inspector responsibilities for the selected vertical wall element
ANCHOR HOLE DRILLING
Review Contractor’s survey results and confirm location of anchor holes based on Plans
Confirm consistency of length, orientation, and diameter of each hole with Plans and
Specifications
Confirm consistency of drilling method used with Specifications
Identify soil and rock types from samples obtained from each hole and report to the
Engineer
If used, confirm consistency of drilling muds and/or foams used during drilling holes
with Specifications
Confirm placement of centralizers according to Plans
TENDON INSERTION
Inspect cleanliness of drilled holes
Inspect all metallic units prior to installation to evaluate corrosion protection
Inspect all tendons for damage prior to insertion
Confirm the consistency of the dimensions of each tendon with Specifications prior to
insertion
Confirm consistency of tendon insertion with Specifications and be aware for any
indications of drill hole collapse
Confirm that tendon can be completely inserted without difficulty
ANCHOR GROUTING
Confirm that grouting equipment produces a grout, free of lumps and undispersed cement
Confirm consistency of grout injection pressures with Specifications and record the
pressures
Confirm that grout is injected at the lowest point in the drill hole.
Record grout volume placed in each hole
Confirm that withdrawal rate is less than grout placement rate
Cantilever sheet pile walls derive their support from the structural stiffness of the wall
sheeting and the passive soil resistance developed below the exposed base of the wall. This
type of wall is generally limited to heights up to about 15 ft and is generally suitable only in
granular soils or stiff clays. For cantilever walls greater than about 15 ft in height the
maximum bending moment in the wall may exceed the structural resistance of commonly
available sheeting, or ground displacement behind the wall may become excessive.
Typically, the required penetration of the sheeting will be 1.0 to 1.5 times the exposed height
of the wall, depending on soil and groundwater conditions, ground slope, and surcharge
loads.
Soldier pile and lagging walls derive their lateral resistance and bending moment capacity
through the embedment of vertical elements (soldier piles). The soil behind the wall is
retained by lagging. The spacing of the lagging varies from 6.5 to 10 ft with a common
spacing of 8 ft. Soldier pile and lagging walls are considered flexible walls with discrete
vertical wall elements. A portion of the load from the retained soil is transferred to the
The lateral displacement of a cantilever wall penetrating a sand layer is shown in Figure 8-
56. The wall rotates about point O, resulting in the reversal of active and passive earth
pressures in the three different zones shown. The corresponding net earth pressure
distribution on the wall is shown in Figure 8-56(b) and 8-56(c).
While the net earth pressure diagram is instructive in understanding the link between
deformations and earth pressures, the net pressure diagram cannot be used directly in LRFD
calculations because active earth pressures (and resultant forces) are treated using factored
loads whereas passive earth pressures are considered to be a resistance. For example, active
earth pressures at the Strength I limit state may use a load factor of 1.5 whereas the passive
resistance may use a resistance factor of 0.75, thus the active side and passive side of the wall
must be “factored” individually.
Table 8-10 summarizes the major design steps for a flexible nongravity cantilevered wall.
Herein, it is assumed that Step 1 has been completed and a nongravity cantilevered wall has
been deemed appropriate for the project and Steps 2 and 3 have been completed to establish
soil and/or rock parameters for design.
The earth pressure distribution that develops on nongravity cantilevered walls (e.g., sheet-
pile or soldier pile and lagging walls which are not anchored) can be expected to undergo
lateral deformations sufficiently large to induce active earth pressures for the entire wall
height. For design of these systems, theoretical active earth pressure diagrams using
Coulomb analysis methods to evaluate the active earth pressure coefficient are used.
AASHTO Figures 3.11.5.6-1 through 3.11.5.6-3 may be used to calculate the unfactored
lateral earth pressure distributions for permanent nongravity cantilevered walls. AASHTO
Figures 3.11.5.6-4 through 3.11.5.6-7 may be used to calculate the unfactored lateral earth
pressure distributions for temporary walls supporting or supported on cohesive soils with the
following restrictions:
Active
Passive pressure Zone B
pressure
Active o Passive
Zone C
pressure pressure
• The active earth pressure shall not be less than 0.25 times the effective overburden
pressure at any depth or 0.035 ksf/ft of wall height, whichever is greater.
For nongravity cantilevered wall designs involving clayey soils, the strength of the clay
changes with time and consequently the earth pressure changes with time also. Immediately
after the sheet piling is installed, the lateral earth pressures can be calculated based on the
total stress method of analysis, i.e., using undrained shear strength parameters. For
permanent wall design, both the short and long-term conditions need to be considered.
For the long term condition, sheet piling in clay should be analyzed using effective strength
parameters, c΄ and φ΄, obtained from triaxial shear strength tests. Assuming that the effective
cohesion value is small, the value of c΄ can conservatively be taken as zero. The long term
condition can then be analyzed based solely on the effective friction angle, φ΄, of the clay.
Information on the maximum wall friction (δ) for design is provided in Section 3.6. Where
the wall may settle relative to the retained ground, assume δ=0.
Other sources of lateral pressure including lateral pressures due to surcharges, seismic
pressures and water should be added to these diagrams using corresponding load factors for
each applicable limit state (as described in Section 8.9.2.3).
On the passive side of the excavation, for nongravity cantilevered walls with discrete
elements in competent ground, it is assumed that loading is resisted over a width equal to 3
times the width of the section, i.e., 3b, where the width of the section is the width of the
flange or diameter of the element for driven sections and the diameter of the concrete-filled
hole for sections encased in structural concrete. The maximum width of 3b can be used
when material in which the vertical element is embedded does not contain discontinuities that
would affect the failure geometry. The width should be reduced if planes or zones of
weakness would prevent mobilization of resistance through this entire width, or if the passive
resistance zones of adjacent elements overlap. For drilled-in soldier piles, the effective width
b is assumed to be equal to the diameter of the drill hole provided the embedded portion of
the drill hole is backfilled with structural concrete (as indicated previously) or lean mix
concrete with a compressive strength of no less than 50 psi.
Above the bottom of the excavation, discrete elements (e.g., soldier piles) are designed for
the total lateral force in the span distance between adjacent soldier piles. Below the bottom
of the excavation, the active earth pressure is assumed to act only on the soldier pile width, b.
For this step, the active earth pressures, surcharge pressures, seismic pressures, and other
sources of lateral pressure are factored using the load factors in AASHTO Tables 3.4.1-1 and
3.4.1-2 for all appropriate limit states. Passive resistance developed on the excavation side of
the wall uses a resistance factor of 0.75 (see Table 11.5.6-1 in AASHTO 2007).
8.9.2.4 Step 6: Evaluate Embedment Depth and Factored (Maximum) Bending Moment
With the factored total lateral pressure diagrams evaluated for the active and passive side of
the wall, the following calculation steps are used to design the vertical wall element:
These steps are most easily performed using a spreadsheet where forces and moments are
calculated at various assumed embedment depths. Figure 8-57 shows an example for a
nongravity cantilevered sheet-pile wall with a height of 10 ft retaining a granular soil. The
retained ground has a friction angle of 30 degrees which results in a calculated KA = 0.33 and
Kp =3.0 (assuming δ = 0 and using equations 3-2 and 3- 3 for KA and KP, respectively). In
this example, it is assumed that only lateral pressures resulting from active earth pressures
are present with a load factor of 1.5 (i.e., maximum load factor for horizontal earth pressure
for active conditions based on AASHTO Table 3.4.1-2). In current practice, sheet pile wall
designs may be performed using computer programs such as CWALSHT (1995), Pile Buck
(2004), and others. These programs, however, are not LRFD based.
AASHTO Figures 3.11.5.6-3, -6, and -7 address nongravity cantilevered wall designs for
continuous elements, such as sheet piles. These figures indicate that the length required to
obtain moment equilibrium (based on the procedure described above using factored loads and
resistances) should be increased by 20 percent. The design method used to evaluate the
required wall embedment is based on achieving moment equilibrium. This method is
referred to as the simplified method in that it is assumed that the difference between the
passive resistance at the back of the wall and the active pressure at the front acts as a
concentrated force, F, at the toe. By taking moments about the toe (thereby eliminating the
force F from the equation), the depth of embedment, do, is found. Because of this
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-90 June 2008
Cantilevered Sheet Pile Wall design LRFD
γ= 125 pcf
10' Ka = 0.33
γp = 1.5
Pa
L
Pp La
Lp
A
Kp = 3
ϕp = 0.75
2
Factored Pa = γp * 0.5 * (L+10) * Ka * γ
2
Factored Pp = ϕp * 0.5 * L * Kp * γ
Based on above:
Required embedment for (factored resist = factored load) = 15.2 ft
Maximum factored moment occurs at 8.9 ft
Maximum factored moment = -36577.0 ft-lbs
Figure 8-57. Design Analysis for Nongravity Cantilevered (Sheet Pile) Wall.
It is noted that for ASD design, the minimum factor of safety with respect to passive
resistance is 1.5. For LRFD (for both continuous and discrete element walls), the resistance
factor used for passive resistance evaluations is 0.75 (see Table 11.5.6-1 in AASHTO 2007).
Assuming only active earth pressure loadings on the active side with a load factor of 1.5
indicates that the LRFD approach provides a corresponding ASD FS value of approximately
2.0 (i.e., 1.5/0.75). It is noted that in the 2002 Interim AASHTO LRFD Specifications, the
resistance factor for passive resistance was 1.00. This information indicates that embedment
depths computed for nongravity cantilevered walls with continuous wall elements using
LRFD will likely be greater than those computed using ASD.
In general, nongravity cantilevered walls comprise either a structural steel section (e.g., steel
sheet-pile or soldier beam) or reinforced concrete (e.g., secant shafts). To date, LRFD-based
formulations for the structural evaluation of vinyl or composite sheet piles are not available.
For structural steel vertical wall elements (which are not subject to significant axial loads),
the design check for flexural resistance can be written as:
M max ≤ ϕM n = ϕ Fy Z (8-1)
where:
8.9.2.6 Step 8: Select Temporary Lagging for Soldier Pile and Lagging Wall
Timber lagging is only used for support of temporary loads applied during excavation,
however, pressure-treated timber lagging has been used to support permanent loads. The
contribution of the temporary lagging is not included in the structural designs of the final
wall face. Temporary lagging is not designed by traditional methods, rather lagging is sized
from charts developed based on previous project experience which accounts for soil arching
between adjacent soldier piles.
Table 8-11 presents experience-based recommendations for selecting wood lagging thickness
based on soil type, depth of excavation and clear distance between soldier piles.
As noted, permanent timber lagging has been used in lieu of a concrete face to carry
permanent wall loads. For permanent applications, the timber grade and dimensions should
be designed according to structural guidelines. Several problems may exist for permanent
timber lagging including: (1) need to provide fire protection for the lagging; and (2) limited
service life for timber.
For nongravity anchored walls with discrete vertical elements, permanent facing should be
designed to resist factored earth pressures, surcharges, water pressures, and seismic pressures
for all appropriate limit states. The factored bending moments in the permanent facings can
be estimated using Table 8-12.
Permanent facings that are cast-in-place (CIP) are typically 8 to 12 in. thick. This thickness
will typically ensure that the wall is structurally sound and allow for some deviations in
soldier beam placement. Significant deviations, however, in soldier beam alignment may
require that additional concrete in excess of that required for the nominal thickness of the
wall be used so that the finished wall face is properly aligned. Precast concrete facing may
be cost-effective if there is a local fabricator and if there is adequate on-site storage. Precast
panels are designed as simple spans between the soldier beams.
Factored moment
Support and soil condition on a unit width or
height of facing
Simple span
No soil arching (e.g., soft cohesive soils; rigid concrete facing p l 2/8
placed tightly against soil)
Simple span
Soil arching (e.g., granular soil or stiff cohesive soil with flexible
p l 2/12
facing; rigid facing where space is available to allow in place soil
to arch)
Continuous facing
No soil arching (e.g., soft cohesive soils; rigid concrete facing p l 2/10
placed tightly against soil)
Continuous facing
Soil arching (e.g., granular soil or stiff cohesive soil with flexible
p l 2/12
facing; rigid facing where space is available to allow in place soil
to arch)
Note: p = average factored lateral pressure, including earth, surcharge, and water pressure acting
on the section of facing being considered
l = spacing between vertical elements or other facing supports
8.9.2.8 Step 10: Evaluate Lateral Displacements at the Service Limit State
Lateral wall displacements may be estimated using beam on elastic foundation analyses or
other soil-structure interaction analysis methods. These methods are discussed in Section
8.12.2.
8.10.1 Overview
Anchored bulk head walls derive their lateral support from the passive resistance on the front
of the embedded portion of the wall, and the pull out resistance of the anchors. Anchors
reduce the required depth of penetration of vertical wall face elements and also permit wall
heights to be increased to about 35 ft, depending on the soil and groundwater conditions. For
higher walls the use of high- strength steel sheet piling, reinforced sheet piling, or additional
levels of anchors may be necessary. The methods of analysis presented in this section apply
for walls with a single anchorage level and which are built from the bottom-up. LRFD for
these wall types is not addressed in AASHTO (2007) and therefore all design information
presented herein is in ASD.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-95 June 2008
Anchored walls with one level of anchors and which are built from the bottom-up are
designed using triangular active and passive earth pressure distributions. Using these applied
pressures, there are two basic methods for design of anchored sheet pile walls: (1) the free
earth support method; and (2) the fixed earth support method. In selecting the method to be
used for wall design, consideration must be given to a number of factors such as the relative
stiffness of the sheet piles, the depth of pile penetration, the relative compressibility and
strength of the soil, and the amount of anchor deflection. Figure 8-58 shows the assumed
deflection and bending moment distribution for the two methods.
Figure 8-58. Variation of Deflection and Bending Moment (a) Free Earth Support Method
and (b) Fixed Earth Support Method (after Das, 1990).
Typically, the fixed earth method is used in granular soils and stiff cohesive soils and for
walls of relatively low stiffness; the free earth method is used in cohesive soils, for walls of
relatively higher stiffness, and for situations where the sheeting penetration may be restricted
because of obstructions, a shallow rock surface, or sheeting length limitations. The fixed
earth method generally results in smaller anchor loads and smaller wall bending moments,
but a greater depth of sheeting penetration than the free earth method.
This method is based on the following assumptions (Teng, 1962 and USS, 1975):
• The soil into which the piling is placed is incapable of producing effective restraint
from passive resistance to the extent necessary to induce negative bending moments;
With the above assumptions, the problem can be solved by considering static equilibrium.
Figure 8-59 presents analysis by the free earth support method for two common cases.
The free earth support method assumes a rigid piling and therefore uses lateral pressures
according to a triangular distribution as shown in Figure 8-59. However, in reality, many
wall systems are quite flexible causing the earth pressure to redistribute and thus differ from
this triangular distribution. With increasing flexibility, the embedded portion of the wall
rotates about its lower edge (Figure 8-58a) causing the center of the passive resistance to
move closer to the exposed wall base. This in turn decreases the maximum bending moment.
Figure 8-60 presents a procedure to reduce the maximum design moment obtained from the
free earth support method based on wall flexibility considerations. The moment reduction
factor (Md/Mmax in Figure 8-60a) should be used where a factor of safety was applied to the
soil shear strength parameters prior to computing the penetration depth. Also, it is
recommended that the reduced bending moment not be lower than that computed by the fixed
earth method.
The fixed earth support method is based on the assumption that the toe of the wall is
restrained from rotation as shown in Figure 8-58b. The deflected shape reverses its curvature
at the point of contraflexure, c, and becomes vertical at the toe of the wall.
The problem may be solved using the theory of beam on elastic foundation, but this
procedure is laborious. In practice, a procedure known as the equivalent beam method
proposed by Blum is used. This method utilizes a theoretical relationship between the angle
of internal friction (φ) and the distance below the exposed wall base to the point of
contraflexure.
Figure 8-61 presents the equivalent beam method, which is limited in its use to granular soils.
For cohesive soils, use methods based on the theory of beam on elastic foundation (Hetenyi,
1946).
Figure 8-59a. Analysis by Free Earth Support Method for Sheet Piling in Granular Soils
(after Teng, 1962 and USS, 1975).
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-98 June 2008
7. Select pile section for the maximum moment or use the moment reduction theory, Figure 8-
60b.
8. Add 20 to 40 percent D’ to provide for safety margin, or divide qu by a factor of safety of 1.5
to 2.0 in above equations.
Figure 8-59b. Analysis by Free Earth Support Method for Sheet Piling Backfilled in
Granular Soil and Embedded in Cohesive Soil (after Teng, 1962 and USS, 1975).
Figure 8-60a. Moment Reduction for Anchored Sheet Pile Wall Analyzed by Free Earth
Support Method for Granular Soils (after Rowe, 1952; Rowe, 1957; and Das, 1990).
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-100 June 2008
1. Compute the Stability Number, Sn, as follows:
Sn = (cu / γeqH)(1 + ca / cu)0.5
where cu(=qu/2) is the undrained cohesion below the exposed base of wall, ca is wall
adhesion (See Table 3-2).
For design purposes, Sn =1.25(cu/γeqH) may be used. Any sheet piling driven into
cohesive soils should have a minimum Sn of 0.3 times a desired factor of safety;
otherwise failure may occur.
2. Compute α as α = H / (H+D).
3. Compute ρ as per Equation in Figure 8-60a.
4. For the magnitudes of α and Sn obtained above, determine the Md/Mmax for various
values of log ρ from above charts and plot Md/Mmax against log ρ. Values of log ρ can
be interpolated.
5. Verify that the plotted point for the selected pile section falls above the curve
developed in Step 4. If the plotted point falls below the curve, select another sheet
pile section and repeat the above Steps 1 through 5.
Figure 8-60b. Moment Reduction for Anchored Sheet Pile Wall Analyzed by Free Earth
Support Method for Cohesive Soils (after Rowe, 1952; Rowe, 1957; and Das, 1990).
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-101 June 2008
8.10.4 Deadman Anchor Systems
A deadman anchor, if used, must be located outside the potential active failure zone
developed behind the sheet pile wall as shown in Figure 8-62. If the anchor system must be
located closer to the wall (see Figure 8-62), anchorage resistance is decreased and an
additional passive reaction is required for stability of the wall base.
The tie rods must be protected to resist corrosion. Typically, the rods are coated with one or
more applications of tar-base paint, and spirally wrapped with a durable fabric or fiber glass
tape after the application of the first coat of paint. The tie rods may also be enclosed in a
rigid casing or supported vertically to eliminate sag for cases where the backfill will settle
significantly or unevenly.
The design methods presented in this section have been developed primarily for anchored
walls with relatively flexible vertical wall elements (i.e., sheet piles or soldier piles) (see
GEC No. 4 (Sabatini et al., 1999). A key distinction in the design of flexible anchored walls
(as compared to stiff anchored walls) is in the earth pressure diagram used for design
analyses. For flexible wall systems constructed from the “top-down”, the wall deformation
pattern is complex and may not be consistent with the development of a theoretical Rankine
or Coulomb (i.e., triangular) earth pressure distribution over the full height of the wall. For
example, higher than active earth pressures develop at the upper anchor location since the
upper anchor restrains the wall from moving outward sufficiently to locally cause a reduction
of earth pressures to the active state. For this reason, flexible anchored walls may be
designed using apparent earth pressure diagrams.
The use of apparent earth pressure diagrams has resulted in reasonable estimates of ground
anchor loads and conservative estimates of wall bending moments between anchors for
flexible walls constructed in competent soils. In general, this approach may also be used for
the design of stiff anchored wall systems; however, experience indicates that stiff anchored
wall designs based on apparent earth pressure diagrams may be overly conservative and
therefore expensive.
The major design steps for an anchored wall are outlined in Table 8-13 and Steps 1 and 2
have been discussed in previous portions of this manual. Soil and rock parameter evaluation
(as part of Step 3) has been discussed in Section 2.
8.11.1.1Overview
The corrosion protection system for a ground anchor consists of components that combine to
provide an unbroken barrier for each part of the tendon and the transitions between them.
Steel components of the anchor include the anchor head, bearing plate, trumpet, prestressing
steel, and couplers (where used). Components of the corrosion protection system include: (1)
for the anchorage, a cover or concrete embedment, a trumpet, and corrosion inhibiting
compounds or grout; (2) for the unbonded (free) length, grout and a sheath filled with a
corrosion inhibiting compound or grout; and (3) for the bond length, grout and
encapsulations with centralizers and/or epoxy coatings.
The components of corrosion protection for bar and strand tendons are shown in Figures 8-63
to 8-65 and brief descriptions are provided below.
The free length is protected by a variety of means since the protection is not required to
transfer stresses from the tendon to the ground. The protection must:
The most common methods of protecting the free length include the use of smooth sheaths
filled with anti-corrosion grease, heat shrink sleeves, and secondary grouting after stressing.
Except for secondary grouting, the protection is usually in place prior to inserting the tendons
in the hole.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-106 June 2008
Figure 8-63. Examples of Corrosion Protection for Anchorages (a) Strand Tendon and (b)
Bar Tendon (after Sabatini et al., 1999).
(b)
Figure 8-64. Simple Corrosion Protected Tendons (a) Strand Tendon and (b) Bar Tendon
(after Sabatini et al., 1999).
(b)
Figure 8-65. Encapsulated Double Corrosion Protection (a) Strand Tendon and (b) Bar
Tendon (after Sabatini et al., 1999).
Secondary grout, in the true sense, refers to grout placement after stressing and lock-off
around a tendon with an unsheathed free length. After set, the grout surrounding the free
length becomes bonded to the ground and the tendon over the initially unbonded length.
These anchors are only normally recommended for semi-permanent, low-risk applications.
When secondary grouting is used for these anchors, extreme care must be taken to insure no
voids in the grout exist beneath the anchor bearing plate. Such areas are the most critical to
protect against corrosion and require complete encapsulation by cement grout.
Figures 8-64 and 8-65 show several schemes for corrosion protection of the tendon bond
length. These are briefly described below:
a. Simple (Class II) Protection: The use of simple protection (Figure 8-64) relies on
Portland cement grout to protect the tendon, bar or strand, in the bond length.
Steel will not corrode in a high pH environment such as Portland cement which
possesses pH values up to 12.6. When Portland cement is used for protection, it is
assumed that the pH will not be lowered with time. When the simple protection is
used, care should be taken that the tendon has at least 0.5 in. of grout cover and
the anchor grout extends 24 in. over the bottom of the free length sheathing.
b. Double (Class I) Protection: Complete encapsulation of the anchor tendon is
accomplished by uniformly corrugated plastic or steel tube (Figure 8-65). The
tube must be capable of withstanding the deformations associated with
transportation, installation, stressing and testing of the anchor, and transferring the
load applied to the tendon. Regardless of the encapsulating medium, the annular
space between the corrugated tube and tendon is usually filled with neat cement
grout containing admixtures to control bleed of water from the grout. Shorter
tendons are grouted before insertion in the hole.
The most critical area to protect from corrosion is in the vicinity of the anchor head
connection. Below the bearing plate, the corrosion protection over the free length is usually
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-110 June 2008
terminated to expose the bare tendon. Above the bearing plate, the bare tendon is gripped by
either wedges, nuts, or deformed in the case of wires. Regardless of the type of tendon, the
gripping mechanism creates stress concentrations at the connection. In addition, a very
aggressive corrosion environment may exist at the anchor head since oxygen is readily
available. The vulnerability of this area is demonstrated by the fact that most anchor failures
that have been reported have occurred within a short distance of the anchor head.
Typical protection system for the anchor head includes the use of a trumpet. A “trumpet,”
usually of steel or strong durable plastic, is used to overlap with the free length corrosion
protection, and to protect the short exposed length of the tendon below the anchor plate. One
end of the trumpet is fastened and sealed to the bearing plate and the other fitted with a
deformable seal that fits tightly around the protective tube, but allows free tendon movement
within the trumpet. The annular space between the trumpet and the tendon is filled with anti-
corrosion grease.
The anchor head, including exposed tendon and friction grips or locking nuts above the
bearing plate, is protected by a cap filled with anti-corrosive grease or by embedment in
concrete. Covering the anchor head with a grease-filled cap allows future lift-off tests and/or
load adjustment. When filling the trumpet and cap, care is required to ensure that the grease
or grout fills the entire space.
SMALL SIGNIFICANT
CLASS I CLASS II
PRO- PRO-
TECTION TECTION
Note: Class I also means Simple Protection and Class II also means Double Protection.
Figure 8-66. Decision Tree for Selection of Corrosion Protection Level (after PTI, 1996).
Service life is used to distinguish between a temporary support system and a permanent
anchor application. The service life of a temporary earth support system is based on the time
required to support the ground until a permanent earth retention system is constructed. The
time period for temporary systems is commonly stated be 18 to 36 months but may be shorter
or longer based on actual project conditions. If the service life of a temporary support of
excavation anchor is likely to be extended, an evaluation should be made to determine
whether or not to provide additional corrosion protection for the tendon, particularly in
aggressive ground conditions.
In general, ground environments may be classified as aggressive if any one of the following
conditions are present in the ground or may be present during the service life of the ground
anchor (PTI, 1996) (Table 8-14).
Table 8-14. Criteria for Electrochemical Properties of Soils for Ground Anchor Applications
(after Cheney, 1988).
The presence of stray currents or buried concrete structures adjacent to the anchored system
project which have experienced corrosion or direct chemical (acid) attack are also indication
of aggressiveness of the ground. Tests from a nearby site can be used to evaluate the
aggressivity of the site if the designer can establish that the ground conditions are similar.
Otherwise, if site-specific aggressivity tests are not performed, then the ground should be
assumed to be aggressive.
The following ground environments are always considered aggressive: (1) soil or
groundwater with a low pH; (2) salt water or tidal marshes; (3) cinder, ash, or slag fills; (4)
organic fills containing humic acid; (5) peat bogs; and (6) acid mine drainage or industrial
waste. Classification of ground aggressivity should consider the possibility of changes
during the service life of the ground anchor, which may cause the ground to become
aggressive, such as, might occur near mining operations, chemical plants, or chemical storage
areas.
For permanent anchors, if failure of the anchored system could result in serious consequences
such as loss of life or significant financial loss, a minimum of Class I protection is required.
The consequences of failure are considered serious for: (1) anchored systems used in urban
areas where there are nearby structures behind the wall; (2) anchored systems used for a
The final criterion for selecting the class of corrosion protection is the increased cost for
changing from Class II (Simple) protection to Class I (Double) protection. For the same
tendon, Class I protected anchors require a larger drill hole as compared to a Class II
protected anchor. Encapsulating an anchor tendon increases the required drill hole size
which may result in increased installation costs. In an uncased drill hole, the additional
drilling costs can be small, and the owner may elect to use Class I protection. In a cased hole
or in rock, the additional drilling costs can be higher, and the owner will decide if the benefit
of providing a higher level of corrosion protection is worth the additional cost. The increase
in drill hole diameter may result in a need to increase bearing plate dimensions, trumpet
diameter, and the opening in the soldier beam to insert the tendon.
Recommended apparent earth pressure diagrams for sands, stiff to hard clays, and soft to
medium clays are provided in Figures 8-67, 8-68, and 8-69. These “apparent earth pressure
diagrams” are semi-empirical diagrams developed to account for the relatively complex
deformation pattern associated with relatively flexible anchored systems.
For the apparent earth pressure diagram for sands (i.e., Figure 8-67), the value for KA is the
Rankine earth pressure coefficient, γ is the effective unit weight of the retained ground and H
is the total height of the excavation.
The apparent earth pressure diagram for stiff to hard clays (Figure 8-68) should only be used
where for support of a temporary excavation (i.e., relatively short duration) and there is no
available free water. If these either of these conditions are not met, an apparent earth
pressure diagram for long-term (permanent) conditions using drained strength parameters
should be evaluated. In this case, use the drained friction angle of the stiff to hard clay with
the sand apparent earth pressure diagram shown in Figure 8-67 and explicitly add in water
pressures.
When using Figure 8-68, a maximum pressure ordinate of 0.3γH should be used where
anchors are locked off at 75 percent or less of the anchor load (based on unfactored loads)
and 0.4γH where anchors are locked off at 100 percent of the anchor load (based on
unfactored loads).
The apparent earth pressure diagram for use in designing temporary wall systems in soft to
medium clays (Figure 8-69) is based on the well-known Terzaghi and Peck (1967) diagram
for soft to medium clays. The Terzaghi and Peck (1967) diagrams, however, did not account
for the development of soil failure below the bottom of the excavation. Observations and
finite element studies have demonstrated that soil failure below the bottom of the excavation
can lead to very large movements for temporary retaining walls in soft clays. Henkel (1971)
Figure 8-69. Apparent Earth Pressure Diagram for Soft to Medium Clay.
For the apparent earth pressure diagram for soft to medium clays (Figure 8-69), the
maximum pressure ordinate, pa, is given as pa=KAγH where KA is given as:
4S u d 1 − 5.14S ub
K A =1 − + 2 2 ≥ 0.22 (8-2)
γH H γH
where:
This method is based on a total stress approach, i.e., water pressures are already accounted
for in the diagram. Additional information on the use of this diagram is provided in Sabatini
et al. (1999). A computation showing the evaluation of KA according to Eq. 8-2 is provided
in Student Exercise 7.
For this step, the apparent earth pressures, water, surcharge loads, seismic pressures, and
other sources of lateral load on the wall are factored using the load factors in AASHTO
(2007) Tables 3.4.1-1 and 3.4.1-2 for all appropriate limit states.
Apparent earth pressure diagrams were originally developed based on wall pressures back-
calculated from field measurements of strut loads in internally braced excavations.
Application of these diagrams will generally produce conservative design loads, implying
that if a strut (or ground anchor) load would be equivalent to the calculated load from the
apparent pressure diagram at that location, the other strut (or ground anchor) loads would
necessarily be less than that calculated from the apparent pressure diagrams. That is, the
apparent pressure diagrams represent an envelope of maximum pressures.
Using AASHTO (2007), a load factor of 1.35 is applied to these apparent earth pressure
distributions. The use of this load factor represents a relatively large amount of conservatism
for LRFD-based designs because of the manner in which the unfactored apparent earth
pressure diagrams were developed.
8.11.4 Step 6: Evaluate Individual Anchor Loads and Subgrade Reaction Force
Factored horizontal anchor loads may be calculated using either the tributary area method or
the hinge method as shown in Figure 8-70 for a one-level wall and in Figure 8-71 for a multi-
(i.e., 2 or more) level anchored wall. Both methods, as shown, assume that a hinge (i.e., zero
bending moment) develops at the elevation of the base of the excavation and that the
excavation subgrade acts as a support. This latter assumption is reasonable for walls that
penetrate into competent materials.
Lateral load resistance is limited below the base of the excavation in very loose to loose
granular soils or soft to medium clays. In both of these material types, the wall elements
must experience relatively large movements to fully develop passive resistance. The vertical
wall element may become overstressed (in bending) prior to achieving these movements.
When undrained conditions exist in soft to medium clays (e.g., as for a temporary excavation
in these soils) with stability number NS (=γH/Su)> 4, it can be shown that the passive
resistance that develops on the excavation side of the wall will not be greater than the active
pressure applied on the retained side of the excavation and therefore, regardless of how long
the wall element is, lateral load resistance is not developed. Referring to Figure 8-72 for a
temporary wall with continuous elements in soft to medium clay, it can be noted that, to
provide lateral resistance, the quantity 2Su (on passive side) must be greater than γH – 2Su
(on the active side):
γH
4 S u ≥ γH ⇒ if N s = then
Su
N s ≤ 4 for passive resis tan ce to develop
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-119 June 2008
Figure 8-71. Calculation of Nominal Anchor Loads For Multi-Level Wall.
Referring to the above discussion, if Ns > 4, from a loading point of view, increasing
embedment has the effect of essentially creating a longer cantilever beam fixed at the
location of the lowest anchor support. Therefore, as this length increases, the bending
moments in the wall would increase considerably. If this condition exists, it is generally
recommended to limit the depth of wall penetration to 20 percent of the excavation depth
unless deeper embedment is necessary to develop sufficient capacity to resist vertical loads,
provide basal stability or groundwater cutoff, or limit ground movements. This condition
would also result in greater amounts of load being shifted to the lowermost anchor supports
since the soils below the elevation of the excavation base cannot provide adequate support.
For the case described above, the embedded portion of the wall should be designed as a
cantilever beam fixed at the lowest anchor. For design, the wall section should be selected
based on the maximum bending moment evaluated, i.e., either the maximum bending
moment in the exposed portion of the wall above the lowermost anchor or the calculated
cantilever bending moment about the lowermost anchor. Alternatively, soil-structure
interaction analyses (e.g., beam on elastic foundation) (see Section 8.12.2) may be used to
design continuous beams with small toe reactions as it may be overly conservative to assume
that all load is carried by the lowest anchor.
The values calculated using Figures 8-70 and 8-71 (for the unfactored (nominal) anchor
loads) are the horizontal component of the unfactored (nominal) anchor load per unit width
of wall, Thi. The total horizontal anchor load, Th, is calculated as:
Th = Thi s (8-3)
where s is the horizontal spacing between adjacent anchors. The anchor load, T, to be used
in designing the anchor bond zone is calculated as:
Th
T= (8-4)
cos θ
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-121 June 2008
where θ is the angle of inclination of the anchor below the horizontal. The vertical
component of the total anchor load, Tv, is calculated as:
Tv = T sin θ (8-5)
These unfactored (nominal) anchor loads should be multiplied by the appropriate load factor
for apparent earth pressures, i.e., γAEP = 1.35. Other sources of loading (e.g., surcharges)
should be distributed to individual anchors using the same procedures noted above (e.g.,
tributary area) and multiplied by the appropriate load factor for the limit state being
considered.
Ground anchors are commonly installed at angles of 15 to 30 degrees below the horizontal
although angles of 10 to 45 degrees are within the capabilities of most contractors.
Regardless of the anchor inclination, the anchor bond zone must be developed behind
potential slip surfaces and in soil or rock layers that can develop the necessary design load.
Steep inclinations may be necessary to avoid underground utilities, adjacent foundations,
right-of-way constraints, or weak soil or rock layers. Anchors should be installed as close to
horizontal as possible to minimize vertical loads resulting from anchor lock-off loads,
however grouting of anchors installed at angles less than 10 degrees is not common unless
special grouting techniques are used.
The prestressing steel element of the tendon (i.e., strand or bar) must be capable of safely
transmitting load in the anchor bond zone to the structure without tendon breakage. The
maximum proof test load shall not exceed 80 percent of guaranteed ultimate tensile strength
(GUTS) of the prestressing steel bar or strand (i.e., tendon breakage resistance factor = 0.8)
(see Table 11.5.6-1 in AASHTO 2007 for resistance factors for other than high strength
steel). Also, the unfactored (nominal) load in each anchor shall not exceed 60 percent of
GUTS of the prestressing steel bar or strand.
Dimensions and strengths of bars and strands commonly used in the U.S. for highway
applications are provided in Tables 8-15 and 8-16, respectively. Larger size strand tendons
(i.e., strand tendons with more strands than those shown in Table 8-16) are available for
applications requiring greater ground anchor design loads. Specific property values should
be obtained by strand and bar providers.
Nominal Ultimate
Ultimate
Steel Nominal cross section strength
stress Prestressing force
grade diameter area fpu Aps
fpu
Aps (= GUTS)
0.8 fpu Aps 0.7 fpu Aps 0.6 fpu Aps
(ksi) (in.) (ksi) (in.2) (kips) (kips) (kips) (kips)
1 150 0.85 127.5 102.0 89.3 76.5
1-1/4 150 1.25 187.5 150.0 131.3 112.5
150 1-3/8 150 1.58 237.0 189.6 165.9 142.2
1-3/4 150 2.62 400.0 320.0 280.0 240.0
2-1/2 150 5.19 780.0 624.0 546.0 468.0
1 160 0.85 136.0 108.8 95.2 81.6
160 1-1/4 160 1.25 200.0 160.0 140.0 120.0
1-3/8 160 1.58 252.8 202.3 177.0 151.7
Table 8-16. Properties of 0.6 in. Diameter Prestressing Steel Strands (ASTM A416, Grade
270).
Ultimate
Number of 0.6 in.
Cross section area strength Prestressing force
diameter strands
(=GUTS)
0.8 fpuAps 0.7 fpuAps 0.6 fpuAps
The type and size of the anchors should be evaluated prior to design of the anchor bond zone
because the required hole diameter varies as a function of the tendon size. Table 8-17 can be
used to estimate the minimum trumpet opening for strand or bar tendons.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-123 June 2008
Table 8-17. Guidance Relationship between Tendon Size and Trumpet Opening Size (after
Sabatini et. al., 1999).
Estimates of load transfer in the anchor bond length are typically based on previous field
experience. When estimating resistance using previous field results, potential variations in
resistance due to differing installation and grouting methods must be considered. In a given
soil deposit, the actual resistance achieved in the field will depend on the method of drilling
including quality of drill hole cleaning and period of time that the drill hole is left open, the
diameter of the drill hole, the method and pressure used in grouting, and the length of the
anchor bond zone. Except for certain minimum values, the selection of these items should be
left to the discretion of the specialty anchor contractor. The main responsibility for the
designer is to define a minimum anchor resistance that can be achieved in a given ground
type.
Because of the similarity of many projects, some fairly typical anchor characteristics can be
summarized. These are intended to provide a range of typical design values to engineers who
are unfamiliar with anchor design.
• Maximum Test Load Between 60 and 240 kips: Anchor tendons to resist these loads
can be handled without the need for unusually heavy or specialized equipment. In
addition, stressing equipment can be handled by one or two workers without the aid
of mechanical lifting equipment. The drill hole diameter is generally 6 in. or less,
Soil Anchors
Anchor bond lengths for gravity-grouted, pressure-grouted, and post-grouted soil anchors are
typically 15 to 40 ft since significant increases in load carrying capacity for bond lengths
greater than approximately 40 ft cannot be achieved unless specialized methods are used to
transfer load from the top of the anchor bond zone towards the end of the anchor.
Pressure grouting in cohesionless soils significantly increases the normal stresses acting on
the grout body (i.e., increases confinement). Small increases may also be observed in the
effective diameter of the anchor bond zone, but capacity estimates should be based on the as-
drilled hole diameter.
