You are on page 1of 28

5-1

Chapter 5 Walls for Excavations



Introduction
Excavation is necessary when constructing shallow footing, mat foundation, or subbasements. It
is a legal necessity to assure no loss of bearing capacity, excessive settlements, or excessive
lateral movements to adjacent properties. Thus, this retaining structure during excavation stage
is one of the important issues.

Failure in Excavations
The earth-retaining structures in excavations can be failed by structural collapse, excessive
deformation, basal instability, and inadequate groundwater exclusion.

Control of groundwater
There are four main methods used to exclude groundwater from excavations:
1. Stopping surface water from entering the excavation by using cut-off ditches
2. Allowing water to flow into the excavation and subsequently pumping it from drainage
systems
3. Pre-draining the soil by lowing the groundwater level ahead of the excavation
4. Stopping the groundwater from entering the excavation by a cut-off wall within the soil

This chapter will address the design issue of earth-retaining structures and the estimation of
deformation.

Wall Construction
The types of retaining walls used in excavation are:
1. Sheet piling
2. Soldier pile walls (Soldier beams with or without lagging)
3. Drilled-in-place concrete piles (or piers) walls
4. Caisson wall
5. Diaphragm (Slurry) walls
6. Soil cement mixed walls
7. Soil nail walls

Systems to hold the retaining wall in place include:
1. Cantilevered or unbraced wall (it is less economic except for shallow excavation)
2. Wales and struts or rakers (braced walls)
3. Compression rings
4. Tie-back (anchored walls) or soil nails

Sheet piling
Sheet piling is commonly used for retaining excavations because the highest strength/weight
ratio, reusable, easily installed and removed. It is not usable, however, when the subsoil contains
many boulders or is dense and the excavation is deep. Noise and vibration may be an issue when
existing buildings, especially medical facilities, are sensitive to disturbance by vibration.
5-2
Seepage may be expected to pass through interlocking steel sheet piling which supports a
difference in hydraulic head.

Sheet Pile Drivability
Little is known about the necessary sections to allow sheet piles to be driven without damage.
However, Williams and Waite (1993) suggested that in granular soils the minimum section could
be judged on the basis of SPT resistance (N value), as shown in Table 5.1.

Table 5.1 Sheet Pile Drivability (CIRIA, 1993)
Dominant
SPT
Minimum wall modulus (cm
3
/m)
Remarks A570 Gr 40
(265 MPa)
A572 Gr 50
(345 MPa)
0 - 10 450 A572 Gr 50 for length > 10 m
11-20 450
21-25 850
26-30 850 Length > 15 m is not advisable
31-35 1300 Penetration > 5 m not advisable
36-40 1300 Penetration > 8 m not advisable
41-45 2300
46-50 2300
51-60 3000
61-70 3000 Some declutching may occur
71-80 4200 Some declutching may occur for L > 15m
81-140 4200 Increase risk of declutching

Soldier Pile Walls
Soldier pile walls have two basic components, soldier piles (vertical component) and lagging
(horizontal component). It is essential that the soldier piles be maintained in full contact with
the soil. For this reason, they are either driven or placed in pre-drilled holes, which are
backfilled to the ground surface with lean concrete. Drilled-in-place piles may be used if
problems of driving exist (noise and vibration).

H piles are driving on a convenient spacing of 2 to 3 m for using standard-length timber. As
excavation proceeds, 50 to 100 mm thick boards are inserted behind the front flanges. The
boards can be placed against the H-pile and clipped to the front flange using patented fasteners.
The method is suitable for overconsolidated clays, all soils above the water table if they have at
least some cohesion, and free-draining soils that can be dewatered effectively.

When water is not a problem, lagging for soldier piles or drilled-in-place piles is not required, as
arching or bridging action of the soil from the lateral pressure developed by the pile will retain
the soil across the open space. A spacing of 2 to 3 m is commonly used in strong soils, where
no sensitive structures are present. The spacing is reduced to 1 to 2 m in weaker soils or near
sensitive buildings. The pile will, of course, have to be adequately braced to provide necessary
lateral soil resistance. In addition, the embedment depth has to be sufficient.

5-3
Drilled-in-place concrete (bored) piles (or piers) walls
The use of low-cost augers and, more particularly, continuous flight auger rigs to drill successive
unconnected piles provides an economical wall for both temporary and permanent use for
excavations to medium depth where soil conditions are favorable. Grouting or jet grouting can
be used to remedy leakage between piles.

When soil and water must be retained, the system will have to be reasonably watertight below
the water table and be capable of resisting both soil and hydrostatic pressures. The secant wall,
consisting of interlocking piles, is most suited. Primary (female) piles, piles at centers slightly
less than twice the nominal pile diameter, are drilled with a concrete guide wall. Before fully
hardening, the secant (male) piles (same or smaller diameter) are drilled: during the process the
drilling removes segments of the primary piles so interlock is obtained. The secant wall
provides a waterproof wall which can be built to a depth ranging 30 to 40 m. In some instances
reinforcement may be bunched at opposite sides of the pile diameter to ensure maximum
effectiveness.

Caisson wall
Hand-dug caisson retaining walls, often referred to as caisson walls, have been widely used in
Hong Kong. The caissons are usually dug in stages of about one meter depth, by husband
miners with their wives serving as winch operators at the ground surface. Each stage of
excavation is lined with a minimum of 75 mm thickness of in-situ concrete using a tapered steel
shutter suitab1y supported and designed for ease of striking. The shutter remains in place to
provide support for the fresh concrete and surrounding ground while the next stage of excavation
proceeds. Submersible electric pumps are commonly used for dewatering at the base of the
caissons. Excavation through core stones and other obstructions is carried cut by pneumatic
drilling. The main advantages of a caisson wall over a conventional retaining wall are
(a) It can be constructed without temporary soil cuts or shoring.
(b) It requires a small plan area and can be used close to site boundaries and existing
structures.
(c) It can act as both temporary and permanent retaining structures.
(d) Obstructions e.g. core stones and boulders, can be overcome without too much
difficulty.

