You are on page 1of 122

IACINT MANOLIU

SOIL MECHANICS I

LECTURE NOTES

CHAPTER 3:

THE CONTROL OF GROUNDWATER


The occurrence of groundwater on site is considered as one of the most difficult problem in excavation work. Reduction of groundwater within an excavation is needed for access by workers and machines and for performing construction works in dry conditions. Methods of control of groundwater in excavations can be grouped in two main categories: Methods of dewatering by pumping water Methods of restraint of flow In this chapter, methods of the first category will be dealt with. 3.1 METHODS OF DEWATERING

3.1.1 SUMP PUMPING


Water accumulated in the excavation is collected in a perimeter trench and pumped out of the excavation for discharge. In the case of unsupported excavation (fig. 3.1) or of supported excavation with a pervious wall, water access is possible both through the sides and through the bottom. When the sides of the excavation are protected by impervious walls, for instance a closed sheet piling braced for lateral support (fig. 3.2a), water access is possible only through the bottom, which is provided with slight inclinations and trenches to conduct the water to one or several sumps (fig. 3.2b). To prevent piping, the sump can be protected by an inverted filter (fig. 3.3).

Fig. 3.1

1 sheet pile wall; 2 sump; 3 - pump Fig. 3.2


79

The greatest risk associated with the sump pumping is the occurrence of the quick sand condition (or piping, boiling). In the case of sheet piling, upward seepage of groundwater into the excavation can produce soil liquefaction, when the vertical effective stress within the soil reaches zero (fig. 3.4). When the quick sand condition is occurring, the pumping must immediately cease. To reach the required level of the excavation, one of the following solutions can be adopted: - recharging water into excavation to the original level and, then, performing the excavation and the concreting under water; - lengthening the seepage path by driving the sheet piles to a deeper penetration; - reducing the head of the water causing seepage, by pumping from wellpoints or bored wells placed at or below the level of the bottom of sheet piles (see p. 3.1.2).

1 sump; 2 inverted filter Fig. 3.3

Both silty or sandy clays, due to their low permeability, and coarse sands and gravels, due to the large dimension of both particles and voids, are unlikely to be subjected to piping. Soils most likely to be subjected to piping are soils with no cohesion or very low cohesion, with grain size small enough to be disturbed by the seepage forces, and permeable enough to allow seepage through them. Loose fine sands, silty sands and sandy silts meet these requirements.

Fig. 3.4

If the depth of the excavation is not too big, pumps are placed at the ground surface, as in fig. 3.1. The aspiration height or suction lift is limited to 67 m.
80

In order to determine the total discharge, the following relation can be used: Q = q. A (m3/h) (3.1)

where A is the surface of the bottom of the excavation and q is a specific discharge which can be taken: 0.16 for fine sands; 0.24 for medium sands; 2.00 for cease sands. When the excavation is supported by sheet piles (fig. 3.2) the following relation can be used:
Q=qHk U

(m3/h)

(3.2)

where H is the head, in m; k coefficient of permeability in m/h; U the perimeter of the wall; q is the specific discharge given in the table 3.1 in function of the ratio (H+t)/l and H/(H+t) where t is the embedment and l the distance from the ground water table to the impervious layer. Table 3.1
H H+t

H+t i

0,10 0,20 0,30 0,40 0,50 0,60 0,70 0,80 0,90 0,95 1,39 1,20 1,12 1,08 1,02 1,13 0,95 0,89 0,84 0,80 0,98 0,81 0,74 0,70 0,67 0,88 0,70 0,64 0,60 0,58 0,78 0,61 0,56 0,52 0,50 0,70 0,53 0,48 0,45 0,42 0,61 0,46 0,41 0,39 0,38 0,52 0,39 0,34 0,32 0,31 0,42 0,30 0,27 0,25 0,24 0,36 0,23 0,22 0,21 0,20

1,00 0,75 0,50 0,25 0,00

3.1.2 GROUNDWATER LOWERING


3.1.2.1 General conditions The main methods used for groundwater lowering are the wellpoint systems and the bored wells. The scheme for the groundwater lowering is illustrated in fig. 3.5.

Fig. 3.5
81

Wellpoints or bored wells systems are installed prior to excavation inside or outside the excavation area. Once in function, they will cause groundwater to flow away from the excavation, improving the stability to its side batters and base and allowing construction works to proceed in the dry. Another use of groundwater lowering is illustrated in fig. 3.6, and is intended to prevent the hydraulic failure of the base of the excavation. Ground conditions on the site are characterized by the presence of two water layers, the upper one with free level and the lower one under pressure, separated by a layer of impervious soil, a clay. When the excavation reaches the clay layer, this one is subjected to a pressure w H , corresponding to the difference in the elevations of the two water layers. If the thickness h of the clay layer below the base of the excavation is not sufficient, there will be a heave of the layer of clay followed by its rupture under the pressure w H . The phenomenon is called hydraulic failure of the base of the excavation. In order to prevent it, the lowering of the groundwater is necessary.

1 excavation; 2 impervious layer; 3 pervious layer Fig. 3.6

In fig. 3.7 is put into evidence the advantage of using the groundwater lowering with bored wells or well points, as compared to the sump pumping, in the case of an excavation with a vertical soil support. When the sump pumping is used (fig. 3.7a), the wall should be impervious and to resist both the active earth pressure and the water pressure. There is a risk of piping, particularly in soils such as fine sands, or silty sands or sands silty. When the groundwater lowering is done (fig. 3.7b), water is drawn away from the excavation and, being filtered as it is removed from the ground, carries little or no soil particles with it once steady discharge conditions have been attained. At the same time, the wall is subjected only to the active earth pressure and should not be impervious. There is, however, a shortcoming of this method, too, namely the occurrence of settlement due to an increase in density as result of the ground lowering of the water table. Indeed, in a point A adjacent to the dewatering system (fig. 3.5), the effective overburden pressure before lowering the ground water table is:
82

peff = 'b + sat a

(3.3)

After ground water lowering, peff becomes:


p'eff = sat ( a + b )

(3.4)

The increase in pressure is:


p'eff peff = sat a + sat b ' b sat a = sat ' b = w b

(3.5)

Fig. 3.7

3.1.2.2 Bored wells A 4060 cm diameter borehole is performed, under the protection of a casing, until the impervious layer is reached or, if this is not possible, deep enough below the bottom of the excavation. Inside the casing a tube of 1530 cm diameter is inserted, provided with a perforated screen over the length where dewatering of the soil is required and it terminates in a 23 m length of unperforated pipe to act as a sump to collect any fine material which may be drawn through the filter. After the well casing is installed, graded gravel filter material is placed between it and the outer borehole casing over the length to be dewatered. The outer casing is withdrawn in stages as the filter material is placed and the remaining space above the screen is backfilled with any available material. Pumping from bored wells can be undertaken by surface pumps, with their suction pipes installed in bored wells. The depth of draw-down in this case is maximum 8 m. When a great depth of water lowering is required or when an artesian head must be lowered in permeable strata at a considerable depth below excavation level (see fig. 3.6), electrically powered submersible pumps, with a rising main to the surface. In the
83

case of submersible pump there is no limitation on amount of draw-down as there is for suction pumping. A pump in a 350 mm borehole can raise 7500 l/min. against 30 m head. In fig. 3.8 is shown a complete installation of a bored well.

