Sei sulla pagina 1di 62

CHAPTER 5.

RETAINING STRUCTURES
5.1 Introductory notions. Classification
Retaining structures are construction works for which the main load is
represented by the earth pressure. The aim of a retaining structure is to
support the side of an excavation or a fill or, in the case of cofferdams, to
support the pressure of the water.
According to their destination, retaining structures can be grouped in
temporary works and permanent works.
Temporary works serve to support the sides of excavation in which a
foundation or an underground structure is built or, in the case of cofferdams,
to create an area in which work below water level can be carried out in dry.
Based on the construction system, the temporary retaining works can be
classified as:
- works with recoverable elements (timbering, sheet piling)
- works with non-recoverable elements (diaphragm walls)
Permanent works can be retaining structures above the ground level
(retaining walls, quays etc.) or underground structures (basement walls,
subway lines and stations etc.).
5.2 Timbering
Timbering implies the support of the sides of excavation by timber boards, the
stability being maintained by means of horizontal struts or, in the case of large
excavations, by raking struts or shores.
There are three main variants of timbering:
- timbering with horizontal boards,
- timbering with vertical boards or runners,
- timbering with soldier piles and horizontal laggings.
5.2.1 Timbering with horizontal boards
The use of this kind of temporary support is restricted to soil having a
cohesion large enough to enable the side of the excavation to stand
unsupported for a certain length of time until the elements of the timbering are
put in place. In this category enter firm to stiff cohesive soils.
105

There are three main elements which form this kind of timbering:
horizontal boards,
walings, placed vertically,
struts.
The fig. 5.1 shows the phases of the construction of a timbering using
horizontal boards:
- I: the untimbered excavation is performed, on a depth depending on
the cohesion of the soil;
- II: boards, walings and struts corresponding to the first step of
excavation are put in place; the struts are tightened by cutting them
slightly too long and then driving one end with the other held in
position, until they are at right angles to the waling;
- III: a new untimbered excavation is performed, followed by the
installation of boards, walings and struts.

1 board; 2 waling; 3 strut


Fig. 5.1
The cycle is repeated until the final excavation level is reached.
In the case of large excavations, in order to prevent the buckling of long
struts, vertical piles are used (fig. 5.2).

106

1 vertical pile
Fig. 5.2
Another solution, which has the advantage of improving the working space, is
to use raking struts (fig. 5.3).

1 raking struts
Fig. 5.3
5.2.2 Timbering with vertical boards or runners
This kind of support is used in non-cohesive soils (sands, gravels) and in soft
clays and silts. Unlike the case when horizontal boards are used, in these
ground conditions the timbering is placed in position ahead of the excavation.
Fig. 5.4 shows the phases of the construction of a support using vertical
boards.
- I: On the ground surface a first row of frames is placed, consisting on
guide walings and struts. Timber runners are pitched behind the
walings. The usual dimensions of these boards are 175 mm by 38
mm or 175 mm by 50 mm in lengths up to 4.8m. After driving the
runners at the depth of the first step, wedges are placed between
them and the walings.
107

- II: The excavation begins in panels of 11,5 m, observing strictly the


following rule: before proceeding to the excavation of a panel, the
runners belonging to the respective panel must be lowered one step
further. For that purpose, the wedges which kept the runners tight
against the walings are removed and the runners are driven. Then, a
second row of frames is installed and wedges between the walings
and runners are placed. The procedure is repeated for the next panel
and so on, until the second row of frames is unstalled on the entire
area.
- III: The same cycle is used for the new step of excavation and then
repeated until the final level of excavation is reached.
Using the described procedure, the depth of the excavation is limited by the
length of the vertical boards (usually no more than 45 m). If the excavation
has to be taken deeper than the runners, a second setting of runners is
placed within the walings of the top setting and an inner guide waling is also
placed. The new runners are then driven down and bracing frames of walings
and struts are put in place as the excavation proceeds. If still deeper
excavation is required, a third setting of runners can be used. In this way, the
excavation at the top setting will be about 1 m wider than at the third setting,
leading to an increase in the volume of excavation and in the consumption of
timber (fig. 5.5).

1 frame; 2 puncheons ; 3 vertical boards; 4 wedges;


5 contour of the excavation for a panel of 11,5 m
Fig. 9.4
108

Fig. 5.5
Another procedure suitable for deep excavation is the one using poling
boards, known also as the marciavante procedure. Construction phases
using marciavante method are shown in fig. 5.6.
- I the first set of bracing frames is put in place, on whose contour
poling boards of 1.5-2 m in length are driven; between them and the
walings of the bracing frames are put wedges;
- II excavation is performed up to a level of 0.3-0.4 m above the tip of
the poling boards; the second set of bracing frames is installed;
between the poling boards and the second set of bracing frames are
placed guiding wedges which will insure the required inclination of
the second set of boards;
- III poling boards of the second set are driven; between them and the
second set of bracing frames are placed wedges; then, phases II
and III are repeated until the final level of excavation is reached.

109

1 poling boards; 2 horizontal frames; 3 puncheon ;


4 guide wedges; 5 wedges
Fig. 5.6
For large excavations (fig. 5.7) a bracing system made of struts and liners is
needed, in order to increase the stiffness of the structure, to reduce the span
of the elements working in bending (walings) and the buckling length of struts.

1 vertical boards; 2 wedges;


3 horizontal frames with liners; 4 clamps
Fig. 5.7
For narrow trenches, of relatively small depths, such as those needed for
installing pipes, sewers etc., a movable support can be provided consisting
generally of vertical sheeting members and struts adjustable either by
110

hydraulic rams or screw jacks, allowing sheeting members to be forced tightly


against the soil (fig. 5.8).

a transversal section; b detail of the horizontal steel frame;


1 vertical board; 2 horizontal steel frame; 3 clamping sleeve;
4 key
Fig. 5.8
5.2.3 Timbering with soldier piles and horizontal laggings
This is a method normally applied to deep excavations and which is some
times called Berlinese method since it was first used at the construction of
the Berlin subway. The soldier piles are rolled steel H section profiles,
driven before excavations from ground level to a level 1-2 m below the bottom
of the excavation. As the excavation is taken down, timber boards are
inserted horizontally between the flanges of the piles and held against the
face by wedges (fig. 9.9). The system is completed with walings and struts.
The distance between soldier piles is usually between 2.5 and 4 m.

1 horizontal board; 2 waling; 3 strut; 4 steel H profile


Fig. 5.9
111

5.3 Sheet piling


Sheet piles are elements made of timber, steel or reinforced concrete,
installed in ground by driving or vibrating, used for the construction of walls
which, besides the strength and stability requirements, should also fulfill the
condition of water tightness.
Sheet piles can be used in temporary retaining structures or in permanent
structures, as for example the harbour quay shown in fig. 5.10.

1 sheet pile; 2 piles; 3 reinforced concrete slab


Fig. 5.10
5.3.1 Timber sheet piles
Timber sheet piles are made of fir or oak, representing boards of 510 cm
thickness or planks up to 25 cm thick and 2030 cm width.
When good water tightness is not required, contiguous (fig. 5.11a) or
overlapped boards (fig. 5.11b,c) can be used. When watertightness is
required, various kinds of joints are used such as in half-wood joint (fig.
5.11d) in birdsmouth joint (fig. 9.11e) or tongue and groove joint (fig. 5.11f).
The tongue and groove type of joint may be obtained by bolting together
three boards (fig. 5.11g).

112

Fig. 5.11
Sheet piles are inserted into ground by driving. In order to ease the
penetration and insure the closing of joints, timber sheet piles are sharpened
at the lower end on the groove side (fig. 5.12).

Fig. 5.12
Sheet piles are driven with the tongue in front and the groove sliding in the
tongue of the previously installed sheet piles. Otherwise, the groove could be
stucked with soil grains obstructing the penetration of the sheet pile and
causing the opening of the joint.
To make better use of the driving capacity of the available equipment, in
some situations 2 or 3 timber sheet piles are driven simultaneously, after
joining them by means of clamps (fig. 5.13).
113

1 tongue; 2 groove
Fig. 5.13

1 pile; 2 sheet pile


Fig. 5.14

To support the walls of large excavation by using timber sheet piles, at


intervals of 44.5 m and, in any case, at the corners of the wall, timber piles
are first driven, serving as guides for the sheet piles (fig. 5.14). To facilitate
the driving of the sheet piles in vertical direction, external guide walings made
of two boards each are placed at the ground level and 1 m above the ground
level and bolted to the piles (fig. 5.15).