Diaphragm (Slurry) walls
Diaphragm (slurry) walls are more expensive than other wall systems. However, frequently they
provide the best solution to many underground construction problems, and the technique can
usually be adjusted to cope with most conditions of difficult ground and adjacent surcharge by
adjustment of the panel length, properties of the slurry, level of slurry in the trench and
groundwater lowering. It can be used as a permanent part of the structural system (basement
walls). Concrete poured in a cavity retained with slurry (a dense liquid) producing a slurry wall.
This method will be discussed in later chapter.

Support Systems
The lateral loads due to earth pressures, surcharge loads, hydrostatic pressures, and seismic loads
are constantly trying to push the wall and must be restrained. The restraint can be developed
5-4
from the inside the excavation (strut and/or rake systems) or outside (tieback systems). A strut
or raker system creates obstructions in the excavation area (working space). For narrow
excavation (less than 20 m wide), it is more economical than tiebacks. In addition, you dont
need to bring in drilling machines, to solve the problem of encountering existing obstructions
such as tunnels, basement wall, or public utilities, and to obtain permission to trespass into the
adjacent property owners subsoil. However, for most excavation involving large area, tiebacks
are preferable than strutted system.

Rakers, sloping compression units, are attached to the wall and braced against either the structure
being constructed, or a footing specifically cast for the purpose of resisting the raker forces.
They impart both a lateral force and uplift force.

Struts, horizontal compression units, are braced against either an existing structure or another
portion of the shoring system.

Tiebacks, tension units, restrain the applied load from outside the excavation. Since tiebacks are
installed at a downward dipping angle, they impart a downward force to the wall as well as a
lateral force.

To insure very little lateral movement, it is essential that
1. The wall fit securely, snugly, and closely against the sides of the excavation.
2. Some elastic movement always occurs but the amount should be limited.
3. Use sufficient rigid wales so that very little movement between supported points.
4. Use vertical bracing to ensure little amount of bulging between brace points.


Analytical Methods for Support Systems
Three methods have been used for analyzing earth-retaining structures in excavation. They are:
(a) limit analysis, (b) beam on elastic foundation, and (c) finite element method (full soil-
structure interaction)

(a) Limit Analysis

To design braced excavations using limit analysis, one must estimate the lateral earth pressure to
which the braced cuts wall will be subjected. The braced system subjected to the same earth
pressure forces as other retaining structures that may be calculated by Rankine or Coulomb
methods. However, the design pressures are different from those because of the manner in
which the pressures are developed.

Peck (1969), from measurements, proposed empirical pressure diagrams for wall and strut
design, as shown in Figure 5.1 and Table 5.2. The pressure envelope for soft to medium clay is
applicable for the condition when H/c > 4. For stiff clay, H/c s 4. Based on the observation
of Tschebotarioff (1973), slightly different apparent design pressure envelops are suggested.
5-5

Figure 5.1 Apparent Pressure Distribution for Braced Cut

Table 5.2 Parameters for Apparent Pressure Distribution
Soil Type Author Z
1
Z
2
Z
3
|
Sand
Peck

0 1 0 0.65K
a

Soft/Medium Clay 0.25 0.75 0 0.3*
Stiff Clay 0.25 0.5 0.25 0.2~0.4
Sand
Tsch.
+


0.1 0.7 0.2 0.25
Short Term (Medium Clay) 0.6 0 0.4 0.3
Long Term (Medium Clay) 0.75 0 0.25 0.375
*or
4
[1 ( )]
c
H
|

= whichever is larger;
+
Tschebotarioff (1973)

Note that these diagrams are not intended to represent actual earth pressure or its distribution
with depth but load envelopes from which strut loads can be evaluated. Clay is assumed to be
undrained and only total stresses are considered. Sands are assumed to be drained (through the
sheeting) with zero pore pressure. Where drainage is precluded behind a non-permeable wall,
hydrostatically distributed water pressure should be added to strut load.

If you design a strut force based on the apparent pressure diagram and uses simple supported
beams for the sheeting as proposed by Terzaghi and Peck, the sheeting may be somewhat over
designed, this over design was part of the intent of using these apparent pressure diagrams.

Tschebotarioff (1973) proposed another apparent pressure diagrams as shown in Figure 5.1 and
Table 5.2. It may be more nearly correct comparing with some case studies when the excavation
depth exceeds 16 m. It is found that when tieback is used, the loads developed resemble those of
5-6
a conventional triangular earth pressure distribution, instead of those usually developed in a strut
wall.

Layered Soils
A major shortcoming of these apparent pressure diagrams is how to apply when the retained soil
is stratified.

For sand-clay layer using o = 0 concept, the average unit weight and cohesion of an equivalent
clay layer can be calculated as (Peck, 1943):

( ) | |
clay sand sand sand
H H H
H
+ =
1
a
(5.1)

| |
u sand sand sand sand a
q n H H H K
H
C ' ) ( tan
2
1
2
+ = o (5.2)

where H = total height of the sand-clay layer
H
sand
= height of the sand layer

sand
= unit weight of sand

clay
= unit weight of clay
o = friction angle of sand
K
sand
= a lateral earth pressure coefficient of sand ( ~1)
n' = a coefficient of progressive failure ( 0.5 ~ 1; average value 0.75)
q
u
= unconfined compressive strength of clay

There is another available method by Bowles (1996) after Liao and Neff (1990).
1. Compute two Rankine-type pressure diagrams using the Rankine K
a
and K
o
and using
effective unit weights. Make a second pressure diagram for the GWT if applicable.
2. Plot the two pressure diagrams (use 0 for any tensile zones) on the same plot.
3. Compute the resultant R
a
and R
o
for the two pressure plots.
4. Average the two R-values.
5. Pick the shape of the apparent pressure envelope (a rectangle or a trapezoid) and calculate
the ordinate so that the area of the pressure envelope equals average R value (e.g. for
rectangle, ordinate = R/H).
6. Include the water pressure as a separate profile that is added to the preceding soil
pressures below the ground water table depending on the inside water level.

