You are on page 1of 56

Design of vertical soil support

5
This chapter addresses the key design items of earth (and water) pressure on
a vertical wall and the analysis of the wall to withstand these pressures.
Cantilevered, single-propped (or anchored) and multi-propped walls will be
considered for both temporary and permanent works.
Geotechnical categories for excavation works are defined in Eurocode 7
(EC7)1 as follows.
. Category 1. Walls that are small and relatively simple structures with a
retained height not exceeding 2 m.
. Category 2. Conventional structures with no abnormal risks or
unusually difficult ground or loading conditions. These walls require
site-specific geotechnical data to be obtained and analyses to be made.
. Category 3. Walls or parts of structures which fall outside categories 1
and 2 and include very large structures, ‘one-off’ structures and those
involving high risk such as unusual ground conditions or loading (for
instance, seismicity).
The design of a typically deep wall for a basement, cofferdam or cut-and-cover
structure will require decisions on walling method, use of props or anchors,
methods of dewatering/recharge, consideration of repropping during con-
struction, acceptable safety factors and allowable settlement and deformation.
The wall design will require the consideration of both ultimate and service-
ability limit states. In practical terms the following matters will need to be
addressed.
(a) At ultimate limit state:
(i) overall stability; strut anchor or waling failure, bending stress
failure in the sheeting, passive failure of soil below successive
stages of excavation, failure by lack of vertical equilibrium of the
wall
(ii) foundation heave – in soft clays, risk of failure by unloading–
bearing capacity failure
(iii) hydraulic failure – piping in cohesionless soils with high external
groundwater table.
Walls should be designed, wherever possible, so that adequate warning
is given in advance of collapse by visible signs. The risk of progressive
failure of retaining structures needs to be particularly addressed,
especially at the temporary works stage.
(b) At serviceability limit state:
(i) deformation of sheeting – avoidance of excessive reduction in
working space or space for permanent structure at construction
stage
(ii) excessive soil movements behind the wall or sheeting; vertical
settlements not to exceed the permitted settlements of existing
structures, services or highways

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 147

(iii) unacceptable leakage through or beneath the wall


(iv) unacceptable flexural cracking of the structure
(v) unacceptable cracking in reinforced concrete walls. Crack control
requirements in the UK are specified in BS 81102 or where more
rigorous water resistance is needed in BS 80073 . For highway
structures in the UK BS 54004 and Highways Agency BD 42/005
apply.
The following practical points arise concerning wall and soil movement:
. The magnitude of movements in walls and retained soil is dependent
in large measure on workmanship standards during wall installation,
excavation and wall support works.
. Generally, flexible walls with many propping levels will give similar
displacements to stiff walls with few props.
. Wall embedment depth below formation level influences soil deforma-
tion, particularly in soft clays.
. Stiffness of propping systems influences soil deformation.
. Top-downwards construction minimizes soil deformation and settle-
ments.
. Installation of a temporary or permanent support of high stiffness at
shallow depth in an excavation is an effective control of wall deforma-
tion and soil deformation.
. Ground movement cannot be estimated with precision. As a broad
indication of movements due to wall installation the following may be
helpful.
In stiff clay: For bored piles: vertical settlement at wall, 0.04% of depth.
For diaphragm walls: vertical settlement at wall, 0.05% of
wall depth.
The distance behind the wall to negligible settlement is approximately twice
wall depth. Similarly, broad indications of patterns of wall deflection and
settlements can be obtained from empirical results published by Peck6 ,
Clough et al.7 , Clough and O’Rourke8 , St John et al.9 , Carder10 , Fernie and
Suckling11 , Carder et al.12 and Long13 .
For excavations in stiff clay, observed soil settlement at the wall due to
excavation in front of the wall is generally within the range 0.1% of
excavation depth for high-stiffness walls to 0.3% of excavation depth for
low-stiffness walls. In sands, a limited number of observations show a similar
vertical settlement at the wall of 0.1 to 0.3% of excavation depth. (The
Author’s experience of several excavations to 20 m depth in medium dense
silty sands in Saudi Arabia using anchored king post walls as support
shows consistent vertical settlement at the wall of less than 0.1% following
excavation to full depth.)

Earth pressures: Mohr’s circle of stress can be used to illustrate graphically the limiting
limiting horizontal horizontal soil pressures which can be generated in active or passive states
pressure at either side of a retaining wall. The pressures will be considered in terms
of effective stress, ignoring, for the time being, the effects of wall friction or
wall adhesion.
The problem of assessing earth pressures can be solved using the Coulomb
equation for shear strength  0 ¼ c0 þ 0 tan 0 , where c0 represents effective
soil cohesion, 0 represents effective vertical stress and 0 represents the
angle of shearing resistance in terms of effective stress. Figure 5.1 shows the
introduction of effective stress parameters, the total vertical stresses and

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
148 Deep excavations

Fig. 5.1. Mohr’s circle values of


limit pressures and values of Ka
and Kp

water pressures (and hence effective vertical stresses) to calculate active and
passive earth pressure according to whether horizontal stress is less than or
greater than vertical stress. Mohr’s circles are seen to touch the failure envel-
ope to give limiting values in the extremes of minimum and maximum hori-
zontal stresses.
Using active and passive earth pressures coefficients,
0H ¼ Ka 0v ð39Þ
and
0H ¼ Kp 0v ð40Þ
where
1  sin 0
Ka ¼ ð41Þ
1 þ sin 0
and
1 þ sin 0
Kp ¼ ð42Þ
1  sin 0
The values of 0 and c0 can be obtained for clay samples from drained triaxial
0
Table 5.1 Estimation of  for tests or undrained tests with pore-water pressure measurement. The effects of
cohesive soils17 stress level, rate of strain for testing, the degree of weathering and of sample
Plasticity 0 swelling before the test, will all influence the test result. When test results are
index not available for 0 , the relationship in Table 5.1 may be used for conservative
values. The design parameters 0 and c0 will cause considerable variation in
15 308 values of limiting soil pressure, and while it is not unusual to assume c0 ¼ 0,
20 288 the effect of this assumption is to reduce to low values the limiting passive
25 278 pressures immediately below dredge or formation level. In turn, this tends
30 258 to produce high wall or sheeting moments and increased strut or anchor
40 228
loads near this level.
50 208
80 158 The values of 0 in cohesionless soils should be based on in situ tests where
possible. The relationship between standard penetration resistance and 0 is

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 149

Fig. 5.2. Relationship between


standard penetration
resistance and 0 17

Table 5.2 Estimation of 0 for


weak rocks19 Stratum 0

Chalk 358
Clayey marl 288
Sandy marl 338
Weak sandstone 428
Weak siltstone 358
Weak mudstone 288
Notes. The presence of a preferred orientation of joints, bedding or cleavage
in a direction near that of a possible failure plane may require a reduction in
the above values, especially if the discontinuities are filled with weaker
materials.
Chalk is defined here as unweathered medium-to-hard, rubbly or blocky
chalk.

shown in Fig. 5.2. Table 5.2 gives approximate values for the effective angle of
shearing resistance of soft rocks considered, somewhat conservatively, as a
mass of granular fragments.

Horizontal earth pressure: at rest


In an undisturbed soil mass the horizontal earth pressure, the at rest pressure,
is K0 0v , where K0 is the coefficient of earth pressure at rest and 0v is the
effective vertical stress. For normally-consolidated soils K0 is approximately
equal to the empirical expression
0H
KONC ¼ ¼ 1  sin 0 : ð43Þ
0v
For normally-consolidated clay KONC lies in the range 0.55 to 0.65. For over-
consolidated clays the removal of considerable overburden pressures within
geological time leaves horizontal stress as the major principal stress. The
reduction in vertical stress is shown at any depth by the over-consolidation
ratio (OCR), which is the ratio of the maximum vertical effective stress ever
experienced by the soil to the current vertical effective stress at that depth.
The value of 0v can be obtained from laboratory consolidation tests and
the general relationship between K0 and the OCR is given in Fig. 5.3 after
Brooker and Ireland14 . The maximum value of K0 may lie in the range 2 to

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
150 Deep excavations

Fig. 5.3. Relationship between


coefficient of earth pressure
at rest K0 and
over-consolidation ratio14

3 for some clays, reducing with depth since OCR reduces with depth. K0 can
also be estimated from pressuremeter test results.

Horizontal earth pressure based on long-term, drained, effective


stress values
The limiting horizontal active and passive earth pressures acting on the wall at
any depth z are given, respectively, by
p0a ¼ Ka 0v ¼ Ka ðz þ q  uÞ ð44Þ
and
p0p ¼ Kp 0v ¼ Kp ðz þ q  uÞ ð45Þ
and for a soil with a cohesion intercept
p0a ¼ Ka ðz þ q  uÞ  2c0 ðKa Þ1=2 ð46Þ
and
p0p ¼ Kp ðz þ q  uÞ  2c0 ðKp Þ1=2 ð47Þ
where  is the bulk soil density (saturated density if below water level), q is the
surcharge on ground surface, and u is the pore-water pressure. These values,
known as Rankine values, are subject to change when the effects of wall fric-
tion and wall adhesion are considered. Walls are not made of perfectly smooth
material and wall friction  and adhesion cw vary soil stresses proportionately.
The limiting effective active and passive pressures acting horizontally at a
depth z are:
p0a ¼ Ka ðz þ q  uÞ  Kac c0 ð48Þ
and
p0p ¼ Kp ðz þ q  uÞ  Kpc c0 ð49Þ
where
  1=2
cw
Kac ¼ 2 Ka 1  0 ð50Þ
c
and
  1=2
cw
Kpc ¼ 2 Kp 1  0 ð51Þ
c
and c0 is the effective shear strength.
The pore-water pressure is added to the effective horizontal earth pressure
to give the sum of earth and water pressure: pa ¼ p0a þ u and pp ¼ p0p þ u.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 151

Fig. 5.4. Coefficient of active earth pressure (horizontal Fig. 5.5. Coefficient of passive earth pressure (horizontal
component) for a horizontal retained surface15 component) for a horizontal retained surface15

The earth pressure coefficients Ka and Kp have been calculated for various
curved failure surfaces; those frequently used are due to Caquot and Kerisel15
and are based on a logarithmic spiral surface as shown in Figs. 5.4 and 5.5.
The coefficients derived from these figures have been corrected to give
horizontal pressures although the actual pressures are inclined at the angle
of wall friction  to the horizontal. Value of Kac and Kpc can be calculated
from the expressions above.

Horizontal earth pressure based on short-term undrained stress


values
In fine-grained soils such as fine silts and clays, the relatively low permeability
only allows slow changes in moisture content, pore pressure and overall
volume. The instant that load is applied, pore pressure increases and the
shear strength of the soil, which depends on the locked-in effective stress, is
only able to increase slowly as pore pressures reduce. This immediate shear
strength, the undrained value cu , applies before any volume change and
pore pressures alter or, expressed more exactly, the value of cu can only be
used where a new system of boundary stresses is imposed but prior to any dis-
sipation of the new excess pore-water pressures. The undrained shear strength
cu is only correctly used immediately load is applied and it is strictly illogical to
use it as soon as pore pressures change. As effective stresses change with pore
pressure, shear strength also changes. In clay soils, where the retaining wall
structure deforms and attempts to move away from the retained soil bulk,
negative pore pressures are generated in the retained soil as excavation pro-
ceeds in front of the wall. In highly fissured or laminated clays the reduction
in negative pore pressure may proceed relatively quickly and the original value
of cu quickly becomes inapplicable. The use of cu in such clays can therefore
become over-optimistic in retaining wall analysis when the effective stress (on
which soil strength depends) reduces as negative pore pressure relaxes.
In soft and very soft homogeneous clays which do not benefit from drainage
paths due to fissuring and jointing, low overall permeability of the soil fabric
may prevent more efficient change in pore pressure, and the use of cu ,
although strictly applicable at the instant of stress change prior to pore
pressure change, can be more satisfactorily relied upon. Indeed, for retaining
walls founded in soft clays it is prudent to obtain both drained and undrained
soil strengths and undertake analysis of both total and effective stress

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
152 Deep excavations

conditions. Temporary works design in these soils, in which the wall structure
may have a limited design life, say, of six months, may therefore use total
stress methods with some confidence, and the design parameter of undrained
shear strength can be conveniently and accurately obtained from in situ shear
tests, such as vane tests. Where such short-term conditions do not apply,
either by the greater permanence of the temporary wall or due to its role in
the permanent structure, analysis of drained effective stress becomes neces-
sary. The choice between the two methods, undrained or drained, depends
strictly on the rate of change of pore pressure which, in turn, depends on
the permeability of the structure, although it must be restated that the
undrained analysis is only strictly applicable immediately before any dissipa-
tion commences. In practice, especially for small schemes, sufficient undrained
shear measurements may be more readily available than test results from
drained samples. However, the availability of test results should not excuse
the application of an illogical method of analysis. The short-term undrained
method is therefore presented here with the recommendation that its use
should be restricted to soft clay conditions for temporary works of short
duration. Schemes of longer duration should be checked by both drained
and undrained analyses.
In terms of undrained conditions the limiting horizontal active and passive
earth pressures acting on the wall at any depth z are given by
pa ¼ Ka v  Kac cu ¼ Ka ðz þ qÞ  Kac cu ð52Þ
and
pp ¼ Kp v  Kpc cu ¼ Kp ðv þ qÞ  Kpc cu : ð53Þ
The pressure coefficients are
 
cw 1=2
Kac ¼ 2 1 þ ð54Þ
cu
and
 
cw 1=2
Kpc ¼2 1þ : ð55Þ
cu

Tension cracks The active pressure near ground level will be a minimum for the retaining wall
and, in cohesive soils, surface tensions within the clay can cause cracks which
can fill with groundwater or rain water. In this case, the water pressure within
the potential crack should be allowed as a pressure on the back of the wall.
Wall adhesion will not apply over the depth of the crack, which is ð2cu  qÞ=.