Rock Anchors
Typical anchor bond lengths in rock range from 10 to 30 ft. The load transferred from the
bond length to competent sound rock may be estimated from the rock type, however, reduced
values may be recommended after input from a geologist especially if the rock mass strength
is controlled by discontinuities. In an allowable stress design (ASD) approach, the maximum
allowable anchor design load in competent rock was determined by multiplying the bond
length by the ultimate transfer load and dividing by a factor of safety of 3.0. This relatively
high value of the factor of safety (compared to a factor of safety of 2.0 for anchors in soil)
was used to account for uncertainties associated with potential discontinuities in the rock
mass such as joints, fractures, and clay-filled fissures. In LRFD, this difference between soil
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-125 June 2008
and rock is evidenced by the resistance factors for rock being less than that for soil (as shown
subsequently).
In weak rocks such as clay shales, bond stress transfer is relatively uniform as compared to
bond stress transfer in more competent rock. These weak rocks may be termed “intermediate
geomaterials” and have unconfined compressive strengths defined as varying from 69 to 694
psi.
For each anchor, the minimum estimated anchor bond length, Lb, is computed as:
Lb ≥ T ϕ Qn (8-6)
where:
T = Factored anchor load (calculated according to Eq. 8-4 with all appropriate load
factors for applicable limit states)
Q n = π × d × τa × L b (8-7)
where:
As described in Sabatini et al. (1999), various procedures are available for estimating the
nominal resistance of ground anchors in soil and rock using semi-empirical correlations or
in-situ testing. As a guide, Table 8-18 may be used to estimate the nominal bond for anchors
installed in cohesive soils, cohesionless soil, and rock, respectively. The values in this table
8.11.8 Step 10: Evaluate Factored Bending Moment and Flexural Resistance of Wall
Discrete vertical wall elements must be designed to resist all applicable earth and water
pressure, surcharge, anchor and seismic loadings, and the vertical component of the anchor
loads and other vertical loads within the tributary area between adjacent vertical wall
elements. For continuous wall elements, the analysis is performed on a per unit length basis.
In designing these elements, fixed horizontal support can be assumed at each anchor level
and at the bottom of the wall if the elements are sufficiently embedded below the base of the
wall. If the soils at the base of the wall are relatively weak, the bending moment about the
lowest anchor may be calculated assuming that the wall element acts as a cantilever below
the elevation of the lowest anchor.
Unless beam on elastic foundation, finite element, or other soil-structure interaction analysis
methods are used, the factored bending moment in the vertical wall element for an anchored
wall may be calculated using the hinge method or tributary area method (see Figures 8-73
and 8-74).
For walls constructed in competent soils such as most sands and stiff clays, the maximum
(factored) bending moment, Mmax, occurs in the exposed portion of the wall. For walls that
penetrate deep deposits of weak material, the maximum bending moment may occur in the
embedded portion of the wall. The embedded portion of the wall refers to the length of wall
that is below the base of the excavation.
The negative bending moment calculated at the location of the uppermost anchor is evaluated
by summing moments about the first anchor location. The vertical wall elements are
commonly assumed to be continuous between each vertical support location. The maximum
positive bending moment between each ground anchor is, for the tributary area method,
assumed equal to 1/10 p l 2 where p is the maximum ordinate of the apparent pressure
envelope (with a load factor of 1.35) and l is the vertical spacing between adjacent anchors.
For the hinge method, the maximum positive bending moment between each ground anchor
corresponds to the point of zero shear. Additional information on computations using the
hinge method is provided in Weatherby (1998). These methods provide conservative
estimates of the calculated bending moments, but may not accurately predict the specific
location. For continuous sheet-pile walls, the bending moment per unit of wall is used to
select an appropriate sheet-pile section. To evaluate the maximum bending moment for
design of a soldier pile, the maximum bending moment per unit of wall calculated from
Figure 8-73 and 8-74 is multiplied by the tributary length of each wall element. Soil-structure
interaction analyses may also be used for evaluating wall bending moments (see Section
8.12.2.2).
In most cases, wall elements of an anchored wall will be subject to both flexure and axial
compression loads. Axial compression loads result from vertical anchor forces, weight of
lagging, and, weight of permanent facing rigidly attached to the vertical wall elements.
AASHTO (2007) Section 6.9.2.2 should be used to evaluate combined axial load and flexure
The bearing resistance for continuous wall elements (e.g., slurry walls) is discussed in
Section 8.12.3.3, 8.12.4.3, and 8.12.6.3.
Since soldier piles will not typically be subject to dynamic analysis and/or load testing (as for
deep foundation elements), the AASHTO LRFD code therefore requires resistance factors for
bearing resistance of soldier piles to range from approximately 0.25 to 0.55 (i.e., the range
reported above for static analysis of single piles and shafts). The corresponding ASD factor
of safety value (for the lower end resistance factor) would be on the order of FS = 4.0.
Because potential settlements of an anchored wall will not typically result in undesirable or
unsafe performance (as compared to a similar amount of settlement for a deep foundation
element used), GEC No. 4 (Sabatini et al., 1999) recommends factor of safety values of 2 to
2.5 for calculating bearing resistance of driven and drilled-in soldier piles for permanent
walls.
• the vertical component of the anchor lock-off loads typically provides the largest source
of vertical loading on the wall;
• a lock-off load is a prescribed load that is imparted to the wall and therefore, in the
context of LRFD, the uncertainty in this load is relatively small resulting in an
appropriate load factor of 1.0; and
• other sources of vertical load (which would use a maximum load factor of 1.25) include
dead weight of the wall element itself, weight of lagging, and weight of facing.
Using the equations shown in Figure 4-1 and assuming that only lock-off loads impose
vertical loads on the wall, results in a resistance factor of 0.5 (for a corresponding ASD
FS=2.0). If other vertical loads are acting on the wall (such as weight of lagging, etc),
because those loads would be factored using a load factor greater than 1.0, the
corresponding resistance factor (for FS = 2.0) would be slightly greater than 0.50, therefore
implying that the use of a resistance factor of 0.5 is reasonable.
8.11.10 Step 12: Evaluate Overall Stability of Anchored Wall at Service Limit
State
Significant basal heave and substantial increases in lateral earth pressures result when the
weight of the retained soil (and surcharge loads) exceeds or approaches the soil bearing
resistance at the base of the excavation. Basal heave should be checked for excavations
made through soft to medium clay.
Traditional methods for assessing the potential for bottom heave are based on the
performance of braced excavations in soft to medium clays. Figure 8-75 shows a cut in soft
clay H deep and B wide. The block of retained soil exerts a vertical pressure qapplied on strip
cd equal to its weight minus the shear resistance of the soil along plane bd. The bearing
capacity of a cohesive soil is equal to NcSu where Nc is the bearing capacity factor.
It is noted that this analysis does need to be written in an LRFD format. The calculated FS
value provides an index of likely lateral deformations (and hence potential basal instability)
and is not associated with a particular strength limit state.
N c Su
FS = (8-8)
S 2
H γ − u
B
The width, B', is restricted if a stiff stratum is near the bottom of the cut (Figure 8-75). For
this case, B' is equal to depth D. Substituting D for B' in Equation 8-8, results in:
N c Su
FS = (8-9)
Su
H γ −
D
In relation to anchored wall designs in shallow deposits, Equation 8-9 may be used.
However in moderate to deep soil deposits where the width of the excavation is very large,
the contribution of the shearing resistance along the exterior of the failure block is negligible
and equation 8-8 reduces to:
Nc N
FS = = c (8-10)
γH Ns
Su
where Ns is the stability number defined as γH/Su. The bearing capacity factor used in
Equation 8-10 is affected by the height/width ratio (H/B), and the plan dimensions of the cut
(B/L). Values of the bearing capacity factor, Nc, proposed by Janbu et al. (1956) for analysis
of footings may be used in Equation 8-10 and these values are shown on Figure 8-75. Note
from Figure 8-75 that Nc values are greater for excavations constructed in short lengths (e.g.,
slotted excavation) as compared to excavation of the entire length of the wall. Unless the
designer specifically requires staged lengths of excavation, the design should be based on the
assumption that the contractor will remove the entire length of each lift of excavation.Current
practice is to use a minimum factor of safety against basal heave of 1.5 for temporary
facilities in soft to medium clays. As the factor of safety decreases, loads on the lowest
ground anchor increase.
b a
Su H HB'
H
H
d c
D
B/ 2
(a) Failure planes, deep deposits of weak clay (b) Failure plane, stiff layer below bottom of excavation
10
Square and
9
circular B/L = 1
B/L = 0.5
8
H = Excavation depth
7 B = Excavation width
Nc
B/L = 0, Rectangular L = Length of excavation
6
4
0 1 2 3 4 5
H/B
Figure 8-75. Analysis of Basal Stability (modified after Terzaghi et al., 1996.).
To evaluate the overall stability of an anchored system at the Service I limit state, potential
failure surfaces passing behind or through the anchors need to be checked. For walls with
multiple levels of anchors, failure surfaces should be checked that pass just behind each
anchor (Figure 8-76). In checking a failure surface that passes behind a level of anchors, the
failure surface may cross in front of or through the anchor bond zone of another level(s) of
anchors. In this case, the analysis may be amended to include a portion of the restraint force
8.11.11 Step 13: Estimate Maximum Lateral Wall Movements and Settlements at the
Service Limit State
The displacement of anchored walls is evaluated at the Service I limit state for all applicable
load combinations. Because evaluations of structure displacements by LRFD are made at the
Service I limit state where γ = 1.0 and φ = 1.0, methods used to estimate settlement and
lateral displacement by LRFD are identical to those used in ASD.
The vertical and lateral displacement of anchored walls is a complex soil-structure interaction
problem, and deformation analyses can be performed using modified forms of beam on
elastic foundation theory or finite element analyses. Depending on project constraints,
requirements with respect to control of wall and ground movements will vary. For example,
permanent anchored walls constructed in granular soils with no nearby structures pose little
concern with respect to movements. Wall and ground movements, however, may be the
primary design issue for a temporary excavation support system located in an urban area.
For earth retaining structures installed using top-down construction methods, numerous
factors influence the amounts of wall and ground movements that may occur. These factors,
identified by Clough et al. (1989) are listed below:
Some of the above factors are not under the control of the designer, but are governed by the
contractor’s selected construction procedure and on the quality of their workmanship.
Nevertheless, some semi-empirical guidelines have been developed by combining field
experience with analytical tools, as discussed below.
Figure 8-77 presents correlations for estimating vertical soil movements behind externally
supported walls. Several key points associated with this figure are as follows (Clough and
O’Rourke, 1990):
• Only the basic excavation and support process has been considered. Other factors
listed above have to be evaluated separately and could result in more movements.
• Excavations in stiff to very hard clays show variable behaviors as they are influenced
by the in situ horizontal stress, degree of fissuring, degree of weathering, and
plasticity. Heave may also be possible for some conditions. For these materials, the
dimensionless diagram in Figure 8-77(b) should be used as a conservative estimate,
provided that the wall is stable and not affected by poor construction practice. In
• The family of curves shown in Figure 8-77(c) is based on numerical studies and
assumes good workmanship and that cantilever deformation (i.e., lateral wall
movement of the very top of the wall (above the uppermost support level)) represents
a small fraction of the total movement. Curves I and II are commonly used for
permanent anchored walls. Settlements increase rapidly for walls constructed in soft
to medium clays where basal stability is marginal. The cantilever stage movements
can be idealized assuming a point of fixity at an appropriate depth below the ground
surface.
• Use Figure 8-77(d) with caution, especially where the factor of safety against basal
heave is below 1.5. In these conditions, construction variables can cause significant
increases in movements.
Maximum lateral wall movements for anchored walls constructed in sands and stiff clays
average approximately 0.2%H with a maximum of approximately 0.5%H where H is the
height of the wall. Maximum vertical settlements behind a wall constructed in these
materials average approximately 0.15%H with a maximum of approximately 0.5%H.
8.12.1 General
In general, the geotechnical (e.g., wall embedment) and structural design of stiff (e.g., slurry,
secant/tangent pile, jet-grouted, and DMM) wall systems may be performed using the
methods previously discussed in Section 8.9 for flexible nongravity cantilevered walls and
Section 8.11 for flexible anchored walls. It is specifically noted, however, that these
relatively simple methods are becoming more limited in today’s practice because the
simplifying assumptions previously required to perform wall design are no longer required.
Today, relatively simple soil-structure interaction analysis programs are available to model
all stages of wall construction leading to better estimates of wall bending moments, shear
forces, and lateral wall movements. In this section, information on the use of soil-structure
interaction analysis methods (which are particularly well-suited to stiff wall systems) is
described. Specific design issues relative to stiff wall systems are also summarized in this
section.
8.12.2.1 General
Soil-structure interaction (SSI) methods are available and being used to model wall behavior
considering various construction stages (e.g., excavation and brace/anchor installation.). In
general, this is the preferable method of analysis because both the temporary construction
conditions and the final design condition are considered. Example staged analyses methods
include modeling the wall as a beam with springs to represent various layers of soil stiffness
and strength (e.g., beam on elastic foundation) and also the finite element approach. There
are a variety of finite element programs, with both two and three dimension capabilities and
numerous other modeling abilities, such as dewatering, anchor representation, etc. Most of
these more sophisticated methods can provide a more accurate, refined estimate of lateral
bracing loads, wall moments and shears, and displacement in the surrounding soil mass,
assuming that appropriate model parameters and excavation stages are used in the analyses.
As a result, a more economical design can often be achieved.
In most staged analysis methods, the type, spacing and stiffness of the brace or anchor can be
modeled and input in the analysis. Also, the wall stiffness is input in these approaches based
on the modulus of elasticity, E, and the moment of inertia, I. Fine tuning (e.g., cracked vs.
uncracked section, composite, etc.) the bending stiffness value, EI, may be appropriate in
some cases where a calibrated model is used with well defined soil parameters. In general,
the primary influence of the EI value (within typical ranges for stiff wall systems) is on
movements, not wall bending moments. However, there are other parameters (such as soil
shear strength) which are typically known with less certainty that influence the strength and
performance requirements for the wall system to a greater degree.
Herein, the so-called “p-y” analysis method is described as a viable approach to perform SSI
analyses. The p-y analysis method predicts soil resistance along a pile (or wall element in
this case) as a function of lateral pile deflection by treating the soil resistance as a series of
non-linear springs. Common practice is to impose a lateral pressure distribution above the
excavation subgrade and to perform the p-y analysis for the embedded (or passive side)
portion of the wall. Distinct soil and rock layers on the passive side with distinct p-y curves
can be modeled. Anchors and braces can also be treated as springs and walls with either
discrete or continuous wall elements can be modeled. Commercially available computer
programs such as LPILE (ENSOFT, 1999), PYWALL (ENSOFT, 1999), WALLAP
(GEOSOLVE, 2002), GT STRUDL (GATECH, 2003), and others are available to perform
this SSI analysis.
It is noted, that although this method allows for modeling all stages of construction and thus
would require the design engineer to develop an earth pressure diagram for each stage of the
analysis, the method can also be used to just model a final excavation height condition. In
that case, an apparent earth pressure diagram is developed to envelope the earth pressures
imposed by all stages of construction. In many cases, however, stiff wall systems used for
temporary support are incorporated into a final structure such as a building. In that case, a
stage construction analysis is required to evaluate the stresses and strains that are locked into
the final structure at the end of construction.
Semi-empirical formulations have been developed for establishing p-y curves based on the
results of full-scale instrumented load tests. For larger projects, the use of pressuremeter data
can be used to establish site-specific p-y curves. For example, for the Marquette Interchange
Project (ca. 2004) in Milwaukee, Wisconsin, 11 distinct p-y curve categories were developed
based on the results of 37 site-specific pressuremeter tests for a particular segment of the
project. These categories are based on soil type, consistency (or relative density), vertical
effective stress (or overburden pressure) at the testing elevation, and similarity of the p-y
curves. Typically, the predicted wall displacements from a SSI analysis are much more
sensitive to the values used for the p-y curves than the predicted wall bending moments. For
this reason, conservative selection of the parameters for the p-y curves should provide
conservative estimates of ground movements without significantly increasing the structural
demand of the wall (and, if used, bracing system) (Pearlman et al., 2004).
Figure 8-78 shows a conceptual model for a SSI analysis. In this case, only the final
excavation condition is shown.
An LRFD-based procedure for designing walls using SSI is not provided in AASHTO
(2007). Herein, a general sequence of analysis for a cantilevered wall (without bracing or
anchors) is provided:
1. Evaluate unfactored loads (e.g., earth, surcharge, water) for the portion of the wall
above the excavation level.
2. Determine factored loads at all appropriate strength and service limit states.
3. Assume a certain structural section to allow for computation of bending stiffness.
TH1
Lateral Earth
Pressure
Anchor Forces
TH2
Subgrade resistance
Modeled with p-y curves
Since the soil/rock response is linked to deformation (via a p-y curve), it is recommended
that for all strength and service limit state evaluations, soil p-y curves should use a resistance
factor of 1.0. This implies, however, that uncertainty in predicted resistance is not addressed.
The evaluation of overall stability is concerned, in part, with the evaluation of potential heave
into the bottom of the excavation. Procedures described in Section 8.10.10 are used to
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-142 June 2008
evaluate bottom heave potential in clayey soils. For excavations in sand, bottom heave is
only a concern if upward seepage gradients directed towards the excavation become large
enough to liquefy the sand.
Figures 8-79, -80, -81, and -82 identify significant features of excavation geometry in
relation to groundwater control. Seepage can enter the excavation through the walls and at
the bottom of the excavation.
Seepage quantities should be evaluated before construction using flow nets or other
analytical techniques. Seepage through a wall should be assessed on the basis of differential
head, retained soil characteristics, and effective permeability of the wall where the effective
permeability of the wall is controlled primarily by defects resulting from construction. For
example, a torn or jumped interlock of a steel sheet pile wall could allow an order of
magnitude more seepage through the wall (and movement of fines) than indicated by
laboratory or design studies of seepage through “tight” interlocks. Similarly, a local zone of
honeycombed concrete could allow significant seepage compared to properly placed
concrete. Qualitative comparisons for various wall types are shown in Table 8-19 (ASCE,
1997).
For the SSI analysis, the lateral pressures above the excavation line are multiplied by the
center-to-center spacing of the piles. Below the excavation, the passive resistance (as
defined by the p-y curves for the below excavation ground) must be modified to account for
spacing effects.
The resistance of a group of closely spaced piles is less than would be calculated based on
the sum of the resistances of the same number of widely spaced piles. Therefore, p-y curves
may need to be adjusted by a so-called “p-multiplier”, Pm, for closely spaced piles. The
selection of a p-multiplier is site specific and is generally taken as a function of the ratio of
the spacing of the piles (s) to the diameter of the pile (b). For example, for the TREX project
in Denver involving cantilevered drilled shaft walls (i.e., tangent pile walls), a Pm of 1.0 was
used for s/b=3 and 0.5 was used for s/b=1 (Sisson et al., 2004). AASHTO (2007) Section
10.7.2.4 recommends that a Pm value less than 1.0 only be used if the pile spacing is 5b or
less (e.g., use Pm = 0.7 for s/b =3).
According to AASHTO (2007), p-y analysis methods apply to elements that have the ability
to bend and deflect. For large diameter, relatively short elements, the element tends to rotate
rather than bend in which case strain wedge theory may be more applicable (Ashour et al,
1998).
Type Of Wall Potential Leakage Through Potential for Piping beneath Wall
Wall
Wood, concrete, or steel Deflection opens joints between To avoid piping, penetration below
cantilever sheeting of sheets. Leakage typically 0.1 to 4 subgrade must exceed ¼ of exterior
low height gpm per 100 linear ft for low head where there is limited depth of
head. Also, depends on joint pervious layer or more than ½ of
detail. head for deep pervious layer.
Braced, interlocked steel Leakage typically 1 to 10 gpm per As above, but in potentially
sheeting 100 linear ft for low head. Lesser “running” soil (i.e., non-plastic silt,
quantity of leakage if movement silty fine sand, or narrowly graded
is minimized, locks are filled and sand), piping may occur in a path
tensile stress acts along the wall along the face of the sheeting.
in longitudinal direction.
Steel soldier piles with No impediment to leakage; Not intended for cutoff below
wood or concrete exterior drawdown and water subgrade. If exterior head is not
lagging pressure in active wedge of greatly reduced by leakage,
retained soil can be analyzed by dewatering to control uplift may be
flow net. Inflow depends on soil necessary.
permeability and recharge.
Concrete cylinder piles: Equivalent permeability of wall is Could accomplish partial or complete
tangent, secant, or typically 5 x 10-4 to 1 x 10-5 cm/s. cutoff below subgrade, but usually
staggered Leakage typically 0.5 to 8 gpm not seated in rock. Layout
per 100 linear ft per 10 ft of head convenient to work around shallow
depending principally on quality obstructions.
of joints between cylinders.
Slurry trench concrete Equivalent permeability of wall is Cutoff enhanced by chiseling into
wall typically 2 x 10-4 to 1 x 10-6 cm/s. underlying rock or by pre-positioning
Leakage typically 0.1 to 4 gpm grout pipe in wall elements to
per 100 linear ft per 10 ft of head. facilitate grouting in strata beneath
Much influenced by anchor wall.
penetration through wall and joint
quality. Details of penetration are
important.
Note: Leakage through wall is expressed as quantity of inflow per 100 ft length of wall for each 10 ft
difference in head across the wall, assuming leakage is not limited by permeability of the retained
soil.
Weatherby et al. (1998) provides a detailed discussion on the use of SSI for anchored walls.
In that report, a detailed description of a modeling approach for the ground anchor is
provided. A summary is provided below.
T-y curves are used to model ground anchors in a SSI analysis after they have been stressed
and locked off. Figure 8-83 shows how a non-linear T-y curve for a SSI analysis is
developed. Since ground anchors are installed at an angle, the horizontal components of
anchor load and tendon elongations are used in developing the T -y curves. The anchor lock-
off load is the starting point for the ground anchor in the analysis, and the deflection
associated with the lock-off load corresponds to the wall deflection from the previous
construction stage (anchor stressing). If the wall moves out, the ground anchor load will
increase, and if the wall moves back into the soil, the lock-off load will decrease. High
ground anchor loads will move the wall back into the ground. Low anchor loads will result
in the wall deflecting outward until the ground anchor load increases. The initial slope of the
T-y curve is the horizontal component of the anchor tendon stiffness and it is given as:
As Es
k= cos α (8-11)
Lu
where:
k = anchor stiffness
As = area of anchor tendon
Es = Young’s modulus for the anchor tendon
Lu = effective unbonded length
α = anchor inclination
In Equation 8-11, the effective unbonded length of the anchor tendon is assumed to be the
sum of the unbonded length plus half the tendon bond length. This value is assumed to
permit the T-y curve to be constructed, but the actual elastic behavior of the ground anchor
will be different. Bending moments are not sensitive to changes in elastic length. If the
ground anchor load changes during the analysis, wall deflections will vary depending upon
the unbonded length used to construct the T-y curves. At the yield load, the T -y curve
changes slope. The second portion of the anchor curve represents the ground anchor
behavior between the yield and ultimate tendon strength.
The design procedures previously presented for soldier pile and lagging walls are generally
applicable for the design of slurry walls. However, there are important differences in the
design of slurry walls since: (1) they are generally continuous structures for their full depth
of penetration; (2) they are considerably stiffer than soldier pile and lagging systems and are
usually used where strict deformation control is required; and (3) they are typically designed
for full hydrostatic water pressure behind the wall. The design considerations specific to
slurry walls are discussed below.
Earth pressures greater than those corresponding to active conditions may be appropriate if
the wall system is highly preloaded and/or stiff and workmanship is of high quality (ASCE,
1997). Based on this, it is reasonable to assume that most slurry walls should be designed for
higher than active earth pressure loadings. Since slurry walls will likely only be used for
projects where strict deformation control must be maintained, it is reasonable to use earth
pressures consistent with those obtained from apparent earth pressure diagrams. Slurry walls
are often used in urban environments wherein different soil types may exist over the depth of
the excavation. In this case, it may be difficult to select a single apparent earth pressure
diagram for the entire depth of the excavation so more than one apparent pressure diagram
may be used to estimate loads on the wall. Usually, unless a dewatered condition is
achieved, hydrostatic water pressures would be added to the apparent earth pressure diagrams
for slurry walls. Alternatively, water pressures consistent with seepage conditions under and
around the wall maybe assumed.
The previous discussion relates primarily to temporary loadings for slurry walls. Usually, if
the slurry wall is used for a permanent system, the permanent lateral earth pressures are
assumed to be consistent with at-rest earth pressures. For the permanent case, appropriate
long-term water pressures are added.
It is noted that when more sophisticated analyses such as finite element methods are used to
analyze all stages of slurry wall construction (including excavation, support (i.e., anchor)
loads, etc.), that the actual computed earth pressures are less than those from apparent earth
pressure diagrams. These more sophisticated analyses are often used for large projects or
where significant construction cost savings may be realized through the optimization of
structural elements.
Reinforced concrete slurry walls are typically analyzed as a beam spanning vertically
between support levels. For Soldier-Pile-Tremie-Concrete (SPTC) walls, however, the
In many cases, slurry walls are incorporated into a permanent structure and must be designed
to support vertical loads. These loads result from vertical anchor forces as well as permanent
vertical loads (such as from a building if the slurry wall is a basement wall). For design,
these vertical forces may be carried, at least partially, above the excavated subgrade, however
since slurry walls are often used in relatively weak soils, it is conservative to assume that all
load is carried by the excavation subgrade. It is important, however, to consider the specific
ground conditions for the site to evaluate the conservatism of this assumption.
The portion of the load carried below the excavation subgrade is resisted by skin friction over
two sides of the panel and through end bearing. Because the width of the base of a slurry
wall is relatively large (e.g., 2 ft or greater), a relatively significant amount of vertical
movement would be required to mobilize end bearing. For ASD, a reasonable assumption is
that no more than 20 percent of axial resistance is achieved through end bearing with the
remainder of resistance generated in skin friction. A similar distribution of resistance should
be assumed for LRFD-based computations.
The earth pressure distribution diagrams, types of wall support, methods of analysis for
design of tangent pile and secant pile walls are the same as those discussed for slurry walls.
When a steel reinforcing cage or steel structural section is placed in all of the bored piles, the
design load for each pile is determined by multiplying the appropriate soil and water pressure
diagrams by the center-to-center spacing of the piles. When only alternate piles are
reinforced, the design load for the reinforced piles is determined by multiplying the
appropriate pressure diagrams by the center-to-center spacing of the reinforced piles.
Procedures used for the axial resistance evaluation of drilled shafts (see AASHTO Section
10.8.3) may be used to calculate required embedment depths for tangent/secant pile walls
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-152 June 2008
designed to carry vertical forces. In general, an analysis of a single shaft is performed and
this capacity is reduced to account for overlapping between shafts and group effects resulting
from very closely spaced piles. The recommendations provided in Section 8.11.9.2 of this
manual may be used for LRFD-based evaluations.
For intermittent pile walls, individual capacities can be reduced to account for group effects
in the same way that the capacity of a drilled shaft group is reduced from that of the sum of
the capacities of individual shafts.
In cases where these walls (or other stiff walls systems) support permanent vertical loads,
site-specific load testing may be performed. In that case, resistance factors consistent with
the level of static load testing performed can be used for bearing resistance design.
For a jet-grouted wall, steel reinforcements can be inserted into the wall elements before the
soil-grout mix hardens. The resulting structural combination can resist lateral earth
pressures, and when the wall is properly supported laterally (e.g., by ground anchors) it can
support excavations of considerable depth. The overlapped columns can also be used as a
cut-off wall for seepage control.
The jet-grouted earth wall is designed based on conventional wall design procedures. The
width of the wall is usually defined in terms of column diameter, configuration of overlapped
columns, and minimum overlapping length. Jetting parameters, which achieve the required
width, strength and permeability of the wall, are then established based on existing soil and
groundwater conditions. The characteristics of the jet-grouted soil (strength, permeability,
etc.) are influenced by the properties of the in situ material, the composition of the grout mix,
and a number of operating parameters, such as injection pressure, flow rate, withdrawal rate
(lifting speed), rotation rate, etc.
Due to the large number of variables involved in the jet grouting process, it is very difficult
to establish a rigorous design process. A typical design procedure is dependent to a great
extent on the field and laboratory test results, the specific equipment and method used, and
the experience of the equipment operator. The following steps are followed in the design of
a jet-grouted wall:
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-154 June 2008
Table 8-21. Typical Range Of Jet Grouting Parameters And Jet-Grouted Soil Properties
(adapted from Kauschinger and Welsh, 1989)
Note: aHigher grout pressures of 10-18 MPa are used with certain equipment where large separation exists
between the water/air nozzle and the grout nozzle. 1 m = 3.3 ft, 1 MPa = 145 psi, 1 kg/m3 = 0.062 pcf, 1 liters =
0.26 gallons.
1. Establish earth pressure diagrams. For determining the lateral earth pressures, the jet-
grouted wall may be considered as a gravity structure due to its large mass. For
thinner, externally supported and reinforced jet-grouted walls subject to bending, the
earth pressures acting on the wall may be similar to those of a slurry diaphragm wall.
FHWA NHI-07-071 8 – Non-Gravity Cantilevered and Anchored Walls
Earth Retaining Structures 8-155 June 2008
2. Perform stability and wall seepage analyses and determine wall thickness based on
assumed strength, permeability, column diameter and overlapping length. Use Table
8-21 and 8-22 in establishing these assumptions.
3. Conduct a field trial, coupled with laboratory tests, to confirm the design assumptions
and select the parameters to use for production.
Table 8-22. Range of Typical Soilcrete Strengths (Three Fluid System) (after Elias et al.,
2006).
8.12.6.1 General
The design of a DMM wall is influenced, to a certain extent, by its planned function. When
used as a cut-off wall, for instance, the wall is designed to provide a required permeability.
When used for excavation support, on the other hand, vertical reinforcing members are
usually provided to resist bending moments as well as shear stresses along the longitudinal
direction of the wall. Between the reinforcements, the soil-cement structure is designed to
resist and redistribute the horizontal stresses to neighboring reinforcing members.
Table 8-23 summarizes the range of anticipated engineering parameters that can be achieved
by deep mixing techniques. Note that the wide range of reported improved properties is a
function of the in-situ soils, reagent content, equipment used for mixing, and the in-situ water
and organic content. Additional information on engineering properties of wet (and dry) mix
treated soils is provided in Elias et al. (2006).
Strength, permeability and modulus of elasticity are the primary engineering properties of
concern for the DMM wall. These properties are affected by the soil type, cement and other
admixture proportions, water/cement ratio of the grout, degree of soil-cement mixing, curing
environment and age. Based on actual wall projects in various soil types, Taki and Yang
(1991) presented design guidelines, many of which are included in the following sections.
For a given cement proportion, the strength increases with age. The 28-day strength is
typically 1.4 to 1.5 times the 7-day strength for clays, and 2 times the 7-day strength for
sands. The 56-day unconfined compressive strength may be 1.5 times to 28-day strength.
Strength continues to increase with time for perhaps 6 months.
For design purposes, one third of the unconfined compressive strength can be considered as
the shear strength, τf, of the soil-cement, i.e.:
For a well mixed soil-cement, the coefficient of permeability ranges from 10-5 to 10-7 cm/sec.
In general, the coefficient of permeability and the porosity of the soil-cement matrix decrease
with decreasing the sand content and the water-cement ratio, and increasing the curing age.
It should be noted, however, that the above indicated coefficients of permeability are those of
the soil-cement mix and do not necessarily represent the overall permeability of the DMM
wall. This permeability is influenced by the tightness of the wall and the potential presence
of “windows” which may be caused by boulders or column deviations. The overall
permeability of the DMM wall is best determined by performing in situ permeability tests in
the completed wall.
The structural design of DMM walls includes: (a) the design of reinforcing members to resist
bending moments, shear stresses and deflections along the height of the wall; (b) the design
of soil-cement elements to resist and redistribute the lateral pressures to the reinforcing
members; and (c) the determination of the minimum depth of embedment required for base
stability or vertical load support.
The design of reinforcing members and the required depth of embedment are carried out as in
the case of conventional soldier pile and lagging walls. The lateral earth pressures are similar
to those discussed for slurry walls. The design information presented herein is for
informational purposes only and has not, to date, been calibrated for a LRFD format.
wall loading
Q (a)
d
l2
wall loading
(b)
Center of H-pile
D h e
45º Center of DMM Column
L2
If L2 ≤ D + h -2e, no bending failure
(c)
Figure 8-84. DMM Wall Design: (a) Analysis for Punch-Through Shear; (b) Analysis for
Compressive Action of Arching Effects; (c) Empirical Guideline for Avoiding Bending
Failure (after Taki and Yang, 1991)
The need for drainage in cut wall system applications varies with project requirements.
Drainage systems may be omitted in cases where groundwater drawdown in the retained soil
is prohibited or undesirable. In other cases, drainage is used as a means to control surface-
water infiltration and groundwater seepage. Other beneficial effects of drainage include:
For cut wall systems, collection of subsurface flow is usually achieved with prefabricated
drainage elements (i.e., geocomposites) placed between the wall and the permanent facing.
With this type of system, vertical drainage strips are extended over the full height of the wall.
Single strips can be placed at appropriate horizontal spacing along the wall or a continuous
sheet can be placed over the entire wall face, depending on the project drainage requirements
and the expected flow rate. Water intercepted in the drainage elements flows downward to
the base of the wall where it is conveyed through the permanent facing in longitudinal/outlet
pipes or weepholes. In other applications, the drainage elements are extended into the
subgrade to a footing drain. Similar drainage systems may be used for face drainage of soil-
nailed walls where the prefabricated drainage elements are typically placed between the
excavated face and the back of the shotcrete facing layer.
In applications where subsurface flow rates are large, horizontal drains may be used to
remove water from behind the wall. A horizontal drain is a small diameter perforated pipe
that is advanced into a nearly horizontal drill hole in an existing slope. For example, a cut
wall constructed on or at the base of a steep slope will likely interfere with pre-existing
General surface water drainage is discussed in Section 5.5.8.3 in Chapter 5. For cut wall,
dikes can be constructed on the ground surface near the top of the wall or the vertical wall
element can be extended above the ground surface grade to minimize surface water that can
enter the excavation during construction and weaken the soils inside the excavation (Sabatini
et al., 1997).
9.1 INTRODUCTION
In-situ reinforced walls are constructed from the top-down to support temporary and
permanent excavations. Construction of these walls involves insertion of reinforcing
elements in the in-situ soils to create a composite earth structure. This chapter presents
information on two different types of in-situ walls: (1) soil nail walls and (2) micropile walls
(a less common wall type).
The main components of an in-situ reinforced wall are the in-situ material, the reinforcing
inclusions, and the wall facing. The reinforcing inclusions typically consist of metal bars
(soil nail walls) and small diameter steel pipe piles (micropile walls). Shotcrete, welded-wire
mesh, cast-in-place concrete, or precast concrete panels are typically used for the facing. In-
situ reinforced walls can be used in a wide range of ground conditions.
In-situ reinforced earth walls have been successfully used for a variety of applications
including:
9.2.1 General
Soil nail walls are one of the in-situ reinforced walls that are constructed to support
temporary and permanent excavations (Figure 9-1). This chapter presents information on
feasibility, construction materials and methods, cost, design, load testing, and construction
inspection of soil nail walls. A more comprehensive coverage of soil nail walls is provided
in Geotechnical Engineering Circular No. 7 “Soil Nail Walls” (Lazarte et al., 2003).
Nails can be installed by (1) driving; (2) drilling and grouting; (3) jet grouting; and (4)
launching. Conventional U.S. practice includes constructing a drill hole in which the bar is
placed and then grouted under gravity (Figure 9-2). Driven soil nails have been more
commonly used in France and Germany than in the U.S. and are typically used for temporary
applications. Jet-grouted and launched nails are not common in the United States however
launched nails have been used for a small number of temporary slope stabilization projects.
Also, so called “hollow-core” soil nails are becoming increasingly used. In this method,
drilling and grouting of the nail is accomplished in one step (i.e., Steps 2 and 3 shown in
Figure 9-2 are combined into one step).
The components of a soil nail wall installed using the techniques listed above vary from one
technique to another. Although the concept of soil nail walls is not proprietary, several
specialized components or procedures are under U.S. or international patents. Patented
components include, but are not limited to, some threaded bars, corrosion-protection systems,
and nail installation systems (e.g., self-drilling, jet-grouted, and launched soil nails). Hollow
bars are proprietary and have been used primarily for temporary walls because of concerns
regarding the consistency of the corrosion protection; however, research is on-going with a
focus towards identifying uses of hollow bars for permanent applications.
Common applications for soil nail walls include: (1) temporary shoring for construction of
CIP walls or other permanent facilities; (2) a permanent wall for roadway widening at a
bridge abutment; (3) a permanent retaining wall for a roadway cut; and (4) slope
stabilization. Examples of each of these applications are shown in Figure 9-3.
The feasibility evaluation of a soil nail wall should encompass technical and economical
considerations and include: (1) an evaluation of the prevailing ground conditions; (2) an
assessment of the advantages and disadvantages of a soil nail wall for the particular
application being considered; and (3) comparison with alternative systems (e.g., ground
Figure 9-2. Typical Nail Wall Construction Sequence (after Porterfield et al., 1994).
(c) (d)
Figure 9-3. Soil Nail Wall Applications (a) Temporary Shoring, (b) Roadway Widening Under Existing Bridge, (c) Roadway Cut, and (d)
Slope Stabilization.
Soil nail walls can be used for a wide range of soil types and conditions. Project experience
has shown that certain favorable ground conditions make soil nailing cost effective over
other techniques. Conversely, certain soil conditions can be considered marginal for soil
nailing applications and may make the use of soil nails too costly when compared with other
techniques. Soil nail walls can generally be constructed without complications in a mixed
stratigraphy, as long as the individual layers of the soil profile consist of suitable materials.
The following two sections present the soil conditions that are considered most and least
suitable for soil nail walls. Intermediate soil conditions, for which the feasibility of soil
nailing is not readily apparent, are also described.