Conventional Design of Braced Excavation Walls
The conventional method, as shown in Figure 5.2, was for any multi-braced walls originally,
however, it is not for pile walls that may be due to the rigidity of piles.

The basic procedure is:
1. Sketch given conditions and indicate all known soil data stratification, water level, etc.
2. Compute the lateral pressure diagram by Tschebotarioff or Terzaghi and Peck method.
5-7
3. Calculate the strut loads based on simple beams (see Figure 5.2) by assuming hinges at
the strut support locations.
4. Design struts or tiebacks

You can also calculate the strut loads at each successive construction stage. The highest value at
each strut level is used for strut and waling design purposes. Similar calculation is also applied
to maximum moment and shear.

The struts are actually horizontal columns subject to bending. The load-carrying capacity of
columns will depend on the slenderness ratio, l/r. The l/r can be reduced by providing vertical
and horizontal supports at intermediate points.

Figure 5.2 Simplified Method of Analyzing the Sheeting and Strut Forces


Location of the First Support
The location of the first wale can be estimated numerically by making a cantilever wall analysis.
For cohesive soils, the depth should not exceed the depth of the potential tension crack. Where
lateral movement and resulting ground subsidence can be tolerated, the depth to the first strut in
sandy soils may be where the allowable bending stress in the sheeting is reached from a
cantilever wall analysis

Effect of Wall Embedment
The apparent pressure diagrams do not include the effects of the toe of the sheeting or walling
extending below the final formation level. There is an empirical design method that allows this
penetration of walling to be taken into account in strut load calculation. It has been used since
the mid-1950s and has been adequately confirmed. The procedure is shown in Figure 5.3.

5-8
1. Construct gross earth pressure diagram using Rankine theory.
2. Calculate value of total active P
a
and passive forces P
p
.
3. Calculate trapezoidal apparent pressure envelope with an ordinate 1.6 P
a
/H.
4. Calculate strut forces (F
1
, F
2
, F
3
, and F
4
) from the apparent pressure envelope by
splitting the distance between supports.
5. Check moment balance by taking moments about the F
1
. Rebalance support loads to
achieve moment equilibrium.
6. Calculate the factor of safety, mobilized passive resistance =
4
F P
p
(> 1.4)
7. Calculate the simply supported moment (M = wl
2
/8) between each support where w is
the apparent pressure envelope.


Figure 5.3 Empirical Method for Wall Embedment
Beam on elastic foundation and finite element method
Numerical analyses based on beam on elastic foundation are popular due to the advent of
powerful personal computers. They are efficient means to obtain the strut force and moment in
the sheeting. The soil is represented by springs attached to the piling. With the easy access of
finite element program, finite element method has been commonly used to analyze walls in
excavation. For linear FEM, the soil is characterized with an elastic modulus and a Poissons
ratio. Nonlinear soil models have been used. Although they cannot produce accurate results,
good designs are achieved constantly with good experience and engineering judgment.
5-9
Instability Due to Heave of Bottom of Excavation
Braced cuts in clays may become unstable as a result of the clay flows beneath the sheeting into
the excavation. The failure is analogous to a bearing capacity failure of foundation, only in
reverse; the failure is a shear failure in the soil below formation level, but caused by relief of
load and not by the application of load as occurs in a conventional foundation bearing failure.

For deep excavations with H/B > 1, Bjerrum and Eide (1956) can be applied to calculate the
factor of safety against base failure. For shallow or wide excavation with H/B < 1, the method
by Terzaghi (1943) can be used. These methods are shown in the Figure 5.4.

If the factor of safety against base heave is less than about 1.5, substantial soil deformation is
likely. Therefore, the embedment depth should be deeper. Many designers believe the FS
cannot be achieved by the use of a flexible retaining system, and that it may require a more rigid
wall (slurry walls or secant walls). Usually the embedment depth is kept less than or equal to
B/2. If lesser soil movement is necessary, a minimum factor of safety of 2 is required.



Figure 5.4 Instability of Braced Cuts
(a) Bjerrum and Eide (1956)
(b) Terzaghi (1943)
5-10
Stability of the Bottom of the Cut in Sand
Although it is unlikely, instability cohesionless soils can also be analyzed using NAVFAC (DM-
7, 1986). The stability is independent of the depth and the width of the excavation. The factor of
safety against base instability can be estimated.

1
1
2
tan 2 o

a
k N FS
|
|
.
|

\
|
= (5.3)

where N

is the bearing capacity factor of soil beneath the excavation level

1
is the unit weight of soil above the excavation level
o
1
is the friction angle of soil above the excavation level
k
a
is the active earth pressure coefficient of soil above the excavation level

2
is the unit weight of soil below the excavation level


Stability of Braced Cut in Sand Due to Piping
Instability of cut in sand mostly will be due to piping (hydraulic failure). When the ground water
table is encountered and the water level inside the cut is lowered below the ground water level by
pumping, instability may be created as a result of the upward seepage of water into the cut. The
designer must ensure that basal instability will not occur because of the flow of water.

As shown in Figure 5.5, design charts (DM-7, 1986) for penetration of cut-off walls to prevent
piping in sand can be used to estimate the stability.

Factor of safety against piping can be calculated as:

w w
w sub
h
d H
FS

) 2 ( +
= (5.4)

Where
H
w
= the height of water in the wall above the dredge line
d = the embedment depth

sub
= submerge unit weight of the soil beneath the base (formation level)
h
w
= the total hydraulic head loss

For stratified soils, the method of flow net and numerical analysis (e.g. program SEEP) are more
appropriate to estimate the FS against piping.
5-11



Figure 5.5 Instability Due to Piping
5-12
Ground Movements of In-situ Walls
In earlier portion of this chapter the analysis focuses on the adequacy of the strut loads,
anchoring, and sheeting or walling. The risk of excessive deformation of the walls is higher
when the excavation is deeper and the site has a greater plan area. This section will address
those factors which cause ground movement around an excavation and the amount of ground
movement. An excellent review on this topic was given by Clough & Schmidt (1981). Ground
movements caused by pile driving and dewatering were discussed in the report by D'Appolonia
(1971).