Minimum values of calculated active pressure


In the UK the prolonged use, from 1951 until 1994, of the Civil Engineering
Code of Practice for earth-retaining structures16 indicated both the sound
principles on which it was based and the difficulties which have been experi-
enced in agreeing revisions to it. Although the Code was based on calculation
of earth pressure in terms of total stress and undrained shear values, it
prudently recommended the use of minimum active pressures for design
irrespective of those calculated from test values of cu using the expression
pan ¼ Ka v  Kac cu .
The Code stated that for cohesive soils the methods of calculation were not
final. In the design of structures to retain cohesive soils, a suitable addition
should be made to the calculated active pressure of the soil in order to provide
for uncertainties. Where, as may be the case with stiffer clays, the total

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 153

pressure on the wall as calculated according to the Code was small, the value
of the total pressure to be assumed for the purposes of design should not be
less than that found by assuming the horizontal pressure at any depth to be
that due to a fluid with a density of 5 kN/m3 . This safety net for minimum
pressure at the active side of the wall was most prudent but did not find
mention in some later Codes.

Softening of clays The process of clay softening is accelerated over time by tension cracks near
the ground surface, by the joints and fissures which frequently occur in stiff,
over-consolidated clays, and by the laminations of silt or sandy silt partings
that can occur in all clays. These routes for easier passage of water not only
increase the permeability of the soil fabric but can form failure surfaces them-
selves of softened planes of clay dividing stiffer, unsoftened blocks of clay
bounded by softened, lubricated surfaces. The changes in pore-water pressure
as these pressures equalize with hydrostatic groundwater levels are therefore
also associated with a loss of clay strength due to softening along surfaces
such as joints, fissures, pre-existing shear surfaces, laminations and tensile
zones. For this reason it is often prudent to assume a complete loss of cohesive
strength in terms of effective stress as the clay softens with time along
preferred drainage surfaces.

Wall friction and wall The effect of assumptions regarding wall friction in granular soils and adhesion
adhesion in clay soils on calculated values of active and passive pressure is considerable,
and while it is self-evident that wall surfaces are not smooth, design values of
friction and adhesion should be chosen carefully. The mobilization of wall
friction, to reduce earth pressures on the active side and increase earth
pressure on the passive side, necessitates the downwards movement of a soil
wedge on the active side and the upwards movement of a similar soil wedge
on the opposite, passive side, as shown in Fig. 5.6. Important changes to
these relative soil and wall movements are caused by the downwards vertical
wall movement due to vertical loads on the wall and the vertical load compo-
nents of inclined pre-loaded ground anchors, both of which tend to mobilize
passive wall friction/adhesion but tend to reduce active friction/adhesion.
Typical values of active wall friction and adhesion as quoted in Codes and
values for both active and passive friction and adhesion in the CIRIA report
for sheet piled cofferdams17 are shown in Tables 5.3 and 5.4. These values are
all based for wall friction on peak values of  and it is understandable that
some conservatism should apply since the peak values themselves may not
apply. More logically, it may be prudent to use max ¼ 0crit if the critical
state value is known. If the wall surface texture is less than the typical particle
size of the soil ðD50 Þ, lower values of  would apply.
The relative movement between wall and soil necessary to generate wall
friction or adhesion may not always occur; where the toe of the wall is

Fig. 5.6. Wall friction acting on


active and passive soil
wedges48

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
154 Deep excavations

Table 5.3 Values of active wall


friction angle and adhesion Construction Wall friction Wall adhesion
angle, 

Concrete, brick 208 Where cw is less than 50 kN/m2 use


cw ¼ cu
Steel piling, tar/bitumen 308 Where walling or sheeting does not
coated penetrate any appreciable depth use cw ¼ 0
Uncoated sheet piles 158 Where cw > 50 kN/m2 use cw ¼ 50 kN/m2

Table 5.4 Active and passive


values of wall friction/adhesion Wall friction angle,  Wall adhesion

Active 0.67 0.5c0 or 0.5cu >


= 50 kN/m2
Passive 0.5 . 0 .
0 5c or 0 5cu >= 25 kN/m2

founded on hard rock, wall friction/adhesion values quoted in Table 5.4


should be reduced by 50% for dense granular materials and stiff over-
consolidated clay. For overlying loose granular soil, where the wall is founded
on rock, wall friction/adhesion should be ignored. For anchored walls where
there is a tendency for the wall to move downwards relative to the soil, zero
wall friction should be allowed.
Neither wall adhesion nor wall friction should be allowed in granular sub-
soils where the proximity of machinery, railways or vehicular traffic causes
vibrations which could be transmitted through the subsoil to the wall surface.
With cast in situ reinforced concrete diaphragm walls and reinforced con-
crete piled walls cast under bentonite slurry, values of wall friction/adhesion
for active and passive pressure can be assumed to be similar to those for
concrete walls given in Table 5.3.

Magnitude of The variation in horizontal pressure, either active or passive, generated by


movement needed to movement of the wall away from or towards the soil mass, is shown in
mobilize limit Fig. 5.7. The curve, first derived by Terzaghi and Peck18 was based on tests
pressures in dense sand. Two conclusions are evident:
. the passive coefficient is much higher than the active coefficient
. the deformation needed to fully mobilize the limit passive pressure
is much greater than that required to fully mobilize the limit active
pressure.
The relative deformation needed to fully mobilize Ka and Kp by means of the
stress path followed by a soil element behind the wall in both normally- and

Fig. 5.7. Relationship between


strain required to mobilize
active and passive pressure for
medium dense sand48

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 155

Fig. 5.8. Stress paths followed


for elements behind and in
front of a retaining wall in
normally- and
over-consolidated clays48

over-consolidated clay conditions is shown in Fig. 5.8. On the active side there
is no change in vertical stress due to excavation, while on the passive side the
effective stress path in drained loading may take varying directions depending
on the relative importance of the relief of overburden to excavation and the
horizontal pressure exerted on the wall.
The stress change needed to fail a normally-consolidated clay in active
pressure is much less than that required to fail the same clay in passive
pressure. The stress path for over-consolidated clay shows the reverse to be
true, however. Terzaghi and Peck’s18 plot of wall movement against mobilized
pressure for dense sand may not therefore, accurately show the strains
necessary to generate limit active and passive pressures in both normally-
and over-consolidated clays.
BS 800219 followed the concept of limit state design, with greater emphasis
on wall conditions at the serviceability limit state than collapse conditions at
the ultimate limit state. The basis of this method is the reduction of soil peak
shear strength values on both the active and passive sides of the wall, to values

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
156 Deep excavations

which would be representative of mobilized shear strengths with the permitted


wall movement. BS 800219 specified that where wall displacements are
required to be less than 0.5% of the wall height, the representative undrained
shear strength should be divided by a mobilization factor M not less than 1.5.
For designs using effective stress values, the value of M was specified to be
1.2 to reduce horizontal wall movements to 0.5% of wall height. For
effective stress design, the calculation of wall equilibrium and the structural
components of the wall was specified to be based on the lesser of:
(a) the representative peak strength of the soil divided by a factor M ¼ 1:2
representative tan 0max
design tan 0 ¼
M
representative c0
design c0 ¼
M
(b) the representative critical-state strength of the soil.
The earth pressures on the active side are thus increased and similarly those on
the passive side are reduced at serviceability state by the use of the reduced
peak shear strength. The Code concludes that structural design of the wall
components can then usually be made without the application of partial
load factors to bending moments and internal structural forces.
This principle of design using soil strengths that can be mobilized at
working conditions in the serviceability limit state is logically extended in
BS 8002 to consideration of wall friction and wall adhesion. It recommends
that the design value of friction or adhesion at the interface with the wall
should be the lesser of either the values obtained from tests or 75% of the
design shear strength to be mobilized in the soil itself, that is, using
design tan  ¼ 0:75 design tan 0
design cw ¼ 0:75 design cu
and since for the soil mass M ¼ 1:2
representative tan 0
design tan 0 ¼
1:2
this is equivalent to
design  2
0 ffi
representative  3
and similarly, in total stress analysis
design cw
¼ 0:5 after taking M ¼ 1:5:
representative cu
The overall effect of reducing peak shear strengths to shear strengths which
are mobilized in working conditions by using such M values as are specified
in the Code may not be necessary when walls are designed in stiff over-
consolidated clays, where relatively small movements are necessary to
mobilize limiting active and passive pressures. On the other hand, the recom-
mendation that bending moments and internal forces derived from earth
pressures calculated from mobilized soil strengths (peak strengths reduced
by the mobilization factor M), without the application of a further partial
safety factor on earth pressures allowed in designs based on ultimate struc-
tural strengths, may not accurately reflect the variations in load that occur
in practice along the length of a wall.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 157

Fig. 5.9. Influence of wall


flexibility on pressure
distribution: (a) rigid wall;
(b) flexible wall48

BS 8002 also specified that in checking the wall stability and soil defor-
mation all walls should be designed for a minimum design surcharge loading
of 10 kN/m2 and a minimum depth of additional unplanned excavation in
front of the wall. This additional design depth was specified to be not less
than 0.5 m or 10% of the total height retained for cantilever walls or of the
height retained below the lowest support level for propped or anchored
walls. It should be noted that BS 8002 was originally applicable to walls up
to about 8 m high, and no differentiation is made between temporary and
permanent walls. A subsequent revision increased this wall depth to 15 m.
The imposition of such rules for all walls irrespective of design life would
appear to be over-stringent.

Wall flexibility The stiffer the wall, the greater the earth pressure it attracts and, conversely,
the greater flexure of the wall the less pressure (and moment) induced in it by
the soil it retains. This phenomenon, originally demonstrated in model tests
on an anchored wall by Rowe20 will be referred to in greater detail later in
this chapter.
Consider, however, the difference in deformed shape of a relatively stiff
anchored wall and that of an anchored flexible wall, as shown in Fig. 5.9.
The flexible wall distorts outwards by a considerable amount at mid-span
relative to the stiff wall, causing a reduction in pressure and an effective reduc-
tion in vertical span of the wall due to a rise in elevation of the centre of
passive support below excavation level. Anchor loads and pressure below
excavation level therefore both increase, but bending moment in the wall
reduces due to both earth pressure reduction at mid-span and the reduction
in the span itself. Due to the movement of the flexible wall away from the
soil at mid-span, the soil tends to arch vertically and, providing neither the
anchor nor the passive support below formation level yield once mobilized,
the abutments to the arch sustain more load and relieve soil pressure from
mid-span.

Surcharge loads The effects of additional loads, surcharging the wall, must be taken into
account at both ultimate and serviceability limit states. Loads due to adjacent
highways, railways, variations in ground level, storage areas for building
materials and plant loads need to be taken into account.

Highway traffic loading


In the UK, guidance on highway traffic loading is given in the publication
BD 37/0121 . Two categories of loading are specified, HA loading that
normally occurs on highways and HB loading, corresponding to abnormal

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
158 Deep excavations

vehicle loads such as industrial loads. The number of HB units is generally


specified. The loadings may be summarized as follows.
. HB loading (45 units): equivalent surcharge ¼ 20 kN/m2
. HB loading (37 units): equivalent surcharge ¼ 16 kN/m2
. HB loading (30 units): equivalent surcharge ¼ 12 kN/m2
. HA loading: equivalent surcharge ¼ 10 kN/m2

Fig. 5.10. Coefficients of earth pressure (horizontal component) for vertical walls with inclined backfill for varying wall
friction : (a) active; (b) passive15

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 159

Railway loading
Railway universal loading (RU) allows for all combinations of rail traffic
which currently runs on UK track – or is expected to run in the future. RL
is a reduced loading, for lighter traffic, such as in use on underground
railways. Both loads are applied over the track width, overall the length of
sleepers. The loadings are:
. RU loading: equivalent surcharge ¼ 50 kN/m2
. RL loading: equivalent surcharge ¼ 30 kN/m2

Surcharge loading: sloping ground


The effect of a sloping ground surface can be taken into account using curves
due to Caquot and Kerisel15 reproduced in Fig. 5.10, giving values of Ka and
Kp for vertical walls, inclined backfill and wall friction varying from  ¼ 0 to
 ¼ 0 . The effect of layered strata behind the wall should be taken into
account in calculations by using modified coefficients for each layer from
the curves. The values presented assume that the sloping ground extends a
sufficient distance from the wall to exceed the lateral dimension of the poten-
tial critical slip plane of the failure wedge (approximately at an angle of
458 þ =2 to the horizontal from the toe of the wall for the active case, and
458  =2 for the passive case).
The CIRIA report17 gives a simple adjustment to the pressure diagrams
(Fig. 5.11(a) and (b)) to show the effect of berms, on the active and passive
sides of the wall in cohesionless soil. The report notes that the risk for sliding
failure should be checked at the base of the berm, especially where the berm
soil has a higher angle of shearing resistance than the soil beneath it. Where
a gravel berm overlies a stiff clay, the report suggests it may be too optimistic
to use the undrained clay strength in checking sliding.
The NAVFAC design manual22 gives a graphical method for the calcula-
tion of passive soil resistance due to an earth berm, as shown in Fig. 5.11(c).