Soil nail walls have been constructed successfully in various types of soils. Construction
difficulties and long-term complications can generally be avoided when specific favorable
soil conditions prevail. Soil nailing has proven economically attractive and technically
feasible when:
• all soil nails within a cross section are located above the groundwater table; and
• if the soil nails are below the groundwater table, the groundwater does not adversely
affect the face of the excavation, the bond strength of the interface between the grout
and the surrounding ground, or the long-term integrity of the soil nails (e.g., the
chemical characteristics of the ground do not promote corrosion).
Although not an absolute requirement, it is advantageous that the ground conditions allow
drillholes to be advanced without the use of drill casings and for the drillhole to be
unsupported for a few hours until the nail bars are installed and the drillhole is grouted.
Alternatively, soil nails have been installed with success using the hollow-stem drilling
method in fully and temporarily cased drillholes. It is important to note that the selection of
the drilling method is typically left to the discretion of the soil nail installation contractor.
Soil conditions are presumed to be favorable for the construction of soil nail walls when
results from field tests indicate competent soils. The Standard Penetration Test (SPT, see
next chapter) provides the SPT value, N, which can be used to preliminarily identify
favorable soil conditions.
• Stiff to hard fine-grained soils. Fine-grained (or cohesive) soils may include stiff to
hard clays, clayey silts, silty clays, sandy clays, sandy silts, and combinations thereof.
Fine-grained soils can be tentatively classified as stiff if they have SPT N-values of at
least 9 blows/ft. However, the consistency characterization of fine-grained soils
should not rely solely on SPT N-values. Instead, the consistency (and thereby shear
strength) characterization should be supplemented with other field and/or laboratory
testing. To minimize potential long-term lateral displacements of the soil nail wall,
fine-grained soils should be of relatively low plasticity (i.e., in general, plasticity
index (PI) < 15).
• Dense to very dense granular soils with some apparent cohesion. These soils include
sand and gravel with SPT N-values larger than 30 (Terzaghi et al., 1996), and with
some fines (typically no more than about 10 to 15 percent of fines) or with weak
natural cementation that provide cohesion. Capillary forces in moist fine sands may
also provide an apparent cohesion. To avoid excessive breakage of capillary forces
and thereby significant reduction of this apparent cohesion, the movement of water
toward the excavation face needs to be minimized including by redirecting surface
water away from the excavation face.
• Weathered rock with no weakness planes. Weathered rock may provide a suitable
supporting material for soil nails as long as weakness planes occurring in unfavorable
orientations are not prevalent (e.g., weakness planes dipping into the excavation). It
is also desirable that the degree of weathering be approximately uniform throughout
the rock so that only one drilling and installation method will be required.
Conversely, a highly variable degree of rock weathering at a site may require changes
in drilling equipment and/or installation techniques and thereby cause a costly and
prolonged soil nail installation.
• Glacial soils. Glacial outwash and glacial till materials are typically suitable for soil
nailing applications as these soils are typically dense, well-graded granular materials
with a limited amount of fines.
Examples of unfavorable soil types and ground conditions are provided below:
• Dry, poorly graded cohesionless soils. When poorly graded cohesionless soils are
completely dry, contain no fines, or do not exhibit any natural cementation, apparent
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-7 June 2008
cohesion is not available. Therefore, the required vertical or nearly vertical cuts are
difficult to achieve.
• Soils with high groundwater. Perched groundwater occurring behind the proposed
soil nail wall will require significant drainage, which is necessary to stabilize the
mass of soil in this location. Additionally, large amounts of groundwater can cause
drillholes (particularly in loose granular soils) to collapse easily, thus requiring a
costly soil nail installation. Excessive groundwater seeping out to the excavation face
may cause significant difficulties for shotcrete application.
• Soils with cobbles and boulders. A large proportion of cobbles and boulders present
in the soil may cause excessive difficulties for drilling and may lead to significant
construction costs and delays. When only a few boulders and cobbles are present,
modifying the drilling orientation from place to place may minimize or eliminate
most of the difficult drilling. However, this approach has practical limitations when
too many boulders are present.
• Soft to very soft fine-grained soils. These soils typically have SPT N-values less than
4 and are unfavorable for soil nailing because they develop relatively low bond
strengths at the nail-grout-soil interface, thereby requiring unreasonably long nail
lengths to develop the required resistance. Long-term deformations (creep) of the
soils may be a concern for highly plastic clays. Concerns for creep deformations are
generally less critical for temporary applications. As with any retaining system
constructed in a top-down manner, the potential for instability at the bottom of the
excavation is high in soft fine-grained soils. Additionally, high-plasticity soils may
be expansive and may induce additional localized pressure on the facing due to
swelling.
• Organic soils. Some organic soils such as organic silts, organic clays and peat
typically exhibit very low shear strengths and thereby low bond strengths, which
causes uneconomical nail lengths. While some organic soils can exhibit acceptable
shear strengths, other organic soils like fibrous peat may be highly heterogeneous and
highly anisotropic. In this case, while the soil shear strength can be reasonable along
some orientations, it may be significantly low along other orientations. These
unfavorable orientations may have a detrimental impact on the wall stability and very
long soil nails will be required. In addition, organic soils tend to be more corrosive
than inorganic soils.
• Highly corrosive soil (cinder, slag) or groundwater. These conditions may lead to the
need of providing expensive corrosion protection. These conditions are obviously
more disadvantageous for permanent applications of soil nail walls.
• Loess. When it is dry, loess may exhibit acceptable strengths that would allow
economical installation of soil nails. However, when sizable amounts of water
ingress behind the proposed soil nail wall, the structure of the loess may collapse and
a significant loss of soil strength may take place. Therefore, the collapse potential
upon wetting of these soils must be evaluated. Appropriate measures to avoid excess
water migration to the soil nail area must be provided in loess exhibiting significant
collapse potential. Additionally, considerably low soil shear strengths may arise for
the wetted condition. In these cases, unusually long soil nail lengths may result in
using conventional methods of nail installation. Regrouting (an atypical and more
costly step) has been used to increase bond strengths in loess.
In addition to the difficulties described above, other aspects related to soil conditions must be
considered when assessing the feasibility of soil nail walls:
• The prolonged exposure to ambient freezing temperatures may cause frost action in
saturated, granular soils and silt; as a result, increased pressures will be applied to the
temporary and permanent facings.
• Repeated freeze-and-thaw cycles taking place in the soil retained by the soil nail wall
may reduce the bond strength at the soil nail grout-ground interface and the adhesion
between the shotcrete and the soil. To minimize these detrimental effects, a suitable
protection against frost penetration and an appropriate shotcrete mix must be
provided.
• Granular soils that are very loose (N ≤ 4) and loose (4 < N ≤ 10) may undergo
excessive settlement due to vibrations caused by construction equipment and traffic.
• Loose and very loose saturated granular soil can be susceptible to liquefaction in
seismically exposed regions. Several ground modification techniques (typically with
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-9 June 2008
significant associated costs) may be utilized to densify granular soils and thereby
minimize these damaging effects.
Despite the difficulties associated with unfavorable soil conditions described above, soil nail
walls may still be built. It should be recognized that these wall systems would typically be
more expensive to design and construct when compared to conventional walls in a more
suitable soil. It is likely that significant extra effort and cost is needed in the design and
construction of soil nail walls in these marginal conditions and that more strict long-term
performance requirements might be necessary to allow soil nailing in such challenging
conditions.
There exists some soil conditions that are intermediate to the two conditions described
previously. Examples of intermediate soil conditions are presented below:
• Engineered fill. Soil nails can be installed in engineered fill if it is a mixture of well-
graded granular material and fine-grained soil with low plasticity (typically, PI < 15).
• Residual soils. Residual soils (i.e., those soils created from the in-place weathering of
the parent rock material) may be an acceptable material for soil nailing. Similarly,
lateritic soil, a highly weathered tropical soil, may be acceptable. For these types of
soil, specific consideration should be given to the soil spatial variability and its ability
to drain.
Soil nail walls exhibit numerous advantages when compared to ground anchors and
alternative top-down construction techniques. Some of these advantages are described
below:
Construction
• requires smaller ROW than ground anchors as soil nails are typically shorter;
• less disruptive to traffic and causes less environmental impact compared to other
construction techniques;
• installation of soil nail walls is relatively rapid and uses typically less construction
materials than ground anchor walls;
• easy adjustments of nail inclination and location can be made when obstructions (e.g.,
cobbles or boulders, piles or underground utilities) are encountered; on the other
hand, the horizontal position of ground anchors is more difficult to modify almost
making adjustments in the field costly;
• because significantly more soil nails are used than ground anchors, adjustments to the
design layout of the soil nails are more easily accomplished in the field without
compromising the level of safety;
• overhead construction requirements are smaller than those for ground anchor walls
because soil nail walls do not require the installation of soldier beams; this is
particularly important when construction occurs under a bridge; and
• soil nailing is advantageous at sites with remote access because smaller equipment is
generally needed.
Performance
• soil nail walls are relatively flexible and can accommodate relatively large total and
differential settlements;
• measured total deflections of soil nail walls are usually within tolerable limits; and
• soil nail walls have performed well during seismic events owing to overall system
flexibility.
Cost
• soil nail walls are more economical than conventional concrete gravity walls when
conventional soil nailing construction procedures are used;
• soil nail walls are typically equivalent in cost or more cost-effective than ground
anchor walls when conventional soil nailing construction procedures are used; and
• shotcrete facing is typically less costly than the structural facing required for other
wall systems.
Some of the potential limitations of soil nail walls are listed below:
• soil nail walls may not be appropriate for applications where very strict deformation
control is required for structures and utilities located behind the proposed wall, as the
system requires some soil deformation to mobilize resistance; post tensioning of soil
nails can overcome this shortcoming in most cases, but this step increases the project
cost;
• the occurrence of utilities may place restrictions on the location, inclination, and
length of soil nails in the upper rows;
• soil nail walls are not well-suited where large amounts of groundwater seeps into the
excavation because of the requirement to maintain a temporary unsupported
excavation face;
To gain further insight into the soil nail wall concept, it is useful to compare the main
features of a soil nail wall with those of a ground anchor wall, which is a commonly used
top-to-bottom system for retaining wall construction. Detailed information on ground anchor
walls can be found in Sabatini et al. (1999).
• Nail/Anchor Density: Under similar project conditions, the number of required soil
nails per wall unit area is larger than the number of ground anchors per wall unit area.
The use of more reinforcing elements in a soil nail wall adds a degree of redundancy
that can contribute to the stability of a soil nail wall. Consequently, the failure of one
reinforcing element will have a smaller effect on the stability of a soil nail wall than
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-12 June 2008
that of a ground anchor wall. Typically, only five percent of production soil nails are
load tested, whereas all ground anchors are tested for acceptance.
• Load on Wall Facing: The density of soil nails implies that the facing in soil nail
walls supports a smaller portion of the soil pressure due to a smaller tributary area
compared to the facing in ground anchor walls, which supports a much greater soil
pressure. This difference is more due to the dissimilar design approaches in the two
systems rather than differences in the controlling load transfer mechanisms.
• Load Transfer: Soil nail transfer load along the entire length of the nails, whereas
ground anchors are designed to transfer load only in the anchor zone behind the
potential failure surface.
• Load Distribution: The resisting force provided by soil nails is variable along its
entire length. In a ground anchor, one portion of the anchor length is unbonded while
the remaining portion is bonded. The load in a ground anchor is approximately
constant in the unbonded length and variable in the bonded zone.
• Stability of Excavation Bottom: In ground anchor walls, soldier beams are embedded
to elevations below the bottom of the excavation. The shear resistance derived from
the embedded portion of the soldier beams provides additional stability of the bottom
of the excavation in ground anchor walls. This favorable effect is absent in soil nail
systems.
• Deflection: Field measurements in ground anchor walls indicate that the maximum
wall lateral deflection is generally at midheight of the wall. In soil nail walls, the
maximum lateral deflection takes place at the top of the wall. Also, maximum wall
deformations are generally greater in soil nail walls than in ground anchor walls.
This chapter presents information on construction materials and methods used for the
construction of soil nail walls typically used in U.S. highway applications. Conventional
U.S. practice includes the use of drilled and grouted soil nails. However, other soil nail
installation methods and materials have been developed and are briefly introduced here.
• Drilled and grouted soil nails: These are approximately 4- to 8-in. diameter nail
holes drilled in the foundation soils. These holes are typically spaced about 5 ft apart.
Steel bars are placed and the holes are grouted. Grouted soil nails are the most
commonly used soil nails and they can be used as temporary and permanent
applications (provided that appropriate corrosion protection is in place).
• Driven soil nails: These soil nails are relatively small in diameter [¾ to 1 in.] and
are mechanically driven into the ground. They are usually spaced approximately 3 to
4 ft apart. The use of driven soil nails allows for a faster installation (as compared to
drilled and grouted soil nails); however, this method of installation cannot provide
good corrosion protection other than by sacrificial bar thickness. For this reason,
driven nails are only used in the United States for temporary applications. At this
time, this method is not recommended for soil nail installations for permanent soil
nail walls.
• Self-drilling soil nails: These soil nails consist of hollow bars that can be drilled and
grouted in one operation. In this technique, the grout is injected through the hollow
bar simultaneously with the drilling. The grout, which exits through ports located in a
sacrificial drill bit, fills the annulus from the top to the bottom of the drillhole. Rotary
percussive drilling techniques are used with this method. This soil nail type allows
for a faster installation than that for drilled grouted nails and, unlike, driven soil nails,
some level of corrosion protection with grout is provided. This system is most
commonly used for temporary nails.
• Jet-grouted soil nails: In this technique, jet grouting is performed to erode the
ground and allow the hole for the nail (subsequently installed) to be advanced to the
final location. The grout provides corrosion protection to the central bar. In a second
step, the bars are typically installed using vibro-percussion drilling methods. At this
time, this method is not recommended for soil nail installations for permanent soil
nail walls.
• Launched soil nails: In this method, bare bars are “launched” into the soil at very
high speeds using a firing mechanism involving compressed air. Bars are ¾ to 1 in.
in diameter and up to 25 ft in length. This technique allows for a fast installation with
little impact to project site; however, it may be difficult to control the length of nail
that penetrates the ground. These types of soil nails are only used for temporary
nails. At this time, this method is not recommended for soil nail installations for
permanent soil nail walls.
The remainder of this section presents a description of the main components of a typical soil
nail used in the U.S. practice (Figure 9-4).
BEARING PLATE
STEEL BAR
BEARING NUT AND
BEVELED WASHER
GROUT / SHOTCRETE CONTACT
HEADED STUD
(TYP)
~
REINFORCEMENT
CENTRALIZER DRILLHOLE
(TYP.) GROUT
WELDED
WIRE MESH
DRILL
TOTAL LENGT
NAIL B H
AR LE
NGTH
Figure 9-4. Main Components of a Typical Soil Nail (after Porterfield et al., 1994).
Nail Bars
Steel reinforcing bars used for soil nails are commonly threaded and may be either solid or
hollow. Bars generally have a nominal tensile strength of 60 ksi (Grade 60) 75 ksi (or Grade
75). Common U.S. practice of soil nailing involves the use of solid steel bars of Grade 60 or
75. All steel bars must be continuous without splices or welds, straight and undamaged.
They can be bare or epoxy coated or encapsulated, as required based on corrosion protection
considerations.
Bars with a tensile strength of Grade 95 and as high as Grade 150 may be considered for soil
nailing. Bars with lower grades are preferred because they are more ductile, less susceptible
to corrosion, and readily available. Grade 150 bars should not be used in conventional soil
Threaded bars for typical soil nail wall applications are available in No. 6, 7, 8, 9, 10, 11, and
14 up to approximately 60 ft in length. Bars having diameters smaller than No. 8 should not
be used or used with great care in applications where long bars are required (e.g., high walls)
because they tend to bend excessively during handling and installation. If needed, couplers
can be used to extend the length of bars in excess of 60 ft; however, soil nails in excess of
this length are typically not required for most highway projects.
As with solid bar nails, steel for hollow-core soil nails must meet the requirements of ASTM
A615. Grade 60 is minimum requirement, however, higher grade steels are available and
commonly used. Currently, two types of hollow core nails are available, the primary
difference between the two being the type of thread on the exterior of the nail. One thread
type is the “rope” thread or the R-thread and is the standard thread for continuously threaded
drill rods. The other thread type is similar to that used on continuously threaded deformed
reinforcement bars. Additional information on hollow core bars is provided in Samtani and
Nowatzki (2006).
Nail Head
The nail head comprises two main components: the bearing-plate, hex nut, and washers; and
the headed-stud (Figure 9-4). The bearing plate is made of Grade 36 (ASTM A36) steel and
is typically square, 8- to 10-in. side dimension and ¾-in. thick. The purpose of the bearing
plate is to distribute the force at the nail end to the temporary shotcrete facing and the ground
behind the facing. The bearing plate has a central hole, which is inserted over the nail bar.
Beveled washers are then placed and the nail bar is secured with a hex nut or with a spherical
seat nut. Washers and nuts are steel with a grade consistent with that of the nail bar
commonly of Grade 60 or 75. Nuts are tightened with a hand-wrench. The head-stud
connection may consist of four headed studs that are welded near the four corners of the
bearing plate to provide anchorage of the nail head into the permanent facing. For temporary
walls, the bearing plate is on the outside face of the shotcrete facing.
Grout for soil nails is commonly a neat cement grout, which fills the annular space between
the nail bar and the surrounding ground. In ground with potential for drillhole caving, a neat
cement grout is always used. Sand-cement grout can also be used in conjunction with open
hole drilling (i.e., for non-caving conditions) for economic reasons. Cement types I, II, III, or
V conforming to ASTM C 150 can be used. Cement Type I (normal) is recommended for
most applications. Cement Type III is grounded finer, hardens faster, and can be used when
a target grout strength is required to be achieved faster than for typical project conditions.
Cement Type II hardens at a slower rate, produces less heat, and is more resistant to the
corrosive action of sulfates than Cement Type I.
The water/cement ratio for grout used in soil nailing applications typically ranges from 0.4 to
0.5. In some cases, a stiffer grout with a slump on the order of 1½ in. may be used. The
need for a stiffer grout may arise when the hollow-stem auger drilling method is used or it is
desired to control leakage of grout into highly permeable granular soils or highly fractured
rock. Occasionally, the stiff consistency of the grout may cause difficulties with the
installation of the centralizers. In this case, the grout itself may provide sufficient support to
centralize the nail bar within the drillhole. Regardless of the ability of the stiff grout to
support the nail bar, centralizers should always be used to assure that a minimum grout cover
around the nail bar is achieved.
The characteristics of the grout have a strong influence on the ultimate bond strength at the
grout-ground interface. The grout should have a minimum 3-day compressive strength of
1,500 psi and a minimum 28-day unconfined compressive strength of 3,000 psi in accordance
with AASHTO T106 or ASTM C109. Admixtures are not typically required for most
applications, but plasticizers can be used to improve grout workability for projects located in
high-temperature climates or where project constraints dictate that the grout must be pumped
over long distances. Typically, the improved workability of grout due to plasticizers can be
extended up to approximately one hour. The use of air entrainment agents can improve
workability and reduce cracking potential, but they cause the grout to develop a more open
matrix and lose some of the chemical corrosion protection provided by cement. Therefore,
its use should be approved only when other corrosion protection methods, other than grout
cover, are present, or the thickness of the grout cover is increased. Some proprietary grouts
contain chemicals that provide zero-volume shrinkage, which is desirable to minimize
cracking and enhance bond strength. Where admixtures are being considered for use, tests
should be performed to verify that the grout and bond properties of the grout are not
adversely affected.
Grout is pumped shortly after the nail bar is placed in the drillhole to reduce the potential for
hole squeezing or caving. In solid nail bar applications, the grout is injected by tremie
methods through a grout pipe, which is previously inserted to the bottom of the drillhole,
until the grout completely fills the drillhole (Figure 9-5). The grout pipe typically consists of
heavy-duty plastic tubing varying between 3/8-in. and 3/4-in. outside diameter (OD). Grout
pipes are removed when used as part of the installation of production nails and commonly
left in place when used for soil nails that are to be load tested. Grout injection must be
conducted smoothly and continuously in such a way that the space between the drillhole and
the nail bar is filled completely, with no voids or gaps. The bottom of the grout pipe must
remain below the grout surface at all times while grout is being pumped into the drillhole.
During grouting operations, the portion of the soil nail near the back of the temporary facing
may not be completely filled with grout. Because this area is the most vulnerable to
corrosion, it is critical that this area be subsequently filled with shotcrete, or less commonly
with a stiff grout to assure complete grout coverage.
Centralizers
Centralizers are devices made of polyvinyl chloride (PVC) or other synthetic materials that
are installed at various locations along the length of each nail bar to ensure that a minimum
thickness of grout completely covers the nail bar (Figure 9-6). They are installed at regular
intervals, typically not exceeding 8 ft, along the length of the nail and at a distance of about
1.5 ft from each end of the nail.
Wall Facing
Nails are connected at the excavation surface (or slope face) to a facing system, which most
commonly consists of a first-stage, temporary facing of shotcrete during construction and, a
second-stage, permanent facing of CIP concrete. The purpose of the temporary facing is to
support the soil exposed between the nails during excavation, provide initial connection
among nails, and provide protection against erosion and sloughing of the soil at the
excavation face. The purpose of the permanent facing is to provide connection among nails,
a more resistant erosion protection, and an aesthetic finish. Temporary facing typically
consists of shotcrete with welded wire mesh (WWM) and additional shorter reinforcement
bars (referred to as waler bars) around the nail heads, which are applied after each row of
nails is installed in the ground. Permanent facing is commonly constructed of CIP reinforced
concrete and WWM-reinforced shotcrete. Prefabricated panels may also be used to construct
the permanent wall facing, especially for projects with special aesthetic requirements or
where prefabricated panels are more cost-effective.
Groundwater is a major concern in both the construction of soil nail retaining walls and in
their long term performance. Soil nail walls are best suited to applications above the water
table. Excess water at the face can result in face stability problems during construction
together with an inability to apply a satisfactory shotcrete construction facing. In addition,
long-term face drainage is required to prevent the generation of localized high groundwater
pressures on the facing.
A commonly adopted design for controlling surface runoff consists of a surface interceptor
ditch, excavated along the crest of the excavation and lined with concrete applied during the
shotcreting of the first excavation lift. The ditch should be contoured to drain away from the
working area, with collector drain pipes installed at appropriate locations, if necessary.
Where larger graded slope areas exist above the wall, installation of plastic film slope
protection sheeting above the interceptor ditch provides another quick and inexpensive
means of controlling surface water during construction. Similar permanent surface drainage
measures are generally required to prevent surface waters from infiltrating behind the facing,
or flowing over the top of the wall, during the operational life of the structure. For stepped or
benched walls vegetation can also be used to inhibit infiltration and lower soil water contents
by evapotranspiration.
• Shallow Drains (Weep Holes): These are typically 16-in. long, 2- to 4-in.
diameter PVC pipes discharging through the face and located where heavier
seepage is encountered.
Excavation
Prior to any excavation, surface water controls should be constructed to prevent surface water
from flowing into the excavation as this condition will adversely affect construction and
potentially cause instability of the excavated face. Collector trenches behind the limits of the
excavation are used to intercept and divert surface water. Subsequently, soil excavation is
performed using conventional earth-moving equipment from a platform, and final trimming
of the excavation face is typically carried out using a backhoe or excavator from a platform.
The initial lift is typically 3 to 4 feet high (Figure 9-7). The excavated face profile should be
reasonably smooth and not too irregular to minimize excessive shotcrete quantities. Soil
profiles containing cobbles and/or boulders may require hand excavation. A level working
bench on the order of 30-ft wide is required to accommodate the conventional drilling
equipment used for nail installation. Track drills smaller than the conventional drilling
equipment can work on benches as narrow as 15 ft and with headroom clearance as low as 9
ft.
For cases where the excavated slope face cannot stand unsupported for the required period of
time, a continuous berm may be employed to stabilize the unsupported face section (Figure
9-8a). In this case the soil nails are installed and grouted first through the stabilizing berm.
Subsequently, the berm is excavated and shotcrete is applied along the entire excavation
level.
construction costs and project schedule. If temporary excavation stand up time is a major
concern and represents a potential risk, alternative top-down construction methods such as
ground anchors may be considered.
Nail holes are drilled at predetermined locations using one of several available drilling
methods, including rotary (Figure 9-9), percussion, auger, and rotary/percussion drilling.
Open-hole installation using auger drilling (in particular, hollow-stem augers) is most
commonly used on soil nailing projects in the U.S. because no casing of the drillhole is
necessary, high installation rates can be obtained, and costs are relatively low. Nail holes
drilled using auger drilling can range between 4 to 12 in. in diameter. More commonly,
drillholes are 6 to 8 in. in diameter. Contractors will usually select a relatively large drillhole
diameter (e.g., 8 in.) to reasonably assure that the ultimate soil nail bond strength required in
the construction specifications can be achieved without difficulties minimizing drilling
equipment costs.
SOIL NAIL
NAIL DRILLHOLE
(a)
SH
SV
SV
WWM
STABILIZING BERMS
MAX: SH (APPROX)
(b)
Figure 9-8. Examples of Alternative Temporary Excavation Support (a) Stabilizing Berm
and (b) Slot Excavation (after Porterfield et al., 1994).
The selection of the drilling method may be controlled by the local availability of equipment
and the specific ground conditions to be encountered. Typical soil nail wall contract
documents allow the contractor to select the drilling method. However, the design engineer
may occasionally restrict the choice of drilling methods and/or procedures based on the
subsurface conditions or other project needs.
The most common practice for placing nail bars is inserting them into a predrilled, straight-
shafted drillhole. After the nail bar is inserted in the drillhole, the drillhole is filled with
clean cement grout, as discussed previously. This method is referred to as open-hole
installation. As the grout sets, it bonds to the nail bar and the surrounding ground. The
open-hole installation is by far the most commonly used method in soil nail wall
construction.
The most common U.S. practice of grouting by gravity provides bond strengths that are
sufficient for soil nailing to be a feasible and cost-effective solution. However, in cases
where poor soil conditions are encountered, higher bond strengths might be required. High
bond strength may be achieved in granular soils and weak fissured rocks by injecting grout or
Drainage Installation
The groundwater collected at strip drains is removed by a series of footing drains at the
bottom of the excavation. The footing drain consists of a trench at the bottom of the
excavation, which is filled with aggregate free of fines and has a PVC slotted collection pipe.
The drainage geotextile must envelope the footing drain aggregate and pipe and conform to
the dimensions of the trench. Additionally, weep holes can be installed through the wall
facing at the lower portions of the wall. In special situation when the groundwater behind the
proposed soil nail wall is high, conventional, deeper horizontal pipe drains are necessary.
Temporary wall facings for soil nail wall applications are usually constructed using shotcrete.
The thickness of the temporary shotcrete facing is typically between 3 and 4 in. It is noted
that this thickness range is based on the typical soil nail spacings described herein. Wider
nail spacings would require the use of a thicker shotcrete facing. Shotcrete provides a
continuous supporting layer over the excavated face that can also serve to fill voids and
cracks on the excavated face. Temporary shotcrete applications have been constructed using
both WWM or fiber reinforcement and bars. WWM is the preferred method among
contractors because it requires less time to install while the excavated face is unsupported. A
shotcrete facing for a wall under construction is shown in Figure 9-10.
Two types of shotcrete methods are commonly used: dry mix and wet mix. In the dry mix
method, the aggregate and cement are blended in the dry and fed into the shotcrete gun while
the mix water is added at the nozzle. Admixtures can be added at the mix plant or with the
water. The addition of water at the nozzle allows the plasticity of the shotcrete to be adjusted
at the nozzle, if required. In the wet mix method, the aggregate, cement, water, and
admixtures are mixed in a batch plant and conveyed to the nozzle by a hydraulic pump. The
plastic mix is applied at higher velocities by compressed air.
Both shotcrete methods produce a mix suitable for wall facings. Dry mix and wet mix
shotcrete use a water-cement ratio of about 0.4 and produce roughly the same mix quality,
although shotcrete obtained with the wet mix process yields a slightly greater flexural
strength. Keeping water cement ratios at about 0.4 and using air entrainment, which is
difficult with the dry-mix process, enhances the durability of shotcrete. Low water-cement
ratios (i.e., < 0.45) result in high strength, high durability, and low permeability as long as
proper in-situ compaction (i.e., elimination of entrapped air in the shotcrete) is achieved.
Steel fiber reinforcement has been added to shotcrete as part of a wet mix to increase
ductility, toughness, and impact resistance. Fibers tend to reduce the shotcrete brittleness and
thereby reduce crack propagation, but they have little effect on compressive strength and
produce only a modest increase in flexural strength. Wet mix is often preferred for the
construction of shotcrete facing walls because:
• the shotcrete rebound (i.e., loss of material due to lack of “stick”) for a wet mix is
typically only about 5 percent, compared to 15 percent for a dry mix;
• there is no need to add water at the nozzle, as in the case of a dry mix, thus it is less
dependent on the nozzle operator’s experience;
• equipment (e.g., concrete pump) is more readily available because shotcrete gun and
moisturizer are not needed as with dry mix; and
• supply of ready-mix concrete from commercial batch plants is readily available and
convenient.
Welded wire mesh is commonly used as reinforcement for temporary facing but occasionally
is also used in permanent facing. The cross-sectional area and mesh opening of the WWM
are selected to satisfy structural requirements (i.e., flexural and punching shear capacities)
and constructibility constraints. The selected WWM must have a width that is consistent
with the excavation lift height (equivalent to the vertical nail spacing), plus an overlap of at
least 8 in. The dimensions of the WWM (i.e., bar size and spacing) are evaluated as part of
the soil nail wall design.
Several methods to provide a permanent facing have been used by contractors. These
methods include reinforced shotcrete, cast-in-place (CIP) reinforced concrete, and precast
concrete.
The reinforcement of permanent facing using CIP concrete typically consists of a mesh
(standard reinforcing bars) and occasionally waler bars placed over the nail head.
Reinforcement is placed approximately at the center of the facing section thickness. Because
bonding between temporary shotcrete and permanent CIP concrete facings cannot be assured,
the temporary facing is typically disregarded as a resisting element in the section design. A
variety of finishes can be implemented by using commercially available form liners, as
shown in Figure 9-12.
The advantage of the CIP reinforced concrete is that the finish is more aesthetically pleasing
and the quality of the concrete tends to be more homogeneous. The main disadvantage is the
need for formwork and potentially longer construction time for facing installation. The use
of reinforced shotcrete as permanent facing has the potential benefit of cost savings and
efficiency as the same shotcrete equipment used for the temporary facing can be utilized. A
major limitation of this technique is that the conventional finish of shotcreted walls is
typically relatively rough and may not meet aesthetic requirements for a finished wall face.
Precast concrete facing has been used in permanent applications to meet a variety of
aesthetic, environmental, and durability criteria. A project using this method is shown in
Figure 9-13. Precast facings also provide a means of integrating a continuous drainage
blanket behind the facing and a frost protection barrier in cold climates.
Subsurface conditions exhibiting high corrosion potential usually do not preclude the use of
soil nails, providing the design life, type of structure, and proper corrosion protection for the
soil nail bars are properly considered. Various ground conditions promote corrosion
including: (1) low electrical resistivity of soil; (2) high concentration of chlorides or sulfates;
and (3) too low or too high hydrogen potential (pH) of soil or groundwater. Examples of
soils with corrosion potential include: (1) acidic soils; (2) organic soil; and (3) soils with
materials of industrial origin (slag, fly ash, fills with construction debris, mine tailings, and
acid mine waste).
The test-based criteria listed in Table 9-1 are used to classify the corrosion potential of the
ground. The ground is classified with a strong corrosion potential or aggressive if any one of
the conditions listed in the first column of Table 9-1 exceeds the limits listed in the second
column of the table during the service life of the soil nail wall, otherwise, the ground is
classified as non-aggressive.
Grout Protection
This method of corrosion protection involves fully covering the bar with neat cement grout.
After the bar is centered in the drillhole, neat grout is injected and fills up the annular space
around the steel bar. Grout encapsulation provides both physical and chemical corrosion
protection. When a minimum grout cover is in place, components such as carbonates and
chlorides in the soil, and oxygen and humidity in the air are prevented or delayed in reaching
the bar due to passivation. Additionally, the grout must have low permeability to ensure the
effectiveness of the encapsulation. The grout provides an alkaline environment that reduces
corrosion potential. A minimum grout cover of 1 in. between the bar and the soil should be
specified.
Corrosion protection with epoxy (Figure 9-14) consists of coating the nail bar with a fusion-
bonded epoxy that is applied by the manufacturer prior to shipment to the construction site.
Cement grout is placed around all epoxy-coated nail bars. The minimum required thickness
of epoxy coatings is 16 mils. The epoxy coating provides physical and chemical protection,
as epoxy is a dielectric material. In transporting and handling bars, the epoxy coating may be
damaged before nail installation. Therefore, it is not uncommon to spray epoxy coating in
the field on chipped or nicked surfaces. Applicable standards for epoxy coating are found in
ASTM A-775.
Encapsulation
TEMPORARY FACING
CENTRALIZER
2 in.
50 mm (2 in.) 1 in.
25 mm (1 in.)
MINIMUM COVER MINIMUM COVER
6 in.(6 in.)
150 mm
MINIMUM
NAIL GROUT
2.5 8mft
(8 ft)
MAXIM
UM 0.5 1m
.5(1
ft.5
MAXIM ft)
UM
FACING REINFORCEMENT
NOT SHOWN
sheathing is corrugated to transfer the effect of anchorage to the surrounding grout. Grout
must completely fill the annular spaces inside and outside the sheathing. The minimum grout
cover between the sheathing and the nail bar is 0.4 in. This distance allows the injected grout
to flow without difficulty and provides sufficient physical protection. Outside the sheathing,
the minimum grout cover between the sheathing and the drillhole wall must be 0.8 in.
In some systems, the inner annular space is grouted in the shop and the whole assembly
transported to the project site. The sheathing must be sufficiently strong to resist
transportation, handling, and installation. Additionally, sheathing must be non-reactive with
concrete, chemically stable, ultra-violet-light resistant, and impermeable. The minimum
sheathing wall thickness is typically 0.04 in. Certain sheathing techniques may be
proprietary.
The use of materials made of galvanized steel and a minimum cover of 2 in. of concrete or
permanent shotcrete provide corrosion protection of bearing plates, washers, and nuts.
Epoxy coating can be applied on bearing plates and nuts.
TEMPORARY FACING
6 in.
150 mm (6 in.)
MINIMUM
2.5 m
8 ft
(8 ft)
MAXIM
UM 0.5 1m
.5(1ft.5
MAXIM ft)
BAR UM
0.4(0.4
10 mm in. in.) INNER GROUT
MINIMUM COVER
PVC SHEATHING
FACING REINFORCEMENT
NOT SHOWN OUTER GROUT
Table 9-2 presents the protection levels and the protection systems commonly used in soil
nail applications.
Table 9-2. Recommendations for Minimum Levels of Corrosion Protection for Soil Nails.
Note 1: Since temporary soil nails are oftentimes used for temporary support of excavation
projects in which design may be the responsibility of the Contractor, it is
recommended that the specific corrosion protection system be evaluated on project-
specific basis. If the temporary condition can be assured to be of a controlled
duration (say less than 18 months) and if the incremental costs associated with
encapsulation are prohibitive, the Owner may wish to consider, as a minimum,
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-33 June 2008
either grout and epoxy coating or grout and galvanization. If there is a potential for
the service life of the soil nails to extend beyond 18 months and/or the incremental
costs of providing encapsulation are not prohibitive, then the additional protection
offered by encapsulation may be considered.
When using corrosion protection measures shown in Table 9-2, it is not necessary to
incorporate a sacrificial thickness into the design. In general, current U.S. public-sector
projects do not use the approach of sacrificial thickness as a means to address potential
corrosion for soil nail applications, especially for permanent applications. In temporary
applications, unprotected, bare bars can be driven, as long as the soil corrosion potential is
mild or insignificant. A preliminary and safe (for most conditions) estimate of the required
sacrificial total thickness for unprotected bars is 0.08 in over the entire surface area.
9.2.5 Cost
The cost range for walls up to 60 ft in height is typically $15 to $35/ft2 of exposed wall face.
A typical cost of a soil nail is $35/linear ft, which is less than the per unit length cost of a
ground anchor. Reasons for the cost difference include:
• soil nails are usually not as long as ground anchors because soil nails develop load
capacity starting right behind the wall whereas ground anchors develop load in a bond
zone which may be located at a relatively large distance behind the wall;
• soil nails are usually smaller in diameter than ground anchors; and
• load testing is only performed on a small number of soil nails (typically 5 percent of
the total production soil nails) compared to all of the ground anchors.
Soil nail walls are also less expensive than anchored walls because, unlike anchored walls,
soil nail walls do not need a vertical structural wall element.
The cost range provided above does not include the cost for the permanent wall facing.
Typically, permanent concrete wall facing can add $20 to $30/ft2 to the cost of a soil nail
wall.
If casing of the drillhole is required for the proper installation of the soil nails, the costs for
the soil nail wall system can increase significantly.
The following wall response and load transfer mechanisms take place during a conventional
soil nail construction:
• Soil excavation is initiated from the ground surface and the Excavation Phase 1 is
completed (Figure 9-16). Because of the soil ability to stand unsupported, the upper
portion of the soil behind the excavation is stable (or at least marginally stable) before
the first row of nails (Nails 1) is installed. Soil strength is mobilized along the
uppermost potential critical failure surface to allow the unsupported soil wall to stand.
• As Nails 1 and the temporary facing are installed, some load derived from the
deformation of the upper soil is transferred to these nails through shear stresses along
the nails and translate into and axial forces. The top portion of Figure 9-16 shows
schematically the axial force distribution in Nails 1 at the end of excavation Phase 1.
At this point, the temporary facing supports the excavation surface and provides
connectivity between adjacent nails in row of Nails 1.
• Nails 2 are then installed. Subsequently the temporary facing between the bottom of
excavation Phases 1 and 2 is installed and integrated to the facing constructed in
Phase 1. Subsequent movements of the soil above the Phase 2 depth will cause
additional loads to be transferred to Nails 1 and generate loads in Nails 2. Note the
increased nail force distribution for Nails 1 at the end of excavation Phase 2.