1 guide piles; 2 tongs; 3 set of 2-sheet piles; 4 wedge;


5 timber piles; 6 wedges; 7 clamps; 8 spacer
Fig. 5.15
114

In order to avoid the opening of the joints between the sheet piles in case
obstacles are met in the ground, the driving is not done for the entire depth
for each sheet pile but in ladder, observing to keep the difference in level of
the tips of two consecutive sheet piles (or bunches of 2-3 sheet piles) less
than 11.5 m.
Timber sheet piles are manufactured of green wood, because the dry wood in
contact with water would expand, deforming the wall.
The advantages of the timber sheet piles are: easiness in manufacturing;
reduced weight; easy to be installed; low cost. The disadvantages are: limited
length (max. 68 m, from which 3...4 m the embedment; complete
recovering is practically not possible; cannot be driven deeply into dense
granular soils or stiff clays without risk of splitting.
5.3.2 Steel sheet piles
Steel sheet piles are used when deep penetration or penetration in hard soils
are required. There are many types of rolled steel sections available for sheet
piles, which differ among them by the shape and by their interlocking
sections. The most common types are:
U type of steel sheet piles, at which joints are placed along the wall
axis at each intersection of the axis with the profile (fig. 9.16a);
S type of steel sheet piles, at which joints are also along the wall axis
but at every second intersection with the profile (fig. 9.16b);
Z type of steel sheet piles, at which joints are placed outside the wall
axis, alternating from one side to the other (fig. 9.16c).

1 sheet pile; 2 interlocking section


Fig. 5.16
115

At equal mass and identical disposition of the material against the vertical
plane, sheet pile walls made of the three types of sections (profiles) have
different stiffnesses, as a result of the position of joints. Thus, at the U type
of walls, the principal axis x ' x ' of each element is parallel to the axis x-x of
the wall. Slight relative rotations of the sheet piles under the pressure normal
to the x-x axis are possible, hence the real stiffness, considering the
interlocking, should be diminished. When stresses in the wall are checked,
one should consider in this case w x 0.75 w x , where wx is the resistance
modulus against the x-x axis and w x is the reduced resistance modulus,
considering the joints. At the S-type of walls, the principal axes x ' x ' of each
element are parallel among them but inclined in respect to the axis x-x of the
wall. The possibilities of rotations are smaller than in the first case and
w x 0.85 w x . At the Z-type of walls, the principal axes x ' x ' of adjacent
sheet piles are normal one to the other, the rotation tendencies of the
adjacent sheet piles are cancelling; in this case w x 1.0 w x .
There is a great diversity of interlocking sections (fig. 5.17). The joints or
interlocking sections must be strong enough to support the tensile stresses
occurring in exploitation, must be watertight and must insure an easy
penetration and extraction of the sheet piles.
Steel sheet piles can reach lengths up to 30 m.

Fig. 5.17
5.3.3 Reinforced concrete sheet piles
Reinforced concrete sheet piles are precast elements, of square or
rectangular cross-section, with sides of max. 5060 cm and thicknesses of
1050 cm. Their length is limited to 1820 m, because of the large weight
which would make the driving very difficult or impossible.
The interlocking system can be similar to the one used at timber sheet piles.
In fig. 5.18 a is shown a reinforced concrete sheet pile having a tongue and
groove type of joint. In order to reduce the friction during the driving process,
the tongue is provided only in the lower third of the sheet pile, in the upper
116

part there are two grooves. The gap between grooves is filled with jute and
mortar (fig. 9.18b).

Fig. 5.18
9.4 Embedded walls
Embedded walls made of reinforced concrete are intensively used as
retaining structures for deep excavations, particularly in urban, congested
areas. Unlike sheet piles walls, embedded walls made of reinforced concrete
cannot be recovered and reused.
From the constructive point of view, embedded walls can be formed by bored
piles or by panels.
5.4.1 Embedded walls made of bored piles
There are two types of pile walls:
- secant pile walls
- contiguous pile walls
Secant pile walls are of two kinds:
- hard/hard secant pile walls
- hard/soft secant pile walls.
Hard/hard secant pile walls consist of overlapping concrete bored piles. The
secant pile wall is constructed in two stages (fig. 9.19). All piles constructed
during stage 1, defined as primary piles, are spaced at specified spacing. All
piles constructed during stage 2, defined as secondary piles, are positioned
between the primary piles and secant with the primary piles.

Fig. 5.19
117

Guide walls are placed at the ground surface to ensure the designed position
and the verticality of each pile.

Fig. 5.20
Usually, only the secondary piles are reinforced (fig. 9.20). However, primary
piles can also be reinforced with reinforcing cages of rectangular shape or
with steel sections (fig. 5.21a,b).
If proper construction is insured, hard/hard secant pile walls offer a high water
tightness.
Hard/soft secant pile walls (fig. 5.22) are constructed in a similar way as the
hard/hard secant pile walls. The difference is that primary piles are formed of
a low strength cement/bentonite mix, where bentonite is an active clay rich in
montmorillonite.
This kind of walls are used when the conditions for watertightness are less
severe.

Fig. 5.21

Fig. 5.22
Contiguous pile walls
118

When the wall is not required to retain water, contiguous bored piles are used
(fig. 5.23). The piles are constructed at centres equal to the pile diameter plus
an allowance for temporary casing width and tolerance which can vary
between 70 and 150 mm. Guide walls may be used at ground level to ensure
positional tolerance.

Fig. 5.23
5.4.2 Embedded walls made of panels
Embedded walls made of panels will be named in what follow as diaphragm
walls.
Panels are rectangular trenches with lengths depending on the excavation
equipment and on the position within the wall.
The common feature of all types of diaphragm walls is that the excavation of
the trench is made under the protection of a supporting fluid or slurry. This is
a clay suspension of 1.031.10 g/cm3 density, formed by mixing water with
bentonite. Additives are also used to improve the flow characteristics of the
fluid and the gelling or blocking action of the fluid. The clay suspension
infiltrates through the sides and the bottom of the trench, filling on a certain
distance the pores of the soil. Through this layer of soil with low permeability
only water can pass, while clay particles are accumulating at the face of the
layer forming a shield called cake. The layer of soil enriched in clay together
with the cake form a screen. The presence of this screen and of the pressure
exerted by the suspension on the sides of the trench ensure the stability of
the walls if the required properties of the suspension are met, such as the
density, viscosity, shear strength, pH, sand content etc.
The corresponding tests and compliance values are specified in various
codes.
Based on the criterion of the material met in a vertical section and of the role
played by the wall, embedded walls made of panels (diaphragm walls) can be
divided into two categories:
- homogeneous walls
119

- composite walls
Homogenous walls are diaphragm walls at which both the material used and
the role played are unchanged along the same vertical.
In function of the material, homogeneous walls are classified in:
- cast in situ concrete diaphragm walls
- precast concrete diaphragm walls
Composite walls are diaphragm walls at which the functions of strength and
water tightness are separated on the vertical.
Cast in situ concrete diaphragm walls
The basic sequences of the construction of a panel are:
I. construction of guide-walls
II. excavation, with a bentonite suspension
III. placing the reinforcement
IV. concreting
V. trimming
A distinct phase is represented by forming the joints. Its position between
the above specified basic sequences depends on the type of installation used
for excavating the trench.
I.

Guide walls are small, parallel temporary walls, usually made of


reinforced concrete (fig. 5.24).
They have multiple roles:
to materialize on the ground surface the position of the wall;
to provide a guide for the excavating tool;
to secure the sides of the trench against collapse in the vicinity of the
fluctuating level of the supporting fluid;
to provide temporary support for the reinforcement cages;
to provide reaction for the hydraulic jacks when extracting the tubes
used for forming the joints;
to serve as rolling tracks for some installations used for the
excavation.
The distance between the parallel guide walls is equal to the width of the
excavation tool, plus 510 cm. In order to keep this distance, both in front
and behind the excavation tool, struts are placed.