Surcharge loading: point loads and line loads


The CIRIA report17 reproduced a simple method due to Krey for use in
granular soils. The horizontal pressures caused by concentrated point and
line loads as estimated by Krey are shown in Fig. 5.12 and should be added
to the earth pressure diagram.
An alternative method of computing the effects of point and line loads by
Terzaghi was reported in the NAVFAC design manual22 and was commented
upon in the CIRIA report17 . According to the CIRIA report the method,
a modified Boussinesq distribution taking into account field and model
measurements, possibly underpredicts the pressure and tends to place the
centre of pressure influence too low down the wall. Some caution should be
applied, therefore, in the use of the method where the applied loads are
high in relation to the weight of soil retained. It should be noted, however,
that the method presumes an unyielding rigid wall and the lateral pressures
are already approximately double the values obtained from elastic equations.
Design charts from NAVFAC22 for line and point loads are reproduced in
Fig. 5.13.

Surcharge loading: uniform rectangular surface load


The NAVFAC design guide also provides the means to estimate the lateral
pressure on an unyielding wall due to a uniform rectangular load on the
surface at the rear of the wall, after Goldberg et al.23 (Fig. 5.14). The guide
comments that any yield of the wall during application of the load will tend
to reduce the pressures calculated from the values given by Goldberg et al.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
160 Deep excavations

Fig. 5.11. (a), (b) Pressure diagrams showing the effect of sloping ground on the active and passive sides of the wall17 .
(c) Culmann method for determining passive resistance of earth berm22

Wall movement The effect of wall movement on lateral earth pressures should be restated
because of its vital effect on both active and passive pressures. Earth pressures
at rest are reduced to limiting active values by movement of the wall away
from the soil; much larger movements are needed to increase the at-rest
pressure to limiting passive values. The relationship between wall rotation
and the coefficient of earth pressure for sands is reproduced in Fig. 5.15.
The curve can be used to assess the movement needed to obtain limit pressures
in sands of varying compaction state. The NAVFAC guide22 also suggests

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 161

Fig. 5.12. Effect on surcharge loading of point and line loads17

that the effects of wall translation on limit pressures can be assessed by


assuming the effect of translation is equivalent to the movement at the top
of the wall due to rotation, as shown in Fig. 5.15.
The extent of wall movement relative to the retained soil therefore deter-
mines the proportion of the limit pressure which applies. This movement
may be derived from elastic deflection of the wall or sheeting itself, the
compression which occurs in strutting or raking shores (or the curtailment
of such movement by pre-loaded supports or anchors), the movement
needed to mobilize passive resistance of a soil at an intermediate stage of
excavation in front of the wall, or the rotation or translation of a rigid wall.
With the exception of the methods specified in BS 800219 the above means
of calculating earth pressures have led to designs based on limit pressures
where these do not necessarily apply if the extent of proposed wall movements
do not justify them. Soil pressures based on the stiffness of the wall, its support
and the soil itself are evaluated more realistically by soil/structure analyses
based either on the Winkler spring theory or by finite element methods.
These methods are referred to again later in this chapter.

Design calculation Eurocode defines design procedures under four headings:


according to Eurocode 7 . by calculation
. by prescriptive measures
. from results of load tests or tests on experimental models
. by observational techniques.
New meanings are given to words within Eurocode 7 as follows.
. actions – foundation loadings
. permanent actions – dead loads, typically weights of structures and
installations
. variable actions – imposed loads, typically wind or snow loads
. accidental actions – vehicle impacts or explosions
. transient actions – a classification related to soil response, for example,
wind loading
. persistent actions – a classification related to soil response, for example,
structural or construction loads.
Eurocode 7 makes use of partial safety factors to allow for uncertainties in the
magnitude and combination of loads (actions) and characteristic values of soil
and rock parameters obtained from field and laboratory testing.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
162 Deep excavations

Fig. 5.13. Horizontal pressures on a rigid wall from line loads (left) and point loads (right)22

The design values of vertically applied actions Vd to be used in geotechnical


design are calculated from the expression Vd ¼ Fk  F , where Fk is
the characteristic load and F is the partial factor of safety for the applied
action.
Vd must be equal to or less than the foundation design bearing resistance
Rd . The value of Vd is calculated from characteristic soil strengths also with
a partial factor applied. The design value of the material properties Xd is

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 163

Fig. 5.14. Lateral pressure on


an unyielding wall due to
uniform surface load22

calculated from the expression Xd ¼ Xk =m , where Xk is the characteristic


value and m is the particular factor of safety for the material properties.
The specified values of these partial factors ðm Þ together with the partial
factors F to be applied to the actions are shown in Table 5.5.
The three cases shown in this table refer to the overall stability of the struc-
ture, typically buoyancy for an underground structure (case A), the strength
of the structural members is the substructure (case B) and the geotechnical
design of foundations and earthworks (case C). The partial factors for the
ground properties for case C are higher than in cases A and B, allowing for
variations in the ground and uncertainties in the analytical work.
Eurocode 7 (EC7) gives some guidance in the choice of characteristic soil
properties, defining them as those that give the probability of a more unfavour-
able value governing the limit state occurrence of not more than 5%, and
defines a characteristic value of soil properties as a cautious estimate of mean
values within the ground influenced by the foundation loads.

Selection of design parameters


The choice of design parameters in terms of strength and stiffness is particu-
larly important because of the considerable effect on wall design at both
ultimate and serviceability states. Three grades of assessment should be
considered:
(a) moderately conservative soil parameters (case A)

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
164 Deep excavations

Fig. 5.15. Effect of wall rotation on active and passive pressure in loose and dense sand22

(b) worst credible soil parameters (case B)


(c) most probable soil parameters (case C)
The first category, moderately conservative values, is a relatively cautious
estimate at the particular limit state. It is a similar value to the representative
value as used in BS 8002 and to characteristic values as defined in EC7.

Table 5.5 Partial factors –


ultimate limit states in Case F for actions M for ground properties
persistent and transient
situations1 Permanent Variable tan  c0 cu qu a

Unfavourable Favourable Unfavourable

Case A 1.00 0.95 1.50 1. 1 1.3 1. 2 1. 2


Case B 1.35 1.00 1.50 1. 0 1.0 1. 0 1. 0
Case C 1.00 1.00 1.30 1.25 1.6 1. 4 1. 4
a
Compressive strength of soil or rock.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 165

Fig. 5.16. Assessment of moderately conservative strength vs. depth profile for site on London clay. (a) Undrained shear
strengths at site; (b) SPT test results from same site; (c) SPT results from (b) and results from adjacent sites compared with the
undrained shear strengths; (d) final assessment of all results24

Simpson and Driscoll24 (Fig. 5.16) used a comparison between undrained


cohesion values of London clay, SPT values on the same site and plots of
undrained strength vs. depth from nearby sites to assess a moderately conserva-
tive strength vs. depth profile. The construction is probably best not made on a
statistical basis but following a comparison of data from other reliable sources.
The second category, the worst credible value is the worst that can be
reasonably assessed could occur, an unlikely circumstance. Simpson and
Driscoll24 suggested that these values can be regarded as 0.1% fractile.
In the third category, the most probable values have a 50% chance of
exceedance. It should not be used as the only basis of design but can be
used with the observational approach providing the second design approach
using worst credible values are used to define contingency measures should
trigger values be exceeded.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
166 Deep excavations

Certain practical matters arise in preparing a testing schedule which is


appropriate to the design on a specific site. These are as follows.
(a) An adequate number of classification, moisture content and density
tests should be made.
(b) Test correlations should be made with other tests, other sites and where
possible by back analysis in similar conditions.
(c) In stiff clays, care should be exercised in the choice of strength tests,
peak, critical state, residual and drained or undrained.
(d ) Laboratory tests for unconsolidated undrained triaxial values should be
made on 100 mm dia. undisturbed samples as single-stage tests.
(e) Laboratory tests for drained shear strength should be made on con-
solidated drained and consolidated undrained triaxial tests and shear
box tests (for cohesionless soils) and peak, critical state and residual
angles of shearing resistance should be measured. Appropriate rates
of shear should ensure complete dissipation of pore pressure.
( f ) Inspection of the soil fabric should be made visually especially for soft
estuarine clays when continuous sampling may prove necessary.
(g) Soil stiffness may be measured with the best expectancy of accuracy in
self-boring or Menard-type pressuremeter tests in stiff clays and weak
to moderately strong rocks.
(h) Measurement of groundwater levels should be applied stringently,
using standpipes and piezometers to critically assess piezometric
levels, their variation and risks associated with changes in groundwater
level.

Temporary works Important differences exist between the design approach to soil-retaining
structures for permanent and temporary works. The dissimilarities between
temporary works and permanent works design may be summarized as follows.
In temporary works . . .
(a) In clays, negative pore-water pressures are present in the soil. These
dissipate with time but while they act they cause the soil to have a
greater shear strength than it has in the long term.
(b) Support conditions are often different in the temporary works before
base slabs, floor slabs or other permanent supports are constructed.
Temporary works often rely more on the soil for their support than
do permanent works.
(c) Loading conditions are often different. Applied surcharges during
temporary works construction are often not present in the long term.
(d) Ground movements are often a matter of concern during construction,
but there is not usually time for long-term movements to develop.
(e) A greater proportion of soil strength may be mobilized during tempor-
ary works depending on constraints operating during construction.
However, the consequences of excessive ground movements or instabil-
ity can be as severe as for permanent works.
( f ) Higher stresses may be permitted in the design of the wall in temporary
works.
Temporary works may be designed either: by assuming full equilibrium (full
drained) pore-water pressures (i.e. long-term strength), the expected loads
in the temporary phase and a very low safety factor; or by taking the higher
short-term strength and a higher safety factor. (The choice is essentially one
of modifying the parameters used and the safety factor adopted in passing
from temporary to permanent works phases.)

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 167

Mixed total and Since the original ground surface and the soil below it on the active side of the
effective stress design wall are usually more susceptible to softening than the soil below final forma-
tion level, a method of calculating earth pressures using effective stress values
on the active side and undrained values on the passive side has gained some
credibility. The benefit, in terms of calculated earth pressures, is obtained
principally from passive pressures based on the limit pressure expression
pp ¼ v þ 2c, that part of the expression relying on cohesion applying at
formation level on the passive side. The risk of clay softening on the passive
side below formation level can be estimated by reducing the undrained
strength/depth profile by 20% to 30% (using 0.7 to 0.8 of the unsoftened
design line for cu ) and ignoring the clay strength in the upper 1 m thickness
of clay at the surface of the excavation. Overall, the loss of safety for the
wall as built will depend on several design matters, which include the choice
of parameters, the choice of total or effective design method and the factors
of safety that are applied to each component of the wall, the bracing or
anchor that supports it, and the extent of embedment of the wall in the
soil below formation level. If more optimistic methods are chosen for wall
analyses (by using total or mixed total and effective stress methods) these
should reflect more stringent factors of safety in the derivation of wall
member dimensions. Conversely, if effective stress methods are used through-
out, say for temporary works design of limited design life, the final factors of
safety in the dimensioning of wall and strutting may be chosen more optimis-
tically. Recommended factors of safety will be reviewed later in this chapter.