1 inner casing; 2 rising main; 3 submersible pump; 4 silt, collecting in sump; 5 mesh filter screen; 6 outer well casing withdrawn; 7 soil backfill; 8 graded filter material Fig. 3.8

The design of a dewatering system using bored wells is based on relations established in hydraulics of underground works, of the kind shown in the chapter 3, in connection with the determination of the coefficient of permeability k by a wellpumping test. As it was shown, when the well reaches the impervious layer (see ch. 2), the equation of the draw-down curve is:
z2 h 2 = q x ln k r

(3.6)

At a distance R from the axis of the wells named radius of influence, there is no effect of the dewatering, the draw-down curve meets the original groundwater level. By replacing x = R, z = H in the relation (3.3):
q= k R ln r

( H2 h 2 )
84

(3.7)

If the draw-down at the well face is so: h = H - so so = H h Replacing (3.5) in (3.4):


q= k ( 2H so ) so R ln r

(3.8)

(3.9)

Equation (3.6) gives the relation between the discharge q of water pumped from the bored well and the resulting lowering so. By experiments, it was found the following relation for R: R = 3000 so k1/2 where so is in m and k in m/s. For an individual well of radius rw, the discharge quantity given by Darcys law is:
qi = 2 rw h w k ie

(3.10)

(3.11)

where hw is the height of well screen and ie is the average entry gradient. According to empirical findings, ie should not exceed 1/(15 k1/2) to avoid turbulence and filter unstability. Thus, the capacity of an individual well should be limited to:
q max = 2 rw h w k 15 K , m3 / s

(3.12)

1 original ground water level; 2 lowered groundwater level; 3 bored well; 4 fictitious well Fig. 3.9
85

The relation (3.12), established for the pumping from a bored well was extended to the case of pumping simultaneously from several wells, assuming that these wells are
BL located around a perimeter of a circle of radius R1, where R1 = , B and L being the length and the width of the rectangular excavation. It is also assumed that the cumulated effect of the wells is equal to the one of a fictitious wells of radius R1 and having the same radius of influence R as the single well (fig. 3.9). By this way is obtained the total flow Q needed for the required lowering so:
Q= K ( 2H so ) so R ln R1
1/ 2

(3.13)

The number of wells is n, where


n= Q q max

(3.14)

The required pump capacity is:


N= Qh w

(3.15)

where w is the density of water to be pumped, and is the system efficiency which, considering friction loss in the delivery pipe work is usually in the range 0.3 to 0.5. For = 0.3 and w =10kN / m3 ,
N= Qh 40

( kW )

(3.16)

where Q is in litres/s and h is in metres

3.1.2.3 Wellpoints A wellpoint consists of a 1 m long and 50 75 mm diameter gauze screen surrounding a central riser pipe. Wellpoints are jetted down by water at a pressure of up to 15 bar, penetrating in the ground by their own weight and requiring only to be guided by the worker (fig. 3.10). Once the desired level is reached, the jetting water supply is cut down to a low velocity sufficient to keep the hole around the point open. A coarse sand is then fed
86

around the annular space to form a supplementary filter around the point, after which the water is cut off.

1 water jet; 2 impervious seal of clay; 3 header Fig. 3.10

In fig. 3.11 are given details of the tip of a wellpoint. A particular feature is a rubber ball acting as a valve, which is lowered when jetting (fig. 3.11a) and raised when pumping (fig. 3.11.b). Wellpoints act most effectively in sands and sandy gravels of moderate permeability. The draw-down is slow in silty sands but these soils can be also effectively drained with wellpoints.

1 riser pipe; 2 gauze screen; 3 rubber ball by jetting; 4 rubber ball by pumping; 5 natural filter Fig. 3.11
87

In soils of lower permeability, the effectiveness of the wellpoints installation is increased by including a vacuum pump which creates in the wellpoints a negative pressure of 0.70.8 daN/cm2. The groundwater which is at the atmospheric pressure is drained forcefully to the wellpoints where the pressure is lower. When using the vacuum, wellpoints have to be provided with a clay seal, in order to maintain a high vacuum at the well screen (fig. 3.12).

1 lowered groundwater level by gravitational dewatering; 2 lowered groundwater level by vacuum dewatering; 3 impervious seal of clay; 4 wellpoint Fig. 3.12

In soils of finer particle size, such as silts, a further improvement of the effectiveness of wellpoints can be obtained by using the drainage by electro osmosis. For that purpose, between the well points and at close distance are driven in the ground steel rods connected to positive pole of a source of electrical current, becoming anodes, while the header main of the well points installation is connected to the pole minus, making the wellpoints to act as cathodes. The water particles surrounded by cations (see ch. 2) flow through the pores in the soil and collect at the cathodes, increasing the rate of flow obtained by gravitational drainage. A typical layout of an installation combining gravitational drainage and electro osmosis is shown in fig. 3.13.

1 wellpoint as cathode; 2 steel rod as anode Fig. 3.13

88

In fig. 3.14 is given an example of the use of dewatering by electro-osmosis in the case of a battered excavation. The anodes are placed nearest to the excavation causing the ground water to flow away from the slopes, which effectively stabilizes them and permits steep slopes. Wellpointing equipment comprises usually wellpoints, header mains, centrifugal pump, vacuum pump, electrical engine etc. Wellpoints are normally spaced from 1 to 4 m apart, depending on soil conditions and drawdown requirements.

1 wellpoint as cathode; 2 steel rod as anode; 3 stream lines Fig. 3.14

Unlike bored wells, which can be equipped with submersible pumps, wellpoints are operating with pumps at the ground surface providing a limited suction lift. A lowering of 5 - 5,5 m below pump level is a practical limit. For deeper excavation, the wellpoints must be installed in two or more stages. In fig. 3.15 is shown a cross-section of a large open excavation with a wellpoint installation applied in two stages.

Fig. 3.15

89

4.3 STRESSES DUE TO SELF-WEIGHT OF SOIL (GEOSTATIC STRESSES)


Vertical stresses Consider a soil mass having a horizontal surface. If the unit weight of soil is constant with depth
gv = z (4.7)

when z is the depth and is the bulk unit weight of the soil. The vertical stress will vary linearly with depth as shown in figure 4.7. If the soil is stratified and the bulk unit weight is different for each stratum (fig. 4.8), then the vertical stress can be computed by means of summation.

Fig. 4.7

Fig. 4.8

gv = i zi

(4.8)

Consider now a soil deposit with the water table at a depth h1. The water
' bearing stratum has the thickness h1 (fig. 4.9). ' At the depth (h1 + h1 ), the value of total stress vg is equal to the weight of the column of soil and water above that level.
' gv = 1 h1 + sat.1 h1

Where 1 and sat.1 are the unit weights of the soil 1 above and below the water table, respectively. The pore water pressure at the same level is:
' u = w h1

Then the effective vertical stress due to self-weight of soil is:


' ' ' 'gv = gv u = 1 h1 + ( sat w ) h1 = 1 h1 + 1 h1

(4.9)

' where 1 is the submerged (buoyant) unit weight of the soil 1. If the soil deposit above the water table is saturated by capillary action, on a height hc the porewater pressure above the water table is negative: - hc w . With the same amount is increased the effective stress.
' If under the water bearing stratum of thickness h1 there is a layer of impervious soil (clay), when computing the effective stress in a point within this layer (which is not subjected to the buoyancy effect) the pressure of the column of

' , must be added (fig. 4.10). The effective water at the roof of the clay layer, w h1 ' stress at a depth (h1 + h1 + h2) is: ' ' ' 'g = 1 h1 + 1 h1 + u h + 2 h 2 = 1 h1 + sat1 h1 + 2 h2

(4.10)

Fig. 4.9

Fig. 4.10

Horizontal stresses
The ratio of horizontal to vertical stress is expressed by a factor called the coefficient of lateral stress (or lateral stress ratio) K: K = h / v (4.11)

The definition is used whether or not the stresses are geostatic. In the special case where there has been no lateral strain within the ground, the ratio of horizontal geostatic stress to vertical geostatic stress is called the coefficient of lateral stress at rest (or lateral stress ratio at rest) Ko: K o =gh / gv (4.12)

The value of Ko can vary over a wide range depending on the stress history of the soil deposit. For sands and normally consolidated clays, Ko will typically have a value between 0.35 and 0.6 while for overconsolidated clays Ko may be as great as 3. The range of horizontal geostatic stresses for the at rest condition are shown in figure 4.11.