Movements of in-situ walls are a function of many factors:
- soil and groundwater conditions
- groundwater level
- initial horizontal Stress
- depth and shape of excavation
- preloading of the support system
- type and stiffness of the wall and its support
- methods of construction of the wall and adjacent facilities

Soil and Groundwater Conditions
In clays, the maximum lateral wall movement could be correlated with the factor of safety
against basal heave (Clough et al, 1979; Mana & Clough, 1981), which in turn depends on the
shear strength of the soil. The rate and magnitude of movement increase rapidly as the factor of
safety approaches one. If a clay is anisotropic instead of the presumed isotropic, the basal heave
factor of safety may be much lower and lateral wall movements as well as ground surface
settlements may be larger than expected (Clough & Hansen, 1981).

Wall movements and ground settlements are smaller in stiff soils, such as granular soils and stiff
clays, than in soft soils, such as soft and medium clays and compressible silts (Peck, 1969).

Initial Horizontal Stress
For excavated walls in soils with a high initial k
0
value, large soil and wall movements are
experienced even at shallow depths of excavation. The wall behavior is dominated by vertical
unloading caused by the excavation process

and large movements still occur even if the wall is
fully restrained from horizontal movement. For walls in soils with a low k
0
value

the
displacements are much smaller in magnitude (Potts and Fourie, 1984).

Groundwater Conditions
In practice, a perfectly watertight wall penetrating an impermeable soil layer at the bottom of the
excavation does not exist. Where water flows into the excavation, a decrease in groundwater
pressure will occur. This will cause an increase in effective stress and settlement of the soil
surrounding the excavation. Where groundwater has not been brought under complete control,
large, erratic and damaging settlement due to the flow or the migration of fines into the
excavation is not uncommon.

Depth and shape of the excavation
The deeper the excavation, the greater is the decrease in total stress, and thus the larger are the
5-13
movements of the surrounding soil. The shape of the excavation affects the basal stability in
clays. Thus, it affects the movement of surrounding soils.

Preloading of the support system
Preloading removes slack from the support system, and thus eliminates this potential source of
movement. Each application of preload reduces the shear stresses set up in the soil due to
previous excavation activities. This means that the soil is partially unloaded, and its stress-strain
response is stiffened until the next excavation step generates shear stresses that reload the soil
beyond its maximum previous level of shear stress. This temporary stiffening of the soil also
leads to reduce movements. Use of preloads in the struts or tiebacks reduces movement,
although there are diminishing returns at higher preloads. Very high preloads may, in fact, be
counter-productive, since local outward movements at support levels can damage adjacent
utilities.

Prestress loads calculated using the apparent trapezoidal pressure diagrams (Terzaghi and Peck,
1967) is more effective in reducing movements than the prestress loads in tiebacks calculated in
accordance with a triangular at-rest pressure diagram (Clough and Tsui, 1974). Using levels
slightly greater than those recommended by Terzaghi and Peck (1967) reduced the movements
Clough (1975). Hanna & Kurdi (1974) supported this finding in their model tests. Clough &
Tsui (1974) showed that movements of the tied-back wall, and the soil settlements behind the
wall, could be substantially reduced by a judicious choice of prestress load and support system
stiffness. In most cases, cross-lot struts prestressed to 50% of their design load will be
sufficiently rigid to restrict further movement at the level of the supports, and sufficiently low in
load to avoid being overstressed as additional excavation occurs (ORourke, 1981).


Type and Stiffness of the Wall and Its Support

The support system stiffness depends on the stiffness of the wall and its supports, the spacing
between supports, and the length of wall embedded below the excavation bottom. Goldberg et al
(1976) produced a valuable study into the effect of wall stiffness and support spacing and the
results are presented in Figure 5.6. The instability number of the base (H/C
u
) is plotted versus
the stiffness parameter (E
w
I
w
/h
4
), where E
w
I
w
is the flexural stiffness of the wall per unit length,
'h' is the vertical spacing between supports, H is the effective overburden pressure, and C
u
is the
undrained shear strength of the soil. A non-dimensional effective stiffness parameter was also
proposed (Koutsoftas et al. 2000, Clough at el. 1989):

4
EI
S
H
= (5.5)

where is either the unit weight of soil or unit weight of water.

Secant and tangent pile walls and structural slurry walls (diaphragm walls) are considered stiff on the
basis of the rigidity of the wall element. Walls that are considered flexible include steel sheet pile
walls and soldier pile and lagging walls
5-14



Figure 5.6 Effect of Basal Stability and Walls Stiffness on Lateral Wall Movement
Various boundary lines are drawn to establish zones of expected lateral wall movements. These
data suggested that an increase in E
w
I
w
/h
4
has a significant effect in reducing movements. The
movement is also a function of the factor of safety against basal heave, being more significant at
lower factors of safety than at higher ones. Increasing the stiffness of the strut or the tieback
decreases movements, but this effect shows diminishing returns at very high values of strut or
tieback stiffness.

Movements are also reduced as the depth to an underlying firm layer decreases and when the
wall toe is embedded into the underlying firm layer.

Methods of Construction of Walls and Adjacent Facilities

Associated Site Preparation Works
Site preparation may include the following activities
(a) Relocation and underpinning of utilities,
(b) Dewatering of aquifers above and below the base of the excavation,
(c) Construction of the excavation support system, and
(d) The installation of deep foundations

In some cases, the movements associated with the site preparation works will exceed those that
occur as a result of the excavation and support process. The relocation of utilities, for example,
5-15
may have a locally severe impact on an adjacent property, especially when trenching is carried
out close to pipelines and communication conduits.