Design water pressures The design of walls to support soil will include consideration of the support of
water pressures where there is groundwater above the final excavation level in
front of the wall. Where support walls are designed for excavations to coffer-
dams and shafts below river or sea bed, the pressures from river or sea water
levels are evident to the designer and would comprise the major part of
pressures on the wall. Where earth support walls are required for excavations
on land, the designer has to evaluate the highest ground water levels during
the design life of the structure and incorporate these with earth pressures
into design pressures for the wall. In excavations on land where groundwater
is high, the water pressure will constitute a large proportion of the horizontal
force on the active side of the wall, but is a lesser proportion than the earth
pressure on the passive side. If water levels rise, the total force on the active
side of the wall increases, while if water pressure rise below formation level
on the passive side, in most cases total passive resistance from water and
soil reduces.
Where vertical soil support walls on land penetrate granular soils with a
high water table but obtain a cut-off at depth, the water pressure on each
side of the wall is calculated without complication (Fig. 5.17). The hydrostatic
pressures on each side of the wall are calculated from v ¼ w x where w is
the density of water (in fresh water it is 9.81 kN/m3 , in salt water it is
10.00 kN/m3 ), and x is the depth below the groundwater table.
Additional pressures due to heavy rain may need to be considered in defin-
ing groundwater levels, and where surfaces are unpaved, full water pressures
should be taken in tension cracks in cohesive soils. In cohesionless soils or
where drainage is introduced at the rear of the wall, an increase in soil density
to its fully saturated value should be allowed for. If the retained cohesionless
soil has low permeability but drainage holes through the wall are adequate to
discharge the water flow through the retained soil, it is recommended that the
effect of the flow of water towards the wall and the increased total active thrust
on the wall should be taken into account by calculating total active thrust on a

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
168 Deep excavations

Fig. 5.17. Hydrostatic water


pressures on walls in
permeable ground with
effective cut-off: (a) gross water
pressures; (b) net water
pressures17

vertical wall:
H2 : H2
Pa ¼ Ka b þ 0 5w ð56Þ
2 2
where b is the submerged density of the soil, and H is the vertical height of the
earth retained.
It is also prudent to add that where drainage through the wall is not
adequate to avoid an increase in head of groundwater in the retained soil
during storm conditions, this increase should be allowed for and, if in extreme
conditions, the head can reach the top of the wall, this should also be allowed
for.
Other conditions that may affect static groundwater levels considered in
designs include the effect of waves on retaining walls for marine structures
and river and sea cofferdam walls, and the dynamic forces of waves breaking
on beaches which can be transmitted to retaining walls for deep excavations
below beach level. Some guidance on these matters may be obtained from
the work of Sainflou25 and Minikin26 .
Where walls are situated in tidal conditions the pore pressure will vary with
the tide; the effect of rapid drawdown, of cycles of loading and any tidal lag
may require special consideration. Guidance is available in reference27 .
Where the vertical wall for soil retention penetrates cohesionless soil but
does not secure a cut-off in an impermeable stratum at the toe of the wall,

Fig. 5.18. Flow net: flow into


base of a twin-wall cofferdam in
sand17

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 169

the effect of water seepage from the retained soil on the active side of the wall
to the passive side should be taken into account. Over time a steady seepage
state will develop. In design, the effect on pressure in these conditions may be
conveniently assessed by use of flow nets, the construction of which was
explained in Chapter 2. A flow net illustrating these conditions is shown in
Fig. 5.18. The total head at any point is
d
H ¼ H1  ðH1  H2 Þ ð57Þ
nd
where H1 and H2 are the heads of groundwater on either side of the wall, d is
the number of flow net equipotential drops to the point considered, and nd is
the total number of drops. Also, at any point the total head is
u
H¼ þz ð58Þ
w
where u is the pore-water pressure, and z is the height of point considered
above datum. Since H can be calculated by flow net and w and z are
known, u can be calculated. The flow net is also used to check the factor of
safety against piping (see Chapter 2).
A simple short-cut for calculating pore-water pressure at any point can be
made by assuming that the head difference between each side of the wall,
active and passive, is dissipated uniformly along the flow path; referring
to Fig. 5.19, head difference ¼ ðh þ i þ jÞ and flow path length is
ð2d þ h  i  jÞ.
Therefore, the head at any point x below the water table on the active side is
x
Hx ¼ ðh þ i  jÞ ð59Þ
2d þ h  i  j
and since u ¼ Hw  z, pore-water pressure at the toe of the wall is
2ðd þ h  jÞðd  iÞ
utoe ¼ w : ð60Þ
2d þ h  i  j
An even simpler first approximation is that in which the water pressure
at the toe of the wall is taken as the average at that depth for active and
passive sides of the wall. All uses of steady-state flow are best checked by
flow net; the reduction of excess head linearly along the length of flow path
at the wall can lead to over-optimistic calculated pore pressures in the
narrow flow paths that occur between the retaining walls of a narrow coffer-
dam. In wider cofferdams (where width is greater than four times the
differential water head) and single retaining walls where the seepage pressure
is free to dissipate horizontally through the passive soils, the first approxima-
tion is satisfactory.
The reduction in pore pressure from one side of the wall to the other in the
steady-state condition can be assessed by such methods of calculation. More
importantly, however, is the effect of preferential drainage paths which can be
caused by sand or silt partings in laminated groundwater flow below a vertical
wall. Where the pore-water pressure does not reduce gradually because of a
horizontal drainage route below the toe of the wall, an excessive pore-water
pressure occurs on the passive side, resulting in high net pressures and
forces on the wall. Figure 5.20 illustrates this risk.
Where groundwater flow is impeded by the periphery wall to a very large
excavation the groundwater level near the wall may rise with time. Where
the risk of this increase in pressure exists piezometers should be installed
and monitored close to the wall.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
170 Deep excavations

Fig. 5.19. Water pressures – simplifying assumptions for steady-state conditions: (a) uniform ground; (b) layered ground48

Fig. 5.20. Excessive pore-water


pressure on passive side due
to underflow: (a) soil
cross-section showing pervious
interlayers; (b) water pressure
diagram48

Design methodology Wall design must satisfy the ultimate limit state and the serviceability limit
state with the specified partial safety factors applied.
The partial factor on soil strength at each limit state for both total and
effective stress designs for each of the design categories, moderately conser-
vative, worst credible and most probable, is shown in Table 5.6. The design

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 171

Table 5.6. Overall stability, cantilever and single-prop walls. Partial factors for use in design calculations49

Design approach Ultimate limit state Serviceability limit state

Effective stress Total stress Effective stress Total stress

Fsc0 Fs0 Fssu Fsc0 Fs0 Fssu

A. Moderately Conservative 1. 2 1. 2 1. 5 1. 0 1. 0 1. 0
B. Worst Credible 1. 0 1. 0 1. 0 Not Not Not
appropriate appropriate appropriate
C. Most Probable 1. 2 1. 2 1. 5 1. 0 1. 0 1. 0
 
tan 0
design 0d ¼ tan 1
Fs0
c0
design c0d ¼
Fsc0
s
design sud ¼ u
Fssu

parameters derived from the application of these safety factors are then used
to obtain earth pressure coefficients, from charts such as those by Caquot and
Kerisel15 .
Either the design approach using moderately conservative parameters (with
Fs ¼ 1:2 applied to effective stress values c0 and 0 or Fs ¼ 1:5 applied to
total stress su ) or the design approach based on worst credible parameters
(with Fs ¼ 1:0) is used. An early check of the two sets of calculated design
parameters shows which design approach gives the lowest design values.
These soil strengths are used with the worst combination of applied loading
in both the short term and the long term.
Groundwater conditions which include water pressure and seepage pressure
with the most unfavourable values that could occur in the extreme or acciden-
tal situation are applied at ultimate limit stage. The accidental situation
includes an allowance of 10% of the retained height or 0.5 m, whichever is
greater, for unplanned excavation. The most unfavourable values applying
in normal conditions are then used to check the serviceability state.
Choice is made then of the analytical method, either based on limiting
values of earth pressure or on soil–structure interaction analysis. For simple
structures, such as cantilever walls, the limit equilibrium can be applied
(with reasonable agreement of the designed wall with more complex soil–
structure analysis) although for propped and anchored walls the soil structure
analysis will usually show economies in terms of wall penetration, bending
moment and prop/anchor loads. These economies may be significant.
For serviceability checks soil structure interaction analysis is necessary to
check wall deformation, soil settlements and compliance with crack criteria
for R.C. walls.
As a general basis of permanent wall design the following steps are made to
establish values of bending moments and shear force, prop/anchor loads and
wall penetration.
(a) Establish ultimate limit states (overall stability and wall strength) and
serviceability limit states (limiting values of wall deflexion, settlements,
crack width criteria).
(b) Collect soil and groundwater data.
(c) Determine wall construction method and wall type.
(d) Determine soil profile and soil parameters.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
172 Deep excavations

(e) Determine load case combinations and wall geometry for ultimate limit
state (ULS) calculation.
( f ) Determine the wall depth for vertical stability. (Check on vertical load
carrying capacity or hydraulic cut-off ). Check wall friction/adhesion.
(g) Make ULS calculation by limit equilibrium or soil–structure inter-
action. Check wall penetration for overall stability.
(h) Use the deeper of wall depths for vertical stability ( f ) or for lateral
stability (g) as design depth.
(i) If wall depth is determined by vertical stability ( f ) recalculate bending
moment (BM), and shear force (SF) in wall and prop loads by limit
equilibrium or soil-structure interaction method.
( j) If serviceability limit state (SLS) calculations are required, select
appropriate soil parameters, groundwater pressures, load case
combinations and design geometry.
(k) Carry out SLS calculations to determine wall deflection, BM, SF and
prop loads.
(l ) Compare calculated wall deflections with initial wall design require-
ments.
(m) Make assessment of settlement risk to adjacent structures and services.
(n) Check risk of progressive failure.
(o) Determine ULS wall moments and shear force for structural design as
the greater of:
(i) BM and SF from step (g), ULS
(ii) 1.4 times the BM and SF from step (k), SLS
(iii) BM and SF from step (n), progressive collapse check
( p) Calculate prop loads, see later this chapter.

Structural design of wall


Ultimate limit state values of bending moments and shear forces for design of
the wall section are taken as the greatest of those calculated from application
of factored soil strengths (the worst values for case A, moderately conserva-
tive or case B, worst credible) or 1.4 times the SLS calculations.

Crack control
The application of both BS and Eurocodes at serviceability state to limiting
crack widths in retaining walls due to early age shrinkage and flexure can
be most onerous. Although the Eurocodes contain a ‘deemed to satisfy’ pro-
cedure in the absence of detailed calculations this is based on providing
sufficient reinforcement to have a greater tensile capacity than the immature
concrete whilst limiting the steel stress to a value which limits the crack
width to the required width.
The draft I.Struct.E report on basement and cut-and-cover works51 recom-
mends that different criteria are set for flexural crack widths at various stages
and suggests the maximum flexural crack widths shown in Table 5.7.
Table 5.7. Maximum flexural
crack widths51 Design stage Suggested maximum
flexural crack width (mm)

During construction and recoverable 0.35


In service 0.30
Short-term event 0.35
Drawdown of groundwater to underside of base slab 0.40
Note. Cracking caused by flexure of the entire structure, such as a tunnel built over a patch of
soft (or hard) ground or subjected to localized surcharge, passes right through the section and
should be limited in width to 0.2 mm at all times.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 173

Fig. 5.21. Pile cross-section


showing crack width calculation
principles to EC2 if tension area
in cover thickness is ignored49

Fig. 5.22. Pile cross-section


showing calculation principles
for crack width in accordance
with BS 811049

The quantity of reinforcement required for crack control may be consider-


able and strict application of code requirements may be needed to keep steel
density to practical levels.
Referring to steel required to resist a defined crack width in bending as
calculated according to EC2 (applied to a circular section cast against the
ground) the critical parameter is the area of concrete in tension A (EC2
Part 1, 1992, Clause 4.4.2.2. (3)). Figure 5.21 shows the net area in tension
which may be used for calculation purposes if the thickness of concrete
cover is assumed to be the minimum required for durability.
Referring to crack control requirements in BS 8110: Part I, 1997, the specific
items for a typical circular pile section are:
Nominal cover to earth face: 75 mm (Part I Clause 3.3.1.4).
Durability requirements: 35 mm for C35 concrete (Part 1 Table 3.3)
(concrete in moderate exposure Table 3.2, non-aggressive soil).
Crack width (Part 2, section 3.8) should be assessed in accordance with acr , the
distance from the point considered to the surface of the nearest longitudinal
bar (Part 1, Clause 3.8.2). Often assessed as the nominal cover required for
durability and not the 75 mm cover provided for constructability. Refer to
the cross-section shown in Fig. 5.22.

Cantilever walls and The design of cantilever and single-prop walls requires that the serviceability
single-prop walls limit state is considered in terms of wall deformation and soil deformation at
the rear of the wall. The ultimate limit state is as follows.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
174 Deep excavations

(a) Overall instability – the provision of sufficient embedment depth to


prevent overturning of the wall. For a single-propped wall the minimum
penetration is just sufficient to avoid forward movement of the wall with-
out preventing rotation of the wall at its toe. Such a condition is known
as the free earth support condition. As wall penetration is increased
beyond this minimum, fixity of the wall is progressively increased until
full fixity is obtained and the toe is unable to rotate when earth and
water loading is applied to the wall. This state is known as the fixed
earth condition. For the cantilever wall without prop support, full
fixity must be obtained to prevent rotation and collapse of the wall.
(b) Foundation failure – in cohesive soils, the wall penetration depth must
be sufficient to prevent basal failure after excavation to formation
level in front of the wall.
(c) Hydraulic failure – in cohesionless soils, the penetration of the wall must
be sufficient to avoid piping in front of the wall after excavation to
formation level.