Fig. 4.11

4.4 STRESSES AT A POINT IN A SOIL MASS


The general state of stress on an element within a soil mass is shown with reference to total stresses in figure 4.12a. On each face the resultant stress is represented by one component of normal stress and two components of shear stress . For any given state of stress there must be exist three mutually perpendicular planes in the element on which the resultant stress is the normal stress, with the components of shear stress being zero. These are the principal planes and the associated normal stresses are the principal stresses. In descending order of magnitude we have the major principal stress 1 which acts on the major principal plane, the intermediate principal stress 2 acting on the intermediate principal plane and the minor principal stress 3 which acts on the minor principal plane. If the faces of the element are oriented in the directions of the principal planes, the state of stress on the element is as represented in figure 4.12b. If the element is considered to be infinitesimally small, the stresses on the faces of the element can be taken to represent the stresses acting on different planes through a point in the soil mass.

Fig. 4.12

Two-dimensional stress analysis: Mohr circle of stress


In geotechnical engineering there are many situations in which the mass of soil under stress is very long in one direction, say y direction, and loads do not vary with this direction. Three situations of this kind are illustrated in figure 4.13. a. retaining wall b. embankment c. strip foundation For this typical geometry, deformations of the soil mass in the y direction occur only locally at the ends of the structure and conditions over most of the soil mass approximate to those of plane strain, with y being the intermediate principal stress. Thus, every section in the x-z plane is the same as every other and they are no shear stresses on the plane perpendicular to the y axis. There is a normal stress y in the y-plane but the shear stresses yx and yz are zero. By treating unit thickness of the soil mass in the y direction we can reduce the problem to one involving two-dimensional stress analysis, in which we need to consider stresses in the x-z plane only.

Fig. 4.13

The two-dimensional state of stress of an element of soil is shown in figure 4.14a. To analyse the stress conditions in the element consider the equilibrium of the prism abc in figure 4.14b. Let and denote the normal and shear components of stress acting on plane ab. Let l denote the length to ab. Thus: Resolving forces normal to ab l = x lsin sin + z cos cos xz lsin cos zx l sin cos Now xz = zx and therefore: = x sin 2 + z cos 2 xz sin 2

= (1/ 2 ) x (1 cos 2 ) + (1/ 2 ) z (1+ cos 2 xz sin 2 ) = (1/ 2 )( x + z ) (1/ 2 )( x z ) cos 2 xz sin 2 and

(4.13)

1/ 2 ( x + z ) = 1/ 2 ( x z ) cos 2 + xz sin 2

(4.13)

Fig. 4.14

Resolving forces parallel to ab l = x lsin cos z lcos sin + xz lsin sin zx lcos cos

= (1/ 2 )( x z ) sin 2 xz cos 2

(4.14)

2 = 1/ 2 ( x z ) sin 2 xz cos 2

(4.14)

Adding eq. (4.12) and (4.13)


2 2 1/ 2 ( x + z ) + = 1/ 2 ( x z ) + xz 2 2

(4.15)

A plot of shear stress against normal stress is thus defined by a circle of radius
2 R = 1/ 2 ( x z ) + xz 2

with its centre on the axis at =1/ 2 ( x + z ) This is illustrated in figure 4.15 with the sign convention that compressive normal stresses and anti-clockwise shear stresses are positive. The plot of shear

stress against normal stress in known as a Mohr diagram and the circle as the Mohr circle of stress. If can be seen that the angle may be chosen such that is 0. This means to write eq. (4.14) =1/ 2 ( x z ) sin 2 xz cos 2 = 0 tan 2 = ( sin 2 ) / ( cos 2 ) = ( 2 xz ) / ( x z ) (4.16)

Thus for every set of x , z and xz there are two values of that satisfy the above conditions. These are the directions of principal planes. Substitution of eq. (4.16) into (4.13) gives the principal stresses:
2 1,3 = ( x + z ) / 2 ( x z ) / 2 + xz

1/ 2

(4.17)

If the x and z direction are the directions of principal planes then: x =1 :z =3 : xz = 0 and eq. (4.12, 4.13) reduce to:

= ( 1 + ) / 2 + ( ) / 2 cos 2 = ( 1 3 ) / 2 sin 2

(4.18) (4.19)

In conclusion, in order to build the Mohr circle of stress one has to know: a. Stresses x , z and xz acting on a vertical and horizontal plane passing through a point in the soil mass. Points H ( z zx ) and K ( x xz ) are plotted in a ( ) diagram. The intersection of the line KH with the axis gives the center of the circle (fig. 4.15). b. Stresses 1 and 3 on two principal planes. The center of the circle is located at a distance ( 1 + 3 ) / 2 from the origin: the radius of the circle is ( 1 3 ) / 2 .

By constructing in the circle the angle 2 , the point N on the circle thus defined has as coordinates the components and of the resultant p on a plane making with the plane of minor principal stress the angle (fig. 4.16).

Fig. 4.15

Fig. 4.16

On the circumference of every Mohr circle there exists a particular point called the pole point which has the following property: A line drawn from the pole point parallel to a given soil will intersect the circle again at a point the coordinates of which represent the normal and shear components of stress on that plane.

Thus, there is a correlation between: - the stress conditions on any plane in the soil element - the direction of that plane - the position of pole point. If any two of these are known, a simple construction on the Mohr circle yields the third. For example, in figure 4.15 point H has coordinates ( z zx ) defining the stress conditions on plane cb in the soil element, and point K has coordinates ( x xz ) defining the stress conditions on ac. Thus, the pole P is found by drawing a line through H parallel to plane cb in the element (or a line through K parallel to ac) to intersect the circle at point P. By definition, the principal planes are planes of zero shear stress and must therefore be represented by the points where the circle cuts the normal stress axis. The Mohr circle of stress is a very useful tool in performing two-dimensional stress analysis. By considering the soil element in figure 4.14a to be extremely small, we can use a Mohr-circle to represent the stress conditions at a particular point in a soil mass, each position around the circumference of the circle giving the components of stress on a different plane through the point. The concept of Mohr circles is equally applicable to two-dimensional total stress or effective stress analysis.

5.4 SETTLEMENT CALCULATIONS


Components of settlements
When a soil deposit is loaded, for example by a structure or a man-made fil1, deformations will occur. The total vertical deformation at the surface resulting from the load is called settlement. The total settlement st of a loaded soil has three components, or: st = si + sc + ss (5.32)

where: si = the immediate settlement; sc = the consolidation settlement; ss = the secondary compression. The immediate settlement, called sometimes also initial settlement, shear settlement or distorsion settlement, is a settlement without drainage for fully saturated soil due to shear deformation (distorsion) at constant volume. The

immediate settlement is produced by short term actions, e.g. construction loads, which act for a period during which drainage of the soil may be negligible. An example of immediate settlement of a saturated clay is given in figure 5.34. The prism of soil immediately below the uniform, flexible load, shortens and bulges elastically. The loaded area and the adjacent surface deform in a sagging profile.