120

Fig. 5.24
II.

Excavation of panels
- Excavation with the Kelly-type installations. The characteristic
feature of these installations, which are attached to an existing
equipment (such as a crane or excavator) is the presence of a heavy
rod named kelly having at its end a hydraulically operated grab. In
fig. 5.25 is shown a kelly-type installation, equipped with a
hydraulic grab made in Romania. The grab has two jaws which are
opened and closed by means of hydraulic jacks. The lifting and the
lowering of the kelly is done by cables. For the excavation, the grab
with open jaws is brought above the guide-walls and is inserted
slowly in the trench to avoid waves of the slurry which could affect
the walls. Under its own weight and the weight of the kelly, the grab
is let to fall from a height of 13 m above the bottom and penetrates
in the soil. Loaded with soil, the grab is then lifted above the guidewall and, after a rotation of the excavator (or crane), is discharged in
a truck. During the excavation, the level of the supporting fluid is kept
permanently 0.51.0 m below the ground level, but at least 1.0 m
above the groundwater table. Samples are taken periodically from
the slurry and tested in the laboratory at the construction site to
determine the properties of the suspension, which are checked
against the required ones, to ensure that the suspension does not
become excessively diluted or contaminated by soil particles.

With kelly-type installations, trenches of 0.600.80.1.0 m width can be


excavated, at depths up to 35 m. The minimum length of a panel is equal to
the maximum opening of the jaws and varies between 2.20 and 2.80 m. The
maximum length of a panel should not exceed 7 m.
In fig. 5.26 is shown the excavation of a 7.0 m length panel, using the grab
E.S.G.H. made in Romania. In the first two phases, two shafts are excavated
at both extremities of the panel, with a length of 2.80 m each. Between the
shafts remains a core of soil which is excavated in the last phase. The reason
121

for doing so is the following: during the excavation, the stability of the
installation, whose center of gravity is high due to the very long kelly, is
ensured by the reaction opposed by the soil encountered by the jaw of the
grab; if this reaction is not uniform, there is a risk for the installation to rotate
and overturn; if the shafts would have been excavated in sequence, one of
the jaw would rest on the ground while the other would enter in the shaft
previously excavated and filled with slurry, situation which should be avoided.

1 grab; 2 Kelly; 3 guide; 4 telescopic arm; 5 roller


a - equipment ESGH 20-30; b - grab
Fig. 5.25

Fig. 5.26

122

1 cup; 2 mast; 3 winch; 4 transporting bin


Fig. 5.27
- Excavation with the installations E.L.S.E. The work of this installation, of
Italian origin, is based upon the principle of the excavator with straight
cup; the cup makes both a translation movement in vertical direction by
descending along a heavy mast and a translation movement around a
point of support (fig. 5.27). The installation uses the guide-wall as rolling
tracks. The moving mast is placed at the end of the panel to be
excavated, the excavation being performed by the retirement of the
installation. The cup in vertical position bolted in the mast, is lowered
together with the mast, with teeth downward, and penetrates in the soil.
Then, the cup is brought in horizontal position and lifted along the mast
above the ground level, where is discharged. The cycle is repeated until
the excavation level specified in the design is reached. The length of the
excavation in one stage, dictated by the amplitude of the cups
movement, is 3.80 m. For the next stage, the installation is retired further,
in order to ensure the excavations of the rest of the panel. The length of
the panel is usually taken 57 m, which means a two-stage excavation.
The width of the excavation made with E.L.S.E. installation is of
0.6..0.81.0 m and the depth up to 30 m. Soils suitable to be excavated
with E.L.S.E. are non-cohesive or weak cohesive soils. In soils with high
plasticity and cohesion (clays), both the excavation and the discharge of
the soil are difficult, due to the large adhesion between the soil and the
cup. Also, E.L.S.E. cannot excavate in hard ground, such as a
sandstones, limestones, etc.
123

- Excavation with the installation C.I.S. Soletanche. This installation, of


French origin, is using the reverse-circulation of the suspension. The
slurry, mixed with the excavated material (detritus) is absorbed through
the pipe of the installation; after the coarse particles are removed, the
slurry is sent back in the trench. Usually, the excavating tool placed at
the end of the pipe is performing a rotary drilling. However, when the
material to be excavated is hard, a percussion type of drilling can be
used, and the excavation tool is adopted in consequence. The
excavation of a panel by using C.I.S. Soletanche installation begins with
the excavation at one of the extremities of the panel of a shaft with a
diameter equal to the width of the panel. Then, moving on the guidewalls, used as rolling tracks, to the other extremity, the installation
excavates a second shaft. The excavation of the core between the two
shafts is made in layers of 3050 cm in thickness, by the displacement
back and forth of the installation (fig. 5.28). With the installation C.I.S.
Soletanche, trenches of 0.401.0 m width and up to 50 m in depth can
be excavated in any kind of ground, including in hard or very hard
materials. Due to this advantage, the installation C.I.S. Soletanche can
be used in combination with the installations Kelly and E.L.S.E. for the
socketing of the wall into the bedrock.

1 engine; 2 winch; 3 slurry pump; 4 stem for boring and


absorption; 5 excavating tool; 6 slurry
Fig. 5.28
III.

Placing the reinforcement

Panels are reinforced with reinforcement cages which include vertical and
horizontal bars, forming two parallel nets linked with stirrups and inclined bars
(fig. 5.29). Each cage is provided also with suspension and lifting bars,
124

bracing bars to improve the stiffness for the handling operations and spacers,
usually made of rollers, to ensure that the correct concrete cover is
maintained. On the vertical, the cage can be made of one piece or of several
pieces assembled by welding as the cage is lowered in the trench. When two
cages are installed in horizontal direction, the minimum distance between
them shall be 200 mm.

1 longitudinal bars; 2 horizontal bars; 3 bracing bars;


4 lifting bar; 5 spacers
Fig. 5.29
The vertical length of a reinforcement cage shall be such that the distance
between its base and the bottom of excavation is at least 0.2 m. The cage
shall not rest on the bottom, but shall be suspended from the guide-walls by
means of the suspension bars.
IV.

Concreting

Concrete is placed beneath the supporting fluid through one or more


concreting pipes or tremie pipes which are pipes equipped with a hopper at
the top but may also be pipes connected directly to concrete pumps. The
inner diameter of the concreting pipe shall be at least 0.15 m and 6 times the
maximum aggregate size. Its outer diameter shall be such that it passes
freely through the reinforcement cage. For panels with length less than 5 m,
one concreting pipe can be used.
A concrete of low consistency is used. The consistency of the fresh concrete
just before concreting shall correspond to a slump value between 180 mm
and 210 mm. Retarding admixtures are used to prolong the workability as
required for the duration of the concreting process.
125

When starting concerting, the supporting fluid and the concrete in the
concreting pipe shall be kept separate by a plug of material of by other
suitable means. To start concreting, the concreting pipe shall be lowered to
the bottom of the trench and then raised approximately 1 m. After concreting
has started, the concreting pipe shall always remain immersed in the fresh
concrete. The minimum immersion should be 2 m.
Since the top of the cast concrete is contaminated because of the contact
with the slurry and may not be of the required quality, sufficient concrete shall
be placed in the panel to ensure that the concrete below the cut-off level pass
the specified properties. This is achieved by providing an additional height of
concrete above the cut-off level.
Fig. 5.30 shows the operations of placing of the reinforcement cages in the
trench and of the concreting of a panel.

1 reinforcement cage; 2 slurry; 3 joint tube; 4 previously


concreted panel; 5 concrete poured in the panel using the tremie pipe
Fig. 5.30
V.