Overall stability
The overall stability of both cantilever and single-prop walls has previously been
assessed using limit equilibrium methods in which a factor of safety is applied
to an analysis of failure conditions to ensure their avoidance. Such factors of
safety were applied in a number of ways and have attracted comment from
Potts and Burland28 , and Symons29 . The various factor methods are as follows.
(a) Factor on embedment. A factor of safety is applied to embedment depth.
The method is described in references 30 and 31.
(b) Factor on moments of gross pressure. This method applies as a factor of
safety to moments of gross pressure on the passive side only. Water
pressure is not factored. The method is described in references 16 and 22.
(c) Factors on moments of net pressure. The net pressure diagram is calcu-
lated and the factor of safety is defined as the ratio of moments of the
net passive and active forces about the toe of the wall.
(d ) Factors on net passive resistance. Potts and Burland28 developed this
method which defines the factor of safety as the ratio of the moment
of the net available passive resistance to the moment activated by the
retained material including water and surcharge. When using effective
stress methods with c0 ¼ 0 and where there is no surcharge on the
passive side, the active pressure diagram is modified so that no active
pressure increase occurs below formation level and any such increase
as calculated is deducted from the passive pressure diagram. Where
total stress or effective stress methods are used, reference should be
made to Potts and Burland28 .
(e) Factor on shear strength on both active and passive sides. Soil shear
strengths are reduced by dividing c0 and tan 0 by the factor of safety,
and the active and passive pressure diagrams are calculated using
these reduced values. The reduced values approximate to mobilized
values. Bending moments and prop loads derived from the calculation
can be used for wall design if they are treated as ultimate limit state
values. This method is recommended in BS 800219 .
( f ) Factor on shear strength of passive side only. The passive resistance is
factored but no factor is applied to the active side.
Comparative studies show that there is no relationship between factors of
safety applied to the methods listed. There is little doubt that the preferred
method is really defined by personal use and experience, but some useful
observations were made in reference 32 as follows.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 175

(a) The factor on embedment is empirical and should be checked by


applying a second method.
(b) The method of factoring moments on gross pressure may give excessive
penetration at low angles of shearing resistance 0 less than 208 so use
varying factors for different ranges of 0 .
(c) The factoring of net passive pressure moments tends to give high
penetration values.
(d) The factors on modified net passive resistance as recommended by
Burland and Potts28 appear to give consistent results in a reasonable
range of soils and wall dimensions.
Factors of safety recommended by Padfield and Mair48 for use in stiff clays
with methods for factoring embedment, moments of gross pressure, net
passive resistance and shear strength on both active and passive sides are
reproduced in Table 5.8. Two approaches are used: approach A is based on
moderately conservative parameters, and approach B uses worst credible
soil parameters, geometry and loading in the design. The values in Table
5.8 had been originally intended for walls embedded in stiff clays but were
also used for walls in granular or mixed soils.

Table 5.8 Factors of safety for methods of analysing embedment48

Method Design approach A: recommended Design approach B: recommended Comments


range for moderately conservative minimum values for worst credible
parameters (c0 , 0 , or cu ) parameters (c0 ¼ 0, 0 )

Temporary Permanent Temporary Permanent


works works works works

Factor on Effective stress 1.1 to 1.2 1.2 to 1.6 Not 1.2 This method is empirical.
embedment, Fd (usually 1.2) (usually 1.5) recommended It should always be
checked against one of
the other methods
Total stressa 2.0 – –
Strength factor Effective stress 1.1 to 1.2 1.2 to 1.5 1.0 1.2 The mobilized angle of
method, Fs (usually 1.2 (usually 1.5 wall friction m and wall
except for except for adhesion cwm should also
0 > 308 when 0 > 308 when be reduced
lower value lower value
may be used) may be used)
Total stressa 1.5 – – –
Factor on moments: Effective stress 1.2 to 1.5 1.5 to 2.0 1.0 1.2 to 1.5 These recommended Fp
CP2 method, Fp 0  308 1.5 2.0 1.0 1.5 values vary with 0 to be
0 ¼ 20 to 308 1.2 to 1.5 1.5 to 2.0 1.0 1.2 to 1.5 generally consistent with
  208 1.2 1.5 1.0 1.2 usual values of Fs and Fr
Total stressa 2.0 – – –
Factor on moments: Effective stress 1.3 to 1.5 1.5 to 2.0 1.0 1.5 Not yet tested for
Burland–Potts (usually 1.5) (usually 2.0) cantilevers. A relatively
method, Fr new method with which
little design experience
has been obtained
Total stressa 2.0 – – –
a
Speculative, treat with caution.
Note 1. In any situation where significant uncertainty exists, whether design approach A or B is adopted, a sensitivity study is always recommended, so that an
appreciation of the importance of various parameters can be gained.
Note 2. Only a few of the factors of safety recommended are based on extensive practice experience, and even this experience is recent. At present, there is no
well-documented evidence of the long-term performance of walls constructed in stiff clays, particularly in relation to serviceability and movements. However,
the factors recommended are based on the present framework of current knowledge and good practice.
Note 3. Of the four factors of safety recommended, only Fp depends on the value of 0 .

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
176 Deep excavations

Fig. 5.23. Analysis of cantilevered wall17

Calculation can be made by soil–structure interaction analysis or using limit


equilibrium methods by program or by hand.

Fixed earth support: cantilevered wall


Certain simplifying assumptions are made due to the relative complexity of
the limit equilibrium calculation. Figure 5.23 shows an idealized pressure
diagram and typical shear, bending moment and wall deflection diagrams.
The simplified procedure is as follows:
. the pressures at the toe of the wall are replaced by a resultant force F3
acting at C
. forces F1 and F2 act through the centres of pressure of their respective
areas
. depth BC is found by assuming a level for C and calculating the
moments for the forces F1 and F2 about level C; repeat until moments
balance
. increase the depth BC by 20% to give the design penetration BD
. calculate the maximum bending moment (BM) at the point of zero shear.
This method can be used for layered soils; net water pressures should be
included in the active pressure diagram where groundwater occurs. The
factor of safety can be applied by using limit pressures and increasing the
depth of embedment. The bending moment distribution at limiting equili-
brium is assumed to correspond to working conditions. A safety factor of
value 2.0 is then applied to these values for design values of BM and shear
force (SF). A partial factor of safety is applied to characteristic soil strengths
for design of the wall cross-section. Alternatively, for design according
to BS 800219 a mobilization factor is applied to peak shear strengths to
obtain design values of BM and SF. No further safety factor is applied
other than the partial factor to material characteristic strength to obtain the
design strength.
Bending moment and shear force in the cantilever wall may also be analysed
using Winkler spring, finite element or boundary element methods. These
methods produce a more realistic soil interaction solution and provide
values of lateral wall deformation.
The importance of wall flexibility related to wall deformation and depth of
embedment was originally shown by Rowe20 . The earth distribution was
shown by model tests to vary according to the flexibility of the wall. (The
wall flexibility was defined by the expression
d 4y
wall flexibility ¼ EI
dx4

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 177

where E is Young’s modulus of the wall material, I is the section modulus of


the wall transverse cross-section per unit length, and y and x are horizontal
and vertical coordinates of the vertical cross-section of the wall). The reduc-
tion in depth to full fixity caused by increasing wall flexibility produced
reduced bending moments in the wall section and, according to Rowe,
produced savings of up to 20% by weight of the wall compared with design
methods based simply on limit pressures. The method was not widely adopted
for cantilever pile design but served as an early indicator of the importance of
wall and soil interaction.

Free earth support: single-propped walls


The earth pressure, shear and bending moment distribution and wall deflec-
tion diagrams which apply to a typical single-propped (or anchored) wall
are shown in Fig. 5.24. The method of calculation uses limit pressures or
pressures calculated on critical state values of soil parameters or on peak
shear strengths reduced by a mobilization factor as defined in BS 8002, and
is as follows.
(a) Draw earth pressure diagrams using limiting or mobilized pressures for
soil pressure in cohesionless or cohesive homogeneous or layered soils.
Include net water pressure, modified to include the effect of steady flow
beneath the wall and equalization of pore pressures either side of the
wall at the toe. Use tension cracks filled with water where applicable
in clays and apply minimum fluid pressures (5 kN/m3 ) when necessary.
(b) Calculate F1 and F2 to act through the centre of their respective pressure
areas.
(c) Calculate the penetration BD by assuming a trial depth of penetration
D and taking moments of forces F1 and F2 about the tie rod level.
Repeat with trial values of penetration D until these moments balance.

Fig. 5.24. Single-propped or


anchored wall with: (a) free
earth support; (b) fixed earth
support17

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
178 Deep excavations

Table 5.9 Zero moment depths


to ensure fixity Angle of shearing resistance  Depth to zero moment as a proportion of depth
from ground surface to formation level H

208 0.25H
258 0.15H
308 0.08H
358 0.035H
408 0.007H

(d) The prop load T can then be found by balancing forces T ¼ F1  F2 .


(e) Calculate the bending moment in the wall section at the level of zero
shear.

Fixed earth support: single-propped walls


The analysis by hand calculation of the fixed earth, single-propped (or
anchored) wall is made less complicated by use of work due to Blum33 in
the 1930s. He approximated the bending moment in a single-anchored wall
with fixed earth support by reducing it to zero at the point of zero net earth
pressure below formation level. Blum gave approximate depths to this point
for relatively uniform soils where the wall is sufficiently deep to ensure
fixity; these depths are given in Table 5.9. This support may be solved
graphically16 or by calculation (Fig. 5.24) as follows.
(a) Draw earth pressure diagrams using limiting or mobilized pressures for
soil pressure in cohesionless, cohesive or layered soils. Include net water
pressure modified to include the effect of steady flow beneath the wall
and equalization of pore pressures on either side of the wall at the
toe. Use tension cracks filled with water where applicable in clays.
Apply minimum fluid pressures (5 kN/m3 ) where necessary.
(b) Draw the net pressure diagram from ground level to the first point of
zero pressure below formation level.
(c) Calculate active forces F1 and F2 from the gross pressure diagram.
(d) Take moments of forces F1 and F2 and tie rod force T about the first
point of zero pressure from the net pressure diagram and equate to zero.
(e) Calculate value of force T from this expression.
( f ) Assume application of passive force F3 at point C and recalculate F1
and F2 to this depth. Take moments of T, F1 and F2 about this level
and repeat with change in depth to point C until the moments balance.
As with the cantilever wall calculation, the calculated depth to point C
is increased by 20% to give design penetration depth.
(g) Draw the shear force diagram and calculate the maximum bending
moment in the wall at the point of zero shear force.
This design procedure, as for free earth support design, is usually made by
computer program and where soil deformation prediction is required can be
the subject of soil/structure interaction programs or finite element or finite
difference computations.

Anchored bulkheads: design development


The description of a propped wall includes within this generic phrase various
constructions such as sheet piling, diaphragm walls and reinforced concrete
pile walls braced by struts or raking shores, or sheeters and walls of similar
construction with tie rods or stressed ground anchors. One of these types of
constructions, sheet piling to form a wall or bulkhead tied back by steel
rods to an anchor block, known as a deadman, or to anchor piles, was the sub-

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 179

ject of a historical keynote paper by Terzaghi34 . This method of construction,


which is widespread on waterfronts where it is used to achieve a quay line with
reclamation behind it at minimum cost, does not strictly lie within the subject
of deep excavations, but no technical discussion on braced sheeting would be
complete without reference to the paper and the controversial discussion that
followed it. The principal purpose of the paper was to identify and rectify
errors in accepted bulkhead design at that time on the basis of tests and
observations by Terzaghi and model tests by Rowe20;35;36 . The paper reached
the following principal conclusions.
(a) The identification of the type of soils and fills and their in situ properties
of uniformity, relative density and strength are vital matters which are
frequently overlooked in anchored bulkhead design.
(b) The distribution of earth pressure on the bulkhead is unlikely to con-
form to the Coulomb distribution, because of the extent of deformation
of the soil structure. This deformation depends on soil and wall stiffness.
(c) If maximum bending moments are calculated on walls in sand assumed
to extend to sufficient depth to achieve full fixity irrespective of wall
flexibility and sand relative density, errors are likely and these are on
the unsafe side.
(d) For sheet piles driven into clay, the assumption of full fixity at depth
will probably not apply as time elapses and no reduction in the
calculated maximum moment in the wall should be allowed due to
wall section flexibility to compensate for this loss.
(e) Anchor tension depends on several factors other than the properties of
the backfill material and the flexibility of the wall or sheeting. There-
fore, the anchor pull should be computed on the assumption of free
earth support. Anchor pull may be greater than that calculated using
Coulomb’s theory, and may increase due to repetition of loading
and unloading from heavy surcharge. An unequal yield of adjacent
anchorage produces variations in tie rod pull. Given these risks, more
conservative stresses should be used in anchor design than are applied
in sheet pile bulkhead design.
These conclusions broadly still apply, although alternative methods of
analysis have been developed in which soil, wall and anchor stiffness can be
modelled and deformation and induced stress in all three can be calculated.
Terzaghi concluded: ‘Because of the great variety of subsoil conditions
which may be encountered, the subject (anchored bulkheads) does, and
always will, leave a wide margin for judgment – and also for misjudgment.’