Fig. 5.34

The consolidation settlement is a time dependent process that occurs in saturated fine grained soils which have a low coefficient of permeability. The rate of settlement depends on the rate of pore water drainage. Secondary compression (or consolidation) which is also time dependent, occurs at constant effective stress and with no subsequent changes in pore water pressure.

Evaluation of the total settlement Stress-strain method


To use the stress-strain method in order to evaluate the total settlement of a foundation on cohesive or non-cohesive soil, a sufficient number of points within the ground beneath the foundation should be selected and the stresses and strains computed at these points. The process implies three stages: 1. Computing the stress distribution in the ground due to the loading from the foundation. This may be derived on the basis of Elasticity theory, generally assuming homogeneous isotropic soil and linear distribution of bearing pressure.

2. Computing the strains in the ground from the stresses using stiffness moduli or other stress-strain relationships determined from laboratory tests (preferably calibrated against field tests) or field tests. 3. Integrating the vertical strains to find the settlements.

Fig. 5.35

A method in this category, recommended in the Romanian Standard 3300/285, refers to the evaluation of the settlement of a shallow foundation (fig.5.35). The deformation is supposed to be one-dimensional (zero lateral deformation) and is produced solely be the vertical stress . The foundation of width B and depth D is acted upon by a vertical load N = P+G, where P is the load transferred to the foundation by the structure and G is the own weight of the foundations and of the soil above the foundation. The effective pressure on the foundation base is defined as pef = N/A = (P+G)/A where A is the area of the foundation base. The net pressure on the foundation base is: p net = pef gD (5.34) (5.33)

where gD is the geostatic pressure at the depth D of the foundation base. The compressibility of various soil strata is defined by the deformation modulus E. The computation proceeds in phases, as follows: a) A transverse section through the foundation is drawn up, showing also the limit, between geological layers. The ground below the foundation is divided into elementary layers. The limits between geological layers, including the water table, represent limits between elementary layers. The thickness hi of the elementary layer should not exceed 0.4 B, and can vary from one layer to another. b) Vertical stresses z due to the net pressure ppnet and geostatic stresses gz are computed and plotted against the vertical axis drawn through the center of the foundation. The diagram of variation with depth of the stress gz starts at the ground level, while the diagram of z starts at the level of the foundation base. c) Based on the diagrams of z is defined the active zone, which represents that part of the ground in which stresses z are large enough to be considered in the settlement evaluation. As one can see, the two stresses z and gz have opposite tendencies: while z decreases, gz increases. On the other hand, soil deformation modulus E increases normally with depth, as a result of the compaction of the soil under the overburden. Hence, below a certain depth stresses z become so small as compared with gz that settlements induced by them are negligible. According to STAS 3300/2-85, the active zone is limited at the depth zo below the foundation depth at which the following condition is fulfilled (fig. 5.36): zo = 0.2 gzo (5.35)

When the lower limit of the active zone defined by eq. (5.33) finds itself in a layer having E < 50 daN/sq.cm (fig. 5.37); zo is extended to include that layer or to meet the requirement: zo = 0.1gz (5.36)

On the contrary, if within the active zone defined by eq. (5.33) a practically incompressible layer is met (E > 1.000 daN/sq.cm) and the presence within this layer of compressible inclusions is excluded, the active zone extends only to the upper limit of the hard layer (fig.5.38).

Fig. 5.36

d) The settlement si of the elementary layer i is computed. We consider that z is constant within the elementary layer and is defined as: z med i = zi + z i 1 2

(5.37)

Fig. 5.37

Fig. 5.38

This approximation means the replacement of the theoretical diagram of variation with depth of z by a diagram in steps. We can understand now the reason to limit the thickness of the elementary layer (hi 0.4 B). The error induced by considering elementary layers of larger thickness, as in figure 5.39 would be unacceptable. Hookes law is used: = E For the i layer we can take:
=med.i :E = Ei : = si / h i

si being the settlement of the layer i induced by the constant stress z.med.i :

z.med.i = Ei ( si / h i ):si = ( z.med.i h i ) Ei


Settlement s is obtained by summing up the settlements si: s = 0.8 si = 0.8 ( z.med.i h i ) Ei

(5.38)

(5.39)

In the eq. (5.37) 0.8 is a correction empirical factor aimed to reduce the difference between computed and recorded settlements.

Fig. 5.39

Adjusted elasticity method


The total settlement of a foundation on cohesive or non cohesive soil may be evaluated using Elasticity theory and an equation of the form s = (p B f)/Em (5.40) where p is the average bearing pressure linearly distributed on the base of the foundation, which for normally consolidated cohesive soils should be reduced by the weight of the excavated soil above the base. Buoyancy effect should also be taken into account.

Em is the representive drained Young's modulus of the ground. If no useful settlement results measured at neighbouring similar structures in similar conditions are available to evaluate Em it may be estimated from the results of laboratory or in-situ tests. f is a coefficient whose value depends on the shape and dimensions of the foundation area, the variation stiffness with depth, the thickness of the compressible formation, on Poisson's ratio and on the distribution of the pressure and the point for which the settlement is calculated. B is the width of the foundation. In the draft version of EUROCODE 7 Geotechnics, it is stipulated that the adjusted elasticity method should only be used if the stresses in the ground are such that no significant yielding occur and if the stress-strain behaviour of the ground may be considered to be linear. Great caution is required when using the adjusted elasticity methods in the case of non-homogeneous ground. To evaluate settlement components, separate calculations are required for the settlements so and s1. Settlements without drainage so
The settlement of a foundation without drainage of the soil may be evaluated using the adjusted elasticity method. For materials which are approximately homogeneous in stiffness, the settlement without drainage may be found from: so = ( pB ) E u fo f u (5.41)

where pB are the same as in (5.35) ; Eu is the mean value of Young's modulus for the deforming stratum for undrained conditions; fo is the foundations depth factor; fu is a settlement factor whose value depends on the shape and dimension of the foundation area, the variation of stiffness with depth, the thickness of the deforming stratum and on Poisson's ratio for undrained conditions u .

Settlement caused by consolidation


To calculate the settlement caused by consolidation a confined onedimensional deformation of the soil may be assumed and the consolidation test curves then used. Addition of settlement in the undrained state and of consolidation settlement on state often leads to an overestimate of the total settlement.

IACINT MANOLIU

SOIL MECHANICS II

LECTURE NOTES

CHAPTER 9. RETAINING STRUCTURES


9.1 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.).

9.2 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.

9.2.1 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. There are three main elements which form this kind of timbering: horizontal boards, walings, placed vertically, struts. The fig. 9.1 shows the phases of the construction of a timbering using horizontal boards: 222

- 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 boards; 2 waling; 3 strut

Fig. 9.1 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. 9.2).

1 vertical pile

Fig. 9.2 Another solution, which has the advantage of improving the working space, is to use raking struts (fig. 9.3). 223

1 raking struts

Fig. 9.3

9.2.2 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. 9.4 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. - 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 volume of excavation and in consumption of timber (fig. 9.5).

224

1 frame; 2 puncheons ; 3 vertical boards; 4 wedges; 5 contour of the excavation for a panel of 11,5 m

Fig. 9.4

Fig. 9.5

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. 9.6. - 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; 225

- 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.