Dewatering may consolidate the soil over an area, which substantially exceeds the area affected
by excavation-induced movements. Also, it often causes settlements well in excess of
excavation settlements. However, in areas that have been subjected to earlier dewatering
activities, settlements due to further dewatering are smaller than those in virgin ground due to the
stiffer response of the preconsolidated soil.

Wall construction may require predrilling for soldier piles, the use of vibratory hammers to
install sheet piles, or the excavation of slurry panels for concrete diaphragm walls. Each of
these can cause permanent movements, the magnitude and distribution of which will vary
according to the soil conditions, site history and details of the construction procedures.

Influence of Construction Factors
Additional movements of excavations, and even local failure, have been produced by late
installation of supports, over-excavation, pile driving, caisson construction, loss of water through
holes for tie-backs and joints or interlocks of slurry or sheet pile walls, remolding and
undercutting of clay berms, and surcharge loads from spoil heaps and construction equipment. A
review of the influence of construction factors on excavation movements was made by Clough &
Davidson (1977), and useful specific case history accounts can be found in Broms & Stille
(1976), Hansbo et al (1973), Lambe et al (1970), O'Rourke et al (1976), White (1976) and
Zeevart (1972).

Lambe (1970, 1972) demonstrated that variations in wedging techniques between the walings
and struts, and differences in excavation procedure, can result in doubling of the wall and soil
movements. Clough & Tsui (1974) used a finite element analysis to show that over-excavation
could lead to a 100% increase in movement.

Because these factors cannot always be quantified, it is difficult to make accurate prediction of
movements. However, many of the undesirable effects of these factors have been identified, and
they can therefore be anticipated and controlled through good specifications, well-planned
construction procedures and close supervision and monitoring. The designer must consider how
the excavation and subsequent construction should be carried out, identify critical construction
factors and, where possible, allow for them in performance estimates and specifications.


5-16
Prediction of Soil Movements
Prediction of soil movements behind a supported excavation can be made by the following
methods:
(a) Empirical methods
(b) Semi-empirical methods
(c) Numerical methods

Empirical Methods
Figure 5.7 is the first practical approach for estimating movements for in-situ wall systems
proposed by Peck (1969). The data used to derive the three zones shown in this figure were
taken from excavations supported by soldier piles or sheet piles with cross-lot struts or tiebacks.
(1) Zone I Sand and stiff to hard clay, average workmanship
(2) Zone II
(a) Very soft to soft clay
- Limited depth of clay below bottom of excavation.
- Significant depth of clay below bottom of excavation but H/C
u
< 5.14
(b) Settlement affected by construction difficulties.
(3) Zone III Very soft to soft clay to a significant depth below bottom of excavation and
with H/C
u
> 5.14.

Generally, where workmanship is average or above average and soil conditions are not especially
difficult, settlements should not exceed one percent of the excavation depth. In cases where
seepage can occur and soil consolidation results, the one percent figure can be exceeded (Lambe
et a1.,

1970). It should also be noted that construction technique could have a strong influence
on movements of strutted systems.


Figure 5.7 Observed Settlements behind Excavations (Peck, 1969)
Pecks recommendation has been found to be very conservative for stiff clays. Clough and
ORourke (1990) found that the average horizontal and vertical movements of support systems in
stiff clays were roughly 0.2 percent and 0.15 percent of the total excavated depth, respectively. Their
findings agree with guidance established in Canadian Foundation Engineering Manual (1985), which
5-17
states that lateral movements of temporary support systems in stiff clay are less than 0.2 percent of
the excavation depth. This compares to guidance established in NAVFAC DM-7.2 (1982) that
suggests in stiff fissured clays lateral movements may reach 0.5 percent of the total excavated depth
or higher depending on quality of construction.

Semi-empirical methods
Semi-empirical methods have been developed based on the estimated lateral wall deflection
profile obtained from simple numerical programs. The volume of the soil (V
s
) in the lateral
displacement zone is obtained by integrate from the wall deflection profile. Then V
s
is related to
the amount of soil movement vertically near the excavation area. Two methods are presented
here. The Caspes method is for excavation in clays, and the Bauers method is for excavation in
sands.

Caspes method
The lateral distance of the settlement influence is calculated as follow (Caspe, 1966) for the case
of the base soil being clay:
1. Compute a distance H
p
below the excavation level as (B = width of excavation area):
0
1
tan(45 ) ; 0
2 2
p
B when
H
B
o
o
o
=

+ >

(5.6)
2. Compute H' = H + H
p
, where H is the excavation depth.
3. Compute the approximate distance D [= H' tan (45 - o/2)] from the edge of the wall over
which ground loss occurs
4. Compute the surface settlement at the edge of the excavation wall as S
o
= 2 V
s
/D
5. Compute remaining ground loss settlements assuming a parabolic variation of S
i
from the
edge of the wall as S
i
= S
o
(1 - x/D)
2
, where x is the horizontal distance from the edge of
the wall.

Bauers method
A semi-empirical method for excavation in sands was proposed by Bauer (1984) as shown in
Figure 5.8. The settlement is related to the friction angle of the sand. The influence distance is a
function of the friction angle of sand, workmanship, and construction difficulty. The effects of
workmanship and construction difficulty on the influence distance are described by factors f
1
and
f
2
as listed in Table 5.4.

Table 5.4 Factors of workmanship and construction difficulty
Factor Workmanship Factor Construction Difficulty
Excellent Good Average Poor None Average Severe
f
1
0.8 0.9 1.0 1.1 f
2
1.0 1.02 1.05


Numerical methods
Finite element method has been used to study and calculate both the horizontal and vertical soil
movements adjacent to the excavation. Due to the complexity of the soil behavior, the
unpredicted construction factors, quantitative predication of soil movements is very difficult.
5-18
However, it has provided valuable qualitative studies. With the appropriate engineering
judgment, reasonable predication can be obtained.