Anchorage location
Anchorages, deadmen or injected tendons must be located behind potential
failure surfaces at the rear of the wall. Figure 5.25 shows the recommended
geometry for analysis of deadmen locations (from BS 634927 ).

Foundation failure
The risk of base failure to an excavation by upward heave applies particularly
in very soft and soft clays and silty clays, typically, quick estuarine deposits.
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 (the relief of overburden) and not by the application
of load as occurs in a conventional foundation bearing failure.
The methods of Terzaghi37 and Bjerrum and Eide38 can be applied to cal-
culate the factor of safety against base failure; these are shown in Fig. 5.26.
Terzaghi’s method is primarily applicable to shallow or wide excavations,

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
180 Deep excavations

Fig. 5.25. Location of deadman


anchorage in granular
retained fill27

Fig. 5.26. Calculation of factors


of safety against basal heave in
cohesive soils: (a) deep
excavations with H=B > 1;
(b) for shallow or wide
excavations with H=B < 1 32

while the method of Bjerrum and Eide is suitable for deep and narrow excava-
tions with no nearby underlying stiff clay to inhibit failure. Both methods
neglect the effect of wall penetration below formation level and therefore
results may prove to be conservative, especially where stiffer clays exist with
depth. A third method22 for predicting the basal safety factor where stiff
clays exist at depth, is shown in Fig. 5.27.
The factor of safety against basal failure is generally required to be not less
than 1.5. If uncorrected values of in situ vane tests are used, the actual factor
of safety may be close to 1.0 according to Aas39 . (A vane correction described
by Bjerrum40 is necessary to obtain more reliable values of safety factor.)
With a factor of safety, based on corrected vane results, which is less than
1.5, substantial soil deformation is likely. If such soil movement is not accep-
table, a factor of safety not less than 2.0 is recommended. Increase in move-
ment occurs as the basal factor of safety decreases, and increases rapidly as
a factor of safety of 1.0 is approached. Although basal heave is rare within

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 181

Fig. 5.27. Calculation of factors of safety against basal heave in: (a) cohesionless soil; (b) cuts in clay of considerable
depth; (c) cuts in clay limited by hard stratum22

excavations in cohesionless soils, a basal heave analysis is included in


Fig. 5.27(a) for completeness, together with basal heave analysis in clay as
described in NAVFAC22 in Fig. 5.27(b) and (c). Figure 5.28 shows the
values of bearing capacity factors for use in these analyses. Field and finite
element analysis predictions of the correlation between movement and basal
failure factor of safety are shown in Fig. 5.29.
Cantilever and single-prop walls, particularly on sloping sites in soft clays
and loose granular soils, should always be checked against risk of deep-
seated circular slip failure.

Hydraulic failure
The risk of piping failure to the base of an excavation in cohesionless soils
was described in Chapter 2. Design charts for penetration of cut-off walls
to prevent hydraulic failure in sand and stratified soil are reproduced in
Figs 5.30 and 5.31.

Wall flexibility
Rowe’s work20;35;36 in the 1950s and 1960s was initially instrumental in
showing the importance of wall stiffness in design. Following a series of

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
182 Deep excavations

Fig. 5.28. Plot of bearing


capacity factors against angle
of shearing resistance22

Fig. 5.29. Analytical


relationship between maximum
lateral wall movement and
factor of safety against basal
heave from field data, free end
and fixed end walls, various
sites (note is for finite element
analysis data)

model tests on sands of varying relative density in Manchester, UK, Rowe was
able to show that interaction between soil and wall was different for steel sheet
piles and reinforced concrete sheet piles because of the greater flexibility of the
steel sheet pile. This greater flexibility causes a redistribution of earth pressure
which differs considerably from the Coulomb distribution, as shown in
Fig. 5.32. The flexure of the wall causes reduction in pressure at mid-height
and causes the resultant passive force to rise with an increase in fixity for
the flexible pile. These changes reduce the design bending moment for a
flexible pile, although too often such reductions are not applied in practice
to ensure the pile does not crumple during driving.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 183

Fig. 5.30. Penetration of cut-off


wall to prevent hydraulic failure
in homogeneous sand:
(a) in sands of infinite depth;
(b) in dense sand of limited
depth22

Materials and stresses


For cantilever and single-propped walls the component parts are designed
either from limiting equilibrium hand calculations or computer program
outputs using either limiting equilibrium or soil–structure interaction. The
input soil parameters are those based on moderately conservative parameters
with safety factors at ULS as shown in Table 5.6 (BS 8002 mobilization
factors are similar) or at serviceability limit state with a partial safety factor
of 1.4 (a check that moderately conservative parameters are more severe,
that is lower, than worst credible parameters at ULS being made at the
beginning of the design). Crack widths are calculated for serviceability limit
state in reinforced concrete walls. Characteristic strengths of steel used for
cantilever and single-prop sheet pile walls are given in Table 5.10.

Multi-prop walls The above description of design methods for cantilever and single-propped (or
anchored) walls referred to computations using limit pressures and the appli-
cation of factors of safety. The methods due to BS 800219 have introduced
design using earth pressures at the serviceability limit state. Using these
methods the bending moment in the walling can be estimated relatively
quickly by hand calculation (adopting Blum’s methods33 for cantilever and

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
184 Deep excavations

Fig. 5.31. Penetration of cut-off


wall to prevent hydraulic failure
in stratified soil: (a) coarse sand
underlying fine sand; (b) fine
sand underlying coarse sand;
(c) very fine layer in
homogeneous sand22

fixed earth support walls) or even more conveniently using finite element,
finite difference or Winkler spring analytical methods. Soil deformation
behind the wall may be predicted, if needed by the finite element or finite dif-
ference programs. Design requirements and analysis methods for multi-prop
walls are a different matter, however. The method of construction for these
walls is usually sequential, installing the sheeting or walling and excavation
in stages followed by installation of the prop or anchor at each installation
stage. The sheeting or walls will, in all likelihood, penetrate the ground
below the final excavation level. The extent of wall deformation in this
sequence of operations is restricted, although the passive resistance of soil
below excavation level at each stage is mobilized to support the wall prior
to installation of the bracing or the anchor at that level. Despite the frequent

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 185

Fig. 5.32. Deflection of a sheet


pile and redistribution of active
earth pressure50

Table 5.10 Characteristic


strengths for steel sheet piling Designation EN 10027 Minimum Minimum Minimum elongation
yield strength tensile strength on a gauge length
(N/mm2 ) (N/mm2 ) of L0 ¼ 5:65S0 (A%)

EN 10248 S270 GP 270 410 24


EN 10248 S355 GP 355 480 22

support for the wall, therefore, horizontal deformation of the wall occurs at
each passive soil zone prior to installation of the prop or anchor at that
level. The wall distorts inwards to mobilize this passive resistance, the wall
movement occurring below each stage of excavation. Pore pressure dissipa-
tion may occur in cohesive soils during the period needed for strut or
anchor installation at successive levels.
The extent of wall movement also depends on the stiffness of the prop or
anchor once installed at each level. Where the soil is relatively stiff, say
dense sands or gravels, the extent of forward movement of the sheeting at
each excavation stage to mobilize soil passive pressure will be relatively
small, and active earth pressures on the wall will considerably exceed
Coulomb active values above dredge level; redistribution of earth pressure
will occur between the lowest strut and formation level. In wide excavations
in soils such as soft clays or loose sands where stiffness is low, the successive
deformations below excavation level at each propping formation level are

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
186 Deep excavations

considerable. Load is redistributed between the struts, and the sum of the
maximum strut loads considerably exceeds the Coulomb values.
Where pre-stressed ground anchors are used at each excavation stage the
earth pressure on the wall is determined by the pre-stress levels and sub-
sequent relaxation, and by relative wall and soil stiffnesses. Design methods
that have been used for many years have been based on calculations of
anchored walls using a Coulomb distribution assuming no pre-stress applied.
This non-pre-stress value of anchorage at each excavation stage is sub-
sequently used as the actual value of pre-stress applied to the tendon. This
empirical method successfully restricts soil movement but in turn inhibits
the Coulomb active earth pressure distribution, on which the calculation is
based, from developing, the actual pressures on the retained side of the wall
being higher (and nearer K0 or Kp values) than those calculated. More
recent methods using Winkler spring and finite element programs allow an
assumed anchor pre-stress load to be introduced to the analysis, from
which the actual earth pressures are calculated on the basis of the soil
movement permitted by wall and soil stiffness and the extent of the anchor
prestress.
The ultimate limit states for multi-prop walls are similar to those for
cantilever and single-prop walls:
(a) overall stability – risk of strut failure, bending stress failure in sheeting
or passive failure of soil below stage excavation level or final formation
level
(b) foundation heave – in soft clays, risk of failure by unloading; bearing
capacity failure
(c) hydraulic failure – piping in cohesionless soils with high external
groundwater table.
The serviceability limit states are as follows.
(a) Deformation of sheeting – the acceptable limits of sheeting deformation
will depend on the purpose of the excavation and whether the works are
temporary or permanent, or a combination of both. Where walls or
sheeting are temporary the deformation must not exceed that which
would occupy space required for the permanent works nor cause
difficulties with sheeting removal if this is intended. For permanent
works, deformation of the wall or sheeting must neither impair the
durability of the substructure nor cause visual offence.
(b) Soil movement behind wall or sheeting; vertical settlements – the extent of
settlement behind the support for the excavation must not exceed the
permitted settlement of existing structures, highways or services,
unless the consequences of this can be estimated accurately and on
re-assessment are acceptable.
(c) Cracking in reinforced concrete walls – at the serviceability state, crack-
ing will occur on the tension face of reinforced concrete walls due to
application of load, in particular earth and water pressures and sur-
charge loading, and also on each face of the wall due to early thermal
cracking of the concrete. For building substructures the provisions
of BS 81102 will apply to crack control in walls or, where more rigorous
waterproofing is needed, BS 80073 may be specified. For highway struc-
tures in the UK, design flexural and tension cracks complying with the
BS 54004 are specified. Design crack widths of 0.25 mm, complying with
‘severe’ conditions are usual although the pressure of saline or sea water
may reduce this value to 0.15 mm. Additional longitudinal steel may
therefore prove necessary in diaphragm walls to control crack widths

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 187

caused by loading, although the application of rules to minimize


vertical crack widths due to thermal shrinkage of concrete in panels
of limited individual length may prove over-strenuous. Such rules for
reinforced concrete works are referred to in the UK Department of
Transport’s Standard BD 28/8741 . Eurocodes EC2 and EC7 refer to
similar crack control requirements in reinforced concrete walls.
The available methods of design for multi-propped walls are:
(a) empirical methods; based originally on strut load envelopes proposed
by Peck for three categories of soil: sands, soft to medium clays and
stiff clays. Twine and Roscoe42 more recently proposed new strut
load envelopes for soft to firm clays, stiff to very stiff clays and dry
clay or submerged granular soils. The use of empirical methods is
recommended as a check on computed strut loads
(b) limit equilibrium programs are a simple solution without addressing all
the matters of influence. Programs based on Winkler Spring theory are
perhaps nowadays the most widely used methods
(c) full soil–structure interaction analysis by finite element, boundary
element or finite difference methods: used where prediction of soil
deformation and soil settlement requires calculated estimates
(d ) pseudo finite element programs.