Trimming

Trimming of the concrete to cut-off level (removing the concrete of poor


quality, in excess) shall be carried out using equipment which will not damage
the concrete or reinforcement. Final trimming to cut-off level shall only be
carried out after the concrete has gained sufficient strength to avoid damage.
Forming the joints
The method used to form the joints depends on the type of equipment used
for excavating the trench.
In the case of Kelly-type and C.I.S. Soletanche equipments, the joints are
normally formed by using steel stop ends or joint tubes.
126

The steel stop ends are steel tubes with a diameter equal to the width of the
trench which are introduced in vertical position at the ends of a panel,
penetrating 0.51.0 m below the bottom of the trench in order to have
ensured the stability. The tubes are lowered into the trench with cranes and
extracted by use of hydraulic jacks. In the case of stop ends which are
extracted vertically, it is essential to define the optimum time for starting this
operation. If the tube is extracted too soon, the fresh concrete behind the tube
will flow in the space left by the tube. If, on the contrary, the extraction is done
too late, the tube can be sticked to the concrete and the recovering becomes
impossible. In order to avoid sticking, small rotations should be applied to the
tube before commencing the extraction and then the extraction should be
made gradually during the setting of the concrete. Usually, the extraction
starts 46 hours after the concreting is finished. A method to prevent the
contact between the concrete and the tube is to provide at the extremity of
the reinforcement cage a shield made of thin steel plate, on the entire depth
of the trench. In the case of tubes which are extracted laterally, the extraction
shall be made upon the completion of the excavation of the adjacent panel.
Instead of recoverable steel tubes, non-recoverable precast elements can be
used as stop ends (fig. 5.31).

Fig. 5.31

1 previously concreted panel;


2 panel under excavation
Fig. 5.32

In the case of using E.L.S.E. installations for the excavation of the trench, the
joints are formed by cutting into the concrete of the previously cast adjacent
panel. Cutting is done by the teeth of the excavating tool (fig. 9.32).
In special cases, water stops can be incorporated into the joints.
Phases of the construction of a cast in situ diaphragm wall made of panels
The wall is made by a number of panels. The panels disposition, their
dimensions in plane, the construction sequence, are established in the
design, taking into account the peculiarities of the job, the excavating
installations etc.

127

When using E.L.S.E. installations, the wall is formed of consecutive panels,


for each panel the construction cycle (excavation, lowering of the
reinforcement cage, concreting) being complete.
When using Kelly-type or C.I.S. Soletanche installations, there are two
variants for the construction of the walls:
a. wall made of primary and secondary panels (fig. 5.33)
I excavating of the primary panels and placing at their extremities the
joint tubes;
II lowering the reinforcement cage in the primary panels;
III concreting the primary panels;
IV extraction of the joint tubes;
V excavation of the secondary panels; the excavating tool is adapted in
order to properly clean the semi-circular joints between the secondary
and primary panels;
VI lowering the reinforcement cage in the secondary panels;
VII concreting the secondary panels.

1 excavated primary panel; 2 joint tube; 3 reinforcement cage; 4


concreted primary panel; 5 excavated secondary panel;
6 concreted secondary panel
Fig. 5.33
b. wall made of a starter and intermediate panels (fig. 5.34)
I excavation of the starter panel; placing at its extremities the joint
tubes;
II lowering the reinforcement cage in the starter panel;
III concreting the starter panel;
128

IV excavation of the intermediate panel; placing at its extremity of a


joint tube;
V lowering the reinforcement cage in the intermediate panel;
VI concreting the intermediate panel.
Phases IVVI are repeated for each intermediate panel until the whole
length of the wall is reached.

1 excavated primary panel; 2 joint tube; 3 reinforcement cage; 4


concreted primary panel; 5 excavated intermediate panel
Fig. 5.34
Diaphragm walls made of precast concrete
These walls are made of precast elements lowered into a trench containing a
self-hardening slurry.
Fig. 5.35 shows a wall made of precast panels which imitate the timber sheet
piles with tongue and groove. Fig. 5.36 shows a wall which imitate a Berlintype wall, with the difference that the horizontal lagging boards are replaced
by continuous vertical precast sheets. Fig. 5.37 shows a solution used at a
wall inside an industrial hall in Bucharest, designed by the Center for
Geotechnical Engineering of the Technical University of Civil Engineering
Bucharest. The length of the panel had to be reduced to a minimum,
representing the opening of the jaws of the grab (2.20 m). Three double T
precast elements were used, provided with steel profiles H and U attached at
the web of the precast elements, serving as guides and to improve the
watertightness.

129

1 panel; 2 guide wall; 3 trench filled with slurry


Fig. 5.35

1 joint panel; 2 field panel; 3 guide wall; 4 trench filled with


slurry; 5 self-hardening slurry
Fig. 5.36

130

Fig. 5.37
The phases of the construction of a diaphragm wall made of precast concrete
are:
- I - excavation of the trench; this is usually done under a bentonite
suspension but, with a carefully prepared slurry, can be done also directly
under a self-hardening slurry;
- I bis - when excavation is made under bentonite suspension, after
completing the excavation the bentonite suspension is replaced by the selfhardening slurry using the tremie pipes, as in the case of concreting;
- II - lowering into the self-hardening slurry of the precast elements.
When the excavation under the protection of the wall is performed, the
hardened slurry on the exposed face of the wall is trimmed.
The characteristic feature of the diaphragm walls made of precast concrete is
the use of self-hardening slurries.
A self-hardening slurry is a bentonite suspension in which a certain amount of
cement and additives are added. The self-hardening slurry is setting and then
hardening, like a plastic mortar, in the excavated trench, producing a firm
binding between the precast element and the surrounding soil and closing the
joints between the elements.
The receipt of the self-hardening slurry is established by laboratory tests.
Requirements are different in the case when the slurry is used also as a
supporting fluid in the excavation phase, as compared to the slurry which is
replacing a bentonite suspension.
The retarder additive should ensure the starting of the setting after the
lowering of the precast elements into the slurry. After that, the hardening
process should be quick enough, in order to ensure a good binding between
the precast elements and the soil.
Composite walls
The self-hardening slurry is widely used in the case of composite walls.
Indeed, in the lower part of a wall for which only the water tightness is
131

required, the reinforced concrete (cast in situ or precast) can be replaced by


the self-hardening slurry.

1 grab; 2 hose for sending the self-hardening slurry; 3 pump for


removing
the bentonite suspension; 4 plastic mortar; 5 concrete; 6 tremie
pipe; 7 reinforcement cage
Fig. 5.38
Fig. 5.38 shows the phases of the construction of a panel in a composite wall
with the bottom in an impervious layer:
a. excavation under bentonite suspension to the final level;
b. b replacing the bentonite suspension in the lower part of the trench
with the self-hardening slurry with higher density (1.201.25 g/cm 3);
c. excavating the self-hardening slurry, after hardening, on a depth of
1.0 m;
d. lowering the reinforcement cage in the upper zone of the panel and
concreting with a tremie pipe of the upper zone of the panel.
The use of embedded walls as retaining structures
There are two main methods in the construction of underground structures
with the use of embedded walls:
- the open excavation or the cut and cover method
- the top-down or Milanese method
Embedded walls for underground structures constructed in open excavation

132

Fig. 5.39
Fig. 5.39 shows the main stages for the construction of an underground
structure (for instance a subway gallery):
- construction of the embedded walls;
- first excavation phase of excavation;
- placement of the first row of struts;
- second phase of excavation;
- placement of the second row of struts;
- third phase of excavation;
- placement of the third row of struts;
- concreting the base of the gallery;
- after the hardening of the concrete in the base of the gallery,
dismantling the lower row of struts;
- concreting the walls and the slab of the gallery;
- after the hardening of the concrete in the gallery, dismantling the
intermediate row of struts;
- gradual filling of the space above the gallery, including the
dismantling of the upper row of struts.
Embedded walls for underground structures constructed with the top - down
method
This method is sometimes called Milanese method, since it was first applied
at the construction of the Milano subway. It consists in construction from the
beginning, near the ground surface, a reinforced concrete slab connected to
the embedded walls, in which holes are left for the access of people and
equipments and evacuation of the excavated soil. The excavation takes place
under this roof. The advantage of the method is the possibility to resume
activities at the ground surface (car traffic etc.) before completing the
133

underground structure. When the span between the embedded walls is large,
intermediate supports for the slab can be built, represented by steel columns
founded on barrettes, which are cast in situ reinforced concrete blocks,
constructed with the same technique as the cast in situ diaphragm walls.