1 poling boards; 2 horizontal frames; 3 puncheon ; 4 guide wedges; 5 wedges

Fig. 9.6 For large excavations (fig. 9.7) 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. 9.7 226

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 hydraulic rams or screw jacks, allowing sheeting members to be forced tightly against the soil (fig. 9.8).

a transversal section; b detail of the horizontal steel frame; 1 vertical board; 2 horizontal steel frame; 3 clamping sleeve; 4 key

Fig. 9.8

9.2.3 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.

227

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

Fig. 9.9

9.3 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. 9.10.

1 sheet pile; 2 piles; 3 reinforced concrete slab

Fig. 9.10

9.3.1 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. 9.11a) or overlapped boards (fig. 9.11b,c) can be used. When watertightness is required, various kinds of joints 228

are used such as in half-wood joint (fig. 9.11d) in birdsmouth joint (fig. 9.11e) or tongue and groove joint (fig. 9.11f). The tongue and groove type of joint may be obtained by bolting together three boards (fig. 9.11g).

Fig. 9.11 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. 9.12).

Fig. 9.12 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. 229

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. 9.13).

1 tongue; 2 groove

1 pile; 2 sheet pile

Fig. 9.13

Fig. 9.14

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. 9.14). 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. 9.15).

1 guide piles; 2 tongs; 3 set of 2-sheet piles; 4 wedge; 5 timber piles; 6 wedges; 7 clamps; 8 spacer

Fig. 9.15 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 230

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 granular soils or stiff clays without risk of splitting.

9.3.2 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. 9.16 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 231

the x-x axis and w x is the reduced resistance modulus, considering the joints. At the Stype 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. 9.17). 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. 9.17

9.3.3 Reinforced concrete sheet piles


Reinforced concrete sheet piles are precast elements, of square or rectangular crosssection, 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 impossible. The interlocking system can be similar to the one used at timber sheet piles. In fig. 9.18 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 part there are two grooves. The gap between grooves is filled with jute and mortar (fig. 9.18b).

Fig. 9.18

232

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.

9.4.1 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. 9.19 Guide walls are placed at the ground surface to ensure the designed position and the verticality of each pile.

Fig. 9.20 233

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. 9.21a,b). If proper construction is insured, hard/hard secant pile walls offer a high water tightness. Hard/soft secant pile walls (fig. 9.22) 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. 9.21

Fig. 9.22 Contiguous pile walls When the wall is not required to retain water, contiguous bored piles are used (fig. 9.23). 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. 9.23

9.4.2 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. 234

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, diaphragm walls, made of panels (diaphragm walls) can be divided into two categories: - homogeneous walls - 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. 9.24). 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; 235

to provide reaction for the hydraulic jacks when extracting the pipes 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.

Fig. 9.24 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. 9.25 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 1..3 m above the bottom and penetrates in the soil. Loaded with soil, the grab is then lifted above the guide-wall 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. 9.26 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 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 236

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. 9.25

Fig. 9.26

237

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

Fig. 9.27 - Excavation with the installations Else. 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. 9.27). The installation uses the guidewall 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 Else installation is of 0.6..0.81.0 m and the depth up to 30 m. Soils suitable to be excavated with Else 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, ELSE cannot excavate in hard ground, such as a sandstones, limestones, etc. - 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. 238

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 guide-walls, 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. 9.28). 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 ELSE 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. 9.28

1 longitudinal bars; 2 horizontal bars; 3 bracing bars; 4 lifting bar; 5 spacers

Fig. 9.29 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. 9.29). Each cage is provided also with suspension and lifting bars, bracing bars to improve the 239

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 piece 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. 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. 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 pas the specified properties. This is achieved by providing an additional height of concrete above the cut-off level. Fig. 9.30 shows the operation 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. 9.30 240

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. 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 thield made of this 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. 9.31).

Fig. 9.31

1 previously concreted panel; 2 panel under excavation

Fig. 9.32 In the case of using ELSE 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.

241

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. When using ELSE 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. 9.33) 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. 9.33 b. wall made of a starter and intermediate panels (fig. 9.34) 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; IV excavation of the intermediate panel; placing at its extremity of a joint tube; 242

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. 9.34 Diaphragm walls made of precast concrete These walls are made of precast elements lowered into a trench containing a selfhardening slurry.

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

Fig. 9.35 Fig. 9.35 shows a wall made of precast panels which imitate the timber sheet piles with tongue and groove. Fig. 9.36 shows a wall which imitate a Berlin-type wall, with the 243

difference that the horizontal lagging boards are replaced by continuous vertical precast sheets. Fig. 9.37 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.

1 joint panel; 2 field panel; 3 guide wall; 4 trench filled with slurry; 5 self-hardening slurry

Fig. 9.36

Fig. 9.37 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 self-hardening 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. 244

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 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. 9.38 Fig. 9.38 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 replacing the bentonite suspension in the lower part of the trench with the self-hardening slurry with higher density (1.201.25 g/cm3); 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; e 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 method 245

- the cut and cover method Embedded walls for underground structures constructed in open excavation

Fig. 9.39 Fig. 9.39 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; - 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 cut and cover 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 246

to resume activities at the ground surface (car traffic etc.) before completing the 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 (barette); 7 ballast; 8 reinforced concrete slab; 9 mat; 10 wall; 11 intermediate slab; 12 column; 13 fill; 14 - pavement

Fig. 9.40 Fig. 9.40 shows the main stages of the construction of a subway station constructed with the cut and cover 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.

9.5 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 problem for the design of a braced excavation is establish the appropriate diagram of earth pressure.

247

9.5.1 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. 9.41, being negligible at the top and increasing with depth. The deformation condition of the Rankinetheory 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. 9.41. 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 pressured diagrams for various types of soils are recommended in practice: for non-cohesive soils in fig. 9.42 a for cohesive soils of low consistency in fig. 9.42 b for cohesive soils of high consistency in fig. 9.42 c

Fig. 9.41

Fig. 9.42

9.5.2 Design of a timbering


248

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. 9.43).

Fig. 9.43 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

(9.1)

The bending moment in the board is:


M=
2 q l1 8

(9.2)

To check of the board section:

249

2 q l1 M = = 8 all W b d2 6

(9.3)

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 (9.3). The analysis of the walings is done as for simply supported beams of span l2, 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

(9.4)

The maximum bending moment for the waling is:


q l2 M1 = 1 2 8

(9.5)

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


= M 1 M1 = all W1 ef 2 6

(9.6)

From the relation (9.6) is determined f. 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. 9.43, the compression load is: N = p aH l1 l 2 (9.7) To check the struts in compression the relation (9.8) is used:
II = N W all II A

(9.8)

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: 250

N all A

(9.9)

where all is the allowable strength in compression normal to the fibres of the timber 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 W and all should be introduced.

9.5.3 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 (can till ever 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 indetermined systems (walls with one level of anchor or prop in the upper part and fixed earth support at the 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. 9.44 are shown various kinds of failure for these walls. In what follows, several types of walls pertaining to the two categories will be considered.