Empirical method based on strut load envelopes


The original empirical method, due to Terzaghi and Peck18 , was applied to
both temporary works (including piled and diaphragm walls permanently
anchored or braced by floor construction, as in top-downwards construction)
and permanent works. The strut load envelopes due to Terzaghi and Peck are
shown in Fig. 5.33. 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 is added to
strut loads. Sheeting or walling was then designed using the Coulomb earth
pressure distribution with hydrostatic water pressure added except where
drainage occurred through sheeting to relieve water pressure.
In 1969, reviewing his empirical method, Peck pointed out that his recom-
mended method for strut design was less satisfactory in soft to medium clays

Fig. 5.33. Apparent pressure


diagrams for computing strut
loads in braced cuts18

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
188 Deep excavations

than for excavation in sands. Peck did point out, however, that the behaviour
of clays and the bracing system depends on the stability number
D
Nc ¼ ð62Þ
cu
where cu represents the clay beside and beneath the cut to the depth that
would be involved if a general failure were to occur due to the excavation.
Peck added that when the depth of excavation corresponds to values of Nc
greater than 6 or 7, extensive plastic zones have developed, at least to the
bottom of the cut, and the assumption of a state of plastic equilibrium is
valid. The movements are essentially plastic and the settlements may be large.
The empirical strut load envelopes reviewed by Peck (Fig. 5.33) do not
include the effects of the toe of the sheeting or walling extending below the
final formation level, and yet from the point of view of reduction of strut
loads in the lower struts (or anchors) and the improvement to ingress of
groundwater at formation level, an extension of sheeting vertically below
formation level was desirable and often achievable. There were two design
methods which allowed this penetration of walling and sheeting to be taken
into account in strut load calculation.
The first method assumed that the passive reaction below formation level
resisted active pressure below that level together with a portion of the load
from the strut load envelope between the lowest strut level and formation
level. This method applied to strutted excavations in uniform soil conditions
of reasonable strength such as medium-dense to dense granular soils and stiff
clays. In less competent soils (loose sands and gravels and soft clays), passive
resistance against the sheeting below formation level might be less effective,
and in such cases Goldberg et al.23 advised that the sheeting should penetrate
to such depth to avoid piping failure (as discussed in Chapter 2) and the
sheeting should be designed as a cantilever from the lowest strut level.
The second empirical method, the source of which is not known, has been
used since the mid 1950s by the Author and has been adequately confirmed.
The procedure is shown in Fig. 5.34. An additional ‘strut’ is assumed to act
on the strut load envelope to represent the passive resistance acting on the
sheeting below formation level. The level of this ‘strut’ is determined from
the net pressure diagram, as shown in Fig. 5.34. The ‘strut’ load calculated
from the strut load envelope is then compared with the available passive resis-
tance using the limiting pressure from the net pressure diagram for the selected
penetration depth. The bending moment in the sheeting is calculated from the
net pressure diagram.
The strut loads calculated for each successive construction stage are
summarized and the highest value at each strut level is used for strut and
waling design purposes. Similarly, maximum moment and shear values are
calculated for the sheeting or walling for each construction stage from the
net pressure diagram and the critical values used for sheeting design.
More recently, empirical data and recommended methods of strut load
estimation have been published by Twine and Roscoe42 . Eighty one case
histories, all from the UK, were reviewed. The walls were supported by
both timber and steel temporary props, not anchors, and the study did not
differentiate between single- and multi-propped walls. Excavations in rock
were not considered. The cases were classified according to the ground
retained as follows:
. Class A: normally- and slightly over-consolidated clay soils (soft and
firm clays)
. Class B: heavily over-consolidated clay soils (stiff and very stiff clays)

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 189

Fig. 5.34. Construction of


trapezoidal strut load envelope
for braced excavation to take
into account passive resistance
below formation level

. Class C: granular, cohesionless soils


. Class D: walls retaining both cohesive and cohesionless soils.
Case histories for soft and firm clays (class A) and for stiff clays (class B) were
further sub-divided into flexible walls (timber, sheet pile and soldier pile walls)
and stiff walls (contiguous, secant pile and diaphragm walls). The character-
istic distributed prop load diagrams proposed by Twine and Roscoe are
shown in Fig. 5.35.
Twine and Roscoe referred to the range of field measurements compared to
values calculated by deformation methods. Field measurements were made on
46 props at four sites. The comparison is shown in Table 5.11. The range of
calculated to required values was generally between 1.4 and 10 times. This
very wide scatter has been referred to by others for measurements in two
deep excavations in London (Batten and Powrie43 and Richards et al.44 ).
Both these site measurements of strut loads showed considerable variation
with temperature.
Twine and Roscoe concluded that, providing the load–strain behaviour of a
prop is ductile, any temperature increase which causes the ultimate capacity of
the prop is unlikely to cause sudden prop failure because of plastic deforma-
tion of the prop and the transfer of load to nearby props. Because of this, they
recommended that the design of steel props should not include temperature
effects provided that certain SLS and ULS criteria are satisfied. This
serviceability criterion ensures that the temperature effect does not cause

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
190 Deep excavations

Fig. 5.35. Characteristic


distributed prop load diagrams
proposed by Twine and
Roscoe42

yielding of the prop steel and thus avoids permanent deformation and the
ULS criterion requires the ultimate capacity of the prop under the tempera-
ture effect to be greater than the total ULS design load.
Potts45 reviewed analysis methods and concluded that all simple methods of
analysis were flawed and empirical methods were only of some value if they
were verified by observation. These conclusions, in 1992, remain currently
valid although the increased use of programs to calculate wall moments,
deflections and strut loads has been rapid and empirical methods appear

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 191

Table 5.11 Comparison of


measured and calculated prop Project No. of Range of Range of ratio:
loads42 monitored measured prop Measured load
props loads (tonnes) Design load
55 Moorgate 6 62.5 to 109 0.16 to 0.28
Almack House, London 15 0 to 118 0.0 to 0.36
A406 Chingford, London
(i) Piled wall 10 45 to 160 0.28 to 0.82
(ii) Wedge excavation 4 45 to 70 0.19 to 0.39
(iii) Diaphragm wall 6 140 to 220 0.7 to 1.37
A12/M11 Hackney, London 5 623 to 1395 0.51 to 0.83

little used. This change of emphasis appears unfortunate in the Author’s view;
the equivalent strut load diagnosis of Peck and Twine and Roscoe should still
be used as a critical comparison compared with calculated results.
A summary of the advantages and drawbacks of the four types of program
is given in Table 5.12, they are based on analytical methods as follows.
(a) Limit equilibrium (e.g. STAWAL and REWARD). A stability analysis
in which calculation is made of the statical equilibrium of arbitrary slip
surfaces by resolving forces or moments and results in the calculated
soil strength mobilized to obtain equilibrium. The internal stress condi-
tions within the soil wedge are not considered.
(b) Subgrade reaction/beam on springs (e.g. WALLAP, LAWALL). Soil
behaviour is modelled by a set of unconnected vertical and horizontal
springs. Any structural support is modelled by simple springs. In
LAWALL the soil is modelled as a simple boundary element using
Mindlin’s equations for horizontal displacement due to a horizontal
force applied within a semi-infinite elastic continuum. Berms are
difficult to model; anchors and raking props are difficult to model
accurately.
(c) Pseudo finite elements (e.g. FREW). Soil is modelled as an elastic
continuum with soil stiffness matrices calculated using a finite element
program. Soil model restricted to linear increasing stiffness with depth.
Only wall deformation is calculated.
(d) Finite element and finite differences (e.g. PLAXIS, CRISP, FLAC). Full
soil–structure interaction with modelled construction sequence. Some
programs are now able to model 3D excavations. Wide choice of soil
models available; Mohr–Coulomb, and more complicated failure
criteria to represent variation of stiffness with strain. Use of low
strain values of soil stiffness are necessary.
Comparative analysis using both empirical methods and computer analyses
shows a comparatively wide scatter of results for strut loads, wall penetration
and wall bending moment. The lack of validation from field observations and
the extensive use of programs nowadays gives cause for concern.

Factors of safety: multi-propped wall design


The attention given to comparative methods of applying factors of safety
to the overall stability of cantilever and single-prop (or anchored) walls has
not extended to multi-prop walls. The design should always consider
hydraulic failure by piping from the base of the excavation and the risk of
failure of the base of the excavation in soft clays. Overall, stability considera-
tion in a multi-propped wall will rely on three matters: the risk of circular slip
instability below the whole excavation in clay conditions on sloping sites;

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
192 Deep excavations

Table 5.12. Summary of advantages and disadvantages of analytical methods49

Method Basis Advantages Disadvantages

Limit equilibrium Equilibrium conditions used Needs only soil strength Does not model soil–structure
(Typical programs: for the failing soil mass. parameters. Simple and interaction, wall flexibility or
STAWAL, Full mobilization of soil straightforward. Estimate of construction sequence. Multi-
REWARD) strength, factors of safety deformation possible by propped walls, non-uniform
applied to lateral stresses at relating mobilized strength, surcharges and berms require
failure soil strength and wall considerable idealization.
rotation Only give an order of wall
deformation and surface
settlement. Better for
cantilever than propped walls
Subgrade reaction/ Soil modelled by a set of Soil–structure interaction Soil behaviour idealization is
beam or springs unconnected vertical and taken into account. Wall likely to be less than perfect.
(Typical programs: horizontal springs. movements are calculated. Results are sensitive to pre-
LAWALL, Alternatively some Simple to apply. excavation stress state and
WALLAP, programs model soil as a Construction sequence spring stiffness. Any failure to
M. SOILS) simple boundary element modelled provide overall stability is
using Mindlin’s equations usually only shown by a
for horizontal displacement failure to converge
Pseudo finite element Soil modelled as an elastic Soil–structure taken into Limited to soil with linearly
(Typical program: continuum with soil stiffness account. Wall movements increasing stiffness at depth.
FREW) matrices calculated using a are calculated. Relatively Results sensitive to pre-
finite element program straightforward. excavation stress state.
Construction sequence Experience shows an
modelled overtendency to predict wall
deformation
Finite element and Full soil–structure Full complex soil models Time-consuming to set up.
finite difference. interaction with modelled can represent variation of High quality data required to
(Typical programs: sequence of construction stiffness with strain and realise full potential. Use of
OASIS SAFE (FE) anistropy. Can model Mohr–Coulomb soil model
PLAXIS (FE) complex wall and excavation may give unrealistic ground
CRISP (FE) geometry. Can model deformation
FLAC (FD) consolidation at construction
ABACUS (FE) stages as soil moves from
undrained to drained
conditions. Some programs
can model 3D conditions

the risk of inadequate prop capacity; wall penetration below final formation
level.
The wall penetration depth should always be regarded as a minimum when
calculated by methods described in this chapter. Adequate embedment depths
will depend on the risk of hydraulic failure as well as passive resistance to the
sheeting and relief of load from the lowest frame of bracing. In all multi-
propped (or anchored) walls the principal risks of failure in soils other than
very soft clays will stem from the adequacy of embedment depth of the wall
or sheeting and the sufficiency of strutting or anchoring. To counter these
risks in temporary works design requires care and experience.

Other design Plastic design


considerations Recent work by Kort46 on sheet pile design for sheet piles has highlighted
potential savings in sheet pile tonnage using plastic design methods. Kort’s

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 193

work includes some validation of the methods by full-scale tests in soft soils
and also includes allowance for the phenomenon of oblique bending that
may occur in U-shaped, Larrsen sheet piles. Kort’s work and its practical
application to sheet pile design is referred to in more detail in Chapter 7.

Construction of long walls in short lengths


It has long been realized that safer construction with less wall deformation
and reduced strut or anchor loads apply when trenchworks and deep
excavations can be taken out in short lengths and the permanent works are
built in stages progressively within these lengths. This avoidance of keeping
excavations for greater length and greater time with only temporary support
than necessary is good construction sense. This method of working in short
lengths can particularly apply at the deepest section of cofferdams with
short length working during excavation and installation of the bottom
frame and then the subsequent replacement of the temporary framing,
bottom upwards, by the permanent support. This method is very often
most desirable at this depth and in some instances, essential.
Until finite element analysis was applied in three dimensions it was not
possible to model short length working and its benefits by calculation.
Recently research work by Gourvenec et al.47 has used the CRISP FE pro-
gram to compare the deformation of a long embedded cantilevered diaphragm
wall excavation using both 3D and plane strain analyses with field measure-
ments of deformation. The wall was 800 m long with a cutting for highway
construction. The upslope side of the cutting was supported by diaphragm
wall panels 22.85 m deep, 1.5 m thick and 7.5 m long. The downslope wall
panels were 12.87 m deep, 1 m thick and 5 m long. An earth berm, 6.36 m in
height and 5.17 m wide at its base with a 60 degree face was used to provide
temporary support to the upslope wall. The wall height at the upslope side
was 7.86 m above formation level. The soil retained was a medium dense
sandy silt overlying a stiff overconsolidated fissured clay. Pore pressures
were hydrostatic below the sand/clay interface. Both diaphragm wall
installation and excavation of the berm were modelled together with realistic
construction times for each part of the staged construction.
The wall deformation measured on site compared favourably with the 3D
FE analysis deformed wall shape but the plane strain analysis predicted
total deformation at the top of the wall as 38 mm, whereas the total from
the 3D analysis was 20 mm and broadly agreed with the field measurement.
The conclusions reached were as follows.
(a) Reduced ground movements during wall installation were calculated
with the 3D analysis compared with the 2D, although the analysis
method did not have a significant effect on subsequent wall movements
during bulk excavation in front of the wall.
(b) Up to 80% of the post-installation wall movement occurred during
bulk excavation to the initial berm profile.
(c) Wall movement when the berm was removed in stages was reduced by
about 50% using the 3D analysis compared to that using the plane
strain analysis.
(d ) Wall movements at the initial base slab prop section during removal of
adjacent sections of the berm were very small. Both in the analysis and
on site the initial section of base slab prop was effective in restricting
wall movement during excavation of the nearby berm sections.
The effect of short length working compared with bulk excavation full length
is the reduced rate of dissipation of negative excess pore water pressure
induced by bulk excavation. This, in turn, makes for greater safety.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
194 Deep excavations