Figure 5.8 Semi-empirical Method to Estimated Settlement (Bauer, 1984)

Base Movement Trends
Maximum movement - Stiff Clays, Residual Soils and Sands
As shown in Figure 5.7, Pecks data suggested that the movements of excavation support
systems in these soils were limited to 1% H. Goldberg et al (1976) showed that the maximum
horizontal movements for in-situ walls and settlements of the retained soil masses in such
materials were usually less than 0.5% H. ORourke (1990) found that the average horizontal and
vertical movements of support systems in stiff clays were roughly 0.2 percent of the total
excavated depth.

Excluding those special points, Figures 5.9 and 5.10 show that:
- The horizontal movements tend to average about 0.2% H
- The vertical movements tend to average about 0.15% H
- There is ample scatter in the data, with the horizontal movements showing more than the
vertical movements.
- There is no significant difference between trends of the maximum movements of
different types of walls, and this includes even the new soil nail and soil cement walls.
These two figures are useful to understand movement patterns and also can be used as design
tools to estimate maximum wall and soil movements. Although we dont know the maximum
movement for a particular soil, the linear trend between movement and excavation depth is
obvious.

A finite element study also predicted maximum lateral wall movements versus H also follow an
approximately linear response with excavation depth, centered around a trend line of 0.2% H.
5-19
This is consistent with the average behavior observed in Figure 5.9. The parameters wall
stiffness and strut spacing were found to have only a small influence on predicted movement
because the soil in these circumstances is stiff enough to minimize the need for the structure.
However, soil modulus and K
o
had a more significant impact (Clough and ORourke, 1990).


Figure 5.9 Observed Maximum Lateral Movements for Insitu Walls in Stiff Clays, Residual
Soils and Sands (Clough and ORouke, 1990)

Figure 5.10 Observed Maximum Soil Settlements by In-situ Walls (Clough and ORouke, 1990)

5-20

It is suggested here that variations in soil stiffness have a more profound effect on wall behavior
than system stiffness.

Maximum Movements - Soft and Medium Clays
As opposed to stiffer soils, basal stability in soft and medium clays may be at issue, and as a
result, movement patterns in these conditions can be dominated by deflections beneath the
excavation. Figure 5.11, a plot of maximum lateral movement of the wall versus FS of basal
heave, shows that as FS falls below 1.5, movements increase rapidly. It also illustrates the
influence that wall stiffness and support spacing. Caution is needed when using this chart for
design.


Figure 5.11 Design Curves to Obtain Maximum Lateral Wall Movement for Soft to Medium
Clays (Clough et al., 1989)

Clough et al (1979) and Mana & Clough (1981) reviewed case histories of sheet pile and soldier
pile walls in clays supported primarily by cross-lot struts. They found that the ratio of maximum
settlement movements ranges from 0.5 to 1.0 times the lateral wall movements.

Figure 5.12 shows that the maximum lateral movement can be correlated with the factor of safety
against basal heave defined by Terzaghi (1943, see Figure 5.4b). The movements increase
rapidly below a factor of safety of 1.4 to 1.5, while at higher factors of safety the non-
dimensional movements lie within a narrow range of 0.2% to 0.8%. Moreover, there do not
appear to be any significant differences in lateral movements between sheet pile walls whose tips
are embedded in an underlying stiff layer and those whose tips remain in the moving clay mass.

5-21

Figure 5.12 Soil Movements versus Basal Stability for Soft to Medium Clays

Displacement Profile

General patterns of ground movement are illustrated in Figure 5.13. They are obtained from
inclinometer and settlement measurements for braced and tied-back excavations (Milligan, 1984,
Clough and ORouke, 1990, Finno et al., 1989).


Figure 2.13 Typical Lateral Movement Profiles

Initial excavation before strutting
The excavation was deepened before supports were installed. Thus, deformation of the wall
occurred primarily as a cantilever-type movement. The horizontal strains produced in this mode
of deformation form a triangular pattern of contours that decrease in magnitude with depth and
5-22
distance away from the wall. Settlements during this stage of construction may be bounded
within the triangular distribution of displacement as shown in Figure 5.13a.

When the excavation advances to deeper elevations, upper wall movement is restrained by
installation of new support or stiffening of existing support members. In the deeper portion of
the excavation, inward bulging of' the wall caused tensile strains, the contours for which were
inclined at approximately 45
o
to the vertical. Deep inward movement of the wall occurs, which
is shown as an incremental component of the total displacement in Figure 5.13b.

The cumulative wall and ground surface displacement profiles are shown in Figure 5.13c.
Therefore, the displacement profile depends on the predominant movement pattern.
- In sand, stiff to very hard clay, cantilever movements predominate and the settlements
tend to follow a triangular pattern.
- In soft to medium clay, the deep inward movement is predominant and the settlements
tend to be bounded by a trapezoidal displacement.

Field measurements of horizontal strains at excavations in different types of soil showed similar
patterns, with triangular contours of strain caused by cantilever wall deformation and deep
concentric contours of strain bounding zones of maximum settlement caused by walls subject to
deep inward displacement.

Excavations in Sand
Figure 5.14 summarized the settlements adjacent to the excavations in predominantly sand and
granular soil profiles. The excavations were in granular soils above the water table by
dewatering. The settlement profiles are very consistent for different kind of supporting systems
(cross-lot struts, soldier pile and lagging with tiebacks, sheet piles with tiebacks, and slurry wall
with cross-lot struts).


Figure 5.14 Observed Settlement Adjacent to Excavations in Sands (Clough and ORourke,
1990)
5-23

Maximum settlements were typically less than 0.3% H. It is a triangular profile and distributes
over 2 times the excavation depth from the edge of the cut.

Excavations in Stiff to Very Hard Clays
Figure 5.15 summarized settlements for excavation sites in stiff to very hard clays for different
retaining systems. The displacements caused by ancillary construction activities were removed.
The settlements are only 0.3% H but are distributed over three times the excavation depth from
the edge of the cut.