251

Fig. 9.44

a. Cantilever walls These walls are free at the upper part and derive their equilibrium from the lower, embedded depth 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 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. 9.45 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

252

Fig. 9.45

Fig. 9.46

By computing along the embedment t the difference between the passive and active earth pressure, a resultant diagram is obtained (fig. 9.46 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. 9.46 b) and which can be approximated by a line (fig. 9.46 c). The stability of the wall is insured of the couple of forces Ep 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. 9.45 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: X=0 (9.10 a) 253

(9.10 b) Equation (9.10 a) represents the horizontal projections of horizontal forces, while equation (9.10 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:
Mc = 0
e f FS

(9.11)

where 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 Ep. 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
1 2 to H ( t + t o ) (K p K a ) t o =0 2 3

(9.12)

Equation (9.12) leads to the following equation:


t3 o 6 H H to 6 h = 0 (K p K a ) (K p K a )

(9.13)

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

20 1.6

25 1.2

30 0.9

35 0.7

Table 9.1 40 0.5

to/h

The embedment depth should be larger than to: t = (1.20.1.25)to

(9.13)

In order to find the maximum bending moment, the pressure diagram used for the computation of t is taken (fig. 9.45 c or fig. 9.46 a) and the depth zo at which the shear force is zero, corresponding to Mmax, is found. For instance, when the diagram in fig. (9.47 a) is applied:
1 2 H zo (K p K a ) = 0 2

(9.14)

254

zo =

2H (K p K a )

(9.15)

1 2 M max = H (h + z o ) (K p K a ) z o 2

(9.16)

Fig. 9.47 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. 9.48). The corresponding diagrams are given in fig. 9.49 a, fig. 9.49 b and fig. 9.49 c. The bending moment diagram is shown in fig. 9.50.

Fig. 9.48

255

Fig. 9.49

Fig. 9.50 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. 9.51 a) or prop (fig. 9.51 b). The second support is represented by the embedment depth t, as a free earth support.

256

Fig. 9.51

Fig. 9.52 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. 9.52). 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.

Fig. 9.53 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. 9.53). 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 257

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. 9.54). The pressure diagram is constructed as in the case of the cantilever wall (fig. 9.43 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. 9.54

Fig. 9.55 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 RA 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.

9.5.4 Solutions for the support in the upper part of embedded walls

258

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. 9.51 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. 9.51 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. 9.56 a); - anchor piles (fig. 9.56 b); - ground anchors (fig. 9.56 c). The deadman may be formed of isolated elements (fig. 9.56 a) or by a continuous plate (fig. 9.56 b).

Fig. 9.56

Fig. 9.57 The stability of the deadman is insured by the passive resistance of the soil in front of the plate or block (fig. 9.57). The stability condition is:
R A + Pa Pp FS

(9.17)

where Pa and Pp are the active and passive forces and FS a factor of safety which can be taken 1.5.
1 1 Pa = d 2 l K a + e d l K a = d l K a ( d + e); 2 2

(9.18)

259

1 1 Pp = d 2 l K p + e d l K p = d l K p ( d + e); 2 2

(9.19)

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 (9.18) and (9.19) 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 + ) in respect to the horizontal. Between this plane and the wall the prism 1 is formed, where is strictly forbidden to place the deadman. At the same time, the passive zone in front of the deadman 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 ) in respect to the horizontal is drawn, defining the prison 2. The deadman should be located above this plane.
2 2

Fig. 9.58

9.6 Ground anchors


A ground anchors 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. 260

Fig. 9.59 shows several examples of permanent structures using ground anchors: a retaining wall (9.59 a), a high rise building with a multiple-level basement, founded below the water table and subjected to uplift forces of the water (fig. 9.59 b), a tower for an electrical line (fig. 9.59 c).

Fig. 59

Fig. 9.60 Examples of temporary ground anchors are given in the fig. 9.60. A ground anchor consists basically of an anchor head, free anchor length and fixed anchor. Fig. 9.61 shows a typical ground anchor.

Fig. 9.61

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. 261

2 introduction in the borehole of a tendon 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 cementitious 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 (Lfixed), 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). 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.

9.7 Retaining walls


262

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.

9.7.1 Gravity walls


9.7.1.1 Mass concrete retaining walls Mass concrete walls are suitable for small retained heights, usually up to 34 m. Fig. 9.62 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 W of the wall; - reaction R on the base. As a rule, the passive force Pp, whose development requires, as shown in the chapter 8, large displacements, is disregarded. In the fig. 9.62 are also given recommendations for a preliminary selection of the dimensions of the wall.

Fig. 9.62 In fig. 9.63 are shown other forms of mass concrete walls.

263

Fig. 9.63 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. 9.64).

Fig. 9.64 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 pall.

264

p max =
min

N M N M N 6e = = (1 + ) A W B 1 1 B 2 B B 6

(9.20)

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

M N

N = Pav + W Pav being the vertical component of the active earth thrust. Three conditions shall be fulfilled:
p med p all p max 1.2 p all

(9.21)

(9.22) (9.23) (9.24)

p min 0

Relation (9.24) expresses the condition that the resultant of forces Pa and W is located within the middle third of the base ( e B 6 ). - stability checks Stability against sliding The friction force S (the component along the base of the reaction R) on the base is compared with the horizontal component Pah 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

(9.25)

where ml is a factor of safety equal to 0.8. When values for obtained by field tests are lacking, values given in tab. 9.2 can be used, at least for a preliminary design. When the base resistance to sliding is inadequate, this can be increased by either widening the base, inclining the foundation (fig. 9.65 a) or providing a shear key (fig. 9.65 b). Stability against overturning The following condition shall be fulfilled:
M r m r MS

(9.26)

where Mr is the overturning moment Mr = Pa a (9.27) 265

a being the arm of the force Pa in respect to the toe of the wall D MS is the stability moment MS = W d d being the arm of the force W is respect to the point D mr is a factor of safety equal to 0.8. (9.28)

Fig. 9.65 9.7.1.2 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. 9.66).

Fig. 9.66 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. 266

Gabion walls are designed on the same principle as a gravity mass wall. In fig. 9.67 are given examples of gabion retaining walls. The density of the stone fill can be taken as 60% of the solid material.

Fig. 9.67 9.7.1.3 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. The units should be so spaced that the fill material is 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. 9.68 is shown an example of assembling the individual units. In fig. 9.69 is given a vertical section on a crib wall.

Fig. 9.68

267

Fig. 9.69 This kind of gravity walls is indicated for use on compressible soils having the ability to adapt to differential settlements.

9.7.2 Reinforced concrete walls


9.7.2.1 Cantilever walls Cantilever or T walls are made of a vertical or inclined slab monolithic with a slab base (fig. 9.70).

Fig. 9.70

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 268

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 2

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. 9.7.2.2 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).

Fig. 9.71

9.8 Reinforced earth


The idea of using the soil itself for structures aimed to resist to the active earth pressure, appered 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, plastics 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 269

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.

Fig. 9.72

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 (9.29) 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 (9.30) 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, Lmin, 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

(9.31)

From the relation (9.31) results that Lmin is constant and independent of z. 270

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. 9.73 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 (9.29) and (9.31) 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 (9.229.24) as for a gravity wall. 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.

9.9 Cofferdams

271

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.

9.9.1 Cofferdams made of earth and rock fill dikes


9.9.1.1 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. 9.74).

Fig. 9.74 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 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 layer than 0.1 m/s, slopes of the dikes should be protected to prevent erosion. 9.9.1.2 Rock fill dikes

272

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. 9.75 a) or clay cores (fig. 9.75 b), when the construction of the dike is made in dry, or sheet pile walls (fig. 9.75 c), when dike is built in water. In all cases, the watertight element should penetrate into an impervious layer of soil.

Fig. 9.75

9.9.2 Sheet piles cofferdams


9.9.2.1 Single skin cofferdams These are formed by one-line timber or steel piles, working as cantilever walls (fig. 9.76 a) or as walls with free earth support and raking struts (fig. 9.76 b), when ground conditions do not allow to drive the sheet piles to a depth required in a fixed-end support.