1 embedded wall; 2 guide wall; 3 short trenche excavated under


slurry;4 steel column; 5 reinforcement cage; 6 foundation of the
steel column (barrette); 7 ballast; 8 reinforced concrete slab;
9 mat;10 wall; 11 intermediate slab; 12 column; 13 fill;
14 - pavement
Fig. 5.40
Fig. 5.40 shows the main stages of the construction of a subway station
constructed with the top-down method:
I construction of the embedded walls;
II construction of interior, shorter trenches, under bentonite suspension;
III lowering of the steel columns connected at their lower part with the
reinforcement cage of the barette;
IV concreting with the tremie pipe of the barrette and filling with gravel the
rest of the short trench;
V constructing the slab near the ground level;
VI excavating under the slab;
VII construction of the underground structure;
VIII placing the fill above the upper slab of the underground structure.
5.5 Design elements for braced excavations and for embedded walls
Supporting the sides of deep, narrow excavations made by timbering or sheet
piling with struts acting across the excavation is called bracing. A basic
134

problem for the design of a braced excavation is to establish the appropriate


diagram of earth pressure.
5.5.1 Earth pressure diagrams
In the chapter 8 it was shown that for the development of the active earth
pressure behind a wall is necessary for the wall to move away from the soil.
This is not the case of a braced excavation. When the first row of struts is
installed, the depth of excavation is small and no significant yielding of the
soil mass will have taken place. As the depth of the excavation increases,
yielding of the soil before strut installation becomes significant but the first
row of struts prevents yielding near the surface. Deformation of the wall will
be of the form shown in fig. 5.41, being negligible at the top and increasing
with depth. The deformation condition of the Rankine theory is not satisfied
and the theory cannot be used for this type of wall. Failure of the soil will take
place along a curved surface as shown in fig. 5.41. Only the lower part of the
soil wedge within this surface reaches a state of plastic equilibrium, the upper
part remaining in a state of elastic equilibrium. Based on numerous
measurements performed on bracing systems with multiple supports, the
following simplified conventional earth pressure diagrams for various types of
soils are recommended in practice:
for non-cohesive soils in fig. 5.42 a
for cohesive soils of low consistency in fig. 5.42 b
for cohesive soils of high consistency in fig. 5.42 c

Fig. 5.41

135

b
Fig. 5.42

5.5.2 Design of a timbering


The structural design of a timbering involves the following phases:
predimensioning the timbering by proposing dimensions for the
elements;
selecting the appropriate earth pressure diagram;
establishing bending moments, shear forces, axial loads in the
elements of the timbering;
checking the sections proposed for the timbering elements (boards,
walings, struts).
Some assumptions are usually taken in the design of a timbering:
a conventional earth pressure diagram, such as previously given, is
adopted;
the continuity of the boards and walings over the supports is
disregarded; boards and walings are treated as simply supported
elements.
In what follows is given, as an example, the analysis of a timbering with
vertical boards in a non-cohesive soil (fig. 5.43).

Fig. 5.43
136

The analysis of the vertical boards is done as for simply supported beams of
span l1, the distance between two consecutive walings or the vertical distance
between struts. For a width b of the board, the load per unit length on the
board is:
q p aH b 0,65 H tan 2 ( 45

)b
2

(5.1)
The bending moment in the board is:
M

q l12
8

(5.2)

To check for the board section:


q l12
M

8 all
W bd2
6

(5.3)
where all is the allowable strength for the material in the board and d is the
thickness of the board, which will result from the relation (5.3).
The analysis of the walings is done as for simply supported beams of span l 2,
the horizontal distance between struts. The waling takes the reaction from
boards pertinent to a field l1. The load per unit length of the waling is:
q1 p aH l1

(5.4)

The maximum bending moment for the waling is:


q l2
M1 1 2
8

(5.5)

If e is the known width of the waling and f is the thickness to be determined:

M1
M1

all
W1 ef 2
6

(5.6)
From the relation (5.6) is determined f.
137

The analysis of struts is done as for elements subjected to compression, with


due consideration for the buckling.
For the most loaded struts of the timbering in the fig. 5.43, the compression
load is:
N p aH l1 l 2

(5.7)
To check the struts in compression the relation (5.8) is used:
II

N
w all II
A

(5.8)

where A is the section of the strut;


w is the buckling factor;
all II is the allowable strength of the timber in compression parallel to
the timber fibres.
To check the struts in crushing normal to the fibres at the strut-waling contact:

where
timber

all

N
all
A

(5.9)

is the allowable strength in compression normal to the fibres of the

Usually, timbering are using fir or pine, for which the following values of the
allowable strengths can be used:
- for bending all = 120 daN/cm2
- for compression along the fibres ac II = 120 daN/cm2
- for compression normal to the fibres ac = 18 daN/cm2
If instead timber, steel elements are used, in the design relations previously
given, appropriate values for all and w should be introduced.
5.5.3 Analysis of sheet pile walls and embedded walls
Sheet pile walls and embedded walls can be grouped, based on the criterion
of the statical system, in two categories:
- walls forming statically determined systems (cantilever walls;
anchored or propped walls with one level of anchor or prop in the
upper part and free earth support at the bottom);
- walls forming statically undetermined systems (walls with one level of
anchor or prop in the upper part and fixed earth support at the
138

bottom; walls with two or more levels of anchor or prop in the upper
part and free or fixed earth support at the bottom).
The analysis of the sheet pile walls and embedded walls has two objectives:
- to determine the depth of penetration of the wall in the soil, taking
into consideration various failure modes;
- to determine the structural design of the wall to resist bending
moments, shear forces and prop or anchor forces derived from
equilibrium calculations.
In fig. 5.44 are shown various kinds of failure for these walls.
In what follows, several types of walls pertaining to the two categories will be
considered.

Fig. 5.44
a. Cantilever walls
These walls are free at the upper part and derive their equilibrium from the
lower, embedded part of wall. Two situations can occur:
- wall acted upon by a horizontal force H;
- wall retaining soil of a height h.
In the first case, the load can be, for instance, the resultant of the pressure
exerted by water on the free upper part of the wall (fig. 9.45 a). Subjected to
the force H, piles bends and rotates. If the deformations by bending are
disregarded, the wall can be treated as an infinitely stiff plate rotating around
a point O (fig. 9.45 b). On the front face of the wall, above the point O, the
139

wall induces a compression on the soil and conditions for developing a


passive resistance are present, while below the point O on the same face the
wall moves away from the soil which is relaxing and the active earth pressure
develop. On the face behind the wall, the situation is opposite.
By neglecting the friction between wall and soil, the pressures diagrams (fig.
5.45 c) have, according to Rankine, the slopes K p and K a where:

K p tan 2 (45 o );
2

K a tan 2 ( 45 o )
2

Fig. 5.45

Fig. 5.46
By computing along the embedment t the difference between the passive and
active earth pressure, a resultant diagram is obtained (fig. 5.46 a), with
ordinates limited on both faces by two lines of (K p K a ) inclination.
The diagram thus obtained is physically not possible, since there are two
pressures in the same point O. In reality, the transition from the left to the
right part of the pressure diagram cannot be done by a jump like in the fig.
9.45.a, but gradually, along a curve passing through the point O (fig. 5.46 b)
and which can be approximated by a line (fig. 5.46 c).
140

The stability of the wall is insured by the couple of forces E p and E 'p ,
representing the resultants of passive pressures developed on the two sides
of the wall is in equilibrium with the overturning moment produced by the
force H.
In order to find the embedment t required for insuring the stability, two
methods can be used:
- in the first method, the final diagram in fig. 5.45 c is used, involving
three unknowns: t, d and e, for which only two equilibrium equations
are available, the horizontal forces equation and the moment
equation. A trial and error approach is then used.
- an embedment depth t is proposed;
- the two equilibrium equations are written:
X0
(5.10 a)
Mc 0
(5.10 b)
Equation (5.10 a) represents the horizontal projections of horizontal forces,
while equation (5.10 b) expresses that the moment about point C is zero.
- the system of equations is solved, to find the unknowns d and e;
- the computation is repeated for a new value of t, until the following
condition is fulfilled:
e

f
FS

(5.11)

where FS is a factor of safety which can be taken 1,52.