Figure 5.15 Observed Soil Movements Adjacent to Excavations in Stiff and Very Hard Clays

In the horizontal movement, the majority falls within a triangular boundary with the same
dimensions as those pertaining to the observed settlements. A second zone was drawn which
contains highway excavations of London clay. They were affected by the low horizontal
stiffness of the support systems.


Excavations in Soft to Medium Clays

Figure 5.16 summarizes settlements for excavations in soft to medium clay, involving cross-lot
struts supporting sheet pile, soldier pile and lagging, and concrete diaphragm walls. Due to the
scatter of the data, it is difficult to identify the maximum settlement. However, if the settlements
5-24
were plotted as fractions of maximum settlement, a relative well-defined grouping of the data is
evident. The settlement distribution is bounded by a trapezoidal envelope in which two zones of
movement can be identified. At 0 s d/H s 0.75, there is a zone in which the maximum
settlement occurs. At 0.75 < d/H s 2, there is a transition zone in which settlements decrease
from maximum to negligible values.


Figure 5.16 Observed Soil Movements Adjacent to Excavation in Soft and Medium Clays

The largest differential soil movements occur near or within the transition zone. In almost all
cases, the maximum angular distortion was located between 0.5 H and 1.25 H from the edge of
cut. There is a linear relationship between the maximum settlement and the logarithmic of
angular distortion. Angular distortion appears to be relatively small when maximum settlement
is less than 50 mm. It increases exponentially with maximum settlement.

Figure 5.17 presents dimensionless settlement profiles recommended as a basis for estimating
vertical movement patterns adjacent to excavations in three different kinds of soils.

5-25



Figure 5.17 Dimensionless Design Settlement Profiles for Excavations in Different Soil Types

Be careful that these diagrams pertain to settlements caused during the excavation and bracing
stages of construction. Movements associated with other activities, such as dewatering, deep
foundation removal or construction, and wall installation, should be estimated separately.

Vertical settlements have been observed during the installation of the diaphragm wall at different
soils (granular soil, soft to medium clay, stiff to very hard clay). The maximum settlement ratio
is less than 0.15% H. The influence distance from the edge of the wall is about 2.5 H.

Driving of sheet-pile in a loose sand layer will also cause the layer to settle. Thus, the
geotechnical engineer should know how in-situ walls are constructed since otherwise he cannot
properly predict how the soil will behave.

General comments
- In stiff clays, in-situ walls will exhibit creep, a prompt installation of supports if movements
are to be minimized.
- Minimize struts spacing (strut stiffness is less important within normal range).
5-26
- Try to enforce the excavation limit such that no over excavation without the installation of
supports.
- It is advisable to penetrate the wall to a bearing layer where possible. Using sheet-pile, every
fifth sheet-pile can be penetrated to the bearing layer.

Summary
System stiffness and basal stiffness affect excavation in soft to medium clays but not in stiff
clays and sands.

Additional movements can also be generated by
(a) poor construction technique
(b) slow installation of supports after the excavation level is reached
(c) construction and removal of foundations within the excavation and,
(d) removal of soil below the design level of a support.

Increasing wall stiffness tends to reduce system movements, but this is most effective in soft to
medium clays. Support spacing is more important than wall stiffness in defining system stiffness
and helping control movements.


References
Bauer, G.E. (1984), Movements Associated with the Construction of A Deep Excavation,
Proc. of the 3
rd
Int. Conf. on Ground Movements and Structures, Cardiff, 694-706.
Boscardin M.D. and Cording, E.J. (1989) Building Response to Excavation-Induced Settlement,
Journal of Geotechnical Engineering, ASCE, Vol.115, No. 1, 1-21.
Broms, B. and Stille, H. (1976), Failure of Anchored Sheet Pile Walls, Journal of Geotechnical
Engineering, ASCE, Vol. 102, 235-251.
Canadian Geotechnical Society (1985), Canadian Foundation Engineering Manual The Canadian
Geotechnical Society, BiTech Publishers Ltd., 902-1030 West Georgia Street,
Vancouver, B.C., 456p.
CIRIA, (1993) Design and construction of sheet piled cofferdams, Special Publication, Thomas
Telford, 95.
Clough, G. W. (1975), Deep Excavations and Retaining Structure, Proc. Of the Short Course
Seminar on Analysis and Design of Building Foundations, Bethlehem, PA, 417-465.
Clough, G. W. and Davidson, R.R. (1977), Effect of Construction on Geotechnical
Performance, Proc. Of Specialty Session 3 on Relationship Between Design and
Construction in Soil Engineering, 9
th
International Conf. on Soil Mechanics and
Foundation Engineering, Tokyo, 15-53.
Clough, G. W. and Denby, G.M. (1977), Stabilizing Berm Design for Temporary Walls in
Clay, Journal of Geotechnical Engineering, ASCE, Vol. 103, 75-90.
Clough, G.W. and Hensen, L.A. (1981), Clay Anisotropy and Braced Wall Behavior, Journal of
Geotechnical Engineering, ASCE, Vol. 107, 893-913.
Clough, G.W., Hensen, L.A., and Mana, A.I. (1979), Prediction of Support Excavation
Movements under Marginal Stability Conditions in Clay, Proc. Of the 3
rd
International
Conf. On Numerical Methods in Goemechanics, Aachen, Vol. 4, 1485-1502.
Clough, G.W. and ORourke, T.D. (1990) Construction Induced Movements of Insitu Walls,
5-27
Proc. Design and Performance of Earth Retaining Structures, ASCE Special
Geotechnical Publication No. 25, ed. Lambe and Hansen, 439-470.
Clough, G.W. and Schmidt, B. (1981), Design and Performance of Excavations and Tunnels in
Soft Clay, Soft Clay Engineering, Ed. Brand and Brenner, Elsevier Scientific Co., 567-
634.
Clough, G.W., Smith, E.M. and Sweeney, B.P. (1989) Movement Control of Excavation
Support Systems by Iterative Design, Proc. ASCE Foundation Engineering: Current
Principles and Practices, Vol. 2, 869-884.
Clough, G.W. and Tsui, Y. (1974), Performance of Tieback Walls in Clay, Journal of
Geotechnical Engineering, ASCE, Vol. 100, 1259-1273.
Cowland, J.W. and Thorley, C.B.B. (1984), Ground and Building Settlement Associated with
Adjacent Slurry Trench Excavation, Proc. of the 3
rd
International Conf. on Ground
Movements and Structures, Cardiff, 723-738.
Davies, R.V. and Henkel, D.J. (1980), Geotechnical Problems Associated with the Construction
of Chater Station, Hong Kong, Proc. of the Conf. on Diaphragm Walling Techniques,
Singapore, paper RDS, 12 pp.
DAppolonia, D. J. (1971), Effects of Foundation Construction on Nearby Structures, Proc. Of
4
th
PanAmerican Conf. On Soil Mechanics and Foundation Engineering, San Juan,
Puerto Rico, Vol. 1, 189-236.
Goldberg, D.T. Jaworski, W.E. and Gordon, M.D. (1976), Lateral Support Systems and
Underpinning, Vol. 1: Design and Construction, Vol. 2: Design Fundamentals, Vol. 3:
Construction Methods, FHWA Reports, FHWA RD-75-128~130.
Hanna, T.H. and Kurdi, I. I. (1974), Studies on Anchored Flexible Retaining Walls in Sand,
Journal of Geotechnical Engineering, ASCE, Vol. 100, 1107-1122.
Hansbo, S., Hofmann, E. and Mosseson, J. (1973), Ostra Nordstaden, Gothenburg, Experiences
Concerning a Difficult Foundation Problem and its Unorthodox Solution, Proc., the 8
th