Fig. 9.76 Other means to ensure the stability of the single skin cofferdams are represented by earth fills placed in front of the wall (fig. 9.77 a) or on both sides of the wall (fig. 9.77 b).

273

Fig. 9.77 9.9.2.2 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. 9.78). 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. 9.78 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. 9.9.2.3 Cellular cofferdams 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 booken 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 274

in the piling, which the straight web piles are designed to resist, unlike through shaped piles sections which are unsuitable.

Fig. 9.79 In fig. 9.79 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 selfsupporting 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. 9.80).

Fig. 9.80

275

CHAPTER 10. STABILITY OF SLOPES


A slope can be defined as the inclined surface of the ground. Slopes of different nature are represented in the fig. 10.1 a. natural slopes, when the inclined surface is the result of geological processes b. slopes formed by excavation c. slopes of embankments and earth dams

Fig. 10.1 A main problem in the design and construction is to ensure the stability of slopes. The principal factors which can cause instability of slopes are: - reduction of the shear strengths of cohesive soils as a result of an increase of the water content, - seepage forces, - surcharge on the upper part of the slope - removal of the soil at the base of the slope - seismic or other dynamic actions. There is a great variety of types of slope failure. The most important ones are illustrated in fig. 10.2

276

a. rotational slip, with failure surface as a circular arc, characteristic for homogenous soil conditions; b. rotational slip, with failure surface as a non-circular arc, characteristic for nonhomogenous soil conditions; c. translational slip, where the form of the failure surface is influenced by the presence of an adjacent stratum of significantly different strength at a relatively shallow depth below the surface; the failure surface could to be plane and roughly parallel to the slope or could be assimilated with as a polygonal one; d. compound slip, where the adjacent stratum of different strength is at greater depth, the failure surface consisting of curved and plane sections.

Fig. 10.2 Defining the forces which provide stability and of those who trigger the instability and of the factors which modify the ratio between these forces represents a compulsory requirement for understanding of natural phenomena, quite often transformed in real disasters, called in Engineering Geology landslides and for adopting appropriate measures to prevent and to control them.

10.1 Stability of slopes in non-cohesive homogenous soil mass


10.1.1 Non-cohesive soil in dry or saturated state
By expressing the limiting equilibrium conditions for a soil particle at the surface of a slope in a non-cohesive soil in dry or saturated state, the following relation is obtained: tan = tan (10.1)

277

where is the angle of stable slope and is the angle of internal friction of the soil. The stability condition to be used in the design is:
tan = tan FS

(10.2)

where FS is the factor of safety. The very important conclusion to be drawn from the relation (10.1) is that for noncohesive soils in dry or saturated state the angle of stable slope is independent of the height of the slope.

10.1.2 The influence of a seepage on the stability of slope in a non-cohesive soil


A slope in a non-cohesive soil is considered, with a flow line tangential to the slope (fig. 10.3). A unit volume of soil at the surface of the slope is subjected to the following forces: - own weight , of two components: - tangential T = sin - normal N = cos h - seepage force: j = w i = w = w sin l - friction force: F = N tan = ' cos tan The stability condition is: T + j F
' sin + w sin ' cos tan

tan

' ' + w

(10.3)

For usual values of s and n , ' is about 10 kN/m3. If w is taken also 10 kN/m3; 1 the relation (10.3) becomes tan tan . 2

278

Fig. 10.3
By comparing (10.1) with (10.3) one can realize that the seepage force reduces to half the magnitude of the stable slope.

10.2 Stability of slopes in cohesive homogenous soil mass


10.2.1 Case of the plane failure surface
A cohesive homogenous soil mass is considered, limited by a slope of height H and angle to the horizontal. It is required to check the stability of the slope. In order to outline the forces involved, one can assume as a working hypothesis the loss of stability along a plane failure surface with an angle to the horizontal (fig. 10.4).

Fig. 10.4
The weight G of the wedge ABC which tends to slip has two components: N = G cos T = G sin The failure condition is fulfilled in each point of the surface BC. f = tan + c By summing up the stresses and along the area BC 1= L 1 (plane strain), the total resistance S opposed by the soil to the tendency of the prison ABC to slip is: S = G cos tan + cL = F + C

279

The stability condition is: T S G sin G cos tan + cL From the sinus rule in the triangle ABC:
AC L L = = o sin ( ) sin (180 ) sin AC = L

(10.4)

sin ( ) sin

G=

1 1 sin ( ) H AC = HL 2 2 sin

(10.5)

By replacing (10.5) in (10.4):


1 sin ( ) sin 1 sin ( ) cos sin HL HL + cL 2 sin 2 sin cos

sin ( ) sin sin ( ) cos sin 1 H c 2 sin sin cos sin ( ) [sin cos sin cos ] 1 H c 2 sin cos sin ( ) sin ( ) 1 H c 2 sin cos

(10.6)

For a given height H, the slope of angle to the horizontal is stable of the condition (10.6) is met for any values of between and . The most unfavorable situation corresponds to the value of for which the first member of the relation (10.6) becomes maximum. If the following notation is introduced: sin ( ) sin ( ) = A The condition of maximum requires:
dA =0 d

280

- cos ( ) sin ( ) + sin ( ) cos ( ) = 0 sin [ ( )]= 0


o = + 2

(10.7)

Considering (10.7), the relation (10.6) becomes:


1 2 c H 2 sin cos sin 2

(10.8)

The stability condition (10.8) shows that, unlike the non-cohesive soils, in the case of cohesive soils the angle of stable slope depends on the height of the slope. At limit, by making equal the two term of the relation (10.8), one can obtain the maximum height for which a slope with an angle to the horizontal can be stable, named critical height Hcr : Hcr =
2c sin cos sin 2 2

(10.9)

Fig. 10.5
In the angle decreases, Hcr increases (fig. 10.5). From the relation (10.9) it results that in the stability condition of a slope of cohesive soil there are involved 5 parameters: shear strength parameters , c unit weight geometrical characteristics of the slope H, The expression (10.9) can be rewritten as:
H 2 sin cos = = Ns c 2 sin 2

(10.10)

281

Ns is a non-dimensional factor named stability number. For given and and for a certain shape of the failure surface, Ns has a defined value. For instance, for the particular case of vertical slope = 90o: Ns = 4 tan (45o + ) 2 Hcr =
4c tan (45o + ) 2

(10.11)

(10.12)

If = 0 Ns = 4 Hcr =
4c

(10.13) (10.14)

Relations (10.12) and (10.14) have been obtained by applying Rankines theory in the case of limiting equilibrium in the soil mass limited by horizontal ground surface. It was established that in the case of cohesive soil, the depth at which the active earth pressure becomes zero is: z0 =
2c tan (45o + ) 2

and the critical height Hcr, the height on which a vertical cut in a cohesive soil can remain unsupported, is: 4c Hcr = 2z0 = tan (45o + ) 2

10.2.2 Case of circular failure surface. Method of friction circle


The potential failure surface is assumed to be a circular arc (fig. 10.6). Forces which should be in equilibrium are: - the weight G of the soil mass which slips ABC; - the resultant C of the cohesion mobilized along the failure surface: C = c BC , BC acting parallel to the chord BC at a distance d = R from the centre 0; BD - the total reaction Q on the failure surface tangential at the -circle of radius r = R sin . There are three ways in which the factor of safety can be expressed.