In the second method, the diagram in fig. (9.46 c) is replaced by the
simplified diagram in fig. (9.47 a), where the passive resistance on the
posterior face of the wall was replaced by the unknown force E p. There are
two unknowns, t and Ep, which can be found with the equilibrium equations. In
fact, only t is of interest, which is obtained by taking the moment about the
point D.
MD = 0
H (t t o )

t
1
( K p K a ) t o2 o 0
2
3

(5.12)
Equation (5.12) leads to the following equation:
t3
o 6

H
H
to 6
h 0
(K p K a )
(K p K a )

(5.13)
141

The solution of eq (9.13) is obtained by trials in order to obtain t o. For the first
trial, values of the ratio to/h in function of given in the table 9.1 can be
used.
Table 5.1

to/h

20
1.6

25
1.2

30
0.9

35
0.7

40
0.5

The embedment depth should be larger than to:


t = (1.20.1.25)to

(5.13)

In order to find the maximum bending moment, the pressure diagram used for
the computation of t is taken (fig. 5.45 c or fig. 5.46 a) and the depth zo at
which the shear force is zero, corresponding to M max, is found. For instance,
when the diagram in fig. (5.47 a) is applied:
H

zo

1 2
z o (K p K a ) 0
2

(5.14)

2H
(K p K a )

M max H ( h z o )

(5.15)
1
( K p K a ) z o2
2

(5.16)

Fig. 5.47
In the case of the cantilever wall retaining a soil of height h, the approach is
similar as in the previous case, with the difference that the horizontal action is
produced by the active earth pressure (fig. 5.48). The corresponding
diagrams are given in fig. 5.49 a, fig. 5.49 b and fig. 5.49 c. The bending
moment diagram is shown in fig. 5.50.
142

Fig. 5.48

Fig. 5.49

Fig. 5.50
b. Walls with one level of anchor or prop in the upper part and free earth
support at the bottom
When the cantilever wall cannot take the loads or when the ground conditions
do not allow to obtain the embedment required by the fixed-end condition, an
additional support is provided in the upper part of the wall by means of an
anchor (fig. 5.51 a) or prop (fig. 5.51 b). The second support is represented
by the embedment depth t, as a free earth support.
143

Fig. 5.51

Fig. 5.52
It is assumed that by the elastic displacement of the prop or anchor and by
the rotation of the wall around the base, conditions for development of
passive earth pressure in front of the wall are met. The two pressure
diagrams can be drawn (fig. 5.52). The problem is statically determined, there
are two unknowns (the embedment depth and the load in the strut or anchor
RA) and two available equilibrium equations ( X 0, M A 0) . A factor of
safety FS of 1.52 can be introduced by dividing the passive pressure force
to FS.

144

Fig. 5.53
Field measurements have shown that the real active earth pressure diagram
is different from the diagram with linear distribution resulting from the
Rankines theory. As a result of the bending of the wall and of an arching
effect, a redistribution of the pressures is occurring, leading to a overloading
of the supports and to a discharge of the field (fig. 5.53). To take into account
this effect, a coefficient of reduction of the maximum bending moment
corresponding to the triangular diagram is introduced. The German Code
EAU 77 recommends to take 0.67 Mmax.
c. Walls with one level of support in the upper part and fixed earth
support at the bottom
Unlike the previous case, the point of rotation is located above the base, the
wall changes in curvature within the embedment depth, leading to the
development of passive resistance on both sides of the penetration depth (fig.
5.54). The pressure diagram is constructed as in the case of the cantilever
wall (fig. 5.43 c). The theoretical diagram can be replaced by the design
diagram in fig. 9.55 and the embedment depth t by the reduced depth to.

Fig. 5.54

145

Fig. 5.55
An approximate method which can be used in this case is named the
method of replacing beam. It is considered known the depth y of the inflexion
point of the deformed shape of the wall. Thus, for 20 o , y = 0.25 h and for
30 o , y = 0.08 h, where h is the height above the excavation level. Taking
the inflexion point as a hinge, the wall is divided into two simply supported
beams: BC and CD. The equilibrium equation for the upper beam BC leads to
the reaction at the level of the strut or anchor R A and the reaction in the
support C, RC. To obtain the value of to, the condition of moment O in the point
D is written for the lower beam CD. Then, t is taken as: (1.201.25)to.
5.5.4 Solutions for the support in the upper part of embedded walls
From the point of view of the support in the upper part, embedded walls can
be classified in two categories:
propped walls, where the support is provided by struts (fig. 5.51 b)
anchored walls, where the support is provided by several means,
such as:
- tendons or tie rods transferring the load to a steel or reinforced
concrete plate or to a concrete block called deadman (fig. 5.51 a);
- tendons or ties transferring the load to a group of raking piles, from
which one pile works in tension T, the other in compression C (fig.
5.56 a);
- anchor piles (fig. 5.56 b);
- ground anchors (fig. 5.56 c).
The deadman may be formed of isolated elements (fig. 9.57a) or by a
continuous plate (fig. 9.57b).

Fig. 5.56

146

Fig. 5.57
The stability of the deadman is insured by the passive resistance of the soil in
front of the plate or block (fig. 9.58). The stability condition is:
R A Pa

Pp
FS

(5.17)
where Pa and Pp are the active and passive forces and F S a factor of safety
which can be taken 1.5.
Pa

1
1
d 2 l K a e d l K a d l K a ( d e);
2
2

Pp

1
1
d 2 l K p e d l K p d l K p ( d e);
2
2

(5.18)

(5.19)
where l is the distance between tendons;
d height of the deadman;
e distance from the upper edge of the deadman to the ground level.
Relations (5.18) and (5.19) are used in the case of continuous plates and in
the case of isolated plates of width b, when l 2 b , and for e 4 d .
As for the position of the deadman, the following procedure is used (fig. 9.58):
from the bottom of the wall (in the case of free earth support) or from the point
of inflexion at the depth y (in the case of fixed earth support) the failure plane

inclined with the angle ( 45 o 2 ) in respect to the horizontal. Between this


plane and the wall the prism I is formed, where is strictly forbidden to place
the deadman. At the same time, the passive zone in front of the deadman
147

should not interfere with the active soil zone behind the retaining wall.

Therefore, from the point C a failure plane with inclination of ( 45 o 2 ) in


respect to the horizontal is drawn, defining the prism II.
The deadman should be located above this plane.

Fig. 5.58
5.6 Ground anchors
A ground anchor is an installation capable of transmitting an applied tensile
load to a load bearing stratum.
Ground anchors can be permanent, when required to ensure the stability and
satisfactory service performance of the permanent structure or excavation
being supported or temporary, when used during the construction phase of a
project to whitstand forces for a known short period of time, usually less than
2 years.
Fig. 5.59 shows several examples of permanent structures using ground
anchors: a retaining wall (fig. 5.59 a), a high rise building with a multiple-level
basement, founded below the water table and subjected to uplift forces of the
water (fig. 5.59 b), a tower for an electrical line (fig. 5.59 c).

148

Fig. 5.59
Examples of temporary ground anchors are given in the fig. 5.60.

Fig. 5.60
A ground anchor consists basically of an anchor head, free anchor length and
fixed anchor.
Fig. 5.61 shows a typical ground anchor.