• To provide global stability, the soil nails must extend beyond the potential failure
surface. As lateral deformation increases due to subsequent excavation, additional
shear stresses along the soil nail/soil interface and axial forces of the previously
installed nails are mobilized. As the depth of excavation increases, the size of the
retained soil mass increases, as shown in Figure 9-16.
• As the size of the retained zone increases, the stresses at the soil/nail interface and the
axial forces in the nails increase. The induced tensile stresses are transferred behind
the retained zone in an anchorage effect. These stresses ultimately tend to stabilize
the potentially sliding mass.
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-35 June 2008
Schematic Distribution of Nail
Axial Force, T, in
Nail 1 after each excavation phase
Deflection
pattern at end
of each phase
N 2 1
Excavation Phase 1
Nail 1
Excavation Phase N
Nail N
Figure 9-16. Potential Failure Surfaces and Soil Nail Tensile Forces (after Lazarte et al.,
2003).
• While the tensile force in the intermediate and lower nails may increase as the
excavation depth increases, the tensile force in some of the upper nails may decrease
due to load redistribution. For example, the upper portion of Figure 9-16 shows
schematically that the axial force distribution for Nails 1 at the end of the last
excavation Phase N does not exhibit the largest values.
• As the critical failure surface becomes deeper and larger, the contribution of the upper
nails to the stabilization of this larger sliding mass diminishes. In some cases, upper
nails may be entirely ineffective in the assessment of deep critical failure surfaces.
However, the upper nails should not be considered superfluous, because they
contribute to the stability during earlier stages of excavation and help reduce lateral
displacements.
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-36 June 2008
The analysis of soil nail walls must consider both “during construction” and “post
construction” loading conditions to establish the most critical case at each soil nail level. The
most critical situation may arise after the wall is completed due to a combination of long-
term design loads (e.g., dead load, live load, and traffic) and extreme loads (e.g., earthquake).
In other situations, the most critical case may occur during construction when the then lowest
excavation surface remains temporarily unsupported and the soil nails and shotcrete are not
yet installed (Figure 9-17). These critical short-term loading conditions can be exacerbated
by temporary seepage conditions.
Figure 9-17. Potential Critical Stability During Construction (after Lazarte et al., 2003).
General
External failure modes refer to the development of potential failure surfaces passing through
or behind the soil nails (i.e., failure surfaces that may or may not intersect the nails). For
external failure modes, the soil nail wall mass is generally treated as a block. Stability
calculations take into account the resisting soil forces acting along the failure surfaces to
establish the equilibrium of this block. If the failure surface intersects one or more soil nails,
the intersected nails contribute to the stability of the block by providing an external
stabilizing force that must be added to the soil resisting forces along the failure surface.
SOIL
STRENGTH
SOIL
NAIL
STRENGTH
RESISTANCE
HEAVE
FAILURE SOIL STRENGTH
SURFACE AT BASE
GROUT BAR
BREAKAGE V
FAILURE M
SURFACE
M = Moment
V = Shear
FAILURE HEADED-STUD
SURFACE BREAKAGE
PLASTIC
MOMENT
Figure 9-18. Principal Modes of Failure of Soil Nail Wall Systems (after Lazarte et al.,
2003).
Global stability refers to the overall stability of the reinforced soil nail wall mass. As shown
in Figure 9-17, the slip surface passes behind and beneath the soil nail wall system. In this
failure mode, the retained mass exceeds the resistance provided by the soil along the slip
surface and the nails, if intersected. The global stability of soil nail walls is commonly
evaluated using two-dimensional limit-equilibrium principles, which are used for
conventional slope stability analyses. As with traditional slope stability analyses, various
potential failure surfaces are evaluated until the most critical surface (i.e., the one
corresponding to the lowest factor of safety) is obtained.
To illustrate the elements of a global stability analysis for soil nail walls, a simple, single-
wedge failure mechanism is shown in Figure 9-19.
i
W
S = R c + Rφ
H T
α
φ' N
LS
Figure 9-19. Global Stability Analysis of Soil Nail Wall using a Single-Wedge Failure
Mechanism (after Lazarte et al., 2003).
∑ resisting forces
FSG = (9-1)
∑ driving forces
where:
S = R c + R f = c * L s + N tan φ * (9-4)
tanφ'
tanφ * = (9-5)
FSG
where:
α = wall face batter angle (from vertical);
β = slope angle;
φ’ = soil effective angle of internal friction;
c’ = soil effective cohesion;
ψ = inclination of failure plane;
i = nail inclination;
Ls = length of failure plane;
W = weight of sliding mass;
Q = surcharge load;
T = equivalent nail force;
N = normal force on failure surface;
S = shear force on failure surface;
Rc = cohesive component of S; and
Rφ = frictional component of S.
and φ* is the mobilized friction angle, and c* is the mobilized cohesion. A single global
factor of safety is used for the cohesive and frictional strength components of the soil (c’ and
tanφ’, respectively). However, it is possible to select different safety factors for each strength
component.
The simplistic analysis presented above only considers force equilibrium. More rigorous
limit equilibrium slope stability analysis methods allow establishing simultaneously moment
and force equilibrium equations. The simple model shown in Figure 9-19 and presented in
the above equations may be used to perform an independent verification of the computer’s
solution.
A global stability analysis can be used to complete either (or both) of the following two tasks
related to the analysis of soil nail walls:
1) calculate the critical (minimum) factor of safety FSG of the sliding mass for a given
soil nail length pattern; or
2) determine the required force T in all nails that will yield a selected target factor of
safety against global failure.
Global stability analyses are performed using computer programs specifically developed for
the design of soil nail walls. The two computer programs most commonly used in the United
States for the analysis and design of soil nail walls are SNAIL and GOLDNAIL. These
programs can consider failure surfaces that are more complex than the simple planar, single-
wedge. SNAIL uses two-part planar wedges; GOLDNAIL uses circular failure surfaces that
consider multiple slices in lieu of wedges. These programs are similar, in many respects, to
general slope stability computer programs (e.g., search routines, closed form force/moment
equilibrium equations, etc.). However, computer programs dedicated to soil nail design
include the iterative and interactive design of the soil nail length and the consideration of
other failure modes (e.g., soil nail tensile force and facing punching shear failure).
Sliding
Sliding stability analysis considers the ability of the soil nail wall to resist sliding along the
base of the retained system in response to lateral earth pressures behind the soil nails.
Sliding failure may occur when additional lateral earth pressures, mobilized by the
Concepts similar to those used to assess sliding stability of gravity retaining structures (in
which Rankine or Coulomb theories of lateral earth pressures are used) can be applied to
assess the sliding stability of a soil nail wall system. Sliding stability analyses are presented
in Section 9.2.7.4 in subsection titled “Sliding Stability”.
Bearing Capacity
Bearing capacity may be a concern when a soil nail wall is excavated in fine-grained, soft
soils. Because the wall facing does not extend below the bottom of the excavation (unlike
soldier piles in cantilever or ground anchor walls), the unbalanced load caused by the
excavation may cause the bottom the excavation to heave resulting in a bearing capacity
failure of the foundation (Figure 9-21). Bearing capacity analyses are presented in Section
9.2.7.4 in subsection titled “Bearing Capacity”.
2H
γ, φ', c' = 0
W
α PA
H1
δ = βeq
H
H/3
θ
B Strength parameters
cb and φ'b
ΣR
Figure 9-20. Sliding Stability of a Soil Nail Wall (after Lazarte et al., 2003).
HB' HB'
H
Su H H Su H soft
fine-grained
soil
soft D
fine-grained B/ 2
soil
Failure surface
a) Deep deposit of soft fine-grained soil b) Shallow deposit of soft fine-grained soil
underlain by stiff layer
10
B/L = 1, Square
9
and Circular
B/L = 0.5 H = Excavation depth
8 B = Excavation width
L = Excavation Length
NC 7
6
B/L = 0, Rectangular
4
0 1 2 3 4 5
H/B
c) Bearing Capacity Factor, NC
Figure 9-21. Bearing Capacity (Heave) Analysis (after Terzaghi et al., 1996).
Factors of safety against heave for soil nail walls should be selected to be consistent with
those typically used for heave analysis at the bottom of excavations. In general, minimum
FSH can be adopted as 2.5 and 3 for temporary and permanent walls, respectively. As the
great majority of soil nail walls are not constructed in soft fine-grained soils, this failure
mode is not critical for most soil nail projects.
General
Internal failure modes refer to failure in the load transfer mechanisms between the soil, the
nail, and the grout. Soil nails mobilize bond strength between the grout and the surrounding
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-43 June 2008
soil as the soil nail wall system deforms during excavation. The bond strength is mobilized
progressively along the entire soil nail with a certain distribution that is affected by numerous
factors. As the bond strength is mobilized, tensile forces in the nail are developed.
Depending on the soil nail tensile strength and length, and the bond strength, bond stress
distributions vary and different internal failure modes can be realized. Typical internal
failure modes related to the soil nail are (Figures 9-18 d–g):
• Nail Pullout Failure: Nail pullout failure is a failure along the soil-grout interface
due to insufficient intrinsic bond strength and/or insufficient nail length, Figure 9-
18d.
• Slippage of the Bar-Grout Interface: The strength against slippage along the grout
and steel bar interface (Figure 9-18e) is derived mainly from mechanical interlocking
of grout between the protrusions and “valleys” of the nail bar surface. Mechanical
interlocking provides significant resistance when threaded bars are used and is
negligible in smooth bars. The most common and recommended practice is the use of
threaded bars, which reduces the potential for slippage between the nail bar and grout.
• Tensile Failure of the Nail: The nail can fail in tension if there is inadequate tensile
strength, Figure 9-18f.
• Bending and Shear of the Nails: Due to relatively modest contribution resulting from
shear and bending stresses in nails, the shear and bending strengths of the soil nails
are conservatively disregarded in the guidelines contained in this document. The
demand in flexure and shear is minimal as well. A discussion of a methodology to
evaluate shear and bending behavior is included in Elias and Juran (1991).
A discussion of the two most common internal failure modes (i.e., nail pullout and nail
tensile failure) is presented in the following two sections.
Pullout failure is the primary internal failure mode in a soil nail wall. This failure mode may
occur when the pullout capacity per unit length is inadequate and/or the nail length is
insufficient. In general, the mobilized pullout per unit length, Q, (also called the load
transfer rate) can be expressed as:
Q = π q D DH (9-7)
q = mobilized shear stress acting around the perimeter of the nail-soil interface; and
DDH = average or effective diameter of the drill hole.
Actual distributions of mobilized bond shear stress (and load transfer rates) are not uniform,
as illustrated in Figure 9-22, and depend on various factors including nail length, magnitude
of applied tensile force, grout characteristics, and soil conditions. As a simplification, the
mobilized bond strength is often assumed to be constant along the nail, which results in a
constant load transfer rate, Q. As a result, the nail force at the end of the pullout length, Lp,
is:
T(L p ) = To = Q L p
(9-8)
The pullout capacity, Rp, is mobilized when the ultimate bond strength is achieved and is
expressed as:
R p = Tmax = Q u L p
(9-9)
with:
Q u = π q u D DH (9-10)
where:
Qu = pullout capacity per unit length (also referred to as load transfer rate capacity); and
qu = ultimate bond strength.
Typical values of ultimate bond strength for various soils and drilling methods are presented
in Table 9-3. The lower and upper bounds provided in Table 9-3 correspond approximately
to the least and most favorable conditions for a particular ground type and construction
method. These values inherently contain some level of conservatism and can be used as
preliminary values for design. It is common practice to require preproduction soil nail load
tests to verify the bond strengths included in the construction specifications and to establish
the minimum required nail length to support a specified nail design load.
dx
q
To (end load)
DDH x
q (x)=constant
Actual distribution
of nail tensile force
T (x)
To
Q
1
Lp
The following allowable values of the bond strength or pullout capacity per unit are used in
design:
qu
q ALL = (9-11)
FS p
Rp
R p ALL = (9-12)
FS P
Ultimate Bond
Material Construction Method Soil/Rock Type
Strength (psi)
Marl/limestone 43.5 – 58.0
Phyllite 14.5 – 43.5
Chalk 72.0 – 86.5
Soft dolomite 58.0 – 86.5
Fissured dolomite 86.5 – 144.5
Rock Rotary Drilled
Weathered sandstone 29.0 – 43.5
Weathered shale 14.5 – 21.5
Weathered schist 14.5 – 25.5
Basalt 72.0 – 86.5
Slate/Hard shale 43.5 – 58.0
Sand/gravel 14.5 – 26.0
Silty sand 14.5 – 21.5
Rotary Drilled Silt 8.5 -11.0
Piedmont residual 6.0 – 17.5
Fine colluvium 11.0 – 21.5
Sand/gravel
low
overburden 27.5 – 34.5
Cohesionless Soils Driven Casing high 40.5 – 62.5
overburden 55.0 – 69.0
Dense Moraine 14.5 – 26.0
Colluvium
Silty sand fill 3.0 – 6.0
Augered Silty fine sand 8.0 – 13.0
Silty clayey sand 8.5 – 20.5
Sand 55.0
Jet Grouted
Sand/gravel 101.0
Rotary Drilled Silty clay 5.0 – 7.0
Driven Casing Clayey silt 13.0 – 20.5
Loess 3.5 - 11
Fine-Grained Soils Soft clay 3.0 – 4.5
Augered Stiff clay 6.0 – 8.5
Stiff clayey silt 6.0 – 14.5
Calcareous sandy clay 13.0 – 20.5
where FSP is the factor of safety against pullout failure. In general, a minimum factor of
safety of 2 is recommended against pullout failure.
The soil-nail interaction that occurs behind the wall facing is complex. The loads applied to
the soil nails originate as reactions to the outward wall movement during excavation of the
soil in front of the wall, as discussed earlier. The portion of the nail behind the failure
Facing q(x)
DDH
(a)
L
x
q(x)
(b) (+)
(-)
T(x)
Tmax
(c)
To
The maximum nail tensile force in the nail bar nail does not necessarily occur at the point
where the nail crosses the failure surface. The mobilized shear stress along the grout-soil
interface, q, is not uniform and, in fact, changes from “positive” to “negative”, as shown in
Figure 9-23a and b. The schematic distribution of the tensile force (T) along the soil nail is
shown in Figure 9-23c.
For design, the tensile force distribution along the nail shown in Figure 9-23 can be
simplified as shown in Figure 9-24. The tensile force the nail increases at a constant slope Qu
(equal to the pullout capacity per unit length), reaches a maximum value, Tmax, and then
decreases at the rate Qu to the value To at the nail head. With reference to Figure 9-24, the
following three conditions related to the maximum tensile force are noted. The value Tmax is
bounded by three limiting conditions: the pullout capacity, RP, the tensile capacity, RT, and
the facing capacity, RF. If RP < RT and RF, pullout failure controls the value of Tmax. If RT
LP
To ~ 0.6-1.0 Tmax
Figure 9-24. Simplified Distribution of Nail Tensile Force (after Lazarte, 2003).
< RP and RF, tensile failure controls Tmax. Finally, if RF < RT and RP, failure of the facing
may control, depending on the ratio of To/Tmax.
To achieve a balanced design, all of the resisting components in a system should have
comparable margins of safety; no component should be significantly oversized or undersized.
In the case of nail tensile forces, a good design should balance the capacities of all resisting
elements; therefore, values of RP, RT, and RF should be reasonably similar.
To achieve a balanced design for all internal failure modes, the soil strength must be fully
mobilized consistently with the full mobilization of the nail tensile strength at the same time.
In other words, when FSG = 1.0 (full soil mobilization), the safety factor for the tensile
strength, is FST = 1.0 (full nail tensile mobilization). The nail tensile force for this condition
is the maximum design force in the nail (Tmax-s). It is intuitive that when the loads are kept
constant, the design force Tmax-s will increase when FSG > 1.0. This is caused because for FSG
The program SNAIL automatically reports the average nail tensile force, but not the
maximum tensile force corresponding to FSG = l. Thus, to estimate the maximum nail tensile
force for a FSG = l without performing an additional stability analysis, the following
simplified method can be used. This procedure is based on the fact that the ratio of the
maximum nail load calculated by SNAIL, Tmax, to the average nail load, Tavg, for FSG > 1, is
similar to the ratio of the maximum nail load for FSG = 1, Tmax-s, to the average nail load,
Tavg-s, for FSG = 1. Therefore, a good approximation of the maximum design nail load (Tmax-s)
can be obtained by the following relationship:
Tmax −s Tavg−s
=
Tmax Tavg
(9-13)
Tavg-s is the average design nail load and is reported by SNAIL in output files as the
“Maximum Average Reinforcement Working Force”. The design nail force Tmax-s is
compared to the tensile capacity of the nail, which is defined as follows.
A tensile failure of a soil nail takes place when the longitudinal force along the soil nail, Tmax-
s, is greater than the nail bar tensile capacity, which is defined as:
R T = Atf y (9-14)
where At is the nail bar cross sectional area and fy is the nail bar yield strength. The tensile
capacity provided by the grout is disregarded, due to the difference in stiffness (i.e., modulus
of elasticity) between the grout and the nail. To take into account uncertainties related to
material strength and applied loads, allowable values of the nail tensile capacity are used in
design as follows:
RT
R T ALL = (9-15)
FST
where FST is the factor of safety against soil nail tensile failure. In general, a minimum
factor of safety of 1.8 is adopted for static loads.
General
The most common potential failure modes at the facing-nail head connection are presented in
Figure 9-18 and are shown in detail in Figure 9-25 as:
• Flexure Failure: This is a failure mode due to excessive bending beyond the facing’s
flexural capacity. This failure mode should be considered separately for both
temporary and permanent facings.
• Punching Shear Failure: This failure mode occurs in the facing around the nails and
should be evaluated for both temporary and permanent facings.
• Headed-Stud Tensile Failure: This is a failure of the headed studs in tension. This
failure mode is only a concern for permanent facings.
For each of these failure modes, the nail head and facing must be designed to provide
capacity in excess of the maximum nail head tensile force (To) at the wall face. Appropriate
dimensions, strength, and reinforcement of the facing and suitable nail head hardware (e.g.,
bearing plate, nut, and headed studs) must be provided to achieve the design capacities with
adequate factors of safety for all potential failure modes.
The nail tensile force at the wall face, To, is less than or equal to the maximum nail tensile
force (Byrne et al., 1996). Discussion on measured nail tensile forces based on several
research studies is provided in GEC No. 7 (Lazarte et al., 2003). For design, the following is
recommended for the nail head tensile force:
where:
VERTICAL
MOMENT
mV
WWM
OR BAR
VERTICAL
WWM MOMENT
mV
BEARING PLATE
HEADED
STUD FLEXURE FAILURE
CONICAL
SURFACE
RFP/2 45°
(PUNCHING SHEAR
RESISTANCE)
TO
BEARING
PLATE
RFP/2
PUNCHING
SHEAR FAILURE
(TEMPORARY
FACING)
COMPOSITE
CONICAL
REINFORCEMENT SURFACE
RFP/2
RFP/2
RFH
HEADED STUD
PUNCHING
SHEAR FAILURE
(PERMANENT
FACING)
HEADED-STUD TENSILE FAILURE
The soil nail wall facing can be considered a continuous reinforced concrete slab where the
loading is the lateral earth pressure acting on the facing and the supports are the tensile forces
in the soil nails (Figure 9-26a and b). The loads from the lateral earth pressure and the
“reaction” in the soil nails induce flexural moments in the facing section. Positive moments
(i.e., tension on the outside of the section) are generated in the midspan between nails;
negative moments (i.e., tension on the inside of the section) are generated around the nails
(Figure 9-26b). If these moments are excessive, a flexural failure of the shotcrete may occur.
In theory, the soil pressure that causes facing failure (i.e., the critical yield line pattern) can
be applied to an influence area around the nail head, and a nail tensile force (“reaction”) is
obtained. This force is designated as the facing flexure capacity, RFF, and is related to the
flexural capacity per unit length of the facing. The flexural capacity per unit length of the
facing is the maximum resisting moment per unit length that can be mobilized in the facing
section. Based on yield-line theory concepts, RFF can be estimated as the minimum of:
S h[ft]
[ ]
R FF [kip] = 3.8 × CF × (a vn + a vm ) in 2 /ft × H × f y [ksi ] (9-17a)
Sv
S h[ft]
[ ]
R FF [kip] = 3.8 × CF × (a hn + a hm ) in 2 /ft × v
S
× f y [ksi ] (9-17b)
H
where:
CF = factor that considers the non-uniform soil pressures behind the facing (Byrne et al.,
1996);
BEARING PLATE
IDEALIZED
DEFLECTION
PATTERN AT
ULTIMATE
LOAD
TO
INITIAL
POSITION
HINGE
(TYP)
FRACTURES ON
INTERNAL FACE
A
(b) ULTIMATE DEFORMATION PROFILE
(a) IDEALIZED YIELD LINE PATTERN
SECTION A-A
TO PROGRESSIVE
CRACKING
To ultimate= FIRST
YIELD
RFF
DEFLECTION,
Figure 9-26. Progressive Flexural Failure in Wall Facings (after Lazarte, 2003).
avn = cross sectional area of reinforcement per unit width in the vertical direction at the
nail head;
avm = cross sectional area of reinforcement per unit width in the vertical direction at
midspan;
ahn = cross sectional area of reinforcement per unit width in the horizontal direction at the
nail head;
ahm = cross sectional area of reinforcement per unit width in the horizontal direction at
midspan;
The factor CF takes into account the non-uniform soil pressures behind the facing (Byrne et
al., 1998) and represents nominally the ratio of soil pressure behind the nail to soil pressure
in the midspan between nails. The soil pressure distribution behind the facing is generally
non-uniform. Soil pressure is affected by soil conditions and the facing stiffness, which in
turn affects the wall displacement. In the midspan between nails, the displacement of the
facing occurs outward and the lateral earth pressure is relatively low. Around the nail heads,
the soil pressure is larger than the soil pressure at midspan between nails.
The pressure distribution in the facing also depends on the stiffness of the facing. When the
facing is relatively thin (as with typical temporary facings), the facing stiffness is relatively
low, causing the facing to deform in the midspan sections. As a result, the soil pressure tends
to be relatively low in the midspan sections. When the facing is relatively thick, the facing
stiffness increases and the resulting wall deformations are smaller than would result from a
thin wall facing. As a result of the increased wall stiffness, the soil pressure is more uniform
throughout. Table 9-4 shows factors (CF) for typical facing thickness. For all permanent
facings and “thick” [≥ 8 in.] temporary facings, the soil pressure is assumed to be relatively
uniform.
In Equations 9-17a and 9-17b, it is assumed that the maximum moments in the facing are
around a horizontal axis and the design of reinforcement in the vertical direction is more
critical than the design of the horizontal reinforcement. In practice, the cross section area of
reinforcement in the horizontal direction is the same as for the vertical direction (i.e., ahm =
avn and ahm = avm); therefore, the most critical case is the one that gives the minimum of
SH/SV and SV/SH.
When the same nail spacing and reinforcement are used in the horizontal and vertical
directions, and Grade 60 is used, Equations 9-17a and 9-17b simplify as:
h
avm avn
Waler Bar (TYP) Waler Bar
SV ahn
WWM
(Temporary
ahm Facing)
h
d = 0.5 h
A
Section A-A
Figure 9-27. Geometry used in Flexural Failure Mode (after Lazarte, 2003).
Nominal Facing
Factor
Thickness
Type of Structure
CF
in.
4 2.0
Temporary 6 1.5
8 1.0
Permanent All 1.0
If (vertical) waler bars are used over the nail heads, the total reinforcement area per unit
length in the vertical direction can be calculated as:
A vw
a vn = a vm + (9-19)
SH
where Avw is the total cross sectional area of waler bars in the vertical direction. Similar
concepts can be applied along the horizontal direction. If rebar is used in permanent facings
instead of WWM, the total area of reinforcement must be converted to a per unit length basis
as:
A vm
a vm = (9-20)
SH
where Avm is the total cross sectional area of rebar reinforcement in the vertical direction (see
Figure 9-27).
Given the tensile force at the soil nail head, To, and the facing flexure capacity, the safety
factor against facing flexural failure can be defined.
R FF
FS FF = (9-21)
To
In general, a minimum factor of safety of 1.35 is adopted for static loads in temporary walls
and 1.5 for static loads in permanent walls.
As with other reinforced concrete structures, the quantity of reinforcement placed in the
facing of soil nail wall generally falls within prescribed limits. The amount of reinforcement
can be expressed as a reinforcement ratio:
a ij
ρ= 100 (9-22)
0.5 h
f c' [psi]
ρ min [%] = 0.24 (9-23)
f y [ksi]
f c' [psi] 90
ρ max [%] = 0.05 (9-24)
f y [ksi] 90 + f y [ksi]
Therefore, the placed reinforcement must be: ρmin ≤ ρ ≤ ρmax. In addition, the ratio of the
reinforcement in the nail and midspan zones should be less than 2.5 to ensure comparable
ratio of flexural capacities in these areas.
Punching shear failure of the facing can occur around the nail head and must be evaluated at:
As the nail head tensile force increases to a critical value, fractures can form a local failure
mechanism around the nail head. This results in a conical failure surface, as shown in Figure
9-28. This failure surface extends behind the bearing plate or headed studs and punches
through the facing at an inclination of about 45 degrees, as shown schematically in Figure 9-
28. The size of the cone depends on the facing thickness and the type of the nail-facing
connection (i.e., bearing-plate or headed-studs).
As is common for concrete structural slabs subjected to concentrated loads, the nail-head
capacity must be assessed in consideration of punching shear, RFP, and can be expressed as:
R FP = C P VF (9-25)
where VF is the punching shear force acting through the facing section and CP is a correction
factor that accounts for the contribution of the support capacity of the soil.
where:
D’C = effective diameter of conical failure surface at the center of section (i.e., an average
cylindrical failure surface is considered); and
hC = effective depth of conical surface.
The correction factor CP is used to take into account the effect of the soil pressure behind the
facing that acts to stabilize the cone. If no subgrade reaction is considered, CP = 1.0. When
the soil reaction is considered, CP can be as high as 1.15. For practical purposes, the
correction is usually omitted and this is considered as CP = 1.0.
These equations can be used for both temporary and permanent facing. However, the size of
the conical surface (values of D’C and hC) must be adjusted to consider the specific type of
facing. For the temporary facing, the dimensions of the bearing plate and facing thickness
must be considered. For the permanent facing, the dimensions of the headed-studs (or anchor
bolts) must be considered. Figure 9-29 shows details of a typical headed-stud connector.
hC = h (9-27b)
LBP
CONICAL
Shear Resistance
FAILURE RF/2 RF/2
SURFACE
h 45°
h/2 (TYP)
To IDEALIZED
SOIL REACTION
DDH
DC
D'C
SHS
COMPOSITE
CONICAL
SURFACE
RF/2
LS
45°
tP hC
(TYP)
h
h/2
To IDEALIZED
SOIL DEFLECTION
DDH
hC = LS - tS + tP (9-28b)
tH
LS
DS
Given the tensile force at the soil nail head, To and the punching shear capacity of the facing,
the safety factor against facing punching shear can be defined as:
R FP
FS FP = (9-29)
To
In general, a minimum factor of safety of 1.35 is adopted for static loads in temporary walls
and 1.5 for static loads in permanent walls.
The tensile capacity of the headed-studs (or anchor bolts) connectors providing anchorage of
the nail into the permanent facing must be verified, as shown in Figure 9-28. The nail head
capacity against tensile failure of the connectors, RHT, is computed as:
R HT = N A S f y (9-30)
Given the tensile force at the soil nail head To and the tensile capacity of the headed-studs,
the factor of safety against tensile failure of the headed-studs can be defined as:
R HT
FS HT = (9-31)
To
A minimum factor of safety of 1.8 is adopted for static loads in temporary walls and 2.0 for
static loads in permanent walls if steel A307 is used for the headed-stud connector. If steel
A325 is used for the headed-stud connector, a minimum factor of safety of 1.5 is adopted for
static loads in temporary walls and 1.7 for static loads in permanent walls (Byrne et al.,
1996). The headed studs may also exert excessive compressive stress on the concrete
bearing surface. The compression on the concrete behind the head of the headed-stud is
assured to be within tolerable limits if the following geometric constraints are met (ACI,
1998):
AH ≥ 2.5 AS (9-32)
where:
2
π 0.9743
AE = D E − (9-34)
4 n t
where:
Geocomposite Drain Strips. These elements are strips of synthetic material approximately
12 to 16 in. wide. They are placed in vertical strips against the excavation face along the
entire depth of the wall (Figure 9-30). The horizontal spacing is generally the same as the
nail horizontal spacing. The lower end of the strips discharges into a pipe drain that runs
along the base of the wall or through weep holes at the bottom of the wall. For highly
irregular excavation faces, the placement of prefabricated drain strips against the excavated
face is difficult and often impractical. In some cases, the prefabricated drain strips may be
sandwiched between the shotcrete construction facing and the permanent CIP facing, with
the drain placed over 2- to 3-in. diameter weep holes passing through the construction facing.
The design engineer needs to provide explicit construction and inspection guidance for this
type application, to assure that the performance of the drainage system is not impacted during
Groundwater
Table
Geodrain
Strips
Weephole
Drains
Toe
Drain
Figure 9-30. Drainage of Soil Nail Walls (after Lazarte et al., 2003).
Shallow Drains (Weep Holes). These are typically 12- to 16-in. long, 2- to 4-in. diameter
PVC pipes discharging through the face and located where localized seepage is encountered
or anticipated. Weep holes are also used as the terminating point of the vertical strip drains
to allow any collected water to pass through the wall.
Drain Pipes. Horizontal or slightly inclined drain pipes may be installed where it is
necessary to control the groundwater pressures imposed on the retained soil mass. Drain
pipes typically consist of 2-in. diameter PVC slotted or perforated tubes, inclined upward at 5
to 10 degrees to the horizontal. Drain pipes are typically longer than the length of the nails
and serve to prevent groundwater from being in contact with the nails or the soil nail wall
mass, as shown in Figure 9-31. The lengths of the drains depend on the application. To
provide drainage of shallow or perched groundwater occurring erratically close to the facing,
drain pipes with lengths varying from 1 to 1.5 ft, and in some cases, up to 3 ft can be
installed. They are installed at a density of approximately one drain per 100 square ft of face.
The PVC pipe should be slotted, as shown in Figure 9-31. Although drain pipes are typically
installed after nails are in place and the shotcrete is applied to avoid either grout or shotcrete
from entering the drain, they can be applied prior to shotcrete application. In this case, a plug
of dry-pack and temporary PVC caps must be used to prevent the shotcrete from coming into
the drain hole and obstructing the drain slots or perforations.
Permanent Surface Water Control. Permanent surface water control measures include
installing an interception ditch behind the wall to prevent surface water runoff from
infiltrating behind the wall or flowing over the wall edge. A vegetative protective cap may
be also be used to reduce or retard water infiltration into the soil.
Design Considerations
Drain pipes require long-term maintenance. Analysis of soil nail walls for long term
conditions may need to take into consideration the potential for clogging. Clogging of
horizontal drains and a corresponding increase in water pressure will reduce the factor of
safety against global stability and/or sliding, and may adversely impact the internal stability
by affecting soil/nail interaction.
During construction and after its completion, a soil nail wall and the soil behind it tend to
deform outwards. The outward movement is initiated by incremental rotation about the toe
of the wall, similar to the movement of a cantilever retaining wall. Most of the movement
occurs during or shortly after excavation of the soil in front of the wall. Post construction
deformation is related to stress relaxation and creep movement, which are caused by post-
construction moderate increases in tensile force in the soil nail described previously.
Maximum horizontal displacements occur at the top of the wall and decrease progressively
toward the toe of the wall. Vertical displacements (i.e., settlements) of the wall at the facing
are generally small, and are on the same order of magnitude as the horizontal movements at
the top of the wall. In general, horizontal and vertical displacements of the facing depend on
the following factors:
~
50 mm (22 in.)
in. MIN
DIA. SCH 40 PVC PIPE
10° - 15°
PROTECTIVE
PVC CAP
(NOTE 1)
ION
SECT
TED
SLOT
(2 ft)
0.6 m ED
m ORAT
50 m PERF
NON ECTION
.) S
(2 in
MIN
NOTES
1. PROTECTIVE CAP NEEDS TO BE REMOVED AFTER FINAL
SHOTCRETE IS APPLIED
10 ft
2. SPACING OF DRAINS IS TYPICALLY 3.3 m (10 ft)
GEOCOMPOSITE GEOCOMPOSITE
DRAIN STRIP CONTINUOUS DRAIN STRIP
SOIL NAIL WALL GRAVEL DRAIN
WEEP HOLE
WEEP HOLE GEOTEXTILE
Figure 9-31. Typical Drain Pipe Details to Provide Groundwater Control in Soil Nail Walls
(after Byrne et al., 1998).
Empirical data show that for soil nail walls utilizing a typical nail-length to wall-height ratio
between 0.7 and 1.0, negligible surcharge loading, and typical global factors of safety (FSG)
values of 1.5, the maximum long-term horizontal and vertical wall displacements at the top
of the wall, δh and δv, can be estimated as follows:
δ (9-35)
δh = h × H
H i
where:
(δh/H)I = is a ratio dependant on the soil conditions “i” indicated in the table below;
H = wall height.
The size of the zone of influence (Figure 9-32), where noticeable ground deformation may
take place, is defined by a horizontal distance behind the soil nail wall (DDEF) and can be
estimated with the following expression:
D DEF
= C (1 − tan α ) (9-36)
H
where:
The movements shown above are considered to be relatively small and comparable to those
obtained with braced systems and anchored walls. These estimates of deformations have
essentially become recommended design values. The adopted tolerable deformation criterion
is project-dependent and should consider not only the magnitude of deformation but also the
extent of the area behind the wall that may be affected by wall movements. As a first
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-67 June 2008
DDEF
EXISTING
STRUCTURE
h
DEFORMED
PATTERN
H INITIAL CONFIGURATION
SOIL NAIL
(TYP)
Figure 9-32. Deformation of Soil Nail Walls (after Byrne et al., 1996).
estimate, horizontal deflections greater than 0.005 H during construction should be a cause
for concern, as they generally represent an upper limit of acceptable performance.
When excessive deformations are considered to be likely with a certain wall configuration,
some modifications to the original design can be considered. Soil nail wall deformations can
be reduced by using a battered wall, installing longer nails in the top portion of the wall,
using a higher safety factor, or even using ground anchors in conjunction with the soil nails.
Post-construction monitoring of soil nail wall displacements indicates that movements tend
to continue after wall construction, sometimes up to 6 months, depending on ground type.
Typically, the post construction deformation increases up to 15 percent of the deformations
observed soon after construction. As a result of this movement, additional tension is
developed in the nails. In general, fine-grained soils of high-plasticity (i.e., approximately PI
> 20) and high water contents (such that LI > 0.2) tend to incur deformation for longer
periods of time.
Frost Protection
The formation of ice lenses in the vicinity of the soil nail wall facing in frost-susceptible soils
may lead to the development of high loads on both the facing and the head of the nail. This
phenomenon may result in damage to the facing. In situations where the facing is designed
to resist frost damage, the nail or to the connection between the nail and the facing can still
be impacted by frost.
The magnitude of the impact to the facing/nail depends on the depth of frost penetration, the
intensity and duration of the freeze period, the availability of water, and the stiffness of the
facing. Kingsbury et al. (2002) report that the force in the nail head caused by frost action
can be as high as 2.5 times larger as the maximum seasonal nail force without frost action.
Increases in nail and facing loads should be anticipated in areas where frost durations are
generally greater than one week, where frost susceptible soils are encountered near the face,
and where the face is in close proximity to a source of water. Soils susceptible to frost action
are those exhibiting the following characteristics: (1) more than 3 percent of the solids
fraction is smaller than 1 mil for non uniform soils (i.e., Cu > 5), or (2) more than 10 percent
of the solids fraction is smaller than 1 mil for uniform soils (i.e., Cu ≤ 5) (Casagrande, 1931).
Cu is the coefficient of uniformity which can be obtained from grain size gradation tests.
External Loads
External loads may be applied at the top of the soil nail wall and may vary from relatively
light highway appurtenance loads (e.g., roadway lighting supports) to significant loads (e.g.,
loads resulting from the integration of a relatively large cantilever retaining structure on top
of the wall). For relatively light loading conditions, the external loads can be used to define
additional shear forces and flexural moments in the section of the wall above the first row of
nails. These loads are then added to the calculated facing loads for subsequent analysis.
For more significant loads (e.g., loads applied by bridge abutments), it may be necessary to
perform a full soil-structure interaction analysis to define how the additional facing and nail
loads are distributed throughout the entire soil nail structure. The magnitude and distribution
of the load transferred to the wall depends on the distance of the load to the wall and the type
of load foundation (shallow or deep). The magnitude of these loads can be significantly
increased if the structure is subject to seismic forces.
The weight of temporary facing must be supported by the installed nails or other
supplementary means until compressive stresses develop at the facing-nail contact. This is
particularly important for the facing of the initial excavation lifts that becomes unsupported
when the next excavation lift is performed. For typical construction facings consisting of 4-
in. thick shotcrete, experience has shown that the soil nails will support the weight of the
facing without major difficulties. For thicker applied shotcrete facings, support for the
shotcrete facing weight by considering the shear capacity of the nails and the bearing
capacity of the soils beneath the nails should be formally evaluated. The maximum thickness
of shotcrete facing that can be supported in this manner is dependent on the strength of the
soils. In competent ground, shotcrete facings up to 8- to 10-in. thick have been successfully
supported.
If necessary, support of the shotcrete facing weight may be achieved by the installation of
additional short, steeply inclined reinforcing elements acting as compression struts. An
analysis method for this case is provided in GEC No. 7 (Lazarte et. al, 2003).
CONCRETE BLOCK
(12" x 12" x 24")
VARIES
PVC SLEEVE
(WEEPHOLES AT 5' O.C.;
4:1 GEOCOMPOSITE LOCATION)
6:1
CONCRETE FOOTING
PERFORATED PVC PIPE
COLLECTION TRENCH
b) Example of soil nail wall with Frost Protection
Figure 9-33. Examples of Frost Protection of Soil Nail Walls (after Lazarte et al., 2003).