International Conf. on Soil Mechanics and Foundation Engineering, Moscow, Vol. 2.2
105-110.
Hsieh, P-G. and Ou, C-Y. (1998). Shape of Ground Surface Settlement Profiles Caused by
Excavation, Canadian Geotechnical Journal, Vol. 35, No. 6, pp. 1004-1017
Koutsoftas, D. C., Frobenius, P., Wu, C. L., Meyersohn, D. and Kulesza, R. (2000).
Deformations During Cut-and-Cover Construction of MUNI Metro Turnback Project,
Journal of Geotechnical and Geoenvironmental Engineering, ASCE, Vol. 126, No. 4,
pp. 344-359
Lamb, T.W. (1970), Braced Excavations, Proc. of the ASCE Specialty Conf. on Lateral Stress
in the Ground and Design of Earth Retaining Structures, Ithica, New York, 149-218.
Lamb, T.W. (1972), Predicted Performance of Braced Excavations, Proc. ASCE Specialty
Conf. on Performance of Earth and Earth-Supported Structures, Lafayette, Indiana, Vol.
3, 255-256.
Lamb, T.W., Wolfskill, L.A. and Wong, I.H. (1970), Measured Performance of Braced
Excavation, Journal of Geotechnical Engineering, ASCE, Vol. 96, 817-836.
Mana, A.I. and Clough, G.W. (1981), Prediction of Movements for Braced Cuts in Clay,
Journal of Geotechnical Engineering, ASCE, Vol. 107, 759-777.
Milligan, G.W.E. (1984), Soil Deformations Behind Retaining Walls, Proc. of the 3
rd

International Conf. on Ground Movements and Structures, Cardiff, 707-722.
5-28
Morton, K., Cater, R.W. and Linney, L. (1980a), Observed Settlements of Buildings Adjacent
to Stations Constructed for the Modified Initial System of the Mass Transit Railway,
Hong Kong, Proc. of the 6
th
SE. Asian Conf. on Soil Engg, Taipei, Vol. 1, 415-429.
Morton, K., Leonard, M.S.M. and Cater, R.W. (1980b), Building Settlements and Ground
Movements Associated with Construction of Two Stations of the Modified Initial
System of the Mass Transit Railway, Hong Kong, Proc. of the 2
nd
International Conf.
on Ground Movements and Structures, Cardiff, 788-802.
NAVFAC (1982), Design Manual 7.2 Foundation and Earth Structure, Department of the
Navy, Naval Facilities Engineering command, Washington, D.C., 244 p.
ORourke, T.D. (1981), Ground Movements Caused by Braced Excavations, Journal of
Geotechnical Engineering, ASCE, Vol. 107, 1159-1178.
ORourke, T.D., Cording, E.J. and Boscardin, M. (1976), The Ground Movements Related to
Braced Excavation and Their Influence on Adjacent Buildings, U.S. Department of
Transportation, Report, DOT-TST 76, T-23.
Peck, R.B. (1969), Deep Excavations and Tunneling in Soft Ground, Proc. Of 7
th
International
Conf. On Soil Mechanics and Foundation Engineering, Mexico City, State-of-the-art
Volume, 225-290.
Potts, D.M. and Fourie, A.B. (1985), The Effect of Wall Stiffness on the Behavior of A Propped
Retaining Wall, Geotechnique, Vol. 35, 347-352.
Stroud, M.A. and Sweeney, D.J. (1977), Discussion (Description of a diaphragm wall test
panel). A Review of Diaphragm Walls, Inst. of Civil Engineers, London, 142-148.
Terzaghi, K. (1943), Theoretical Soil Mechanics, John Wiley and Sons, New York.
Terzaghi, K. and Peck, R.B. (1967), Soil Mechanic in Engineering Practice, 2
nd
Edition, John
Wiley and Sons, New York.
White, E.E. (1976), Foundation Difficulties, Methods of Control and Prevention, Proc. Of the
Short Course Seminar on Analysis and Design of Building Foundations, Bethlehem,
PA, 671-700.
Zeevart, L. (1972), Foundation Engineering for Difficult Subsoil Conditions, Van Nostrand
Reinhold, New York.

You might also like