Fig. 5.61
The main phases in the construction of a ground anchor are:
1 boring of a borehole
Special equipments are used, able to bore holes at any inclination; the kind of
boring (cased or uncased; under a slurry etc.) as well as the boring tool are
adapted in function of the ground conditions.
2 introduction in the borehole of a tendon
149

Tendons usually consist of steel bar, strand or wire, either singly or in groups.
3 grouting
The most common grouts used for ground anchors are cemetious grouts
which usually are water/cement mixes, with a ratio lying in the range of 0.35
to 0.60. Admixtures to improve the properties of the grout, such as the
workability or durability, or to increase the rate of strength development are
sometimes used.
The main functions of the groutings are:
- to form the fixed anchor length (L fixed), which is the designed length of
the anchorage over which the tensile load is capable of being
transmitted to the surrounding ground;
- to reinforce the ground in the close vicinity of the fixed length, in
order to increase the capacity of the anchor;
- to increase the watertightness of the ground in the close vicinity of
the fixed length in order to reduce the losses of the grout.
Grouting is usually made through a grouting tube; the separation between the
fixed anchor length and the free anchor length is made by using a device
called packer, which is an expandable rubber ring surrounding the grouting
tube.
4 stressing
Stressing is required to fulfill the following two functions:
- to ascertain and record the load carrying behavior of the anchor;
- to tension the tendon and to anchor it at its lock-off load.
Stressing equipment is similar to the one used for prestressed concrete
elements.
Stressing should not be carried out until a sufficient hardening of the grout in
the fixed length has been achieved, which normally requires seven days.
The component of the ground anchor which transmits the tensile load from
tendon to bearing plate or structure is called anchor head.
A special problem is the corrosion protection.
All steel components which are stressed (tendon bond length, tendon free
length, anchor head) shall be protected against corrosion for their design life
(less than two years for temporary ground anchors, more than two years for
permanent ground anchors).
150

The load capacity of ground anchors depends primarily on the ground in


which the fixed anchor length is located. Common values are of 1000.2000
kN in sands and gravels and 200300 kN in cohesive soils. In rocks, the
load capacity can reach much larger values, of 5000.10.000 kN.
The load capacity of the ground anchors can be estimated by computation,
but it is compulsory to be then checked by field tests performed in advance of
the working anchors. Stressing each ground anchor represents by itself a
test.
Due to the special character of the work, the construction of the ground
anchors shall be assigned only to contractors with proven experience in this
field.
5.7 Retaining walls
Retaining walls are permanent retaining structures used along the roads or
railroads in hilly or mountain areas, along navigation canals, behind buildings
on a slope etc. Their role is to support the soil placed behind them, thus
enabling a transition on a very short distance between two levels, when it is
not possible to insert a slope between the two levels.
In the past, natural stones were used in the construction of retaining walls. At
present, concrete and reinforced concrete are materials most used for these
retaining structures.
There is a great variety of retaining walls. In the following, some of the most
used types of retaining walls will be presented.
5.7.1 Gravity walls
5.7.1.1 Mass concrete retaining walls
Mass concrete walls are suitable for small retained heights, usually up to 3
4 m.
Fig. 5.62 shows a section through a gravity retaining wall made of concrete
and the forces which are acting:
- active earth thrust Pa on the back of the wall;
- passive earth force Pp on the face of the wall, below the ground level;
- weight G of the wall;
- reaction R on the base.

151

Fig. 5.62
As a rule, the passive force Pp, whose development requires, as shown in the
chapter 8, large displacements, is disregarded.
In the fig. 5.62 are also given recommendations for a preliminary selection of
the dimensions of the wall.
In fig. 5.63 are shown other forms of mass concrete walls.

Fig. 5.63
A filter of coarse permeable material is desirable behind a retaining wall to
prevent the development of high pore water pressures within the backfill. To
allow the water to percolating into the filter to drain out, weep-holes are
provided in the wall (fig. 5.64).

152

Fig. 5.64
The design of the wall has three aspects:
- structural design
Normally mass concrete walls should be designed on a no-tension basis
under the design earth pressure. At least two sections should be checked (fig.
9.62):

at the middle of the elevation AB

at the joint BC between the elevation and the foundation.


- foundation design
The pressure on the soil under the base of the wall should be checked to
ensure that it does not exceed the allowable bearing pressure p all.

p max
min

N M N
M
N
6e

(1 )
A W B 1 1 B 2 B
B
6

(5.20)
where N is the total normal force on the base and e =

M
N

N = Pav + G

(5.21)

Pav being the vertical component of the active earth thrust.


Three conditions shall be fulfilled:
153

p med p all

(5.22)

p max 1.2 p all

(5.23)
p min 0

(5.24)

Relation (9.24) expresses the condition that the resultant of forces P a and W
is located within the middle third of the base ( e B 6 ).
- stability checks
Stability against sliding
The friction force S on the base (the component along the base of the
reaction R) is compared with the horizontal component P ah of the active earth
thrust. The friction force S is equal to the normal force N multiplied by the
friction coefficient between the base and the ground.
Pah m l N

(5.25)
where ml is a factor of safety equal to 0.8.
When values for obtained by field tests are lacking, values given in tab. 5.2
can be used, at least for a preliminary design.

Tab. 5.2 Values of the friction coefficient


Ground type

Clays:
- Ic< 0.75
0.25
- Ic 0.75
0.30
Sandy clays and clayey
0.30
sands
Fine, silty sands
0.40
Coarse sand, gravels
0.50
Rocks
0.60

154

When the base resistance to sliding is inadequate, this can be increased by


either widening the base, inclining the foundation (fig. 5.65 a) or providing a
shear key (fig. 5.65 b).
Stability against overturning
The following condition shall be fulfilled:
M r m r MS

(5.26)

where Mr is the overturning moment


Mr = P a a

(5.27)

a being the arm of the force Pa in respect to the toe of the wall D
MS is the stability moment
MS = G d

(5.28)

d being the arm of the force G in respect to the point D


mr is a factor of safety equal to 0.8.

Fig. 5.65
5.7.1.2 Gabions
Gabions are large cages or baskets usually of steel wire or square welded
mesh, rectangular in shape, filled with stone and used as gravity retaining
walls or anti-erosion works (fig. 5.66).

155

The permeability and flexibility of gabions make them suitable where the
retained material is likely to be saturated and the bearing quality of the soil is
poor.

Fig. 5.66
Gabion walls are designed on the same principle as a gravity mass wall. In
fig. 5.67 are given examples of gabion retaining walls. The density of the
stone fill can be taken as 60% of the solid material.

Fig. 5.67
5.7.1.3 Crib walls
Crib walls are built of individual units assembled to create a series of box-like
structures containing granular free draining fill, to form a gravity retaining wall.
156

The units should be so spaced, that the fill material contained within the crib
is not affected by climatic changes and acts in conjunction with the crib work
to support the retained earth.
The individual units can be made of timber or of precast concrete.
In fig. 5.68 is shown an example of assembling the individual units. In fig.
5.69 is given a vertical section on a crib wall.

Fig. 5.68

Fig. 5.69
This kind of gravity walls is indicated for use on compressible soils having the
ability to adapt to differential settlements.
5.7.2 Reinforced concrete walls
5.7.2.1 Cantilever walls
157

Cantilever or T walls are made of a vertical or inclined slab monolithic with a


slab base (fig. 5.70).

Fig. 5.70
The advantage of this wall is the use of the weight of the retained material,
resting on the base slab, together with the weight of the wall, in order to
ensure the stability against sliding and against overturning. Various structural
elements such as the slabs AB, BC and DE, working as cantilevers, are
designed to resist bending.
For the stability checks of a cantilever wall, the active earth pressure should
be defined. There are two approaches to compute the earth pressure:
- on the polygonal surface AFCD, in which FC represents the failure
surface with an angle
-

45 o

in respect to the horizontal. In this

case, the soil prism FBC is considered as part of the wall,


on the vertical plane CH. In this case, the soil prism AHCB is
considered as part of the wall.

For heights up to about 8 m, a cantilever wall is generally economic. For


greater heights a counterfort wall is more appropriate.
5.7.2.2 Counterfort walls
Counterfort walls are made of a vertical or inclined slab supported by
counterforts monolithic with the back of the wall slab and base slab (fig. 9.71).