Considerations for the effects of seismic forces on soil nail walls is provided in Section 5.4.5
of GEC No. 7 (Lazarte et al., 2003).
9.2.7.1 Introduction
The purpose of this section is to present a step-by-step generalized method for soil nail wall
design. The major design steps are outlined in Table 9-6 and Steps 1 and 2 are discussed in
other chapters.
After completing the design, the design engineer will prepare soil nail wall specifications and
recommendations for construction monitoring.
As part of this step, design factors of safety (see Table 9-7) and corrosion protection
requirements are selected. In addition, the following design elements are selected to permit
preliminary design calculations:
Table 9-7. Minimum Recommended Factors of Safety for the Design of Soil Nail Walls
Using the ASD Method (after Lazarte, 2003).
Notes: (1) For non-critical, permanent structures, some agencies may accept a design for static loads and
long-term conditions with FSG = 1.35 when less uncertainty exists due to sufficient geotechnical
information and successful local experience on soil nailing.
(2) The second set of safety factors for global stability corresponds to the case of temporary
excavation lifts that are unsupported for up to 48 hours before nails are installed. The larger value
may be applied to more critical structures or when more uncertainty exists regarding soil
conditions.
(3) The safety factors for bearing capacity are applicable when using standard bearing-capacity
equations. When using stability analysis programs to evaluate these failures modes, the factors of
safety for global stability apply.
Wall Layout
Establish the layout of the soil nail wall, including: (1) wall height; (2) length of the wall; and
(3) wall face batter (inclination typically ranges from 0 to 10). The evaluation of the wall
layout also includes developing the wall longitudinal profile, locating wall appurtenances
(e.g., traffic barriers, utilities, and drainage systems), and establishing ROW limitations.
Horizontal nail spacing, SH, is typically the same as vertical nail spacing, SV (Figure 9-34).
Nail spacing ranges from 4 to 6.5 ft for conventional drilled and grouted soil nails, and may
be as low as 1.5 ft for driven nails. This reduced spacing for driven nails is required because
driven soil nails develop bond strengths that are lower than those for drilled and grouted
nails. A soil-nail spacing of 5 ft is routinely used and is preferred for conventional drilled
and grouted soil nails. Soil nail spacing may be affected by the presence of existing
underground structures.
Soil nail spacing in horizontal and vertical direction must be such that each nail has an
influence area SH × Sv ≤ 40 ft2 ft. The design engineer should specify a minimum horizontal
soil nail spacing of about 3.3 ft. Design forces from global stability analysis and facing
design are affected by soil nail spacing. In general, the larger the spacing, the greater the
design forces. The purpose of the minimum nail spacing is to reasonably ensure that group
effects between adjacent soil nails are minimized due to potential nail intersection as a result
of drilling deviations. Group effects reduce the load-carrying capacity of individual soil
nails. The maximum soil nail spacing should also be specified. The purpose of a maximum
spacing [usually about 6.5 ft] is to provide for a soil nail system that is relatively easy to
construct and that effectively supports the lateral earth pressures and imposed surcharge
loads.
The soil nail pattern is commonly one of the following (Figure 9-34): (1) square
(rectangular); (2) staggered in a triangular pattern; and (3) irregular (at limited locations).
A square pattern results in a column of aligned soil nails, and facilitates easier construction of
vertical joints in the shotcrete facing (or easier installation of precast concrete panels). Also,
a square pattern enables a continuous vertical installation of geocomposite drain strips behind
the facing to be easily constructed. In practice, a square pattern is commonly adopted.
SVO < SV
NAIL
#1
1 ftM
0.30
(TYP)
SV
i
BOTTOM OF
EXCAVATION
LIFTS
N
BOTTOM
OF
SVN < SV EXCAVATION
SVO
0.30
1 ftM
(TYP)
SV
SVN
Figure 9-34. Soil Nail Patterns on Wall Face (after Lazarte et al., 2003).
A staggered soil nail pattern results in a more uniform distribution of earth pressures in the
soil mass. This effect is beneficial because an enhanced soil arching effect is achieved. This
method should be considered in cases where marginally stable soils are present because such
soils have less margin to redistribute loads. The main disadvantage of the use of a triangular
The use of uniform nail spacing is beneficial because it simplifies construction and quality
control. However, due to project-specific geometric constraints, nail spacing may need to be
irregular, with reduced spacing at some locations; for instance, in areas where the bottom of
the excavation or the top of the wall is not horizontal. In such cases, it is more convenient to
install one or two nail rows parallel to the non-horizontal edge and then establish a transition
zone where nails have a closer vertical spacing until a horizontal nail row is achieved (Figure
9-35a). It is also customary to reduce horizontal spacing at the vertical edges of the wall to
accommodate transition zones (Figure 9-35a).
Soil nails are typically installed at an inclination ranging from 10 to 20 degrees from
horizontal with a typical inclination of 15 degrees. This recommended range of soil nail
inclination assures that grout will flow readily from the bottom of the hole toward the nail
head for typical borehole and soil nail dimensions and conventional grout mixtures. Steeper
nail inclinations may be required, particularly for the upper row of nails, if a significantly
stronger soil zone is located at a greater depth and a more effective anchorage in the stiffer
layer is desired. Such evaluations can be readily made during design. Nail inclination
smaller than about 10 degrees should not be used because the potential for creating voids in
the grout increases significantly. Voids in the grout will affect the load capacity of the nail
and reduce the overall corrosion protection provided by the grout.
Project conditions may, however, require that other nail inclinations be used. For example,
Figure 9-35b shows a case in which utilities or other underground structures are located
within the proposed soil nail zone. In most cases, this situation only occurs for the upper first
and second rows of nails. Another situation where different nail inclinations may be used is
at exterior wall corners. To avoid intersecting nails behind exterior corners of a wall, nail
inclination on one side of the corner could be installed with a different inclination. An
alternative layout for exterior corners is to splay the nails on a plan view (Figure 9-35c).
Overhead space restrictions may require that the nail inclination be smaller than 15 degrees.
This might be the case for road widening at embankment bridge abutments. Logistical
limitations due to location of nailing equipment (i.e., operating at the bottom of a narrow
excavation) may require a steeper nail inclination.
SV
SH
UTILITIES
2 <15°
SV
SH
SH
15°
SH SH SH SH SH
SH
(b) CHANGE OF NAIL DECLINATION AROUND UTILITIES (c) NAIL SPLAYING IN CORNERS
The effect of nail inclination should be considered in global and local stability analyses of the
soil nail wall system because stability factors of safety for the system, particularly for sliding
wedge analyses in the upper portion of the wall, can decrease significantly as the nail
inclination increases below the horizontal.
The distribution of soil nail lengths in a soil nail wall can be selected as either uniform (i.e.,
only one nail length is used for the entire wall), or variable, where different nail lengths may
be used for individual soil nail levels within a wall cross section. Additional information on
nail distribution is provided below.
• Uniform Nail Length: When the potential for excessive wall deformation is not a
concern (e.g., soil nail walls constructed in competent ground or in an area without
nearby structures), it is beneficial to select a uniform length distribution because it
simplifies construction and quality control. Additionally, a slightly smaller total
length of nails is obtained with a uniform soil length pattern. This pattern provides
commonly a high sliding stability safety factor. Uniform patterns should be used in
most projects.
Performance of soil nail walls has shown that larger displacements are observed when the
upper nails are too short. The deformations in soil nail walls can be significantly reduced
when nails at the top of the structure are longer than required by stability analysis. In
general, the higher the global factor of safety of a soil nail wall, the smaller the wall
deformations.
Nail lengths have been installed successfully with a uniform nail length in the upper two-
thirds to three-quarters of the wall, with progressively shorter nails to a minimum value, not
smaller than 0.5 H (H is the wall height), at the bottom of the wall in dense cohesionless soils
that provide relatively large sliding stability. In general practice, nail length in the lower
rows should never be shorter than 0.5 H. Nail lengths less than 0.5 H will not likely satisfy
sliding stability requirements. In all cases, and especially where reducing the nail lengths in
In general, variable nail lengths result in a more complicated installation and require more
nail materials. Nevertheless, as many soil nail projects are specified based on performance
criteria, contractors may prefer to use longer nails in the upper rows to reduce deflections.
Project specifications must provide ROW constraints, locations of underground utilities and
substructures (or requirements that the contractor locate these), and specific deformation
criteria (i.e., maximum wall deflection and location where this deflection is to be measured).
Based on the discussion presented in this section, the following recommendations are made
concerning soil nail length and distribution.
• Select longer nails than required by the target factor of safety as a means to reduce
wall deformations in the upper portions of the wall.
• Avoid the use of too “short” nails in lower portion of wall. Evaluate if shorter nails
in bottom rows installed in competent ground satisfy sliding stability requirements.
Shorter nails at the bottom should be not smaller than 0.5 H.
• Non-uniform nail length patterns may be used if soil layers with very dissimilar
conditions are encountered.
For feasibility evaluations, soil nail length can be initially assumed to be 0.7 H, where H is
the height of the wall. The length of the nails may be greater than 0.7 H if large surcharge
loads are expected or if the wall is very high [greater than 30 ft high.]
Select appropriate grade of steel for the soil nail bar. For most applications Grade 60 steel is
used, however information on the selection of steel grade can be found in Section 9.2.3.1.
Soil Properties
The ultimate bond strength for the grout-ground interface can be selected using Table 9-3.
Introduction
Nail length, diameter, and spacing typically control external and internal stability of a soil
nail wall. Therefore, these parameters may be adjusted during design until all external and
internal stability requirements are satisfied. A series of charts was developed by Lazarte et
al. (2003) as a design aid to provide preliminary nail length and maximum tensile forces
(Figures 9-36 through 9-41). The charts were developed using the computer program
SNAIL, which was selected because it is public domain software, readily available, and free
of charge. In preparing these charts, the following main assumptions were made:
• homogenous soil;
• no surcharge;
• no seismic forces;
• uniform length, spacing and inclination of nails; and
• no groundwater.
When the conditions of a new analysis case do not match the assumptions listed above, it is
recommended that interpolations or extrapolations be made to estimate the soil lengths from
these charts. Alternatively, the use of a preliminary nail length between 0.7 to 1.0 times the
wall height can be made. The upper range of soil nail length is used for less favorable soil
conditions, wall heights greater than 30 ft, and where large surcharge loads need to be
resisted by the wall.
The charts were developed for different values of face batter (α); backslope (β); effective
friction angle (φ’); and ultimate bond strength (qu). Table 9-8 presents the geometric and
material conditions used for the development of the design charts.
The first type of chart was developed to evaluate the nail length (Figures 9-36a through 9-
41a) for combinations of α and β. Using these charts, the required nail length, L,
(normalized with respect to the wall height, H) to achieve a global safety factor of 1.35 is
obtained as a function of the normalized allowable pullout resistance (µ). The normalized
allowable pullout resistance is defined as:
γ, c, φ c* = c / γH
H SV
c* = 0.02 t max-s = Normalized Maximum Design Force in Nails
FS=1.35 = Tmax-s/γ H SH SV
4 in.mm
DDH= 100
For other FS, c*, and DDH,
see Figure B7
1.5
Friction Angle
(degrees)
31
35
1 39
27
L/H
0.5
(a)
00
0.4
Normalized Design Nail Force, tmax-s
0.3
0.2
0.1
γ, c, φ c* = c / γH
H SV
c* = 0.02 t max-s = Normalized Maximum Design Force in Nails
FS=1.35 = Tmax-s/γ H SH SV
4 in.mm
DDH= 100 For other FS, c*, and DDH,
see Figure B7
1.5
Friction Angle
(degrees)
27
31
1 35
39
L/H
0.5
(a)
00
0.4
Normalized Max. Design Force, tmax-s
0.3
0.2
0.1
NOTE:
(b) Nail forces are for FSG = 1.0
0
0 0.1 0.2 0.3 0.4
q aD DH
Normalized Bond Strength, µ =
γ SH SV
γ, c, φ c* = c / γH
H SV
c* = 0.02 t max-s = Normalized Maximum Design Force in Nails
FS=1.35 = Tmax-s/γ H SH SV
4 in.mm
DDH= 100 For other FS, c*, and DDH,
see Figure B7
1.5
Friction Angle
(degrees)
27
31
1 35
39
L/H
0.5
(a)
00
0.4
Normalized Max. Design Force, tmax-s
0.3
0.2
0.1
NOTE:
(b) Nail forces are for FSG = 1.0
0
0 0.1 0.2 0.3 0.4
q aDDH
Normalized Bond Strength, µ =
γ SH SV
γ, c, φ c* = c / γH
H SV
c* = 0.02 t max-s = Normalized Maximum Design Force in Nails
FS=1.35 = Tmax-s/γ H SH SV
DDH= 4
100
in.mm For other FS, c*, and DDH,
see Figure B7
1.5
Friction Angle
(degrees)
27
31
1 35
39
L/H
0.5
(a)
Normalized Max. Design Nail Force, tmax-s
00
0.4
0.3
0.2
0.1
NOTE:
Nail forces are for FSG = 1.0
(b)
0
0 0.1 0.2 0.3 0.4
q aD DH
Normalized Bond Strength, µ =
γ SH SV
Figure 9-39. Batter 10o – Backslope 10o (after Lazarte et al., 2003).
γ, c, φ c* = c / γH
H SV
c* = 0.02 t max-s = Normalized Maximum Design Force in Nails
FS=1.35 = Tmax-s/γ H SH SV
DDH=4100
in. mm
For other FS, c*, and DDH,
see Figure B7
1.5
Friction Angle
(degrees)
39
1
L/H
0.5
(a)
00
0.4
Normalized Max. Design Force, tmax-s
0.3
0.2
0.1
NOTE:
Nail forces are for FSG = 1.0
(b)
0
0 0.1 0.2 0.3 0.4
q aD DH
Normalized Bond Strength, µ =
γ SH SV
γ, c, φ c* = c / γH
H SV
c* = 0.02
t max-s = Normalized Maximum Design Force in Nails
FS=1.35
= Tmax-s/γ H SH SV
DDH=4100
in. mm
For other FS, c*, and DDH,
see Figure B7
1.5
Friction Angle
(degrees)
39
1
L/H
0.5
(a)
00
0.4
Normalized Max. Design Force, tmax-s
0.3
0.2
0.1
NOTE:
Nail forces are for FSG = 1.0
(b)
0
0 0.1 0.2 0.3 0.4
qaD DH
Normalized Bond Strength, µ =
γ SH SV
Figure 9-41. Batter 10o – Backslope 30o (after Lazarte et al., 2003).
q u D DH
µ = (9-37)
FSP γ SH SV
where FSP is the factor of safety against pullout (typically 2.0); DDH is the drillhole diameter;
γ is the total unit weight of the soil behind the wall; and SH and SV are the horizontal and
vertical nail spacing, respectively.
The nail lengths in these charts were computed based on the most critical failure surface (i.e.,
considering base and toe failures) for the selected geometry and material properties, and
assuming that failure of the nail (i.e., tensile breakage) and/or failure of the facing would not
take place. Therefore, the pullout failure is implicitly assumed. Equation 9-37 is based on a
drillhole diameter of 4 in. Also, the use of Equation 9-37 inherently assumes that the soil has
a cohesion intercept c′ such that c* = c′/γH = 0.02. If the drillhole diameter or cohesion
intercept values being considered are different than the assumptions stated here, then
adjustments to the calculated nail length and maximum tensile forces are made in the final
step. These adjustments are discussed subsequently.
The second type of charts (Figures 9-36b through 9-41b) provides the corresponding
maximum normalized design tensile force of all nails (tmax-s) as a function of µ calculated for
a global safety factor of 1.0. The maximum normalized design tensile force in the bar is
defined as:
Tmax − s
t max − s = (9-38)
γ H S H SV
With tmax-s read from the design charts, the maximum nail tensile force, Tmax-s can be
calculated using Equation 9-38. These design charts are developed for the case in which all
nail bars are the same length. These design charts do not provide information on the
distribution of tensile load in individual soil nails or the maximum load in any particular nail.
A step-by-step procedure for preliminary design using the charts shown on Figures 9-36
through 9-41 is presented in this section.
1. For a specific project application, evaluate batter (α), backslope (β), effective
friction angle (φ′), and ultimate bond strength (qu). Calculate normalized pullout
resistance (µ) using Equation 9-37.
2. Obtain normalized length (L/H) from the first set of charts (Figures 9-36a through
9-41a).
3. Obtain normalized force (tmax-s ) from the second set of charts (Figures 9-36b
through 9-41b).
4. Using Figure 9-42, evaluate correction factors for: (a) normalized length to
account for a drillhole diameter other than 4 in. (correction factor C1L), (b) a c*
value other 0.02 (correction factor C2L), and (c) a global factor of safety other than
1.35 (correction factor C3L).
5. Using Figure 9-42, evaluate correction factors for normalized maximum nail force
to account for: (a) a drillhole diameter other than 4 in. (correction factor C1F), and
(b) a c* value other 0.02 (correction factor C2F).
6. Apply correction factors to normalized length and/or normalized force.
Calculation method is provided on Figure 9-42.
7. Multiply the normalized length by the wall height to obtain the soil nail length.
8. Calculate the maximum design load in the nail Tmax-s using the value of tmax-s and
Equation 9-38.
9. Calculate the required cross-sectional area (At) of the nail bar according to:
Tmax − s FST
At = (9-39)
fy
where fy is the steel yield strength and FST is the factor of safety for nail bar tensile
strength.
10. Select closest commercially available bar size using Table 9-9 that has a cross-
sectional area of at least that evaluated in the previous step.
11. Verify that selected bar size fits in the drillhole with a minimum grout cover
thickness of 1 in.
12. If the length and/or nail diameter are not feasible, select another nail spacing
and/or drillhole diameter, recalculate the normalized pullout resistance, and start
the process again.
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-88 June 2008
Corrections Of Normalized Soil Nail Length
L L
(corrected ) = C1L × C 2L × C 3L × ( from charts for D DH = 4 in., c * = 0.02 , FS G = 1.35)
H H
where:
t max - s (corrected ) = C1F × C 2F × t max -s ( from charts for DDH = 4 in., c * = 0.02)
C1F
1.8
0.9 C 0.9
1.8
C1F1F
Force,C 1L
for Length,
forLength,
1.6
0.8 0.8
1.6
Correctionfor
Correction for
Correction
Correction
1.4
0.7 0.7
1.4
0.6
1.2 C1L 1.2
0.6
C1L
0.5
1 10.5
100
4 5
150
6 7
200
8 9
250
10 11
300
12
Drillhole Diameter, DDH (mm)
Drillhole Diameter, DDH (in.)
Figure 9-42. Correction Factors for Use in Design Chart Solutions (after Lazarte, 2003).
Nominal Bar Cross-Sectional Nominal Unit Max. Diameter ASTM Max. Axial
Yield Strength
Designation Area Weight w/Threads Grade Load
English in.2 lbs/ft in. English ksi kips
60 60 26.4
#6 0.44 1.50 0.86
75 75 33.0
60 60 36.0
#7 0.60 2.04 0.99
75 75 45.0
60 60 47.4
#8 0.79 2.67 1.12
75 75 59.3
60 60 60.0
#9 1.00 3.40 1.26
75 75 75.0
60 60 76.2
#10 1.27 4.30 1.43
75 75 95.3
60 60 93.6
#11 1.56 5.31 1.61
75 75 117.0
60 60 135.0
#14 2.25 7.65 1.86
75 75 168.8
For this step (and all subsequent design steps), it is necessary to perform a final design in
which the actual wall geometry, stratigraphy, loads, variation of engineering parameters (if
present), and other conditions are considered. The preliminary design procedure described
earlier should not replace the findings and results obtained with the final design presented
herein.
Global Stability
• Select a well-established computer program for design of soil nail walls that considers
heterogeneous soils, groundwater, general loading conditions, seismic forces, and
diverse nail characteristics. In this section, the computer program SNAIL (see
Lazarte et al., 2003 for a description of input and output capabilities of SNAIL) is
selected.
• Select the factor of safety against pullout failure (FSp) (SNAIL requires the value of
bond stress factor (BSF) equal to 1/FSp (e.g., for FSP = 2, the corresponding BSF is
0.5).
• For the first SNAIL analysis, use nail length estimated previously (or calculated in the
preliminary design) and perform global stability analysis using SNAIL.
• After selecting an initial nail length, perform the following iterative procedure using
SNAIL: (1) calculate the global factor of safety using the selected nail length; (2)
compare the calculated global factor of safety to the recommended minimum factor of
safety; and (3) increase or decrease the nail length if the calculated factor of safety is
lower or higher than the recommended value and start the process again.
• If the length of the nail needs to be reduced without reducing the factor of safety, then
increase the nail hole diameter or reduce the nail spacing.
Sliding Stability
Evaluate the potential for sliding failure using the equations and procedures outlined
below:
• Calculate the horizontal resisting forces (ΣR) as follows (see Figure 9-20).
where cb soil cohesion strength along the base, BL is the length of horizontal failure
surface where cb is effectively acting, W is the weight of soil nail block, QD is
permanent portion of total surcharge load QT, PA is active lateral earth pressure, β is
backslope angle, and φb angle of internal friction of the base.
where Hl is the effective height over which the earth pressure acts [Hl = H +
(β + tan α) tan βeq].
o Assume that the active lateral earth force is applied a distance of H1/3 from
the elevation of the bottom of the soil nail wall (Figure 9-20); and
o Calculate the horizontal driving force (ΣD) as:
∑ D = PA cos β (9-42)
∑R
FSSL = (9-43)
∑D
• If the factor of safety against sliding is lower than the specified minimum, increase
the length of the lower nails and reevaluate sliding stability.
Bearing Capacity
If soil nail wall is constructed in soft soils, evaluate the factor of safety against bearing
capacity failure (FSH) using Equation 9-44.
Su N c
FSH = (9-44)
S
H eq γ − u
B'
where Su is undrained shear strength of soil, Nc is bearing capacity factor (see Figure 9-
21), γ is the unit weight of soil behind wall, Heq is equivalent wall height [Heq = H + ∆H],
B’ is width of influence [B’ = Be / 2 ].
• The SNAIL analysis provides (at the end of the output file) the average nail tensile
force calculated for a case with FSG = 1.0 (Tavg-s).
• Calculate the average nail load (Tavg) as the sum of the individual nail forces
calculated by SNAIL divided by the number of nails in the analyzed cross section.
• The SNAIL analysis provides the maximum nail tensile force (Tmax).
T
Tmax − s = avg − s Tmax (9-45)
Tavg
• With Tmax-s, fy and the factor of safety against tensile failure (FST), calculate the
required cross sectional area of a steel nail bar (At) according to:
Tmax −s FST
At ≥ (9-46)
fy
• Select the closest commercially available nail bar size (see Table 9-9)
• Verify that the bar fits in the drillhole subject to a minimum grout cover thickness of
1 in. and the required corrosion protection.
Steel reinforcement:
Grade (fy), WWM dimensions, and Rebar dimensions (see Tables 9-9 and 9-10)
Select bearing plate geometry: min. 8 × 8 in. and 0.75 in. thick.
f c' [psi]
ρ min [%] = 0.24 (9-48)
f y [ksi]
f c' [psi] 90
ρ max [%] = 0.05 (9-49)
f y [ksi] 90 + f y [ksi]
b. Select reinforcement area per unit length of WWM for temporary/permanent facing at
the nail head (an) and at mid-span (am) in both the vertical and horizontal directions.
Typically, the amount of reinforcement at the nail head is the same as the amount of
reinforcement at the mid-span (i.e., an = am) in both vertical and horizontal directions.
For temporary facing, if waler bars are used at the nail head in addition to the WWM,
recalculate the total area of reinforcement at the nail head in the vertical direction (see
Equation 9-50) and horizontal direction (change Equation 9-50 appropriately).
A vw
an = am + (9-50)
SH
c. Calculate the reinforcement ratio (ρ) at the nail head and the mid span as:
an
ρn = 100 (9-51)
b h/2
am
ρm = 100 (9-52)
b h/2
d. Verify that the reinforcement ratio of the temporary and permanent facing at the mid-
span and the nail head are greater than the minimum reinforcement ratio (i.e., ρmin ≤
ρ), otherwise increase the amount of reinforcement (an and/or am) to satisfy this
criterion.
e. Verify that the reinforcement ratio of the temporary and permanent facing at the mid-
span and the nail head are smaller than the maximum reinforcement ratio (i.e., ρ ≤
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-95 June 2008
Table 9-11. Headed-Stud Dimensions (Metric and English Units) (after Byrne et al.,
1996)
DSH
tSH
LS
DSC
f. Using Table 9-4, select factor CF (typically 1 for permanent facings) to take into
account the non-uniform soil pressures behind facing.
g. Calculate facing flexural resistance (RFF) for the temporary and permanent facing as:
ρ tot = ρ n + ρ m (9-54)
and use Table 9-12a (interpolate for ρtot if necessary) and calculate RFF for the
temporary/permanent facing.
h. Using the recommended factor of safety for facing flexure (FSFF), verify that the
temporary and permanent facing flexural resistance is higher than nail head tensile
force (To):
i. If the capacity of the temporary and/or permanent facing is insufficient, increase the
thickness of facing, steel reinforcement strength, concrete strength, and/or amount of
steel and repeat the facing flexural resistance calculations.
a. Temporary Facing: With the values of concrete strength (fc’), facing thickness (h),
and bearing plate length (LBP), use Table 9-12b to obtain the punching shear
resistance (RFP) for the temporary facing.
b. Permanent Facing: With the values of concrete strength (fc’), headed-stud geometric
characteristics and spacing, use Table 9-12c to obtain the punching shear resistance
(RFP) for the permanent facing.
(c) FACING RESISTANCE FOR SHEAR (5) hc = Ls – tH + tP where: Ls is the effective headed-
PUNCHING, RFP stud length (Table 9-11); tP is the bearing plate
(PERMANENT FACING) thickness [typically 0.75 in.]; tH is the headed-stud
head thickness (Table 9-11).
Headed Stud Spacing, SHS in.
hc(5) f’c(3)
4 5 6
in. ksi RFP in kip
3 21 21 21
4
4 25 25 25
3 30 33 33
5
4 35 39 39
3 40 44 48
6
4 46 51 55
(d) FACING RESISTANCE FOR HEADED STUD,
TENSILE FAILURE, RFH,
(PERMANENT FACING)
d. If capacity for the temporary/permanent facing is not adequate, then implement larger
elements or higher material strengths and repeat the punching shear resistance
calculations.
a. Calculate the maximum tensile resistance due to headed-stud tensile failure (RHT)
using Table 9-12d, or alternatively as:
R FH = N AS f y (9-57)
b. Verify that that capacity is higher than nail head tensile force:
c. Verify that compression on the concrete behind headed-stud is within tolerable limits
by assuring that:
AH ≥ 2.5 AS (9-59)
where AH is the cross-sectional area of the stud head; AS is the cross-sectional area of
the stud shaft; tH is head thickness; DH is diameter of the stud head; and DS is diameter
of the headed-stud shaft.
f. If capacity is not enough, adopt larger elements or higher strengths and recalculate.
To minimize the likelihood of a failure at the nail head connection, use the recommended
minimum specifications for the hardware elements provided below. Additional information
can be found in Chapter 5 of Geotechnical Engineering Circular No. 7 (Lazarte et al., 2003).
• Bearing Plates: Bearing plates should be mild steel with a minimum yield stress, fy,
equal to 36 ksi (ASTM A-36/A36M).
• Nuts: Nuts should be the heavy-duty, hexagonal type, with corrosion protection
(oversized when epoxy-coated bars are used).
• Beveled Washers: Beveled washers (if used) should be steel or galvanized steel. If
the plate and other hardware elements are not within the ranges recommended, a
formal calculation of capacities should be performed. Note that some proprietary
systems employ spherical seat nuts that do not require washers.
1) Use Figure 9-32 as a guide to estimate the magnitude of vertical and horizontal
displacements.
2) Obtain wall height (H) and batter angle (α) (see Figure 9-32 for a description of
variables).
3) Identify ground conditions (i.e., weathered rock/stiff soil, sandy soil, clayey soil).
4) Estimate horizontal and vertical displacements δh and δv at the top of the wall using Table
9-5
5) Calculate zone of influence, DDEF, where noticeable ground deformations occur using
Equation
D DEF
= C (1 − tan α )
H (9-61)
The design of a soil nailing system is usually tested in the field to verify that: (a) the design
loads can be carried by the nails without excessive movements, (b) the contractor’s
equipment and installation procedures are adequate, and (c) the long-term behavior of the
nails is as anticipated. Four types of tests are usually performed: an ultimate test, a
verification test, a proof test and a creep test. Table 9-13 presents a brief discussion on
various aspects of these tests (Porterfield et al., 1994 and Clouterre, 1991).
Figure 9-43. Details of Typical Soil Nail Test Set-Up (Porterfield et al., 1994).
Figure 9-43 illustrates a typical set-up for a soil nail test. A hydraulic jack and pump are
used to apply the load to the nail. A jacking frame or reaction block is usually installed
between the shotcrete (or the excavated face) and the jack. Once the jack is centered, an
alignment load is applied to the jack to secure the equipment.
Movement of the nail head is measured by one or preferably two dial gages attached to a
rigid support independent of the jacking set-up. The dial gages are zeroed after the alignment
The load is applied to the nail by a calibrated hydraulic jack. The jack should have a
minimum travel of 6 in. A calibrated load cell, which should be aligned with the axis of the
nail and the jack, is used to detect small changes in load and allow maintenance of constant
load during creep testing.
Figure 9-44 shows a data log sheet that can be used for the load testing of soil nails. Figure
9-45 presents an example of data reduction of soil nail load testing to calculate elastic
movement. Figure 9-46 presents an example of data reduction of soil nail load testing to
calculate creep movement between 1- and 10-minute readings.
Inspection responsibilities for soil nail walls are summarized in Table 9-14.
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor’s schedule
Discuss anticipated ground conditions and potential problems with Contractor
Review Contractor’s methods for surface water control and verify adequacy throughout
construction
Review corrosion protection requirements from the Specifications and confirm that
Contractor is following these requirements
If specified, obtain test samples from steel components, centralizers, and drainage
materials and check all Mill test certificates for compliance with Specifications
NAIL STORAGE AND HANDLING
Nails, cement, and bars must be kept dry and stored in a protected location
Nails and bars should be placed on supports to prevent contact with the ground
EXCAVATION
Prior to starting excavation, check for any variance between actual ground surface along
the wall line and that shown on the Plans
Collect excavated soil samples and perform visual identification. Inform Engineer of the
results for comparison against the assumed soil type for design.
Confirm that stability of excavated face (i.e., stand-up time) is maintained at all stages of
construction
Confirm that excavations are constructed within Specification tolerances of the design
line and grade
For each excavation lift, confirm that Contractor is not over excavating
Enforce specific excavation sequencing plan provided on the Plans as they relate to lift
thickness, length of open unsupported excavation, and, if required use of stabilizing
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-107 June 2008
berms
Identify areas of excessive seepage and report to Engineer
Confirm that excavated face profile is sufficiently smooth to facilitate shotcrete
placement and to minimize overages in shotcrete quantities
DRILLING OF NAIL HOLES
Confirm that drilling technique used is consistent with ground conditions
Document drilling procedures and report to the Engineer if drilling method unsuitable for
actual ground conditions encountered
Confirm that soil nail hole is drilled within acceptable tolerances of the specified
location, length, and minimum diameter
Observe and document locations of excessively hard drilling
Visually inspect for loss of ground or drill hole interconnection and confirm that neither
of them are occurring during drilling; subsidence of ground above drilling location or
large quantities of soil removal with little or no advancement of the drill head should not
be permitted
TENDON INSTALLATION AND GROUTING
Inspect open soil holes for caving or loose cuttings using a high intensity light
Inspect all soil nail bars and reinforcing steel for damage and defects prior to installation
Confirm consistency of epoxy coated or encapsulated tendons and inspect for any
damage to corrosion protection prior to installation into drill hole
Confirm mix design compliance of soil nail grout and take grout samples as required
Record volume of grout placed for each drill hole
Confirm that nail bars are inserted to the minimum specified length
Confirm that centralizers are installed at specified intervals
Confirm that all required hardware is appropriately affixed at the soil nail head
Confirm that no damage occurs to corrosion protection components during installation
Confirm that grout is injected by tremie pipe starting at the bottom of the hole and that
the end of tremie pipe always remains below the level of the grout as it is extracted
Confirm that grout is continued to be pumped as the grout tube, auger, or casing is
removed
Confirm that the Contractor does not reverse the auger rotation while grouting except as
necessary to initially release the tendon
Confirm that grout is batched in accordance with approved mix designs
Observe Contractor’s methods to place grout/shotcrete just behind the soil nail head and
confirm continuous coverage
Confirm that any required testing for grout strength is conducted in accordance with
specified testing methods
LOAD TESTING
Obtain all required calibration certifications of Contractor’s load testing equipment
Check all deformation gauges and confirm movements during load testing
Confirm that load testing of individual nails does not commence until minimum grout
curing time has passed
Confirm that the load test is performed consistently with Specifications and all required
load test data is provided to permit comparison to acceptance criteria outlined in
Specifications
If the soil nail fails, report to the Engineer and do not allow any retesting until the
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-108 June 2008
Contractor modifies the installation procedures
DRAINAGE INSTALLATION
Confirm compliance of drainage materials with Specifications
Confirm that geocomposite drain strips and weep hole outlet pipes are installed as
specified and Plans and that the drain elements are sufficiently interconnected and
provide continuous drainage paths
WALL FACING
Confirm shotcrete mix design consistent with Specifications
Confirm that steel reinforcing is appropriately positioned within temporary shotcrete
facing
Confirm that exposed soil face is covered with shotcrete within specified time limits
Confirm that minimum shotcrete thickness is maintained at all sections of the work
Confirm that shotcrete installation methods used in the field are consistent with the
Specifications and as approved by the Engineer
Confirm that construction joints are clean and acceptable for shotcrete placement
Confirm that shotcrete is batched in accordance with the approved mix design
Confirm that wall finish line and grade is in accordance with Plans and Specifications
If specified, confirm that shotcrete test panels are prepared, cured, and transported to the
Testing Laboratory
POST INSTALLATION
Verify pay quantities
9.3.1 Introduction
Micropiles are small diameter (less than 12 in.) drilled piles constructed with steel
reinforcement, and bonded to the ground with grout using gravity or pressure grouting
techniques. Micropiles may be used for structural support, slope stabilization, and retaining
systems. Information on the design and construction of micropiles for structural support and
slope stabilization is provided in Sabatini and Tanyu (2006).
Micropile walls may be used for temporary shoring and permanent earth retaining systems.
In general, micropiles are relatively expensive compared to other forms of deep foundation
elements such as driven piles or drilled shafts. Inasmuch as drilled shafts and driven pile
elements are used as vertical wall elements (e.g., secant pile walls and driven steel soldier
piles), the use of micropiles for wall systems will likely only be a viable and cost-effective
system where driven piles or drilled shafts cannot be installed.
The principal components of a micropile wall consist of vertical micropile elements installed
from the ground surface at or near the final excavated wall face line and subhorizontal
Micropile retaining walls are constructed from the top-down and generally follow this
sequence:
• at the ground surface, excavate an area wide and deep enough to accommodate the
cap beam;
• install the formwork for the cap beam and place the cap beam steel reinforcement;
• place corrugated plastic sleeves for installation of the micropiles through the cap
beam;
24 ft
4 in.
10 ft
15 ft
Figure 9-47. Micropile Wall Cross for Wall 600, Portland, Oregon.
Excavation in front of the wall is performed in lifts (typically no more than 6 ft thick).
During excavation, shotcrete is applied to the excavation face to temporarily prevent raveling
of the soil face. Connection to the micropiles is performed via head studs that are welded to
the front line micropiles. Following completion of the excavation, a leveling pad is poured to
allow erection of one-sided forms. Once the leveling pad is completed the wall face is
constructed from CIP concrete. Headed studs welded to the micropiles are embedded in the
CIP grade beam and wall face to provide connection of the micropile structure to the CIP
wall face.
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-111 June 2008
The typical construction sequence for simple gravity grouted and pressure-grouted micropiles
(Figure 9-48) includes drilling the pile shaft to the required tip elevation, placing the steel
reinforcement, placing the initial grout by tremie, and placing additional grout under pressure
as applicable. In general, the drilling and grouting equipment and techniques used for the
micropile construction are similar to those used for the installation of soil nails and ground
anchors.
9.3.3.1 Overview
For the Wall 600 project in Portland (i.e., micropile wall shown in Figure 9-47), detailed soil-
structure interaction analyses were performed to verify a more simplified model in which a
design total pressure diagram (including earth pressures, seepage pressures, seismic forces,
and traffic barrier impact loading) was applied proportionally to the two vertical micropiles
and the two battered micropiles for each micropile section (see Ueblacker, 1997).
Micropiles were used to provide temporary excavation support for a project involving
improvements to State Highway 82 near Aspen, Colorado (Macklin et al., 2004). In this
mountainous region, construction is hampered by difficult site access and slope instability
risks. This site consists of loose to very dense silty to gravelly sand with cobbles and
boulders, which were deposited as debris flow, sheet wash, and colluvium over dense alluvial
sandy gravel with cobbles. Originally, temporary shoring using combinations of soil nails
and tiebacks was considered for the project. However, in an effort to improve the
construction schedule and phasing, micropiles were selected as an alternative temporary
shoring system. The micropile shoring implemented on this project was essentially a hybrid
between a soldier pile and lagging system and a soil nail stabilization system (Figure 9-49).
Phasing of the project required constructing a number of bridges in a specific sequence. The
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-113 June 2008
micropile shoring system allowed the contractor to transition access road grades where
convenient, depending upon the excavation and access requirements for any phase of the
work.
For this project, the internal and external stability of the temporary micropile shoring system
was analyzed using a combination of gravity wall calculations, free earth support methods,
lateral pile, and slope stability calculations. Also, finite element and finite difference models
were used to predict deformation behavior, stresses in the micropiles, and to analyze the
potential for soil flow around the micropiles. The design approach for this project is
described in the following six steps.