Fig. 5.36. ‘Heidelberg’ soldier


pile and lagging wall allowing
extraction of steelwork wallings
and soldiers (patented by
Bilfinger and Berger)

King post walls


The use of soldier piles, often supported by tiers of ground anchors with hori-
zontal timbers (or laggings) or in situ concrete placed in successive lifts or
sprayed as shotcrete, provides an economical form of soil support, particu-
larly in dry subsoils. Care should be taken, however, in assessing the passive
resistance below final formation level of such a wall system. The passive

Fig. 5.37. King post wall derivatives (dimensions in metres)

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 195

Fig. 5.38. Cofferdam standard construction details: (a) typical layouts, sheet piled cofferdams; (b) connection details;
(c) temporary anchor head detail; (d) standard details with tie rods to deadmen

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
196 Deep excavations

Fig. 5.39. (a) Schematic arrangements for anchored and braced piled and diaphragm walls

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 197

Fig. 5.39. (b) Typical jackable end detail for a tubular (CHS) prop (after Twine and Roscoe42 )

resistance of soil at the front of the soldier piles within their embedment depth
can only be mobilized by movement of the soldier pile towards the excavation.
Soldier pile walls are relatively flexible constructions, and in relatively dense
granular soils and stiff clays this mobilization of passive resistance without

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
198 Deep excavations

Fig. 5.39. (c) Typical jackable end detail for a UC prop (after Twine and Roscoe42 )

undue lateral movement can reasonably be assumed. In these circumstances it


is also reasonable to allow an effective width of soil acting to provide passive
resistance in excess of the net width of the steel or reinforced concrete pile
acting as the soldier. Where soldier spacing centre-to-centre is greater than
four times the net width of the soldier below formation level it is reasonable
to assume an effective width of soil to provide passive resistance equal to
twice the soldier width below formation level. Where the soldier spacing is
less than four times the soldier width below formation level no such increase
in calculated passive resistance is permissible because of the risk of overlap of
calculated passive zones.
It is essential to check the bearing capacity at the base of such soldiers where
inclined anchors are used. The vertical components of all anchors should be
summated as they are pre-stressed together with the soldier pile self-weight
comprising the vertical load at the soldier base.
Figure 5.36 shows a soldier pile wall constructed by Bilfinger and Berger.
Part of the wall is built in a patented form known as a ‘Heidelberg’ wall
where the walings and soldiers can be extracted for reuse.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 199

Construction details
A selection of construction details from the Author’s files is given in Figs 5.37
to 5.39. The wall types shown include anchored sheet pile walls, king post
walls using timber laggings, cofferdam details and multi-prop walls using
secant piles and diaphragm walls. Details of typical end details where jack
installation is included are also reproduced from Twine and Roscoe’s
CIRIA report42 .

References 1. DD ENV 1997–1 Eurocode 7. Geotechnical design. Part 1: general rules. British
Standards Institution, London, 1995.
2. BS 8110. Structural use of concrete. British Standards Institution, London, 1997.
3. BS 8007. Code of practice for design of concrete structures for retaining aqueous
liquids. British Standards Institution, London, 1981.
4. BS 5400. Steel, concrete and composite bridges. British Standards Institution,
London, 1990.
5. BD 42/00. Design manual for roads and bridges, Vol. 2, section 1, (DMRB 2.1.2),
HMSO, London, 2000.
6. Peck R.B. Deep Excavations and tunneling in soft ground. Proc. 7th Int. Conf.
S.M.F.E., Mexico City, 1969, State of the art volume, 225–290. Sociedad de
Mexicana de Suelos, Mexico City, 1969.
7. Clough G., Smith E. and Sweeney B. Movement control of excavation
support systems by iterative design procedure. ASCE Ground. Engng, 1989, 1,
869–884.
8. Clough G. and O’Rourke T. Construction induced movement of insitu walls.
ASCE Special Publication No. 15. Proc. Design and Performance of Earth Retain-
ing Structures, Cornell University, 1990, 439–470.
9. St John H., Potts D., Jardine R. and Higgins K. Prediction and performance of
ground response of a deep basement at 60 Victoria Embankment. Proc. Wroth
Memorial Symposium, Oxford, 1992. Thomas Telford Ltd., London, 1993.
10. Carder D. Ground movements caused by different embedded retaining wall con-
struction. Transport Research Laboratory, Crowthorne, 1995, TRL report 172.
11. Fernie R. and Suckling T. Simplified approach for estimating lateral wall
movement of embedded walls in UK ground. Proc. Int. Symp. Geo. Aspects of
Underground Construction in Soft Ground, City University, 1996, 131–136. City
University, London, 1996.
12. Carder D., Morley C. and Alderman G. Behaviour during construction of a
propped diaphragm wall in London clay at Aldershot and road underpass. Trans-
port Research Laboratory, Crowthorne, 1997, TRL Report 239.
13. Long M. Database for retaining wall and ground movements due to deep
excavations. ASCE J. Geotech. Eng., 127, March 2001, 203–224.
14. Brooker E.W. and Ireland H.O. Earth pressures at rest related to stress history.
Canadian Geotech. J., 1965, 11, Feb., 1–15.
15. Caquot A. and Kerisel J. Tables for the calculation of passive pressure, active
pressure and bearing capacity of foundations. Gauthier-Villars, Paris, 1948.
16. Civil engineering codes of practice. No. 2. Earth retaining structures. Civil
Engineering Codes of Practice Joint Committee, London, 1951.
17. CIRIA. The design and construction of sheet piled cofferdams. Thomas Telford
Ltd., London, 1993, Special publication 95.
18. Terzaghi K. and Peck R.B. Soil mechanics in engineering practice. Wiley, New
York, 2nd edition, 1967.
19. BS 8002. Earth retaining structures. British Standards Institutions, London,
1994.
20. Rowe P.W. Anchored sheet pile walls. Proc. Instn Civ. Engrs, Part 1, 1952, 1,
Jan., 27–70.
21. BD 37/01. Design manual for roads and bridges, Vol. 1, section 3, (DMRB 1.3).
Highways Agency, London, 2001.
22. Design manual 7.2. Foundation and earth structures. US Navy, Washington, DC,
1982.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
200 Deep excavations

23. Goldberg D.T. et al. Lateral support systems and underpinning, Vol. 1, design and
construction, Vol. 2, design fundamentals, Vol. 3, construction methods. Federal
Highway Administration, Washington, DC, 1976, FAWA-RD-75, 128.
24. Simpson B. and Driscoll R. Eurocode 7: A commentary. Building Research
Establishment, Watford, 1998.
25. Sainflou M. Essai sur les digues maritimes verticales. Ann. des Ponts et Chausse´es,
1928, 98, 5–48, (translated by US Corps of Engineers).
26. Minikin, R. Wind, waves and maritime structures. Griffin, London, 1963.
27. BS 6349 (part 1). Maritime structures. British Standards Institution, London,
1984.
28. Potts D.M. and Burland J.B. A parametric study of the stability of embedded earth
retaining structures. Transport Research Laboratory, Crowthorne, 1983, Report
813.
29. Symons, I.F. Assessing the stability of a propped in-situ retaining wall in over-
consolidated clay. Proc. Instn Civ. Engrs, 1983, 75, Dec., 617–633.
30. Piling handbook. British Steel Corporation, Scunthorpe, 1988.
31. Sheet piling design manual. USS Steel, Pittsburgh, 1974.
32. Review of design methods for excavations in Hong Kong. Geotechnical Control
Office, Hong Kong, 1990.
33. Blum H. Einspannungsverhaltnisse bei Bohlwerken. Ernst and Son, Berlin, 1931.
34. Terzaghi K. Anchored bulkheads. Trans. ASCE, 1954, 119, paper 2720, 1243–
1324.
35. Rowe P.W. A theoretical and experimental analysis of sheet pile walls. Proc. Instn
Civ. Engrs, 1955, 4, Jan., 32–69.
36. Rowe P.W. Sheet pile walls in clay. Proc. Instn Civ. Engrs, 1957, 7, July, 629–654.
37. Terzaghi K. Theoretical soil mechanics. Wiley, New York, 1943.
38. Bjerrum L. and Eide O. Stability of strutted excavations in clay. Ge´otechnique,
1956, 6, 32–47.
39. Aas G. Stability problems in a deep excavation in clay. Norwegian Geotechnical
Institute, Oslo, 1985, Publication 156.
40. Bjerrum, L. Problems of soil mechanics and construction on soft clays. Proc. 8th
Int. Conf. S.M.F.E., Moscow, 1973, Vol. 3, 111–159. USSR National Society for
Soil Mechanics and Foundation Engineering, Moscow, 1973.
41. BD 28/87. Early thermal cracking of concrete. UK Department of Transport,
London, 1987.
42. Twine D. and Roscoe H. Prop loads: guidance on design. CIRIA, London, 1996,
Funders report.
43. Batten M. and Powrie W. Measurement of temporary prop loads at Canary
Wharf. Proc. ICE Geotech. Eng., 143, 2000, 151–163.
44. Richards D., Homes G. and Beadman D. Measurement of temporary prop walls
at Mayfair car park. Proc. ICE Geotech. Eng., 137, 1999, 165–174.
45. Potts D.M. The analysis of earth retaining structures. Proc. Conf. Retaining
Structures. Institution of Civil Engineers, London, 1992, 167–186.
46. Kort D.A. Steel sheet pile walls in soft soil. DUP Science, Delft, 2002.
47. Gourvenec S., Powrie W. and De Moor E. Three dimensional effects in the
construction of a long retaining wall. Proc. ICE Geotech. Eng., 2002, 155, July,
163–173.
48. Padfield C.J. and Mair R.J. Design of retaining walls embedded in stiff clay.
CIRIA, London, 1984, report 104.
49. CIRIA. Embedded retaining walls: guidance for economic design. Gaba A.R. et al.
CIRIA, London, 2002.
50. Bjerrum L. et al. Earth pressures on flexible structures. Proc. 5th Euro. Conf.
SMFE, Madrid, 1972, Vol. 2, 169–196.
51. Institution of Structural Engineers, Design and construction of deep basements
including cut-and-cover structures. Draft, 2003. Inst. Structural Engrs, London.

Bibliography Bjerrum L. et al. Earth pressures on flexible structures. Proc. 5th Euro. Conf.
S.M.F.E., Madrid, 1972, Vol. 2, 169–196. Sociedad Española de Mechánica del
Suelo y Cimentaciones, Madrid, 1972.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.
Design of vertical soil support 201

Burland J.B. et al. The overall stability of free and propped embedded cantilever
retaining walls. Ground Engng, 1981, 14, No. 5, 28–38.
Carder D.R. and Symons L.F. Long term performance of an embedded cantilever
retaining wall in stiff clay. Ge´otechnique, 1989, 39, No. 1, Mar., 55–76.
EAU 90. Recommendations of the Committee for Waterfront Structures, Harbours and
Waterways. Ernst and Son, Berlin, 6th English edition, 1993.
Egger P. Influence of wall stiffness and anchor prestressing on earth pressure distribu-
tions. Proc. 5th Euro. Conf. S.M.F.E., Madrid, 1972, Vol. 1, 259–264. Sociedad
Española de Mechánica del Suelo y Cimentaciones, Madrid, 1972.
Garrett C. and Barnes S.J. Design and performance of the Dunton Green retaining
wall. Ge´otechnique, 1984, 34, Dec., 533–548.
Head J.M. and Wynne C.P. Designing retaining walls embedded in stiff clay. Ground
Engng, 1985, 18, Apr., 30–38.
McRostie G.C. et al. Performance of tied back sheet piling in clay. Canadian Geotech.
J., 1972, 9, 206–218.
Potts D.M. and Burland J.B. A numerical investigation of the retaining walls of the Bell
Common tunnel. Transport Research Laboratory, Crowthorne, 1983, Supplemen-
tary Report 783.
Potts D.M. and Fourier A.B. The effect of wall stiffness on the behaviour of a
propped retaining wall. Ge´otechnique, 1984, 34, Sept. 383–404 (Discussion 35,
119–120).
Simpson B. et al. An approach to limit state calculations in geotechnics. Ground
Engng, 1981, 14, No. 6, 21–28.
Symons L.F. Behaviour of a temporary anchored sheet pile wall on A1(M) at Hatfield.
Transport Research Laboratory, Crowthorne, 1987, Research report 99.
Terzaghi K. Evaluation of coefficients of subgrade reaction. Ge´otechnique, 1955, 5,
No. 4, 297–326.
Winkler, E. Die lehre von elastizitat und festigkeit, Prague, 1867.

Downloaded by [ UNIV OF ILLINOIS] on [13/10/16]. Copyright © ICE Publishing, all rights reserved.

You might also like