158

Fig. 5.71
5.8 Reinforced earth
The idea of using the soil itself for structures aimed to resist to the active
earth pressure, appeared in the construction of crib walls, gabions, cantilever
or counterfort walls, is best expressed by the reinforced earth, a system
patented by the French engineer Henri Vidal and introduced in practice in the
years 60s of the 20th century.
A compacted soil mass is stabilized as a result of frictional forces developed
between the soil and the tensile reinforcing elements, usually in the form of
horizontal strips made preferably of galvanized steel, but also of aluminium
alloys, plastic or geotextiles. The stresses within the soil mass are transferred
to the elements which are thereby placed in tension. The soil used as the fill
material should be predominately coarse-grained and be adequately drained
to prevent it from becoming saturated.
A reinforced earth retaining structure should be made in such a way that
several basic conditions are fulfilled: the facing should resist to the earth
pressure and be sufficiently flexible to withstand any deformation of the fill;
the length L of the strips should be long enough in order to ensure, by the
friction developed on the top and bottom surface of each strip in contact with
the soil, the stability of the structure; reinforcing elements should be able to
carry out the tensile stresses.
The facing is attached to the reinforcing elements to prevent the soil from
spilling out and to satisfy aesthetic requirements. Two of the most used
solutions are the facing consisting of precast concrete units (fig. 9.72 a) or of
pliant U shaped steel sections arranged horizontally (fig. 9.72 b).
In what follows, a simplified method of analysis is presented.
159

b
Fig. 5.72

For given d (distance between strips in the horizontal plane) and H


(distance between strips in the vertical plane), the total earth thrust on the
facing, pertaining to a strip at the depth z, is:
P K z d H

(5.29)
where K is the appropriate earth pressure coefficient at depth z.
The frictional resistance available on the surfaces of the element is given by:
F 2 L B Y z tan

(5.30)
where L is the length of the element, b is the width of the element, the
angle of friction between soil and element.
In order to obtain the minimum length of the element, L min, which is the length
of the element beyond the failure surface AC (fig. 9.72), P should be
multiplied by a factor of safety FS which should not be less than 2.
FS P F
FS K z d H 2 L min B z tan
F K d H
L min S
2 b tan

(5.31)
From the relation (9.31) results that Lmin is constant and independent of z.
160

The thickness of the element is obtained knowing the tensile force P and the
width b, from the condition of tensile resistance of the material. The thickness
of the facing is determined in function of P, H , d and the strength
characteristics of the material used for facing.
The anchoring length Lmin should be ensured outside the failure surface AC,
where the shear stresses on the surfaces of the element are acting in wards
(fig. 9.73). Sometimes, in order to facilitate the construction process, instead
of a linear variation with depth of the length, a step-wise variation can be
adopted (fig. 9.73 b).

Fig. 5.73
When the fill is made of granular material, displacements of the wall are
sufficient to develop the active limit state condition and K in relations (5.29)
and (5.31) can be taken as Ka.
For a preliminary design, the following values can be used: d = 0.7 m; H =
0.250.30 m; b = 75 mm.
In the design of the earth reinforced earth structure, the external stability must
also be considered.
Although behaving as a relatively flexible structure, a reinforced earth
structure should be designed, from the point of view of external stability, as if
it were a gravity wall. The back of the wall should be taken as the vertical
plane through the inner end of the lowest reinforcing element. The total active
thrust on this plane is calculated by the Rankine theory. The factor of safety
against sliding between the reinforced fill and the foundation soil should not
be less than 2. The pressure distribution on the base must be wholly
compressive and fulfill conditions (5.225.24) as for a gravity wall.
161

A basic problem for a permanent structure made of reinforced earth is the


durability of reinforcing elements. Data available on the rate of corrosion of
galvanized steel in soils indicate that elements of this material are likely to
have a minimum service life of 120 years.
5.9 Cofferdams
Many engineering works must be constructed in rivers or in still water or in
the areas prone to flood along rivers. This is the case of bridge foundations,
of docks, locks and other hydraulic structures etc. To make possible the
construction in the dry of the respective object, this should be surrounded by
a temporary structure, creating an area from which the water is removed. The
temporary structure is called cofferdam.
5.9.1 Cofferdams made of earth and rock fill dikes
5.9.1.1 Earth dikes
Earth dikes as cofferdams are used for shallow waters (23 m) and rates of
the flow of the river under 0.5 m/s. Due to the relatively large transversal
area, required by the slopes, the earth dikes lead to a significant narrowing of
the river section, hence they are rarely used for the bridge piers located in the
river. Instead, they can be used for bridge abutments, for which the needed
area for work in the dry is obtained by linking the dike with the bank of the
river (fig. 5.74).

Fig. 5.74
The soil for the construction of the earth dike should fulfill some requirements:
to ensure watertightness, to be compactable, to avoid being easily eroded by
the water flow. Clayey sands with about 25% clay fraction meet these
requirements. But even dikes made of sand could be a reliable solution,
counting on a watertightness reached in short time, as a result of the filling of
the voids by the fine particles carried by the water. When the dike is
constructed on the land or in dry in periods of low waters, clay can be also
162

used, if excavated in dry form, spread and compacted in thin layers. The clay
fill should not be used for the construction of dikes by discharge under water,
since the soil is softening and forming an unstable fill.
At flow rates larger than 0.1 m/s, slopes of the dikes should be protected to
prevent erosion.
5.9.1.2 Rock fill dikes
Rock fill dikes, made of large blocks of stone, have the advantage of steeper
slopes in comparison with the earth dikes and of resisting better to erosion
forces caused by the water flow. Instead, they are permeable and require
special measures to ensure watertightness, such as concrete cores (fig. 5.75
a) or clay cores (fig. 5.75 b), when the construction of the dike is made in dry,
or sheet pile walls (fig. 5.75 c), when dike is built in water. In all cases, the
watertight element should penetrate into an impervious layer of soil.

Fig. 5.75
5.9.2 Sheet piles cofferdams
5.9.2.1 Single skin cofferdams
These are formed by one-line timber or steel piles, working as cantilever walls
(fig. 5.76 a) or as walls with free earth support and raking struts (fig. 5.76 b),
when ground conditions do not allow to drive the sheet piles to a depth
required in a fixed-end support.

Fig. 5.76
163

Other means to ensure the stability of the single skin cofferdams are
represented by earth fills placed in front of the wall (fig. 5.77 a) or on both
sides of the wall (fig. 5.77 b).

Fig. 5.77
5.9.2.2 Earth-filled double-wall cofferdams
Double-wall cofferdams consist of two parallel lines of steel piling connected
together by a system of steel walings and tie rods. The space between the
lines of piling is filled with coarse cohesionless materials such as sand, gravel
or broken rock.
The width b of the cofferdam should be not less than 0.8 of the retained
height h of water (fig. 5.78). The penetration of the piling into the soil below
the bottom of the river should be sufficient to develop the necessary passive
resistance and prevent horizontal sliding of the cofferdam as a gravity
structure.

Fig. 5.78
A major disadvantage of double-wall cofferdams is that if failure takes place in
a certain point, it will propagate on a relatively large distance along the wall.
5.9.2.3 Cellular cofferdams
164

Cellular cofferdams are self-supporting structures, constructed using straight


web steel sheet piles driven to form cells of various shapes and filled with
sand, gravel or broken rock. They can be founded on rock, sand or stiff clay
and utilized as either temporary or permanent structures to retain
considerable heights of soil and/or water.
The stability of a cellular cofferdam depends upon the tensile strength of the
sheet piling, the properties of the filling, the shape and size of the cells and
the foundation materials. The outward pressure of the filling produce a high
circumferential tensile forces in the piling, which the straight web piles are
designed to resist, unlike through shaped piles sections which are unsuitable.

Fig. 5.79
In fig. 5.79a is shown one of the most utilized type of cellular cofferdam, with
circular diaphragm cells. The main advantage of this type is that each cell is a
self-supporting unit, which means that the loss of stability of one cell does not
imply the progressive failure of the entire wall. Also, each cell can be filled
independently of adjacent cells. Circular diaphragm cells can be constructed
in rough and flowing water of maximum velocity about 1.3 m/s, and at large
depths of water, reaching 2025 m.
A good example of the use in Romania of this kind of cellular cofferdam is
represented by the works at the navigation and hydro energetic system at the
Iron Gates, on the Danube (fig. 5.80).

165

Fig. 5.80

166

Potrebbero piacerti anche