Figure 9-49. Cross Section Showing Steep Canyon Slope and Temporary Micropile Shoring
(after Macklin et al., 2004).
• Micropile sections were assumed and the ultimate bending capacity of the micropile
sections was calculated. The flexural rigidity (EI) of the micropiles was calculated
and these values were later used in numerical analyses. Micropiles that consisted of a
centralized reinforcing bar in a drilled and grouted hole were analyzed using LPILE’s
ultimate bending analysis module. The tensile and compressive capacity of each
section was also calculated.
• The micropile wall system was next analyzed using the free earth support method (as
is commonly used for anchored bulkhead design). For this evaluation, the front row
of closely spaced micropiles is considered to be analogous to a sheet pile wall and the
battered rows of micropiles are analogous to the deadman anchors. The analysis was
modified in that it was assumed that one-half of the calculated active earth loads was
applied to the vertical micropile row as a triangular pressure distribution and one-half
of the calculated active earth load was applied to the rear, battered micropile row. A
lateral pile analysis was then completed for both the vertical and battered micropiles
using LPILE.
• The potential for soil flow in-between the relatively closely-spaced micropiles was
evaluated.
Load testing is performed on micropiles in the field to verify that: (a) the design loads can be
carried by the micropiles without excessive movements, (b) the contractor’s equipment and
installation procedures are adequate, and (c) the long-term behavior of the micropiles is as
anticipated.
Micropiles are tested individually using the same conventional static load testing procedures
as are used for driven piles and drilled shafts. These tests include incremental loading (which
may be applied in compression, tension, or laterally and which may be cycled (i.e.,
FHWA NHI-07-071 9 – Soil Nail and Micropile Walls
Earth Retaining Structures 9-115 June 2008
load/unload)) until the micropile reaches the selected maximum test load, structural
displacement limit, or ground creep (i.e., movement under constant load) threshold.
Unlike ground anchors and soil nails for which well-defined testing programs, consistent
with a well-developed design approach, are available, load testing program protocols for
micropiles used for earth retaining systems have not been developed. Regardless, however,
load testing does need to be performed to verify the displacement response and capacity of
micropiles used for wall systems. Such a testing program will need to be developed on a
project-specific basis. Details of micropile load testing can be found in Sabatini and Tanyu
(2006).
It is also recommended that performance data (including micropile load transfer, axial loads,
bending moments, and displacements) be collected for micropile wall systems to enable
design methods to be updated.
The responsibilities of the inspector for micropile wall construction are summarized in Table
9-15.
CONTRACTOR SET UP
Review Plans and Specifications
Review Contractor’s schedule
Review in-situ property test results (i.e., grain size, Atterberg limits, unit weight, and
shear strength)
Discuss anticipated ground conditions and potential problems with Contractor
Confirm that Contractor’s grout pump is consistent with Plans (i.e., positive displacement
pump) and grout equipment is capable of producing uniform grout
Confirm that the dimensions of micropiles consistent with Specifications
Confirm that cement, reinforcement steel, and micropile are handled and stored
consistently with Specifications
Review corrosion protection requirements of metallic units and confirm their consistency
with Specifications
CONCRETE CAP INSTALLATION
Confirm that site preparation for the wall construction is consistent with Specifications
Confirm that the area excavated for the cap beam is consistent with Plans
Confirm that formwork for the cap beam is installed according to Specifications and
Plans
Confirm that corrugated plastic sleeves are installed through the cap beam according to
the Plans
Confirm that the concrete cap is poured according to Specifications and Plans
10.1 INTRODUCTION
Prior to 1970, the predominant types of earth retaining walls for permanent structures were
gravity and cantilever. Both gravity and cantilever wall types had decades of successful use
in both cut and fill situations. Therefore, selection of a wall type consisted of choosing
between gravity and cantilever walls. Selection of a wall system in 2007 is considerably
more complex because the number of wall types available has increased significantly. A
systematic evaluation process should be used to select the most appropriate wall type for the
project.
This chapter presents a systematic wall system evaluation and selection process. The
objective of wall selection is to determine the most appropriate wall type that is cost-
effective, practical to construct, stable, and aesthetically and environmentally consistent with
its surroundings. As part of the evaluation process, walls are scored using a wall selection
matrix based on wall selection factors. The wall with the highest score is chosen as the wall
for the project.
A flowchart illustrating wall selection is presented in Figure 10-1. This flowchart is intended
to serve as a guide for highway design and construction specialists for evaluating and
selecting wall system alternatives for a project application. Aspects of wall selection that are
outside the scope of this flowchart, but which may be part of a formal review and acceptance
program, include review of approved wall system lists by an owner agency, establishment of
technical guidelines and criteria by which feasible wall system alternates can be judged, and
performance of a life cycle cost analysis for candidate wall systems. These aspects are
generally agency-specific and are not discussed in this chapter.
The first step in the selection process is to identify the need for an earth retaining system for
the project. The function of a retaining system is to form a nearly vertical face through
confinement and/or strengthening of a mass of earth material. Typically, an earth retaining
system is needed for projects that require abrupt changes in slope grades that cannot be
achieved by simply grading the slopes. As a rule of thumb, it can be stated that the more
restricted or congested the site, the greater the need for an earth retaining system. This is
because congested sites typically have very limited right-of-way (ROW) which does not
allow abrupt changes in slope grades by just grading the slopes.
The ROW restrictions of the site can be evaluated by visiting the site for preliminary review.
During the site visit, the site geometry can be evaluated and ROW restrictions, such as
existing structures and utility lines, can be noted. Typically, the observations from the site
visit when compared against the proposed project requirements are enough to make an initial
decision on the need for an earth retaining system for the project.
The second step in wall selection involves identifying site specific constraints and project
requirements. This information can be obtained during a preliminary site review. Items
affecting wall selection include, but are not limited to the following: (1) site accessibility and
space restrictions that may include limited ROW and headroom, availability of on-site
storage for wall materials, access for specialized construction equipment, and restrictions on
traffic disruption; (2) location of above-ground utilities and nearby structures; (3) aesthetic
requirements imposed by project surroundings; (4) environmental concerns that may include
local policies concerning construction noise, vibration, and dust, on-site stockpiling and/or
transport and disposal of excavated material, discharge of large volumes of water, and
encroachment on existing waterways; and (5) exposed wall face height. The relative
importance of each of the above items should be assessed for the specific project under
consideration so that the more important items are given priority during the selection process.
For this step, eleven wall selection factors are considered. These factors include: (1) ground
type; (2) groundwater; (3) construction considerations (i.e., availability of material,
equipment, and etc.); (4) speed of construction; (5) ROW; (6) aesthetics; (7) environmental
concerns; (8) durability and maintenance; (9) tradition; (10) contracting practices; and (11)
cost. This list is not project specific and is presented as a tool to help evaluate the
importance of project requirements determined in step 2 in a systematic manner. For a given
project, more specific factors such as, for example, lateral movements, cost of maintenance,
and the availability of a standard design may be critical in the process.
The evaluation is performed by the party that is responsible in wall selection. Depending on
the contracting policy adopted for the project, this evaluation may be performed by the
owner, the consulting engineer representing the owner, or the contractor.
Each wall selection factor is evaluated based on its relevancy and importance to the project
requirements and site constraints and they are assigned a rating number between one (1) and
three (3). This is termed the weighted rating (WR) for a given wall selection factor. Three is
assigned to the most relevant or important factors and one is assigned to the least relevant
ones. The evaluation results should be tabulated as shown in Table 10-1. Typically, cost,
speed of construction, and durability are the most important wall selection factors for
permanent wall systems. In general, any issues which are given a WR = 1 need not be
considered further.
Note: 1Rating of the importance of each wall selection factor based on project requirements and site constraints. Each factor should be rated between 1 and 3, where 1 is the least important or
relevant factor and 3 is the most important factor.
10.5.1 General
From this step forward, various wall systems are evaluated for the project. As a logical first
substep, obviously inappropriate wall systems should be eliminated. For this elimination
process, project constraints related to wall geometry and wall performance (or design
requirements) should be considered. Constraints related to cost should not be addressed at
this stage; cost will be addressed consistently for all wall systems that are not eliminated as
part of this step.
The factors affecting the selection of cut and fill wall systems are summarized below. In
some projects both cut and fill wall systems may be suitable and in these types of projects
both wall systems should be evaluated.
Cut walls are constructed from the top down and most cut walls can be used for both
temporary and permanent applications. The permanent systems are typically designed with
greater corrosion protection measures and are constructed with permanent facing elements
such as cast-in-place or precast concrete panels.
Cut walls are either drilled or cut into the in-situ soil or rock and are constructed with
specialized equipment and labor. For cut walls without ground anchors (i.e., nongravity
cantilever walls) little or no ROW is required. For anchored walls and soil-nailed walls,
significant ROW or permanent easements may be necessary.
Costs can vary significantly for cut walls depending on the specific wall being constructed
(i.e., complexity of the construction process) and the availability of experienced contractors
and equipment in the project location. For cut walls, the unit cost of the wall increases as the
height of the wall increases. For wall heights greater than approximately 15 to 30 ft, an
anchored wall or a soil-nailed wall is necessary. Additional cost results from material
procurement, drilling, installation, corrosion protection, and testing of the anchors or soil
nails. For all permanent wall systems, factors affecting costs include constructing an
aesthetically pleasing wall finish, fabricating and installing special connections for the facing
panels, and, if necessary, providing adequate long-term corrosion protection and constructing
drainage systems.
Fill walls are constructed from the bottom up and are typically used for permanent wall
applications due to the high cost of their facing components, although temporary MSE
systems are used without permanent facings. All fill walls require ROW and their cost-
effective height range is less than the cost-effective height range of cut walls.
For permanent highway applications, fill wall systems generally require granular, nonplastic,
free-draining backfill. The cost effectiveness of a MSE wall system, which typically requires
a greater quantity of select backfill than rigid gravity and semi-gravity wall systems, may be
reduced if select backfill is unusually expensive at a specific project site.
The advantages and disadvantages, cost-effective height range, required ROW and
representative tolerances for differential settlement for fill wall systems are summarized in
Table 10-3.
The performance of wall alternatives that are not eliminated earlier in this step should be
evaluated against wall selection factors that are defined here using a rating between one (1)
and four (4). For each factor, four is assigned to the most suitable factor for the wall being
evaluated and one is assigned to the least suitable factor for the wall being evaluated. At this
stage, constraints related to cost should be addressed for all wall systems that are not
eliminated previously.
The rating for each wall type should be tabulated for use in Step 5. A brief summary of how
each of the wall selection factors affects the performance of each wall type is given below.
An earth-retaining system is influenced by the earth it is designed to retain, and the one on
which it rests. The influence of the earth is particularly important in “earth walls” where the
retained earth itself has a major load-carrying function. In MSE walls, for instance, which
usually involve some sort of reinforcement, the pull-out force in the reinforcement is resisted
by (1) the friction along the soil-reinforcement interface and (2) the passive resistance along
the transverse members of the reinforcement, if any (grid reinforcement). Therefore, these
systems are best suited for soils with high internal friction such as sands and gravels.
FHWA NHI-07-071 10 – Wall Selection
Earth Retaining Structures 10-6 June 2008
Table 10-2. Summary of Evaluation Factors for Cut Walls (after Sabatini et al., 1997).
Gravity-type structures are less influenced by the type of soil than the systems involving soil
reinforcement. For soils with large vertical and horizontal deformations, a very flexible
system such as a gabion wall may be chosen in lieu of a more rigid system that attempts to
resist such deformations.
In DMM and jet grouted walls, the compatibility of the ground and the structure depends
largely on the type of chemicals used, and hence on the specific system employed. While
DMM walls, for instance, are normally used in sandy and liquefiable soils, the lime-column
walls are mostly suitable for use in deep clay deposits rich in pozzolans. In jet-grouted walls,
the in situ properties and structure of the soil are not as important since the concept of jet
grouting is to break down the soil structure and replace it with a self supporting composite
mass of soil and grout sometimes called soilcrete. The strength of the soilcrete, however;
and the permeability of the wall may be influenced by certain soil elements such as peat or
boulders.
10.5.4.2 Groundwater
Generally, the groundwater table behind an earth-retaining structure is lowered for the
following reasons:
To reduce the negative impact of groundwater, a free-draining system such as a MSE wall
can be used. Sometimes, it is desirable to keep the water table high to prevent settlement of
adjacent structures or protect existing untreated timber pile foundations from fungus decay
due to exposure to oxygen. In these cases, a relatively rigid watertight structure is used
(slurry wall, tangent/secant piles, jet-grouted wall, etc.). These structures usually are
designed to support the full hydrostatic pressure.
Site accessibility is also an important factor. Depending on the terrain, the mobilization of
heavy equipment might not be possible. This would limit some of the available wall systems
that could be built. For example, crib walls or bin walls can be preassembled and mobilized
to the sites that are difficult to access with heavy equipment. However, a sheet pile or soldier
pile and lagging wall require the use of a hammer and/or drilling equipment and could not be
built unless such equipment could be mobilized to the site.
Speed of construction is one of the most important factors in wall selection. Precast concrete
module walls and precast concrete crib walls can be constructed relatively quickly. These
walls are shipped to the site assembled. Each wall face panel covers a relatively large area,
which makes the construction even faster. The cells of the bin walls are manufactured and
delivered to the site ready for assembly. Each steel member is bolted together at the site and
then filled with backfill soils.
Speed of construction for all gravity wall systems is affected by the condition of the wall
foundation materials. If the foundation soils consist of unsuitable material, such material will
need to be supported on a deep foundation, removed, or improved (e.g., surcharge fill) before
the construction of the walls, which will increase construction time significantly.
Walls that require a relatively long time to construct include concrete gravity walls, concrete
cantilever walls, and concrete counterfort walls. Preparation of foundation and pouring
concrete are the two important factors that affect the speed of construction for these types of
walls. These walls have very little tolerance to differential settlement, therefore the
foundation has to be prepared to provide a relatively stiff and uniform bearing surface. Once
the foundation is prepared, concrete should to poured carefully to eliminate cold joints and
no further construction should be performed until the concrete is allowed to cure and reach a
specified strength. Concrete cantilever and counterfort walls also require reinforcing
elements.
Gabion walls require a relatively long time to construct because they require significant
labor. Each gabion basket module has to be assembled in the field and tied to one another to
form the face of the wall. After that, each gabion basket has to be filled with stone and the
lid of the gabion basket module has to be closed and tied. As with any other modular gravity
wall, unsuitable foundation material has to be removed or treated before the construction of
the gabion baskets.
In MSE walls, a relatively large space is required behind the structure face as compared to
that needed for construction of conventional walls (the length of the reinforcing elements is
typically 0.7 times the wall height).
To support an excavation in a very tight space, a top-down staged excavation and support
system, such as soil nailing or an anchored wall, may be the most suitable. The feasibility of
such a structure, however, is influenced by the presence of utilities and buried structures
nearby and the additional cost of permanent underground easement for placement of the
reinforcing elements. Soil nail walls may also allow for construction underneath a bridge
without the need to disrupt traffic on the bridge.
Site congestion may be a drawback for some systems such as slurry walls. When low
headroom does not allow the operation of conventional construction equipment, walls which
can be implemented in a limited operating space (i.e., soil nails, micropiles) or from a remote
10.5.4.6 Aesthetics
In addition to being functional and economical, permanent earth retaining structures, in most
cases, have to be aesthetically pleasing. Different types, shapes and color facings are used in
construction of earth walls. The types of facings range from built on-site continuous facings
(shotcrete, welded wire mesh, cast-in-place concrete) to prefabricated concrete or steel
panels. Cast-in-place facing and precast panels usually are more attractive than shotcrete or
soldier pile and lagging walls. For permanent drilled shaft walls (i.e., tangent/secant pile
walls), an architectural wall facing usually is provided, but at an additional cost to the
structure.
The aesthetic factor is very important when building a retaining system in parks, forests and
natural habitat. A number of attractive wall systems (Criblock, Evergreen, etc.) are usually
considered for those areas because of their aesthetic, acoustic and anti-graffiti advantages.
The Evergreen wall, for instance, consists of precast concrete units with open spaces at the
face into which are planted shrubs, vines, etc. With adequate water supply for the foliage,
the concrete facing will no longer be visible a few years after construction.
Like most structures, the selection of an earth retaining system is influenced by its potential
environmental impact during and after construction. Excavation and disposal of
contaminated material at the project site, and discharge of large quantities of water or slurry
fluids generated during jet grouting and slurry wall constructions are of primary concern.
MSE walls which allow construction of roadway embankments with vertical sides to
minimize encroachment on wetlands have positive environmental and ecological benefits.
The environmental advantages of these walls can be further enhanced by the use of wood
chips, and other recycled waste materials in construction.
To reduce noise and vibration impacts, the systems which use pile driving or heavy
construction machinery may be rejected. To reduce traffic noise in environmentally-sensitive
areas, the gravity-type gabion and Evergreen walls offer specific advantages. The open
nature of the face and the presence of foliage covering are effective in absorbing the noise
hitting their facings, making these walls acoustically superior as compared to other earth-
retaining structures where the traffic noise is reflected on hard or smooth continuous
surfaces.
An earth retaining structure built of concrete has a higher durability against corrosion and
deterioration effects than a structure constructed of metal, or which uses metal or synthetics
for reinforcement and/or facing. The durability factor is extremely important when selecting
a maintenance-free earth retaining structure in highly corrosive surroundings, or when the
structure is subjected to attack by non-conventional elements such as waves, chemicals or
marine borers.
Corrosion of the reinforcement, for instance, is one of the major design issues of MSE walls
that use metal reinforcement. Although, these walls are sometimes avoided in higher than
normal corrosive environments, corrosion-protection measures can be provided if necessary,
but at an added cost to the structure. Gabion walls have durability concerns similar to those
of MSE walls with metal reinforcement.
When geosynthetics are used for reinforcement, their long-term creep behavior and resistance
to deterioration due to chemical attack and exposure to ultraviolet light are major
considerations that have to be addressed in the design. Selection of the types of geosynthetic
is sometimes dictated by durability.
The durability of a concrete structure (gravity wall, slurry diaphragm wall, etc.) is influenced
by the quality of the aggregates and water used in the mix, and by the casting procedures. As
indicated before, concrete walls are not recommended in areas where quality aggregates are
not economically available.
10.5.4.9 Tradition
Tradition may dictate or prevent the use of a certain-type of structure, irrespective of its
technical rating. Although earth walls are very popular in certain states they are rarely built
in others. While still considered novelty in certain parts of the U.S.A., slurry walls are
heavily used in construction of underground facilities in other areas such as Boston where
drilled caissons are traditionally used as deep foundations, bored-pile walls are popular and
generally economical since the local contractors are equipped for, and experienced with, that
type of construction. Tradition plays a greater role in construction in underdeveloped
countries.
The contracting policies and procedures followed in the United States may discourage or
even preclude the use of certain types of walls, particularly those involving patented
equipment, materials or procedures. Contracting issues are discussed in detail in Chapter 11.
FHWA NHI-07-071 10 – Wall Selection
Earth Retaining Structures 10-13 June 2008
10.5.4.11 Cost
The total cost of an earth-retaining system has many components including the structure,
ROW, temporary or permanent easement, excavation and disposal of unsuitable material, and
drainage.
The construction costs of specific types of retaining systems have been discussed in previous
chapters and are summarized here in Tables 10-4 and 10-5 for cut and fill wall systems
respectively. It should be noted, however, that a structure with the least construction cost
does not necessarily mean an economical alternative, as the ultimate cost of the system is
influenced by many indirect cost factors such as those listed above as well as schedule,
permitting, maintenance, and wall face requirements.
For cut walls, the most common permanent wall facing is either cast-in-place (CIP) or
precast concrete. As of 2004, the cost of precast concrete facing is $25 to $30 per ft2 of
facing. The cost of CIP concrete facing is typically less than the cost of precast concrete
facing and as the area increases the cost decreases. At about 10,000 to 15,000 ft2, the cost of
CIP concrete wall facing may drop to $15 to $25 per ft2 of facing.
This is the final step where walls are compared to each other in a wall selection matrix
format and the wall (or walls) that have the highest score is selected for the project. The
scoring of each wall type is obtained for each wall selection factor by multiplying WR from
Table 10-1 with the 1 through 4 rating for each wall. The wall that scores the highest may be
developed as the base design and other high scoring walls may be included in the Contract
Documents as acceptable alternates.
An earth retaining wall is needed to allow construction of a road adjacent to a creek. The
road is proposed for construction to provide temporary access to remote areas in a U.S.
national forest for about 5 years while a permanent highway is built nearby. Therefore, the
earth retaining wall is proposed for temporary support for 5 years with minimum
maintenance requirements.
Cost(1) in $
Wall Type
per ft2 of entire wall(2)
Sheet-pile wall(3, 4) 15 – 40
(3, 5)
Soldier pile and lagging wall 20 – 35(6)
Slurry (diaphragm) wall(3) 70 – 120
(3)
Tangent pile wall 25 – 45
Secant pile wall(3) 30 – 50
(3)
DMM wall 45 – 60
(3)
Jet grouted wall 60 – 90
Soil-nailed wall(7) 25 - 60
Anchors(8) 30
Permanent Concrete Facing 20 – 30
1
Note: Total installed costs in 2007 U.S. dollars;
2
Costs include whole wall include embedded portion, unless otherwise noted.
3
Cost shown do not include permanent facing or anchors;
4
Lower cost is associated with renting sheet piles for temporary applications;
5
Cost shown for walls up to 16 ft;
6
Costs shown are based on wall face area above excavation bottom;
7
Cost shown do not include permanent facing; and
8
Assume average 50 ft long ground anchor “affecting” 75 ft2 of wall.
Cost(1) in $
Wall Type
per ft2 of exposed wall face
Concrete gravity wall (10 ft) 25 – 35
Concrete cantilever wall (20 ft) 20 – 35
Concrete counterforted wall 20 – 35
Concrete crib and metal bin walls 25 – 35
Gabion wall 30 – 50
Concrete module wall 30 – 35
MSE wall (precast facing) 20 – 35
MSE wall (cast-in-place facing) 25 – 45
MSE wall (modular block facing) 15 – 25
MSE wall (geotextile/geogrid/welded wire facing) 10 – 25
T-wall 30 – 40
1
Note: Total installed costs in 2007 U.S. dollars.
FHWA NHI-07-071 10 – Wall Selection
Earth Retaining Structures 10-15 June 2008
Based on the roadway profile and the site topography, the required wall height along the
alignment is estimated to be 30 ft in certain sections. The soils at the site consist of medium
dense silty sand, with zones of soft compressible clays that may cause long-term differential
settlement problems, even over the relatively short service life of the wall.
Wall selection for this example is performed following the wall selection flow chart provided
in Figure 10-1.
The owner’s evaluations of the factors that affect the wall selection are presented in Table
10-6. The ground represents a critical factor because the site consists of soils that have a
potential to undergo differential settlement. Environmental concerns are rated high because
the site is located in a U.S. national forest, which requires that any construction satisfy
specific construction permitting requirements (which may involve limitations on noise and
restrictions on working in the creek). Durability and maintenance are not considered
important due to the temporary nature of the wall. Groundwater is rated as one of the least
important factors for the project because no temporary dewatering is needed during
construction. Contracting practice and tradition are not important issues overall, however the
Owner has requested that a concrete cantilever wall be considered since standard design
details are available for this system.
For this project both cut and fill wall systems could be constructed, however due to costs and
the need for specialized equipment and labor associated with cut wall systems, fill wall
systems are chosen as the most viable for this project. Based on an initial review of the
project data, concrete walls can be eliminated immediately because of the required wall
height, relatively slow construction, and poor performance where the potential for differential
settlement exists. Furthermore, MSE walls with modular block or precast concrete facing are
also eliminated because the project calls for a temporary wall and other facing types (for
MSE walls) are more appropriate for temporary applications.
Concrete module wall, gabion wall, bin wall, and MSE wall with a wrap-around geotextile
face are chosen to be further evaluated. As presented in Step 3, the owner decided to keep
tradition as one of the factors affecting the wall selection. Therefore, to satisfy traditional
construction practice in the area, a concrete cantilever wall is also added to the list for further
evaluation.
Ground conditions at the site are known to be problematic due to their potential for
differential settlement. Therefore, walls that can typically tolerate higher differential
settlement are rated higher and walls that are more rigid and susceptible to damage caused by
differential settlement.
Construction considerations are evaluated based on the availability of material and the effort
it takes to deliver these materials to the site. Walls that require no special backfill material
and consist of wall materials that could be delivered to the site fairly easy, such as geotextile,
bin, and concrete modules, are rated high. Walls that require special material such as gabions
or suitable aggregate for concrete are rated lower than the other wall types.
The ROW for the wall is evaluated based on the space required behind the structure. Walls
that require less space are rated higher than the walls that require more space behind the
structure.
The importance of aesthetics for the wall selection is a bit subjective. Walls that have a
concrete facing are typically considered more aesthetic however, because the site is located
in a U.S. national forest, walls with a more natural appearance, such as gabions and
vegetation are rated high.
Walls that could be constructed with less potential for encroachment on the creek and less
potential for noise are rated high.
Concrete walls are known to be more durable and they are rated higher than geotextile walls
where the geotextile may degrade due to exposure to sunlight.
The final wall selection matrix is shown in Table 10-7. The scores range from 58 for
concrete cantilever wall to a high of 77 for MSE wall with a wrap-around geotextile face.
The wall type with the highest score is selected for the project and second and third ranked
wall types are considered as alternates.
WR 3 1 2 3 2 2 3 2 1 1
Durability
Construction Speed of Environmental Contracting
Ground Groundwater ROW Aesthetics and Tradition
Considerations Construction Concerns Practice
Maintenance
Bin Wall 2 3 4 3 3 1 3 4 2 3
Gabion 4 4 2 2 3 3 3 3 2 4
Wall
Concrete
Module 2 3 3 3 3 2 3 4 3 3
Wall
Concrete
Cantilever 1 1 3 1 4 4 3 4 4 4
Wall
MSE Wall
(Geotextile 3 4 4 3 4 2 3 3 1 4
wrap)
10.8.1 General
Occasionally hybrid wall systems are used for an earth retaining system project. Hybrid
walls combine elements from typical fill and cut wall systems and are effective alternatives
to fill or cut walls when they can be built higher, with less ROW restrictions, and/or less
expensively than cut or fill wall systems alone.
Hybrid walls always require special attention by the design engineer because:
• The hybrid wall may combine systems whose components require differing
magnitudes of deformation to develop resistance to loading. These differences may
lead to incompatible deformations at the wall face;
• The relatively well-defined measures available to provide internal drainage of fill and
cut wall systems may require special detailing for a hybrid wall system to provide
drainage continuity;
• While design information and performance requirements are available for individual
components of a given cut or fill wall system, these values may need to be modified
to address the performance of the complete hybrid system.
In the remainder of this section, specific issues related to the design of hybrid wall systems
are identified and selected examples of hybrid wall systems that have been constructed for
highway facilities are provided. It is recognized that other hybrid systems have been
constructed in the U.S.
Systems designed for cut support characteristically require much smaller strains to mobilize
the restraint mechanism than systems designed for fill support. Attachment of multiple
systems with incompatible strain characteristics to a common wall face can result in
overstress of the low strain elements and damaging and/or aesthetically unpleasing
differential movement of the wall face.
and
Cost
ROW
Ground
Practice
Speed of
Tradition
Concerns
Durability
Aesthetics
Contracting
Construction
Construction
Maintenance
Groundwater
Environmental
Considerations
Bin Wall 6 3 8 12 6 2 9 8 2 3 12 71
Gabion Wall 12 4 4 8 6 6 9 6 2 4 6 67
Concrete 6 3 6 12 6 4 9 8 3 3 9 69
Module Wall
Concrete
Cantilever 3 1 6 4 8 8 9 8 4 4 3 58
Wall
MSE Wall
(Geotextile 9 4 8 12 8 4 9 6 1 4 12 77
wrap)
In many cases, hybrid wall systems may consist of two walls effectively stacked on one
another to achieve a given total wall height. Two examples of such systems are shown in
Figures 10-2 and 10-3.
The wall in Figure 10-2 is a combination of a cast-in-place cantilever wall and MSE wall.
MSE walls or cantilever walls may require a slope cut behind the wall to provide for the
necessary area to construct the wall from the bottom-up. In locations where ROW
restrictions would not otherwise permit the construction of a full-height MSE wall, this
hybrid wall system can be built to reduce the ROW requirements of the wall because the
length of the footing for the lower wall can be significantly reduced as compared to a case
where the full wall height requirement is satisfied with one wall. Typically this type of
hybrid wall is built for wall heights greater than 30 ft. This hybrid wall can also be used for
road widening applications where a road exists on top of the cantilever wall.
The hybrid wall shown in Figure 10-3 is a combination of a MSE wall and cast-in-place L-
wall. This type of wall may be used for bridge abutment applications. The MSE wall serves
as the foundation for the L-wall. Without the MSE wall, the full height of the wall would be
reinforced concrete. A cost savings is therefore achieved because less concrete is required
for the wall.
In general, the design of the lower wall system for those shown in Figures 10-2 and 10-3,
requires consideration of the weight of the upper wall system. Typically, the design of the
lower wall system considers the upper wall system as a surcharge loading for stability
computations. The design of the upper wall system must consider the potential effects of
settlement of the lower wall and backfill because the lower wall provides foundation support
for the upper wall.
CANTILEVER
WALL
Figure 10-2. MSE Wall on Top of Cantilever Wall (after CDOT, 2003).
L - WALL
REINFORCEMENT
In many cases, hybrid systems such as these can be analyzed using slope stability limit
equilibrium methods as compared to earth pressure concepts usually used for conventional
cut and fill walls. Limit equilibrium analyses allow numerous potential slip surfaces to be
analyzed and can allow the various restraining forces available in a hybrid system to be
modeled within one analysis. With this, however, the design engineer needs to select an
acceptable factor of safety for the hybrid wall system which may be different from those
commonly specified for individual fill or cut walls.
This hybrid wall is a combination of a gabion wall and geogrid reinforcement of a MSE wall
(Figure 10-4). In this application, the gabion baskets may be considered as “facing” elements
for an MSE wall. Alternatively, stacked stones can also be used as facing elements. In the
stacked stone application, stones are grouted together. These types of walls are typically
used in mountainous areas where there is an abundant source of stone. These walls are used
for projects that have strict ROW requirements (i.e., available base width < 0.5 H) which
would preclude the use of gravity or MSE wall systems alone. With geogrid reinforcement,
gabion baskets can be placed on top of each other to heights greater than 26 ft. The
differential settlement tolerance of this wall is similar to the tolerances for gabion and MSE
walls, which is 1/50.
Figure 10-4. Gabion Wall Anchored with Geogrid (after CDOT, 2003).
Anchored L-Wall
This hybrid wall is a combination of a modular precast L-wall (or modular bin wall) and
geogrid reinforcement of an MSE wall (Figure 10-5). This wall is similar to a MSE wall
with precast facing elements. The modular precast L-wall facing may be a proprietary
product. The L-wall units are 2 ft tall which results in a larger reinforcement spacing than for
a typical MSE wall. Walls with L-wall facings are typically used for permanent applications.
This hybrid wall is a combination of CIP cantilever wall and ground anchors (Figure 10-6).
This type of wall is used for rehabilitation and for roadway widening applications. If the
existing CIP cantilever wall shows any signs of movement, or cracks, ground anchors may be
installed to stabilize the wall. Ground anchors are only lightly post tensioned to prevent
damage to the existing face of the CIP cantilever wall. Depending on the site constraints,
ground anchors may be installed using an over hanging drill. The post tensioning and load
tests are performed from a working platform.
Composite T-wall
This hybrid wall (Figure 10-7) is a combination of several structural systems with elements
of:
• Where existing utilities limit the space behind the wall, a composite T-wall of narrow
cross section can be built in front of the utilities. If built on soil, the wall can consists
of piles, tension anchors, and concrete filled facing units. If built on rock, the wall
can consists of horizontal or vertical rock anchors, and concrete filled facing units.
For some cut wall applications, construction is performed underneath an existing bridge.
Because headroom and equipment access is limited, soil nails can be an effective system to
provide support. However, in some states, permanent soil nails may still not be acceptable.
In that case, an effective hybrid system consists of installing temporary soil nails in the upper
portion of the cut wall followed by permanent ground anchors for the lower reaches where
access is sufficient. If soldier beams cannot be installed, a horizontal wale beam can be used
to support anchor lock-off loads. With this system, the temporary nails are assumed to carry
no loads for the permanent system.
A landslide stabilization system using tiered soil nail walls and a MSE wall was used to
stabilize an unstable slope and allow for the construction of a roadway widening for a project
FHWA NHI-07-071 10 – Wall Selection
Earth Retaining Structures 10-26 June 2008
in Wyoming (Turner and Jensen, 2005). A typical cross section is shown in Figure 10-8.
The design includes two soil nail walls and a MSE wall. The lower soil nail wall reinforces
the existing embankment and provides foundation support for the MSE wall. The upper soil
nail wall provides support for the existing roadway during and after construction of the MSE
wall. The design results in the roadway being partly supported by the upper soil nail wall
and partly by the MSE wall and with soil nails crossing the failure plane of the existing slide.
Figure 10-9 shows a hybrid system constructed for an Arizona DOT project in which a
geofoam backfill and fascia wall was constructed in front of a permanent soil nail wall. For
this project, the existing fill slope was stabilized with soil nailing. The section of the slope
below the bottom elevation of the wall fascia was very loose and unstable. An MSE wall
was considered to achieve the roadway grades, but it would have overstressed the foundation
soils. Rather than improve the foundation soils, the widening was achieved using lightweight
geofoam blocks (with an approximate unit weight of 2 pcf).
TEMPORARY
FACING
SOIL NAIL
PERMANENT
FACING
SLIP SURFACE
Figure 10-8. Soil Nail and MSE Wall (after Turner and Jensen, 2005).
CIP OR PRECAST
FASCIA WALL
SOIL NAIL
GEOFOAM BACKFILL
Figure 10-9. Soil Nail and Fascia Wall with Geofoam Backfill (after Samtani, personal
communication).
Federal Lands Highway (FLH), a program of the FHWA, is responsible for design and
construction of roadways in rugged, mountainous terrain. Where the terrain is steep,
retaining walls are frequently required to accommodate widening of existing roads, or
construction of new roadways. Recently, FLH investigated the design and performance of a
hybrid system that includes a composite MSE and shoring wall, i.e., a SMSE wall. The
report titled “Shored Mechanically Stabilized Earth (SMSE) Wall Systems” (Morrison et al,
2006) provides a design procedure for SMSE systems. Figure 10-10 shows a schematic of a
SMSE wall. It is specifically noted that actual wall dimensions including length of
reinforcement are evaluated on a project-specific basis.
As previously described in Chapter 7, a minimum bench width of 0.7 H is required for MSE
walls. Also, toe embedment is proportional to the steepness of the slope below the wall toe.
In some cases, the excavation requirements for construction of an MSE wall become
substantial and unshored excavation for the MSE wall is not practical, particularly if traffic
must be maintained during construction of the MSE wall. Shoring walls, often soil nail
walls, have been used to stabilize the backslope (or back-cut) for construction of the MSE
wall, with the MSE wall being designed and constructed in front of the shoring wall. When a
composite MSE and shoring wall system is proposed, the MSE wall component of the system
should consider the long-term retaining benefits of the shoring wall, including reduction of
lateral loads on the MSE wall mass and contributions to global stability.
FHWA NHI-07-071 10 – Wall Selection
Earth Retaining Structures 10-28 June 2008
0.6H
Shoring Wall
MSE Wall
Facing
MSE Reinforcement
0.3H
Minimum
11.1 INTRODUCTION
The purpose of this chapter is to describe contracting approaches that are commonly used in
developing construction contract documents for retaining structures. Three contracting
approaches may be used for retaining structures and are described herein. These include: (1)
method approach; (2) performance approach; and (3) contractor design/build approach. The
responsibilities of the owner and the contractor with respect to design, construction, and
performance of the wall vary for each of these approaches. All contracting approaches
should use performance based acceptance criteria for retaining structures.
All contracting approaches are valid for most earth retaining systems, if properly
implemented. Often the approach will be selected based on the experience of the owner and
their engineering consultants with wall systems, the complexity of the project (e.g.,
coordination of wall project with other contracts, high risk if unsatisfactory wall
performance, elaborate design analyses required), the availability of specialty contractors or
material suppliers, and the local highway agency philosophy with respect to contracting
methods.
FHWA NHI-07-071 11 – Contracting Approaches
Earth Retaining Structures 11-1 June 2008
In 2007, most State DOTs have a formal policy with respect to selection, design,
construction, and contracting of wall systems. The general objectives of such a policy are to:
• establish standard policies and procedures for technical review and acceptance of
proprietary and generic earth retaining systems;
• establish internal agency responsibility for the acceptance of new retaining systems
and/or components; and for the design, preparation of bid documents, and
construction control and, if necessary, performance monitoring, of such systems;
• develop uniform design and performance criteria standards and construction and
material specifications for earth retaining systems; and
11.2.1 Introduction
The method contracting approach involves the development of a detailed set of plans and
construction specifications for inclusion in the bidding documents. Depending on the
specific wall type, certain components of the work will be the responsibility of the
Contractor. The advantage of this approach is that it enables the design engineer to examine
various earth retaining system options during design, with impartiality that cannot be
expected from the contractor. The design engineer also has more time to optimize the design
and develop technical details that would minimize uncertainties and disputes during
construction.
A disadvantage of the method approach is that for alternate bids, more systems must be
evaluated, and more sets of design must be developed. Therefore, the owner’s resources may
be expended, even though only one wall system will be constructed. Another disadvantage is
that the designer may be unfamiliar with newer and potentially more cost-effective systems
and thus may not consider them during the design stage. Similarly, the proprietary wall
systems, may have technical details known only to the proprietors, thus, the owner or the
consulting engineer may not feel comfortable enough to use them.
The method contracting approach is best suited for walls supporting fill where the available
technology is either traditional or widely disseminated and reasonably well-established.
Knowledgeable contractors and material suppliers of fill-type wall systems are widespread
throughout the United States. Detailed plans and special technical provisions are often
furnished to the design engineer, at no expense, by specialty contractors and proprietary
material suppliers, especially those involved in construction of MSE wall systems.
The use of a variant to this method, in which the contractor is responsible for developing, for
example, required anchor or soil nail capacity by varying the drilling and grouting methods,
drill hole diameter, and length of anchors or nails from specified minimums, has several
advantages. It empowers contractors to maximize the use of their experience and specialized
equipment and allows the agency to share the major risk (i.e., pile capacity for a specified
length) with the contractor.
The contract documents in the method approach consist of plans, specifications, and bidding
items and quantities. In the method approach, the contract can be bid on a lump sum basis or
following a detailed unit price list.
Bidding documents (i.e., Plans and Specifications) prepared using the method approach
should typically include at least the following items:
• horizontal alignment of the wall identified by stations and offset from the horizontal
control line to the face of the wall and all appurtenances that affect construction of the
wall;
• elevation at the top and bottom of the wall, beginning and end stations for wall
construction, horizontal and vertical positions at points along the wall, and locations
and elevations of the final ground line;
• cross sections showing limits of construction, existing underground interferences such
as utilities or piles supporting adjacent structures, any backfill requirements,
excavation limits, as well as mean high water level, design high water level, and
drawdown conditions, if applicable;
In addition to the general requirements listed above, the following items are required for
specific wall types under the method contracting approach:
• length, size, and type of the gravity unit (concrete module, gabion basket, bin cell,
crib cell, etc.), and positions for which unit dimensions change;
• wall cross-section showing unit arrangements and positions for which different unit
sizes are used;
• footing location, depth, dimensions and details (removing unsuitable materials, proof
rolling, leveling, etc.);
• limits of any required wall excavations;
• properties and methods of placement of infill and backfill materials;
• corrosion protection requirements of metallic units;
• arrangement of wall face and tolerance on alignment; and
• planting or seeding requirements in the facing blocks.
• length, size, and type of soil reinforcement, and positions for which the reinforcing
elements change in length or size;
• layouts, dimensions and elevations of the footings and/or leveling pads;
• backfill soil property requirements and, if required, requirements for backfill placed
just behind the wall facing;
• horizontal alignment of wall face and offset from the horizontal control line to the
face of wall;
• alignment and elevation of internal drainage systems, and method of passing
reinforcing elements around the drainage systems;
• construction constraints, such as staged construction, vertical clearance, right-of-way
limits, etc.;
• details of facing panels and panel connections with reinforcing elements;
• details of wire-mesh reinforcement for shotcrete facing and steel reinforcement for
cast-in-place facing;
• details of architectural treatment or surface finish of the facing;
• details for construction along curved alignments, and around drainage facilities,
overhead sign footings or other structures; and
• corrosion protection requirements/details for reinforcing elements.
Slurry Walls
• dimensions of the slurry wall panel, top and bottom wall elevations, panel joints and
depth of embedment below the bottom of the excavation;
• bentonite slurry and concrete mix and concreting details in slurry walls;
• panel excavation requirements and disposal of excavated material;
• details of reinforcing steel, soldier piles, or precast concrete panels in slurry walls;
• guide wall details, width, and height and requirements for the finished face of the
guide wall
• methods of placing structural steel shapes, reinforcing steel, and concrete;
• field testing and inspection requirements;
• requirements for cleaning, patching, and sealing of leaks during construction; and
• verticality requirements of the slurry wall.
• corrosion protection requirements for the anchor head, the unbonded length, and the
anchor length;
• requirements for anchorage devices, drilling, and tendon insertion;
• requirements for bondbreaker and centralizers;
• acceptable cement grout types, water for mixing grout and grout tubes;
• requirements, details, frequency and acceptance criteria of anchor testing;
• details for facing treatment or permanent facing installation including drainage
requirement and water proofing.
11.3.1 Introduction
For the performance contracting approach, the owner establishes the scope of work and
prepares drawings showing the geometric requirements of the retaining wall, design loadings
and factors of safety, material specifications or components that may be used, performance
requirements, and any instrumentation or monitoring requirements.
The performance approach offers several benefits over the method approach when used with
appropriate specifications and prequalification of suppliers, specialty contractors, and
materials. Design of the structure is the responsibility of the contractor and is usually
performed by a trained and experienced contractor or engineering consultant. This enables
engineering costs and manpower requirements for the owner to be decreased since the
owner’s engineer is not preparing a detailed design, and transfers some of the design cost to
construction.
The disadvantage of the performance approach is that if the owner’s engineer is not
experienced with designed wall system technology, he/she may not be fully qualified to
review and approve the wall design and any construction modifications. Newer and
potentially more cost-effective methods and equipment may be rejected due to the lack of
confidence of owner personnel to review and approve these systems.
Three principal methods have been used to implement the performance approach for walls.
These methods are referred to as pre-bid wall design, pre-bid typical section design, and
post-bid design and are described in subsequent sections. Differences between these
methods are associated with the required time to perform the design.
Contract documents for pre-bid wall designs are prepared to allow for various retaining wall
alternates. With this method, the owner contacts specialty contractors and informs them that
a retaining wall is being proposed for a site. The owner requests that the contractors prepare
detailed wall designs prior to the advertisement of the bid. The designs are based on owner-
provided line and grade information, geotechnical and subsurface information, and design
requirements. Approved designs are then included in the bid documents. This approach
allows the owner to review design details based on submittals from several contractors.
Because of the detail that must be provided with this type of a submission, only those
contractors who have significant expertise and experience in proposed retaining wall are
FHWA NHI-07-071 11 – Contracting Approaches
Earth Retaining Structures 11-8 June 2008
likely to prepare the required submission. The owner should prepare and include a generic
wall system design in the bid documents to enable general contractors to decide whether they
want to use the generic design or a design from a specialty contractor.
With pre-bid typical section design, schematic or conceptual plans are developed by
prequalified specialty contractors based on geometric and performance requirements
specified by the owner. Sufficient detail must be provided by the specialty contractor to
enable the owner to judge whether the approach of the contractor is acceptable. Contractors
will typically exclude details which they believe are unique to their design. The advantage of
this approach compared to pre-bid wall design is that specialty contractors are more likely to
submit their solutions for review and inclusion in the bid documents. With this approach,
only limited preparation effort is required by the contractor, and development of a detailed
design and working drawings is only necessary if they are the successful bidder.
The disadvantage of this approach is that total project requirements are less well defined and
may lead to misunderstandings and claims. In cases where the general contractor will not be
constructing the proposed retaining system, the apparent lack of detail using this approach
may result in problems during construction because the general contractor does not fully
understand the design.
Like pre-bid wall design and pre-bid typical section design, the post-bid wall design
approach allows for various prequalified contractor-designed wall alternates. In the bid
documents, each wall and acceptable alternates are identified. Design requirements for each
wall type are contained in the special provisions or standard agency specifications. General
contractors receive bids from prequalified specialty contractors and subsequently select a
specialty contractor-prepared wall design and wall price to include in their bid. Once the
contract is awarded, if the general contractor decides to build the wall system, he/she then
requests that the selected specialty contractor prepare detailed design calculations and a
complete set of working drawings for owner review and approval. Upon approval, the walls
are built in accordance with the working drawings. When an owner uses this type of
contract, they benefit from the experience of the wall contractors or supplier. However, they
do not have as much control over the finished product as they do when they require the pre-
bid approval of the working drawings. Also, since the general contractor wants to minimize
risk, he/she will likely not select an alternate design unless the construction cost savings is
significant.
Regardless of which performance approach is used, the owner must prepare and include as
part of the contract documents geometric and site data, design guidelines, and performance
requirements. Also, for performance specifications, an instrumentation and monitoring
program may be required in which minimum levels of instrumentation to be used by the
contractor and threshold values against which the monitoring data will be evaluated are
included. Required information for inclusion in the bid package is listed below:
• horizontal alignment of the wall identified by stations and offset from the horizontal
control line to the face of the wall and all appurtenances that affect construction of the
wall;
• elevation at the top and bottom of the wall, beginning and end stations for wall
construction, horizontal and vertical positions at points along the wall, and locations
and elevations of the final ground line;
• cross sections showing limits of construction, any backfill requirements, excavation
limits, as well as mean high water level, design high water level, and drawdown
conditions, if applicable;
• all construction constraints such as staged construction limitations, vertical clearance,
right-of-way limits, construction easements, etc.;
• location of utilities, signs, etc., and any loads that may be imposed by these
appurtenances; and
• data obtained as part of a subsurface investigation and geotechnical testing program;
• design life for the earth retaining system and, if applicable, required corrosion
protection;
• tolerable horizontal and vertical movements of the structure and acceptable methods
of measuring these movements;
• required pile axial resistance and static and dynamic driven pile and drilled shaft
testing requirements for walls supported on deep foundations;
• acceptance criteria, and performance and proof testing requirements for ground
anchors , soil nails, and micropiles;
• anticipated creep behavior of anchors, soil nails, and micropiles, and methods of
measuring creep movements;
• durability requirement of jet grouted or DMM walls, and methods of testing long-
term behavior of soilcrete; and
• permissible range of variation in groundwater levels, and methods of groundwater
level measurement.
With respect to nongravity cantilevered walls, anchored walls, and soil nail walls which may
be used for temporary and/or permanent support in urban or congested areas, there are other
specific project elements that must be considered and addressed in the Contract Documents.
These include:
If a performance contracting approach is used, the contractor’s submittals are reviewed and
approved by the agency, or its consultant, before construction can commence. The
evaluation by the agency’s structural and geotechnical engineers must be rigorous and must
consider, as a minimum, the following items:
For the contracting approaches previously described, the owner and contractor share
responsibility in the design and construction of the wall system. With the contractor
design/build method, the owner outlines the project requirements, obtains complete
subsurface and geotechnical information, and provides construction quality assurance. The
design-build contractor is responsible for the complete design, construction, and performance
of the wall system. A design/build proposal may be submitted either before the bid
advertisement (pre-bid) or after the contract award (post-bid). This method is most often
used for securing bids on temporary retaining wall projects. The key elements for a
successful contract are communication of basic design concepts to the owner and the joint
development of a quality assurance plan prior to construction.
In 1994, the Civil Engineering Research Foundation (CERF) established the Highway
Innovative Technology Evaluation Center (HITEC). HITEC’s purpose was to accelerate the
introduction of technological advances in products, systems, services, materials, and
equipment to the highway and bridge markets. The evaluation of new and more cost-
effective retaining wall systems was performed through HITEC’s nationally-focused, earth
retaining system (ERS) group evaluation program. While the HITEC program is still
available, FHWA funding of the program has been reduced. Wall system suppliers are
encouraged to conduct an independent review of newly developed components and/or
systems related to materials, design, construction, performance, and quality assurance. As
many public agencies, especially state DOTs, require HITEC evaluations or independent
evaluations of wall components or wall systems, suppliers should consider obtaining such
reviews as it will be beneficial in securing acceptance of their system.
Various wall specifications are available from state DOTs either as part of their standard
specifications or from Special Provisions from individual projects. Table 11-1 provides a
partial listing of websites and sources to obtain specifications for wall types covered in this
manual (at the time of the Manual preparation). This list is by no means exhaustive.
Specification Sources
Wall Types
DOT Agencies Others
Cast-in-Place Gravity and
Wisconsin -
Semi-Gravity Walls
New York
Crib Walls Wisconsin -
California
Connecticut
Gabion Walls Montana -
Wisconsin
Contech
Bin Walls Wisconsin
(http://www.contech-cpi.com)
Georgia
Concrete Module Walls (Doublewal) Doublewal Corporation
New York
Illinois
NHI Course Manual
MSE Walls Montana
(FHWA-NHI-00-043)
Wisconsin
Neel Company
T-Walls -
(www.neelco.com)
Illinois
Sheet Pile Walls Wisconsin -
California
Montana
Soldier Pile and Lagging
Wisconsin -
(Post and Panel) Walls
Illinois
Central Artery Project
(www.bigdig.com)
Slurry (Diaphragm) Walls -
Port Authority of New York and
New Jersey
Wisconsin
Tangent/Secant Pile Walls -
(Marquette Interchange Project)
Nicholson Construction
Jet Grouted Wall - www.nicholson-rodio.com
ACI (1998). American Concrete Institute. Code Requirements for Nuclear Safety-Related
Concrete Structures (ACI 349-97) and Commentary. Farmington Hills, MI: p.129.
ASCE (1997). American Society of Civil Engineers. Guidelines of Engineering Practice for
Braced and Tied-Back Excavations. Geotechnical Special Publication No. 74.
ASCE (1994). American Society of Civil Engineers. Guidelines for Braced and Tie-back
Excavations.
Allen, T.M. (2005). “Development of Geotechnical Resistance Factors and Downdrag Load
Factors for LRFD Foundation Strength Limit State Design”, Federal Highway
Administration, Report No. FHWA-NHI-05-052.
Ashour, M., Norris, G.M., and Piling, P. (1998). “Lateral Loading of a Pile in Layered Soil
Using the Strain Wedge Model.” Journal of Geotechnical and Geoenvironmental
Engineering, ASCE, Vol. 124, No. 4, pp. 303-315.
Becker, D.E., Crooks, J.H.A., Been, K. and Jefferies, M.G. (1987). “Work as a Criterion for
Determining In-Situ and Yield Stresses in Clays.” Canadian Geotechnical Journal Vol.
24, No. 4, pp. 549-564.
Bieniawski, Z.T. (1974). “Geomechanics Classification of Rock Masses and its Application
in Tunnelling.” Proceedings of the 3rd International Congress on Rock Mechanics,
Denver, Vol. 2, No. 2, pp 27-32.
Bolton, M.D. (1986). “The Strength and Dilatancy of Sands.” Geotechnique Vol. 36, No. 1,
pp. 65-78.
Bowles, J. E. (1977). Foundation Analysis and Design. New York, McGraw-Hill Publication.
Bowles, J. E. (1988). Foundation Analysis and Design, Fourth Edition. New York, McGraw-
Hill Book Company.
Brandl, H. (1992). Retaining Structures for Rock Masses, Engineering in Rock Masses.
Boston, MA, Buttewroth-Heinemann Limited, Chapter 26, pp. 530-572.
Broms, B.B. (1965). “Design of Laterally Loaded Piles.” Journal of the Soil Mechanics and
Foundations Division Vol. 91, No. SM3, Proceedings Paper 4342. ASCE, pp. 79-99.
Bruce, D. A. (1994). “Jet grouting - Chapter 8.” Ground control and improvement. New
York, NY: John Wiley & Sons, Inc.
Byrne, R.J., Cotton, D., Porterfield, J., Wolschlag, C., and Ueblacker, G. (1998). Manual for
Design and Construction Monitoring of Soil Nail Walls. Federal Highway
Administration, FHWA-SA-96-069.
Caquot, A., and Kerisel, F. (1948). “Tables for the calculation of passive pressures.” Active
Pressure and Bearing Capacity of Foundations, Paris, France: Gautheir-Villars.
CDOT (2003). “Feasibility of a Management System for Retaining Walls and Sound
Barriers.” Colorado Department of Transportation Research Report by George Hearn,
Report No. CDOT-DTD-R-2003-8, p. 106.
Cedergren, H. R. (1989). Seepage, Drainage, and Flow Nets 3rd Edition. John Wiley & Sons.
Cheney, R.S. (1990). “Selection of retaining structures: The Owner’s Perspective.” Proc. of
Conf. on Design and Performance of Earth Retaining Structures Geotechnical Special
Publication No. 25, Cornell University, Ithaca, P. C. Lambe and L. A. Hansen, eds.,
ASCE, New York, NY, pp. 52-65.
Christopher, B. R., Gill, S. A., Giroud, J. P., Juran, I., Mitchell, J. K., Schlosser, F., and
Dunnicliff, J. (1989). Reinforced Soil Structures, Vol. 1, Design and Construction
Guidelines, FHWA-RD-89-043.
Clough, G. W., Smith, E. M., and Sweeney, B. P. (1989). “Movement Control of Excavation
Support Systems by Iterative Design.” Proc., Foundation Engineering: Current
Principles and Practices, Vol. 2. ASCE, pp. 869-884.
Collin, J.G., Berg, R.R. and Meyers, M. (2002). Segmental Retaining Wall Drainage
Manual. Herdon, Virginia: National Concrete Masonry Association.
CWALSHT (1995). Computer Program to Design Sheet Pile Walls, U.S. Army Corps of
Engineers.
Das, B. M. (1990). Principles of Foundation Engineering, Second Edition. Boston, MA: PWS-
KENT Publ.
DePaoli, B., Stella, C., and Perelli, C. A. (1991). “A Monitoring System for the Quality
Assessment of the Jet Grouting Process Through an Energy Approach.” Proc., 4th Int. Conf.
on Piling and Deep Foundations, April 7-12. Stresa, Italy.
Duncan, J.M., Clough, G.W., and Ebeling, R.M. (1990). "Behavior and Design of Gravity Earth
Retaining Structures." Design and Performance of Earth Retaining Structures. American
Society of Civil Engineers, Geotechnical Special Publication No. 25, pp. 251-277.
Duncan, J. M., Williams, G. W., Sehn, A. L. and Seed, R. B. (1991). “Estimated Earth Pressures
Due to Compaction.” Journal of Geotechnical Engineering Division, 117(2), ASCE, New
York, NY, pp. 1833-1847.
Duncan, J. M., Williams, G. W., Sehn, A. L. and Seed, R. B. (1993). “Estimated Earth Pressures
Due to Compaction”, Journal of Geotechnical Engineering Division, 119(7), ASCE, New
York, NY, pp. 1162-1176.
Elias, V., Welsh, J., Warren, J., Lukas, R., Collin, J.G. and Berg, R.R. (2006). “Ground
Improvement Methods” U.S. Department of Transportation, Federal Highway
Administration, FHWA-NHI-06-020, Washington, D.C., 884 pp.
Elias, V., Salman, I., Juran, E., Pearce, E., and Lu, S. (1997). “Testing Protocols for Oxidation
and Hydrolysis of Geosynthetics”, Federal Highway Administration, FHWA RD-97-144,
Washington, D.C.
Elias, V., and Juran, I. (1991). “Soil Nailing for Stabilization of Highway Slopes and
Excavations.” Federal Highway Administration, FHWA-RD-89-198, Washington, D. C.
FHWA (1988). “Checklist and Guidelines for Review of Geotechnical Reports and Preliminary
Plans and Specifications.” Geotechnical Guideline No. 2, Federal Highway Administration,
FHWA ED-88-053 (Revised in February 2003), Washington, D.C.
Franklin, A.G. and Chang, F.K. (1977). “Earthquake Resistance of Earth and Rock-Fill Dams,”
Report 5: Permanent Displacement of Earth Embankments by Newmark Sliding Block
Analysis, Misc. Paper 5-71-17, Soils and Pavements Laboratory, US Army Waterways
Experiment Station, Vicksburg, MS.
Gallavresi, F. (1992). “Grouting improvement of foundation soils.” Proc., ASCE Conf. Grouting,
Soil Improvement and Geosynthetics, 2 Vols., Feb 25-28, New Orleans, LA, pp. 1-38.
Godfrey, K. (1987). “Subsiding Problems with Slurry Walls.” Journal of Civil Engineering,
January, pp. 56-59.
Goldberg, D. T., Jaworski, W. E., and Gordon, M. D. (1976). “Lateral Support Systems and
Underpinning, Design and Construction.” Federal Highway Administration, Vol. 1, FHWA-
RD-75-128, Washington, D. C.
Gould, J. P. (1990). “Earth Retaining Structures - Developments Through 1970.” Proc. of Conf.
on Design and Performance of Earth Retaining Structures, Geotechnical Special Publication
No. 25, Cornell University, Ithaca, P. C. Lambe and L. A. Hansen, eds., ASCE, New York,
NY, pp. 8-21.
Henkel, D.J. (1971). “The Calculation of Earth Pressures in Open Cuts in Soft Clays.” The
Arup Journal, Vol. 6, No. 4, pp. 14-15.
Hetenyi, M. (1946). Beams on Elastic Foundation. Ann Arbor, Michigan: The University of
Michigan Press, University of Michigan Studies.
HITEC (1998). “Guidelines for Evaluating Earth Retaining Systems”, ASCE Publication, CERF
Report 40334, 32 pp.
HKGEO (1993). Guide to retaining wall design - GEOGUIDE 1. Hong Kong, China:
Geotechnical Engineering Office, Civil Engineering Department.
Hoek, E. (1983). “Strength of Jointed Rock Masses.” Geotechnique, Vol. 33, No. 3, pp. 187-
223.
Hoek, E. and Bray, J. W. (1977). “Rock Slope Engineering.” Institution of Mining and
Metallurgy, London, U.K.
Hoek, E. and Brown E. T. (1988). “The Hoek-Brown Failure Criterion – a 1988 Update.”
Proceedings, 15th Canadian Rock Mechanics Symposium, Toronto, Canada.
Hoek, E. and Brown, E. T. (1997). “Practical Estimates of Rock Mass Strength.” International
Journal of Rock Mechanics and Mining Science, www.rocscience.com.
Hoek, E., Carranza-Torres, C., and Corkum, B. (2002). “Hoek-Brown Criterion – 2002 Edition.”
Proceedings NARMS_TAC Conference, Toronto, Vol. 1, pp. 267 - 273.
Holtz, R.D., Christopher, B.R. and Berg, R.R. (2007). “Geosynthetic Design and Construction
Guidelines.” U.S. Department of Transportation, Federal Highway Administration, FHWA-
HI-07-092, Washington, D.C.
ISRM (1981). Suggested Methods for the Quantitative Description of Discontinuities in Rock
Masses. UK: Pergamon Press, 1981.
Jaeger, J.C., and Cook, N.G.W. (1976). Fundamentals of Rock Mechanics. London, U.K:
Chapman & Hall, 99 pp.
Kauschinger, J. L., Hankour, R., and Perry, E. B. (1992). “Methods to estimate composition of
jet grout bodies.” Proc., ASCE Conf. Grouting, Soil Improvement and Geosynthetics, 2, Feb.
25-28, New Orleans, LA, pp. 194-205.
Kauschinger, J. L., and Welsh, J. P. (1989). “Jet grouting for urban construction.” Proc.,
Geotechnical Lecture Series, Boston Society of Civil Engineering: Design Construction and
Performance of Earth Support Systems, MIT, Nov. 18, Boston Society of Civil Engineers
Section (ASCE), Boston, MA,
Kavazanjian E. Jr., Matasovic, N., Hadj-Hamou, T., and Sabatini, P. J. (1997). “Design
Guidance: Geotechnical Earthquake Engineering for Highways.” Geotechnical Engineering
Circular No. 3, Volume I and II, Federal Highway Administration, Report No. FHWA-SA-
97-077, Washington, D.C.
Kerisel, J. (1992). “History of retaining wall designs.” Proc. of Conf. on Retaining Structures,
Institute of Civil Engineers, C.R.I. Clayton, ed., Thomas Telford, Robinson College,
Cambridge, England, pp. 1-16.
Kingsbury, D.W., Sandford, T.C., and Humphrey, D.N. (2002). “Soil Nail Forces Caused by
Frost.” Proceedings of the Transportation Research Board, Annual Meeting, Washington,
D.C.
Koerner, J., Soong, T-Y, and Koerner, R. M. (1998). “Earth Retaining Wall Costs in the USA.”
GRI Report #20, Geosynthetics Institute, Folsom, PA.
Kramer, S.L. (1996). Geotechnical Earthquake Engineering. Upper Saddle River, New Jersey:
Prentice-Hall Inc.
Lazarte, C. A., Elias, V., Espinoza, R. D., Sabatini, P.J. (2003). “Soil Nail Walls”, Geotechnical
Engineering Circular, No. 7, Federal Highway Administration Report, Report No. FHWA-
IF-03-017, Washington, D.C.
Macklin, P. R., Berger, D., Zietlow, W., Herring, W., and Cullen, J. (2004). “Case History:
Micropile Use for Temporary Excavation Support”, Geotechnical Special Publication,
Drilled Shafts, Micropiling, Deep Mixing, Remedial Methods, and Specialty Foundation
Systems, ASCE Publication, No. 124, pp. 653-661.
Marinos, V., Marinos, P., Hoek, E. (2004). “The Geological Strength Index: Applications and
Limitations.” Bulletin of Engineering Geology and the Environment.
Marinos, P. and Hoek, E. (2000). “GSI: A Geologically Friendly Tool for Rock Mass Strength
Estimation.” Proceedings of the GeoEng 2000, International Confernece on Geotechnical and
Geological Engineering, Melbourne, Technomic Publishers, Lancaster, pp. 1422 -1446.
Mesri, G., and Abdel-Ghaffar, M.E.M. (1993). “Cohesion Intercept in Effective Stress Stability
Analysis.” Journal of Geotechnical Engineering, Vol. 119, No. 8, pp. 1229-1249.
FHWA NHI-07-071 12 – References
Earth Retaining Structures 12-8 June 2008
Meyerhof, G. G. (1976). “Bearing Capacity and Settlement of Pile Foundations.” Proceedings of
American Society of Civil Engineers, Journal of Geotechnical Engineering Division, V. 102
pp. 197-228.
Meyerhof, G.G. (1951). “The Ultimate Bearing Capacity of Foundations.” Geotechnique Vol. 2
pp. 301-332.
Morrison, K. F., Harrison, F. E., Collin, J. G., Dodds, A., and Arndt, B. (2006). “Shored
Mechanically Stabilized Earth (SMSE) Wall Systems”, Technical Report, Federal Highway
Administration, Washington, D.C., Report No. FHWA-CFL/TD-06-001, 212 p.
Mueller, C.G., Long, J.H., Weatherby, D.E., Cording, E.J., Powers III, W.F., and Briaud, J.-L.
(1998). “Summary Report of Research on Permanent Ground Anchor Walls, Volume III:
Model-Scale Wall Test and Ground Anchor Tests.” Federal Highway Administration,
Washington DC, Report No. FHWA-RD-98-067, 197 p.
Munfakh, G., Abramson, L., Barksdale, R., and Juran, I. (1987). “In-Situ Ground
Reinforcement.” ASCE Geotechnical Special Publication, No. 12, New York, NY, pp. 1-66.
Munfakh, G., Arman, A., Collin, J.G., Hung, J. C-J. and Brouillette, R.P. (2001). “Shallow
Foundations, NHI Course No. 132037 Module 7 Reference Manual.” Federal Highway
Administration, FHWA NHI-01-023, Washington, D.C., 222 p.
NAVFAC (1986). Foundations and Earth Structures, Naval Facilities Engineering Command,
Design Manual 7.02. Washington, D. C: U.S. Government Printing Office.
O’Neil, M. W. and Reese, L. C. (1999). “Drilled Shafts: Construction Procedures and Design
Methods.” Federal Highway Administration, FHWA-HI-99-025, Washington, D.C.
O’Rourke, T. D., and Jones, C.J.F.P. (1990). “Overview of earth retention systems: 1970-1990.”
Proc. of Conf. on Design and Performance of Earth Retaining Structures, Geotechnical
Special Publication No. 25, Cornell University, Ithaca, P. C. Lambe and L. A. Hansen, eds.,
ASCE, New York, NY, pp. 22-51.
Okabe, S. (1926). “General Theory of Earth Pressures.” Journal of the Japan Society of Civil
Engineering, Vol. 12, No. 1.
Padfield, C. J. and Mair, R. J. (1984). “Design of Retaining Walls Embedded in Stiff Clay”,
Construction Industry Research and Information Association (CIRIA) Publication, Report
No. 104, London, England.
Pearlman, S.L., Walker, M.P., and Boscardin, M.D. (2004). “Deep Underground Basements fro
Major Urban Building Construction.” GeoSupport 2004, Geotechnical Special Publication
No. 124, ASCE, pp. 545-560.
Peck, R. B. (1969). “Deep excavation and tunneling in soft ground.” Proc., 7th Int. Conf. on Soil
Mech. and Foundation Engrg. State-of-the-Art Volume, Mexico City, Mexico, pp. 225-290.
Peck, R., Hanson, W. E., and Thornburn, T. H. (1974). Foundation Engineering. Wiley.
Pile Buck (2004). Computer Program for Sheet Pile Wall Design, SPW911, Version 2.2, Pile
Buck Inc., Vero Beach, FL.
Porterfield, J. A., Cotton, D. M. and Byrne, R. J. (1994). “Soil Nailing Field Inspectors
Manual.” Federal Highway Administration, FHWA-SA-93-068, Washington, D. C.
PTI (1996). Recommendations for Prestressed Rock and Soil Anchors. Post Tensioning Institute
Phoenix, AZ.
Richards R. Jr. and Elms, D.G. (1979). “Seismic Behavior of Gravity Walls.” Journal of the
Geotechnical Engineering Division, ASCE, Vol. 105, No. GT4, pp. 449-464.
Robertson, P.K., and Campanella, R.G. (1983). “Interpretation of Cone Penetration Tests.”
Canadian Geotechnical Journal, Vol. 20, No. 4, pp. 718-754.
Rowe, P. W. (1952). “Sheet pile walls encastre at anchorage.” Proc., Inst. of Civil Engrg., Part
I, Vol. I., London, England.
Sabatini, P. J., Tanyu, B. F., Armour, T., Groneck, P., and Keeley, J. (2005). “Micropile Design
and Construction Manual.” National Highway Institute Manual, Report No. FHWA NHI-05-
039, Federal Highway Administration, Washington, D.C.
Sabatini, P. J., Bachus, R. C., Mayne, P. W., Schneider, J. A., and Zettler, T. E. (2002).
“Evaluation of Soil and Rock Properties.” Geotechnical Engineering Circular, No. 5, U.S.
Department of Transportation, Federal Highway Administration, FHWA-IF-02-034,
Washington, D.C.
Sabatini, P. J., Pass, D. G. and Bachus, R. C. (1999). “Ground Anchors and Anchored Systems.”
Geotechnical Engineering Circular, No. 4, U.S. Department of Transportation, Federal
Highway Administration, FHWA-SA-99-015.
Sabatini, P. J., Elias, V., Schmertmann, G. R., and Bonaparte, R. (1997). “Earth Retaining
Systems.” Geotechnical Engineering Circular, No. 2, U.S. Department of Transportation,
Federal Highway Administration, FHWA-SA-96-038, Washington, D.C.
Samtani, N. C. (2007). “Limit States and Interaction between Structural and Geotechnical
Specialists”, Article # 01-0507-R0, Free copy available for download from
www.ncsconsultants.com (Contact the author at NSamtani@msn.com if a copy cannot be
downloaded from the referenced website).
Schmidt, B. (1966). “Earth Pressures At-Rest Related to Stress History.” Canadian Geotechnical
Journal, V. 3(4), pp. 239-242.
Schofield, A.N., and Wroth, C.P. (1968). Critical State Soil Mechanics. McGraw-Hill, London,
U.K., 310 pp.
FHWA NHI-07-071 12 – References
Earth Retaining Structures 12-11 June 2008
Seed, H.B., Woodward, R.J., and Lundgren, R. (1962). “Prediction of Swelling Potential for
Compacted Clays.” Journal of Soil Mechanics and Foundation Division, ASCE, Vol. 88, No.
3, pp. 53-87.
Simac, M. R., Bathurst, R. J., Berg, R. R., and Lothspeich, S. E. (1997). Design manual for
segmental retaining walls, Second Edition. Ed. Collin, J. G. National Concrete Masonry
Association.
Sisson, R.C., Harris, C.J., and Mokwa, R.L. (2004). “Design of Cantilevered Soldier Pile
Retaining Walls in Stiff Clays and Claystones.” GeoSupport 2004, Geotechnical Special
Publication No. 124, ASCE, pp. 309-321.
Sokolovski, V.V. (1954). Statics of Soil Media (translated from Russian by D.H. Jones and A.N.
Schofield). Butterworths, London.
Spangler, M. G. and Handy, R. L. (1984). Soil Engineering. New York, NY: Harper and Row.
Stark, T.D., and Eid, H.T. (1994). “Drained Residual Strength of Cohesive Soils.” Journal of
Geotechnical Engineering, Vol. 120, No. 5, pp. 856-871.
Taki, O., and Yang, D. S. (1991). “Excavation Support and Groundwater Control Using Soil-
Cement Mixing Wall for Subway Projects.” Proc. RETC, June 11-14, Los Angeles, Calif.,
R. A. Pond and P. B. Kenny, eds., Society for Mining, Metallurgy, and Exploration, Inc
(SME), Littleton, CO, pp. 164-165.
Tamaro, G. J. (1990). “Slurry wall design and construction.” Proc. of Conf. on Design and
Performance of Earth Retaining Structures, Geotechnical Special Publication No. 25, Cornell
University, Ithaca, P. C. Lambe and L. A. Hansen, eds., ASCE, New York, NY, pp. 540-550.
Tamaro, G. J., and Poletto, R. J. (1992). “Slurry Walls - Construction Quality Control.” Slurry
Walls: Design, Construction, and Quality Control. Eds. David B. Paul, Richard R. Davidson
and Nicholas J. Cavalli. Philadelphia, Pennsylvania: ASTM STP 1129, American Society for
Testing and Materials.
Terzaghi, K., Peck, R.G., and Mesri, G. (1996). Soil Mechanics in Engineering Practice. New
York, NY: John Wiley & Sons, Inc., pp. 549.
Technical European Sheet Piling Association (TESPA) (2001). Installation of Sheet Piles.
TRC/Imbsen & Associates (2006). Recommended LRFD Guidelines for the Seismic Design
Highway Bridges. NCHRP Project 20-07, Task 193.
Turner, J.P. and Jensen, W.G. (2005). “Landslide Stabilization Using Soil Nail and
Mechanically Stabilized Earth Walls: Case Study.” Journal of Geotechnical and
Geoenvironmental Engineering, ASCE, Vol. 131, No. 2, pp. 141-150.
Ueblacker, G. (1997). “Portland Westside Lightrail Corridor Project Wall 600 Micropile
Retaining Wall Design and Construction.” ADSC Technical Library Report, TL123.
US Army Corps of Engineers (1993). Engineer Manual for Retaining and Flood Walls, Manual
No. EM 1110-2-2502, Washington, D.C.
Vidal, H. (1966). “La Terre Armee.” Annales de L’Institute Technique du Batiment et Travaux
Publics, Vol. 19, pp. 223-224.
Wang, S-T. and Reese, L.C. (1986). “Study of Design Methods for Vertical Drilled Shaft
Retaining Walls.” Texas State Department of Highways and Public Transportation, Austin,
TX.
Weatherby, D.E. (1998). “Design Manual for Permanent Ground Anchor Walls.” Federal
Highway Administration, Report FHWA-RD-97-130, McLean, VA.
Weatherby, D.E., Chung, M., Kim, N.-K., and Briaud, J.-L. (1998). “Summary Report of
Research on Permanent Ground Anchor Walls, Volume II: Full-Scale Wall Tests and a Soil-
Structure Interaction Model.” Federal Highway Administration, Report FHWA-RD-98-066,
McLean, VA.
Welsh, J. P., Rubright, R. M., and Coomber, D. B. (1986). “Jet grouting for support of
structures.” Proc. ASCE Conf. Grouting for the Support of Structures, April 16, Seattle,
WA.
Wood, D.M. (1990). Soil Behavior and Critical State Soil Mechanics. Cambridge University
Press, U.K., 462 pp.
Wyllie, D. (1999). Foundations on Rock. Routledge, New York, N.Y. 401 pp.
Wyllie, D. and Mah, C. W. (1998). “Rock Slopes.” National Highway Institute Course Manual,
Federal Highway Administration, FHWA HI-99-007, Washington, D.C.
Xanthakos, P. P. (1994). Slurry Walls as Structural Systems. Second Edition. New York, NY,
McGraw-Hill.
Zornberg, J. and Mitchell, J.K. (1992). "Poorly Draining Backfills for Reinforced Soil
Structures-A State of the Art Review." Geotechnical Engineering Report No. UCB/GT/92-
10, University of California, Berkeley.
STUDENT EXERCISES
AND
SOLUTIONS
CONTENTS
# Title
S1 Active Earth Pressure Calculation
S2 Lateral Load Distribution Due to Strip Load
S3 Sliding Resistance
S4 MSE Wall Design External Stability
S5 MSE Wall Internal Stability
S6 Corrosion Calculation
S7 Apparent Earth Pressure Diagrams
S8 Single-Tier Anchored Soldier Beam and Lagging Wall
S9 Design of a Two-Tier Anchored Soldier Pile and Lagging Wall
S10 Lateral Wall Movement
S11 Soil Nail Wall Design
S12 Fill Wall Selection
S13 Cut Wall Selection
Manual Reference:
S1-1
Summary of Equations
S1-2
Top
S1-3
SOLUTION TO STUDENT EXERCISE 1
Top:
σ 'v ,top = γz = 121 pcf (0) = 0
Middle:
σ 'v , mid = γz = 121 pcf (10 ft ) = 1210 psf
Bottom:
σ 'v , bot = ∑ γ i zi − u
σ 'v , bot = γ 1 z1 + γ 2 z2 − γ w zw
σ 'v , bot = 121 pcf (10 ft ) + 120 pcf (10 ft ) − 62.4 pcf (10 ft )
σ 'v , bot = 1786 psf
Layer 1: φ’ = 33°
φ' 33
K a1 = tan (45 −
2
) = tan 2 (45 − )
2 2
K a1 = 0.295
S1-4
Layer 2: φ’= 28°
φ 28
K a2 = tan 2 ( 45 − ) = tan 2 ( 45 − )
2 2
K a2 = 0.36
Top:
σ ' h ,top = σ 'v ,top K a 1
σ 'h,top = 0(0.295) = 0
Middle:
Bottom of layer 1
σ 'h, mid − = σ 'v , mid K a 1
Top of layer 2
σ 'h , mid + = σ 'v , mid K a 2
Bottom:
σ 'h, bot = σ 'v , bot K a2
S1-5
Draw the pressure diagram:
10 ft
1
P1
436 psf
357 psf
PA
10 ft
2
P2
Note: Not
to Scale
643 psf
P1 =
1
(357 psf )(10 ft ) = 1785 lb
2 ft
Resultant force for area 2 by calculating the area of a trapezoid or area
of a triangle and rectangle:
P2 =
1
(436 psf + 643 psf )(10 ft ) = 5395 lb
2 ft
(Area of Trapezoid)
S1-6
Calculate the total resultant active force considering
drained strength parameters:
PA = P1 + P2
lb lb lb
PA = 1785 + 5395 = 7180
ft ft ft
b) Solution (Undrained Strength for Lower
Clay)
Middle:
σ v , mid = σ 'v , mid = 1210 psf
Bottom:
σ v , bot = ∑ γ i zi
σ v , bot = γ 1 z1 + γ 2 z2
σ v , bot = 121 pcf (10 ft ) + 120 pcf (10 ft )
σ v , bot = 2410 psf
S1-7
Middle:
Bottom of layer 1
σ h, mid − = σ v , mid K a = 357 psf
1
(same as part a)
Top of layer 2
σ h , mid + = σ v , mid − 2Su
σ h , mid + = 1210 psf − 2 (630 psf )
σ h , mid + = −50 psf
Bottom:
σ h , bot = σ v , bot − 2Su
σ h , bot = 2410 psf − 2(630 psf )
σ h , bot = 1,150 psf
Pressure diagram:
10 ft
1
P1
-50 psf 0
D
2 357 psf
10 ft
PA
3 P3
Note: Not
to Scale
1150 psf
S1-8
Calculate Distance (D) by using similar triangles:
50 1150
=
D (10 − D)
1150 D = 50(10 − D)
1150 D = 500 − 50 D
1150 D + 50 D = 500
500
D= = 0.42 ft
1200
P1 =
1
(357 psf )(10 ft ) = 1785 lb (area of triangle)
2 ft
Resultant force for area 3:
P3 =
1
(1150 psf )(10 ft − 0.42 ft ) = 5509 lb (area of triangle)
2 ft
PA = P1 + P3
lb lb lb
PA = 1785 + 5509 = 7294
ft ft ft
S1-9
STUDENT EXERCISE 2
Lateral Load Distribution Due to Strip Load
Manual Reference:
S2-1
3 ft 5 ft 800 psf
1
γ1 = 121 pcf φ1′ = 33°, c1′ = 0
10 ft Layer 1
2
γ2 = 120 pcf φ2′ = 28°, c2′ = 0
10 ft Layer 2
3
Figure S2-1. Excavation through two layers of soil with water table located at
10 ft below the ground surface.
S2-2
SOLUTION TO STUDENT EXERCISE 2
σ v = 0 = σ v' 1
1
σ v = 2410 psf
3
σ '
v3 = 2410 − 62.4 pcf (10 ft )
σ v' = 1786 psf
3
Pa 1 = 0
2.- Pa = 357 psf
2−
1. Pw1 = 0
S2-3
2. Pw2 = 0
3. Pw3 = hw ⋅ 62.4 pcf
Pw3 = 10 ft ⋅ 62.4 pcf = 624 psf
a) z = 2 ft
From Figure 3-24
Ph =
2q
(β − sin β cos 2α )
π
S2-4
Therefore,
Ph =
2(800 psf )
[0.34 − sin(0.34) cos(2 ⋅ (1.15))]
π
= 509 [0.34 − (−0.22)]]
= 290 psf
b) Repeat calculations at z = 20 ft
Compute β and α:
⎛ 8 ⎞ ⎛ 3⎞
β = tan −1 ⎜ ⎟ − tan −1 ⎜ ⎟
⎝ 20 ⎠ ⎝ 20 ⎠
= 0.38 - 0.15
= 0.23 radians
⎛ 3 ⎞ β
α = tan −1 ⎜ ⎟ +
⎝ 20 ⎠ 2
0.23
= 0.15 +
2
= 0.26 radians
Therefore,
Ph6 m =
2(800 psf )
[0.23 − sin(0.23) cos(2(0.26))]
π
= 17 psf
S2-5
Pa Ph Pw
0 200 400 600 800 0 100 200 300 400 0 200 400 600 800
0 0 0
2 290 psf
2 2
4 4 4
6 6 6
8 8 8
Depth (ft)
357 psf
Depth (ft)
Depth (ft)
10 10 10
436 psf
12 12 12
14 14 14
16 16 16
18 18 18
17 psf
20 20 20
643 psf 624 psf
Figure S2-2. Pressure diagrams due to lateral soil pressure Pa, strip loading Ph,
and hydrostatic water pressure Pw.
S2-6
Combine active earth pressure and pressure due to strip loading:
Pa+Ph Pw
0 200 400 600 800 0 200 400 600 800
0 0
2 2
4 4
6 6
Depth (ft)
8 8
Depth (ft)
444 psf
10 10
522 psf
12 12
14 14
16 16
18 18
20 20
660 psf 624 psf
Figure S2-3. Combined active earth pressure and pressure due to strip loading
plus the hydrostatic water pressure.
S2-7
STUDENT EXERCISE 3
Manual Reference:
S3 - 1
1.5 ft
β = 10°
Not to Scale
(DCstem)
γb =115pcf ∆h
γconc =145pcf φ’b = 29° (Groups 1 and 2)
φ’b = 32° (Group 3)
h = 20 ft
DC EV PV EH
δ = 10°
PH
θ = 90°
5 ft
S3 - 2
Summary of Equations
Sliding Equations
Active Earth Pressure Coefficient:
sin 2 (θ + φ ' )
Ka = b
Γ sin 2 θ sin(θ − δ ) (Figure 3-5)
2
⎡
⎢
sin(φ ' +δ ) sin(φ ' −β ) ⎤⎥
Γ = ⎢1 + b b
⎥
⎢ sin(θ − δ ) sin(θ + β ) ⎥
⎣ ⎦
β = Angle of backfill with horizontal = δ (Assumed)
δ = Orientation of the resultant active force. It is assumed that resultant
active force is oriented at the same angle as backfill)
φ’b = Effective internal friction angle of backfill
θ = Angle of back wall face to horizontal
Bearing Resistance:
q = ϕ qn (Eq 7-7)
R b
ϕb = Bearing resistance Factor
qn = 0.5γ B' NγmC wγ (B’ is based on AASHTO C10.6.3.1.1)
f
γf = total unit weight of foundation soil
B’= B – 2eB
Nγm = Nγ sγ iγ (From AASHTO section 10.6.3)
Nγ = Bearing capacity factor
⎛B⎞
Sγ = 1− 0.4⎜⎜ ⎟⎟ = footing shape factor
⎝L⎠
L = length of wall
iγ = load inclination factor
Cwγ = ground water table correction factor
S3 - 4
SOLUTION TO STUDENT EXERCISE 3
H = 2.5 ft + 20 ft + ∆h
H = 2.5 ft + 20 ft + 8.0 ft tan10°
H = 23.9 ft
sin 2 (θ + φ ' )
Ka = b β = 10°
Γ sin 2 θ sin(θ − δ ) δ = 10°
2 where: φ ' = 29°
⎡
⎢
sin(φ ' +δ ) sin(φ ' −β ) ⎥
⎤ b
Γ = ⎢1 + b b θ = 90°
⎥
⎢ sin(θ − δ ) sin(θ + β ) ⎥
⎣ ⎦
K a = 0.365
1
EH = K a H 2γ
2 b
1
EH = (0.365)(23.9 ft ) 2 (115 pcf )
2
EH = 11,988 lb
ft
S3 - 5
Resolve active earth pressure into horizontal (PH) and vertical
(PV) components:
P = EH cos δ
H
P = 11,988 lb cos10° = 11,806 lb
H ft ft
(see Figure S3-1)
P = EH sin δ
V
P = 11,988 lb sin 10° = 2,082 lb
V ft ft
Rτ = V tan δ
b
V =γ DC + γ EV + γ P
p( DC ) p( EV ) p( EH ) V
γ = 0.9
p( DC ) (Minimum load factors for DC and EV and maximum
γ = 1.0 load factors for EH were used to evaluate the maximum
p( EV ) force effect for sliding. Load factors determined from
γ = 1.5 Table 4-2)
p( EH )
⎡ 2 ft +1.5 ft ⎤
DC = ⎢2.5 ft ×12 ft + (20 ft )⎥(145 pcf ) = 9425lb
⎢⎣ 2 ⎥⎦ ft
⎡ 1 ⎤
EV = ⎢20 ft + (8 ft ) tan(10°)⎥(8 ft )(115 pcf ) = 19,049 lb
⎣⎢ 2 ⎦⎥
ft
V = (0.9) 9,425lb + (1.0)19,049 lb + (1.5)2,082 lb = 30,655lb
ft ft ft ft
δ = φ ' = 35°
b f
S3 - 6
⎛ ⎞
Rτ = ⎜⎜ 30,655lb ⎟ tan(35°)
⎝ ft ⎟⎠
Rτ = 21,465lb
ft
R = (0.8)(21,465) = 17,172 lb
R ft
H =γ P
p( EH ) H
H = (1.5)(11,806 lb ) = 17,709 lb
ft ft
R > H (criteria)
R
17,172 lb > 17,709 lb ∴ Doesn’t meet criteria
ft ft
S3 - 7
Group 2: Heel width = 10 ft and φ’b = 29° (Figure S3-1)
H = 2.5 ft + 20 ft + ∆h
H = 2.5 ft + 20 ft + 10 ft tan10°
H = 24.3 ft
sin 2 (θ + φ ' )
Ka = b β = 10°
Γ sin 2 θ sin(θ − δ ) δ = 10°
2 where: φ ' = 29°
⎡
⎢
sin(φ ' +δ ) sin(φ ' −β ) ⎥
⎤ b
Γ = ⎢1 + b b θ = 90°
⎥
⎢ sin(θ − δ ) sin(θ + β ) ⎥
⎣ ⎦
K a = 0.365
1
EH = K a H 2γ
2 b
1
EH = (0.365)(24.3 ft ) 2 (115 pcf )
2
EH = 12,393 lb
ft
S3 - 8
Resolve active earth pressure into horizontal (PH) and vertical
(PV) components:
P = EH cos δ
H
P = 12,393 lb cos10° = 12,205 lb
H ft ft
(see Figure S3-1)
P = EH sin δ
V
P = 12,393 lb sin 10° = 2,152 lb
V ft ft
Rτ = V tan δ
b
V =γ DC + γ EV + γ P
p( DC ) p( EV ) p( EH ) V
γ = 0.9
p( DC ) (Minimum load factors for DC and EV and maximum
γ = 1.0 load factors for EH were used to evaluate the maximum
p( EV ) force effect for sliding. Load factors determined from
γ = 1.5 Table 4-2)
p( EH )
⎡ 2 ft + 1.5 ft ⎤
DC = ⎢2.5 ft ×14 ft + (20 ft )⎥(145 pcf ) = 10,150 lb
⎢⎣ 2 ⎥⎦ ft
⎡ 1 ⎤
EV = ⎢20 ft + (10 ft ) tan(10°)⎥(10 ft )(115 pcf ) = 24,014 lb
⎣⎢ 2 ⎦⎥
ft
V = (0.9)10,150 lb + (1.0) 24,014 lb + (1.5)2,152 lb = 36,377 lb
ft ft ft ft
δ = φ ' = 35°
b f
S3 - 9
⎛ ⎞
Rτ = ⎜⎜ 36,377 lb ⎟ tan(35°)
⎝ ft ⎟⎠
Rτ = 25,471lb
ft
R = (0.8)(25,471) = 20,377 lb
R ft
H =γ P
p( EH ) H
H = (1.5)(12,205 lb ) = 18,308 lb
ft ft
R > H (criteria)
R
20,377 lb > 18,308lb ∴ Meets Criteria
ft ft
S3 - 10
Group 3: Heel width = 8 ft and φ’b = 32° (Figure S3-1)
H = 2.5 ft + 20 ft + ∆h
H = 2.5 ft + 20 ft + 8.0 ft tan10°
H = 23.9 ft
sin 2 (θ + φ ' )
Ka = b β = 10°
Γ sin 2 θ sin(θ − δ ) δ = 10°
2 where: φ ' = 32°
⎡
⎢
sin(φ ' +δ ) sin(φ ' −β ) ⎥
⎤ b
Γ = ⎢1 + b b θ = 90°
⎥
⎢ sin(θ − δ ) sin(θ + β ) ⎥
⎣ ⎦
K a = 0.321
1
EH = K a H 2γ
2 b
1
EH = (0.321)(23.9 ft ) 2 (115 pcf )
2
EH = 10,543 lb
ft
S3 - 11
Resolve active earth pressure into horizontal (PH) and vertical
(PV) components:
P = EH cos δ
H
P = 10,543lb cos10° = 10,383 lb
H ft ft
(see Figure S3-1)
P = EH sin δ
V
P = 10,543 lb sin 10° = 1,831lb
V ft ft
Rτ = V tan δ
b
V =γ DC + γ EV + γ P
p( DC ) p( EV ) p( EH ) V
γ = 0.9
p( DC ) (Minimum load factors for DC and EV and maximum
γ = 1.0 load factors for EH were used to evaluate the maximum
p( EV ) force effect for sliding. Load factors determined from
Table 4-2)
γ = 1.5
p( EH )
⎡ 2 ft + 1.5 ft ⎤
DC = ⎢2.5 ft ×12 ft + (20 ft )⎥(145 pcf ) = 9425 lb
⎢⎣ 2 ⎥⎦ ft
⎡ 1 ⎤
EV = ⎢20 ft + (8 ft ) tan(10°)⎥(8 ft )(115 pcf ) = 19,049 lb
⎣⎢ 2 ⎦⎥
ft
V = (0.9) 9,425 lb + (1.0)19,049 lb + (1.5)1,831lb = 30,278 lb
ft ft ft ft
δ = φ ' = 35°
b f
S3 - 12
⎛ ⎞
Rτ = ⎜⎜ 30,278lb ⎟ tan(35°)
⎝ ft ⎟⎠
Rτ = 21,201lb
ft
R = (0.8)(21,201) = 16,961lb
R ft
H =γ P
p( EH ) H
H = (1.5)(10,383 lb ) = 15,575 lb
ft ft
R > H (criteria)
R
16,961lb > 15,575 lb ∴ Meets Criteria
ft ft
S3 - 13
All Groups: Bearing Resistance Check
P = 11,806 lb
H ft
P = 2,082 lb
V ft
Factored Moment
Moment arm Load factor Factored load
Load (lb)
Item about toe (ft) (γp) (lb/ft)
(lb/ft) (Arm x Factored
(Figure S3-1) (Table 4-2) (γp x Load)
Load)
Vertical Loads
DCstem 5,075 3.12 1.25 6,344 19,793
DCftg 4,350 6 1.25 5,438 32,628
PV 2,082 12 1.5 3,123 37,476
EV 19,049 8.14 1.35 25,716 209,328
Horizontal Loads
PH 11,806 7.97 1.5 17,709 141,141
∑V
σV =
B − 2e
B
∑ V = 6,344 + 5,438 + 3,123 + 25,716 = 40,621lb
ft
B
e = − Xo
B 2
M −M
Xo = VTOT HTOT = 299,225 lb ⋅ ft −141,141 lb ⋅ ft = 3.89 ft
V 40,621 lb
S3 - 14
12 ft
e = − 3.89 ft = 2.11 ft
B 2
40,621lb
ft
σV = = 5,221 psf
12 ft − 2(2.11 ft )
qR > σ v (criteria)
q = 11,526 psf > σ = 5,221 psf ∴ Meets Criteria
R V
S3 - 15
STUDENT EXERCISE 4
MSE Wall Design External Stability
Manual Reference:
S4- 1
23 ft
S4 - 2
SOLUTION TO STUDENT EXERCISE 4
Length of Reinforcement:
For preliminary sizing, consider L = 0.7 H
L = 0.7 (23 ft) = 16.1 ft, use L = 16.5 ft
Calculate Moments:
M=Force x Moment arm about toe of wall (see
Figure S4-2)
S4 - 4
Summary of Unfactored Loads and Moments:
Total V 52.4
Total H 13.0
S4- 5
Table: Summary of Factored Loads
PEV PLSV PEH PLSH VTOT HTOT
Group
(kips/ft) (kips/ft) (kips/ft) (kips/ft) (kips/ft) (kips/ft)
Check Eccentricity:
eB = B / 2 − X o = (8.25 ft − X o )
M EV − M hTOT
Xo =
PEV
eB (max) = 3.46
em = B / 4 = 4.13 ft
eB < em ∴ O.K .
S4 - 6
L/2
PEV,PLSV PLSH
PEH
H/2
H/3
S4- 7
Calculate Factored Resistance against Sliding:
S4 - 8
STUDENT EXERCISE 5
MSE Wall Internal Stability
Manual Reference:
Section 7.8.3
Example in Section 7-9
S5 - 1
SOLUTION TO STUDENT EXERCISE 5
S5 - 2
S5 - 3
Figure S5-1. Active and Resistant Zones for MSE Wall
b. Compute K coefficient at Zi = 11.5 ft
therefore by interpolation:
(0.5K a ) = xzi
20 ft 20 ft −11.5 ft
xzi = 0.2125 Ka
S5 - 4
20 ft
therefore by interpolation:
(1.5 − 0.67 ) = x zi
20 ft 20 ft − 11.5 ft
xzi = 0.35
S5 - 5
STUDENT EXERCISE 6
Given:
Compute:
Manual Reference:
Sections 7.8.3
Example from section 7-9
S6 - 1
Summary of Equations
Ac = bt c
S6 - 2
SOLUTION TO STUDENT EXERCISE 6
S6 - 3
Calculate the section loss of steel:
tc = tn – ts
tc = 160 mil – 55.5 mil = 104.5 mil
= ϕ bt c Fy
S6 - 4
STUDENT EXERCISE 7
Apparent Earth Pressure Diagrams
a) Sand
b) Soft Clay
c) Stiff to Hard Clay
Manual Reference:
Section 8.11.2.1
Figures 8-67, 8-68, 8-69
S7-1
SOLUTION TO STUDENT EXERCISE 7
a. Sand Profile
30
K a = tan 2 ( 45 − )
2
K a = tan 2 (30) = 0.33
p = 0.65K a γH
= 0.65(0.33)(115 pcf )(20 ft )
= 493 psf
S7-2
H=20 ft
p
p=493 psf
Figure S7-2. Apparent earth pressure envelope for
excavation through sand.
S7-3
b. Soft Clay Profile
γ = 121 pcf
10 ft Clay Su = 400 psf
γ = 121 pcf
10 ft Clay Su
Su = 400 psf
Hard Clay
S7-4
= 0.55
p = 1.0(0.55)(121 pcf )(20 ft )
p = 1331 psf
H=20 ft
p=1331 psf
S7-5
c. Stiff to Hard Clay
S7-6
Calculate the Apparent Earth Diagram:
Total Force
= 0.65 K a γ H 2
= 0.65 (0.31)(127 pcf )(20 ft ) 2
= 10,236 lb
ft
S7-7
5.33 ft
p=0.4γH
6.67 ft
8 ft
Total Force:
1 1
= (8 ft ) ( p ) +(6.67 ft ) ( p ) + (5.33 ft ) ( p ) = 13.3 p
2 2
= (13.3) (0.4) (127 pcf ) ( 20 ft )
= 13,513 lb
ft
S7-8
STUDENT EXERCISE 8
Single-Tier Anchored Soldier Beam and Lagging
Wall
Manual Reference:
S8-1
Summary of Equations
S8-2
Driven soldier
pile w/ timber
lagging facing
H1
T
Cohesionless Soil
φ=33
φ’ =33°°
H= 25 ft
γm=18 pcf 3
=115kN/m
H-H1
S8-3
(A) Calculation of Anchor force
⎛ φ' ⎞
K A = tan 2 ⎜ 45 - ⎟
⎝ 2⎠
⎛ 33° ⎞
K A = tan 2 ⎜ 45 - ⎟ = 0.295
⎝ 2 ⎠
Tn =
(23H − 10 HH1
2
)
p (spacing )
54(H − H1 )
p = 1.0 K A γ H
p = (1.0)(0.295) (115 pcf) (25 ft) = 848 lb/ft per ft of
wall
Tn =
(23(25 ft )− 10(25 ft )(8.5 ft )
2
)
(848 lb / ft per ft of wall)(8 ft )
54(25 ft − 8.5 ft )
Tn = 93.3 kips
S8-4
B) Select Tendon Type and Check Tensile
Resistance
γ pTn
ϕt Fn ≥
cos (i )
ϕ t = 0.8
γ p = 1.35
(1.35)(93.3 kips )
Fn ≥
(0.8)cos 30°
Fn ≥ 181.8 kips
S8-5
Select a four strand anchor with GUTS = 234.4 kips.
T
Lb min =
ϕ p Qn
γ pTn
T=
cos(i )
1.35(93.3 kips)
T= = 145.4 kips
cos 30°
ϕ p= 0.65
Qn = πdτ a
τa = 4.4 ksf
d = 6 in.
⎛ 1 ft ⎞
Qn = π ⎜⎜ 6 in. × ⎟⎟ × (4.4 ksf )
⎝ 12 in. ⎠
Qn = 6.9 kips / ft bond length
145.4 kips
Lb min = = 32.4 ft (Say 32 ft)
(0.65)(6.9 kips / ft )
S8-6
STUDENT EXERCISE 9
Design of a Two-Tier Anchored Soldier Pile and
Lagging Wall
Manual Reference:
Sections 8.11
Figures 8-67, 8-71, 8-73
AASHTO (2007) Section 6.5.4.2, 6.9.2.2
S9-1
Driven soldier
pile w/timber
lagging facing
H1 q=0.2 γ [[psf ]
T1
Cohesionless Soil
φ=33°
H= 25 ft
H2
γm=115 pcf 3
=18 kN/m
T2
8 ft O.C.
Pile spacing = 2.43 m O.C.
H3
d
S9-2
(A) Calculate individual anchor forces and subgrade
reaction forces
0.65 K A γ H 2
p=
H H (Eq. 9-1)
H- 1 − 3
3 3
⎛ φ⎞
K A = tan 2 ⎜ 45 - ⎟ = 0.295
⎝ 2⎠
S9-3
TOTAL SURCHARGE EARTH
H1
TT1 TS1 T1
pt ps p
H2 = +
TT2 TS2 T2
H3
RT RS R
Horizontal Forces
ps = KA q
ps = 0.295 (115 pcf) 2 ft = 68 psf
S9-4
Figure 8-71. The horizontal forces are summarized
in Table S9-1.
S9-5
(B) Determine Load Factors and Resistance Factors
S9-6
Select ASTM A722 Grade 150 Steel bar with
nominal diameter of 1 in. (GUTS = 127.5 kips) for
upper and lower anchor tendon steel.
S9-7
(D) Calculate Flexural Resistance of Soldier Pile
S9-8
(E) Calculate axial loads in wall
⎡ T +T T + T2 S ⎤
= ⎢ 1 2 (1.35) + 1S (1.75)⎥ sin(30°)
⎣ cos(30°) cos(30°) ⎦
⎡ (55,122 + 46,264) × (1.35) (6800 + 4536) × (1.75) ⎤
=⎢ + ⎥ sin(30°)
⎣ cos(30°) cos(30°) ⎦
= 90.5 kips
S9-9
S9-10
Pu P M
If < 0.2, then u + u ≤ 1.0 AASHTO (2007) Section 6.9.2.2
Pr 2 Pr M r
Pu = 90.5 kips
If λ ≤ 2.25, then Pn = 0.66 Fy As
λ
2
⎛ Kl ⎞ Fy
λ = ⎜⎜ ⎟⎟
⎝ rsπ ⎠ E
2
⎛ ⎞
⎜ ⎟
⎜ (1.0)(8.33 ft ) ⎟ 50 ksi
λ =⎜ ⎟ 29,000 ksi = 0.069 < 2.25
⎜ ⎛⎜ 5.03 in × 1 ft ⎞⎟π ⎟
⎜⎜ 12 in ⎟⎠ ⎟
⎝⎝ ⎠
Pn = (0.66) 0.069 (50 ksi)(15.5 in 2 ) = 753 kips
Pr = φPn = (0.7)(753 kips ) = 527 kips
φ = 0.7 ( AASHTO Section 6.5.4.2)
Pu
= 0.17 < 0.2
Pr
M r = φM n = 1.0 Fy Z = 1.0(50 ksi)(74.0 in 3 ) = 3,700 kip − in = 308 kip − ft
Pu M u 90.5 kips 161 kip − ft
+ = + = 0.61 ≤ 1.0 ∴ OK
2 Pr M r 2(527 kips ) 308 kip − ft
S9-11
STUDENT EXERCISE 10
Lateral Wall Movement
Manual Reference:
Section 8.11.10, 8.11.11
Figures 8-69, 8-76d
S10-1
Sand (SP) 4.7 ft
N=15-26
φ=33°
9.3 ft
γ=121 pcf
3.9 ft
13.1 ft
d = 12 ft
18 ft
Stiff Clay
Su=1500 - 2300 psf Note: Drawing not to scale.
φ=32°
γ=125 pcf
S10-2
Assume the following material properties,
S10-3
S10-4
SOLUTION TO STUDENT EXERCISE 10
Where,
γ w = 62.4 pcf
havg = 13.1 ft (as illustrated in Figure 10-1 )
⎛ 1000 lb 1 ft ⎞
⎜⎜ 445,150 kip − in × × ⎟⎟
=⎝
EI kip 12in ⎠
= 20
γ w (havg )4 62.4 pcf (13.1 ft ) 4
S10-5
Calculate Factor of Safety Against Basal Heave:
Where,
Su = 600 psf
H
= 1 .0
B
B L = 0.1
6.5(600 psf )
FS = = 1.33
⎡ 600 psf ⎤
42 ft ⎢(120 pcf ) −
⎣ 12 ft ⎥⎦
FS = 1.33
EI
= 20
γ w (havg )4
S10-6
δ H MAX
% ≈ 1 .4 %
H
Ec = 3,100 ksi
I=
1
(12in )(30in )3 = 27,000in 4 / ft of wall
12
(
EI = 3,100ksi 27,000in 4 / 12in )
EI = 6,975,000kip − in
EI
Normalized System Stiffness = γ (h )4
w avg
⎛ 1000lb 1 ft ⎞
⎜⎜ 6,975,000kip − in × × ⎟⎟
=⎝
EI kip 12in ⎠
= 316
γ w (havg )4 62.4 pcf ⋅ (13.1 ft )
4
S10-7
Calculate Factor of Safety Against Basal Heave:
FS =1.33
FS = 1.33
EI
= 316
γ w (havg )4
δ HMax
% = 0.6%
H
EI = 445,150 kip-in
S10-8
EI
= 20
γ w (havg )4
S u = 950 psf
6.5(950 psf )
FS = = 3.6
⎡ 950 psf ⎤
42 ft ⎢(120 pcf ) −
⎣ 12 ft ⎥⎦
FS = 3.6
EI
= 20
γ w (havg )4
δ HMAX
% = 0 .3 %
H
S10-9
STUDENT EXERCISE 11
Soil Nail Wall Design
Manual Reference:
Section 9.2.7.3
S11-1
20 ft
10°
Soil Parameters:
Assumptions:
S11-3
Backslope
Face Batter β = 10O
α = 10O µ = (qaDDH)/(γ SH SV)
L
qa = qu/FSP
γ, c, φ c* = c / γH
H SV
c* = 0.02 t max-s = Normalized Maximum Design Force in Nails
FS=1.35 = Tmax-s/γ H SH SV
DDH= 4
100
in.mm For other FS, c*, and DDH,
see Figure B7
1.5
Friction Angle
(degrees)
27
31
1 35
39
L/H
0.5
(a)
Normalized Max. Design Nail Force, tmax-s
00
0.4
0.3
0.2
0.1
NOTE:
Nail forces are for FSG = 1.0
(b)
0
0 0.1 0.2 0.3 0.4
q aD DH
Normalized Bond Strength, µ =
γ SH SV
Where,
qu = 1200 psf
DDH = 4in.
γ = 115 pcf
S H = SV = 5 ft
qa = qu/FSP with FSP = 2.0 = 600 psf
Therefore,
600 psf (4in.)⎛⎜1 ft ⎞⎟
µ= ⎝ 12in. ⎠
115 pcf (5 ft )(5 ft )
µ = 0.070
Choose #6 bar
A = 0.44 in2
Summary
20 ft
10°
3 ft
32 ft
5 ft
(typ)
3 ft
92
3.0fm
t
Nail Spacing: SV = SH = 5 ft
S11-7
STUDENT EXERCISE 12
Fill Wall Selection
Problem Statement
An earth retaining wall is needed to allow construction of a road adjacent to a creek. The
road is proposed for construction to provide temporary access to remote areas in a U.S.
national forest for about 5 years while a permanent highway is built nearby. Therefore, the
earth retaining wall is proposed for temporary support for 5 years with minimum
maintenance requirements.
Based on the roadway profile and the site topography, the required wall height along the
alignment is estimated to be 30 ft in certain sections. The soils at the site consist of medium
dense silty sand, with zones of soft compressible clays that may cause long-term differential
settlement problems, even over the relatively short service life of the wall.
Wall selection for this example is performed following the wall selection flow chart
provided in Figure 10-1.
Assignment
Select the appropriate wall alternative for construction at the above described embankment
in the national forest. Select weighting factors from 1 to 3 for the wall selection factors
listed on the selection matrix in Table S12-1. Using Tables 10-3 and your own table that
you have created during class, perform initial screening and select potential alternatives. For
each wall alternative considered, assign an initial qualitative rating from 1 to 4 based on
each wall selection factor. Calculate the weighted ratings by multiplying the initial rating by
the weighting factors and summarize the results in Table S12-2. Assign a final score for
each wall alternative.
Manual Reference:
Chapter 10
Figure 10-1
Tables 10-3 and 10-5
Example Section 10.7
S12 - 1
Solution
(See Manual Chapter 10)
Step 1 and Step 2 are already given in the problem statement.
The owner’s evaluations of the factors that affect the wall selection are presented in Table
S12-1. The ground represents a critical factor because the site consists of soils that have a
potential to undergo differential settlement. Environmental concerns are rated high because
the site is located in a U.S. national forest, which requires that any construction satisfy
specific construction permitting requirements (which may involve limitations on noise and
restrictions on working in the creek). Durability and maintenance are not considered
important due to the temporary nature of the wall. Groundwater is rated as one of the least
important factors for the project because no temporary dewatering is needed during
construction. Contracting practice and tradition are not important issues overall, however
the Owner has requested that a concrete cantilever wall be considered since standard design
details are available for this system.
Based on an initial review of the project data, concrete walls can be eliminated immediately
because of the required wall height, relatively slow construction, and poor performance
where the potential for differential settlement exists. Furthermore, MSE walls with modular
block or precast concrete facing are also eliminated because the project calls for a temporary
wall and other facing types (for MSE walls) are more appropriate for temporary
applications.
Reinforced soil wall, MSE wall with a wrap-around geotextile face, gabion wall, bin wall,
and concrete module wall are chosen to be further evaluated. To satisfy traditional
construction practice in the area, a cast-in-place concrete cantilever wall is also added to the
list for further evaluation.
The performance of each wall type against the factors previously identified are evaluated
and tabulated in Table S12-1. A brief discussion on the rating of some of the wall selection
factors is presented below.
Ground conditions at the site are known to be problematic due to their potential for
differential settlement. Therefore, walls that can typically tolerate higher differential
settlement are rated higher than walls that are more rigid and susceptible to damage caused
by differential settlement.
S12 - 2
Construction considerations are evaluated based on the availability of material and the effort
it takes to deliver these materials to the site. Walls that require no special backfill material
and consist of wall materials that could be delivered to the site fairly easy are rated high.
Walls that require special material are rated lower than the other wall types.
The ROW for the wall is evaluated based on the space required behind the structure. Walls
that require less space are rated higher than the walls that require more space behind the
structure.
The importance of aesthetics for the wall selection is a bit subjective. Because the site is
located in a U.S. national forest, walls with a more natural appearance are rated high.
Walls that could be constructed with less potential for encroachment on the creek and less
potential for noise are rated high.
Concrete walls are known to be more durable and they are rated higher than geotextile walls
where the geotextile may degrade due to exposure to sunlight.
S12 - 3
Table S12-1. Initial rating of each wall selection factor for each wall alternative.
Owner’s 1 2 3 2 2 3 3 1 1 3
3
WR
Durability
Construction Speed of Environmental Contracting Cost
Ground Groundwater ROW Aesthetics and Tradition
Considerations Construction Concerns Practice
Maintenance
Wall
Initial Rating
Types
Reinforced 2 3 4 3 2 3 3 3 3 4 3
Soil Wall
Geotextile 3 2 4 3 2 3 3 3 3 4 4
Wall
Gabion 2 3 3 2 4 3 3 3 2 3 3
Wall
Bin Wall 2 2 3 3 3 2 2 3 2 3 2
Concrete
Module 3 2 3 3 3 2 2 4 4 3 3
Wall
Cast-in-
Place 3 2 3 1 3 2 2 4 4 4 1
Concrete
S12 - 4
Table S12-2. Wall selection matrix and total score of each wall alternative.
Risk Factors That Affect Wall Selection
Considerations
Environmental
Groundwater
Construction
Construction
Maintenance
Contracting
Durability
Aesthetics
Tradition
Concerns
Speed of
Practice
Ground
Wall Types Total Score
ROW
Cost
and
Reinforced Soil
Wall
6 3 8 9 4 6 9 9 3 4 9 70
Geotextile Wall 9 2 8 9 4 6 9 9 3 4 12 75
Gabion Wall 6 3 6 6 8 6 9 9 2 3 9 67
Bin Wall 6 2 6 9 6 4 2 9 2 3 6 55
Concrete
Module Wall
9 2 6 9 6 4 6 12 4 3 9 70
Cast-in-Place
9 2 6 3 6 4 6 12 4 4 3 59
Concrete
S12 - 5
STUDENT EXERCISE 13
Cut Wall Selection
Problem Statement
Existing structures including pile foundations and utilities are in close proximity to the
wall line. The wall system, therefore, needs to be relatively watertight to prevent
groundwater drawdown and potential resulting settlements of nearby structures. The
contract documents will require that the maximum lateral wall movement be less than 1
percent of the wall height.
The maximum height of the wall is 33 ft as shown in Figure S13-1, and is approximately
200 ft in length. The general soil profile along the alignment of the wall includes a silty
clay and sand fill overlying silty clays that increase in strength with depth. The majority
of the wall will be constructed in a medium stiff to stiff silty clay layer with sands and
gravel. SPT blowcount values range from 20 to 30 blows per ft in this layer. Prior to
wall construction, the upper 10 ft of fill will be excavated to the elevation of the top of
the wall.
Fill
Exposed Face of
Medium Stiff to Future Wall
Stiff Clay ~33 ft
S13-1
Assignment
Select the appropriate wall alternative for construction at the above described interchange
project using the procedure described in the wall selection flow chart in Figure 9-1.
Select weighting factors from 1 to 3 for the wall selection factors listed in the selection
matrix. For each wall alternative assign an initial qualitative rating from 1 to 4 based on
each wall selection factor. Assign a final score for each wall alternative.
Manual Reference:
Chapter 10
Figure 10-1
Tables 10-2 and 10-4
S13-2
Solution
Step 1 and Step 2 are already given in the problem statement.
The owner’s evaluations of the factors that affect the wall selection are presented in Table
S13-1. The ground represents a critical factor because of the relatively hard driving
conditions which would preclude the use of sheet-piles or driven soldier pile sections.
Groundwater is also a critical factor. A cut-off wall is required to permanently maintain
groundwater levels behind the wall. Construction considerations are not rated as an
important factor due to easy access to building materials; however, the effect of
freeze/thaw conditions and the water table on the constructability of the wall system must
be considered. Cycles of freezing and thawing may be detrimental to the durability of
exposed soil mix columns. Because the retaining wall is being constructed in a densely
populated urban environment, ROW and aesthetics are both important factors.
Environmental concerns, tradition, and contracting practices are all non-important issues
in this project, although construction noise should be minimized. Table 10-4 summarizes
cut wall costs per square foot of exposed wall face.
For this project, due to the proximity to existing structures and utilities and the need for a
groundwater cut-off system, cut wall systems were chosen as the most viable option. Soil
nail, soldier pile and lagging, secant pile, structural slurry, sheet pile, and DMM wall
systems are chosen to be further evaluated in this example. The performance of each
wall system against the factors previously identified are evaluated and tabulated in Table
S13-1.
The final wall selection matrix is shown in Table S13-2. The scores range from 48 to 83
with the secant pile wall being the highest rated. At the required maximum height of 33
ft, a relatively large-diameter secant pile wall could likely be used without requiring
ground anchors for additional support.
S13-3
Table S13-1. Initial rating of each wall selection factor for each wall alternative.
WR 3 3 2 3 3 2 1 3 1 1 3
Durability
Ground- Speed of
Ground Construction ROW Aesthetics Environmental and Tradition Contracting Cost
water Construction
Maintenance
Soil-Nail 2 1 3 3 1 3 3 1 2 2 2
Soldier Pile
and 2 1 3 2 1 3 2 2 3 3 3
Lagging
Secant Pile 4 3 4 3 3 3 3 4 3 3 3
Structural
4 4 2 2 3 3 3 4 2 2 1
Slurry
Sheet Pile 1 2 2 1 1 3 2 2 3 3 3
DMM 4 3 2 3 3 3 3 1 1 2 3
S13-4
Table S13-2. Wall selection matrix and total score of each wall alternative.
Risk Factors That Affect Wall Selection
Considerations
Environmental
Groundwater
Construction
Construction
Maintenance
Contracting
Aesthetics
Durability
Concerns
Tradition
Speed of
Practice
Wall Types Total Score
Ground
ROW
Cost
and
Soil Nail 6 3 6 9 3 6 3 3 2 2 6 49
Soldier Pile and
6 3 6 6 3 6 2 6 3 3 9 53
Lagging
Secant Pile 12 9 8 9 9 6 3 12 3 3 9 83
Structural
12 12 4 6 9 6 3 12 2 2 3 71
Slurry
Sheet Pile 3 6 4 3 3 6 2 6 3 3 9 48
DMM 12 9 4 9 9 6 3 3 1 2 9 67
S13-5