282

with respect to the cohesion

Fig. 10.6
For a given , the -circle is drawn and from the intersection of lines of action of forces G and C a tangential to the circle defines the direction of Q. The triangle of forces is drawn by decomposing G to the directions of C and Q. By this way is obtained the total cohesion force required for the limiting equilibrium Creq which, divided by the area BC 1 , gives the required cohesion creq. The factor of safety is: FS c =
c C req

(10.15)

with respect to the internal friction angle For a given c, the forces polygon is built, with known values of G and C, resulting the value and the direction of the force Q. From the point of intersection of the forces W and C a parallel is drawn to the direction of Q. The normal drawn from the centre of the circle to the direction of Q defines the radius of the -circle. By knowing Rsin , tan req can be established.
The factor of safety: FS =
tan tan req

(10.16)

with respect to both cohesion and internal friction For a given , the limiting equilibrium requires creq; for a given c, the limiting equilibrium requires tan req. There is an infinite number of pairs of values (tan req, creq) which in the system of coordinates (c 0 tan ) defines a curve. A point M corresponds to the parameters and c of the shear strength. By linking this point to the origin, the line OM intersects the curve in point P. The factor of safety in respect to both internal friction and cohesion is:

283

FS =

OM OP

(10.17)

The same curve allows to obtain the two other factors of safety:
FS c = ML c = P' L c req MK tan = P' K tan req (FS =1)

FS =

(FS c =1)

Analysis for the case = 0


In the case of a fully saturated clay under undrained conditions, an analysis in terms of total stress is appropriate. Only moment equilibrium is considered in the analysis. The potential failure surface is a circular arc (centre 0, radius R, length La (fig. 10.7). Potential instability is due to the total weight of the soil mass (W) above the failure surface. For the equilibrium, the shear strength required to be mobilized along the failure surface is expressed as:
req = f c u = Fc Fc

(10.18)

where Fc is the factors of safety with respect to shear strength.

Fig. 10.7
Moment about 0 is taken:
c G d = u La R Fc c L R Fc = u a G d

(10.19)

284

To determine the minimum factor of safety Fc, it is necessary to analyse the slope for a number of trial failure surfaces.

Taylors stability coefficients


Using the method of -circle for soil having both internal friction and cohesion and the analysis for the case =0, Taylor established stability coefficients NS for the analysis of homogenous slopes in terms of total stress (fig. 10.8), in the hypothesis of a circular failure surface.

Fig. 10.8 As one can see, for the particular case =90o and for =0, NS=3.85 (as compared to NS=4 for the hypothesis of plane failure surface). For the use of Taylors graphs, there are two problems encountered in practice.

Design of a stable slope , c, , NS are given. The height H for a given factor of safety FS is sought for. The height required for the limiting equilibrium is Hreq. Then, in order to comply with the factor of safety FS, Hreq = FS H
NS =
H req c = FS H c

(10.20)

From the value of NS on the ordinates axis, a horizontal is drawn until intersects the curve corresponding to the given . The vertical drawn from the point of intersection meets the needed abscissa .

Checking the stability of a slope ,c, , H, are given. FS is required From the abscissa , a vertical is drawn until intersects the curve of the given . An horizontal is drawn and the ordinate NS is found. 285

NS =

FS H c c NS H

FS =

(10.21)

10.2.3 The influence of the groundwater on the stability of slopes in cohesive soils
The presence of the groundwater can significantly modify the conditions of stability of slopes in cohesive soils. Depending on the water level inside and in front of the slope, three distinct situations are occurring.

a. Slope completely imersed, the soil in submersed state


There are two ways in which the presence of water can be considered (fig. 10.9).

Fig. 10.9
1. When computing the weight G of the sliding soil mass the unit weight of the soil is taken sat . The weight G is composed with the resultant Pw1 of the water pressure on the side AB and with the resultant Pw 2 of the water pressure on the failure surface BC. The resultant G ' is obtained, which then is decomposed, as shown at p. 10.2.2, on the directions of C and R. 2. The weight of the sliding mass G ' is obtained directly by considering the unit weight in submersed state ' . Then, the procedure outlined in the variant 1 is used.

286

b. Rapid drawdown of the water level in front of the slope


This situation can take place in the case of dikes or earth dams facing an accidental discharge of the water reservoir (fig. 10.10). Since the drawdown is rapid and sudden as a result of the low permeability of the soil, it is not followed immediately by the draw down of the water inside the soil mass. The total weight G of the soil in saturated state is the same as in the previous case, but there is no longer water pressure on the slope and the ' of the water pressure on the failure surface BC is smaller due to the lowering resultant G 1 of the water level in front of the slope.

G and and R. The stability conditions in this case are much more severe than in the previous case.

' G1

Fig. 10.10 are composed and their resultant B is decomposed on the direction of C

c. Stability of slope under conditions of steady seepage


After a sufficient time elapsed to allow the consolidation (dissipation of excess pose water pressure will have taken place) a steady seepage is generated through the soil mass directed toward the base of the slope, accompanied by the lowering of the ground water level. If stability conditions are checked with the Swedish method, the factor of safety can be expressed in terms of effective stress with the following relation:
FS =

{[(G i cos i u i li ) tan ' ] + c'li + G i sin i l }


G i sin i r

(10.22)

287

The pore water pressure ui at the centre of the base of each slice is taken to be w z w , where zw is the vertical distance of the centre point below the water table. This procedure slightly overestimates the pore water pressure, which strictly should be w z c , where zc is the vertical distance below the point of intersection of the water table and the equipotential through the centre of the slice base (fig. 10.11). The error involved is on the safe side.

Fig. 10.11
A series of trial failure surfaces must be chosen in order to obtain the minimum factor of safety.

10.3 Stability of slopes in non-homogenous soil mass


10.3.1 The method of slices
The potential failure surface, in section, is assumed to be a circular arc with centre 0 and radius R. The soil mass (ABCD) above a trial failure surface (BD) is divided by vertical planes into a series of slices of width b (fig. 10.12). The base of each slice is assumed to be a straight line. The inclination of the base to the horizontal is and the height, measured on the centre-line, h. The factor of safety is defined as the ratio of the available shear strength ( f ) to the shear strength ( req ) which is required to be mobilized to maintain a condition of limiting equilibrium: FS =
f req

(10.23)

The factor of safety is taken to be the same for each slide, implying that there must be mutual support between slices, meaning that forces must act between the slices.

288

Fig. 10.12
The forces (per unit dimension normal to the section) acting on a slice are: the total weight of the slice, G = bh ( sat if the case); the total normal force on the base, N, equal to l . In general, this force has two components: the effective normal force N ' , equal to ' l , and the boundary water force U, equal to ul, where u is the pore water pressure at the centre of the base and l is the length of the base; the shear force on the base S = req l

the total normal forces on the slides, E1 and E2; the shear forces on the sides, X1 and X2. Any external forces, such as a surcharge pressure on the ground surface, must be also be included in the analysis. The problem is statically indeterminate and, in order to obtain a solution, assumptions must be made regarding the interslice forces E and X and the resulting solution for factor of safety is not exact. The Fellenius solution
Known also as the Swedish method, this solution implies that for each slice the resultant of the interslice forces is zero. The following example refers to a stratified soil mass (fig. 10.13). When the sliding mass ABCD is divided in slices, it is required that the base of each slice belongs to only one layer. The forces acting on the slice i are considered. The total weight of the slice i, Gi, is decomposed in a normal component Ni and a tangential one Ti. Ni = Gi cos i Ti = Gi sin i The force Ti tends to produce the sliding of the slice i along the failure surface. The force opposing to the sliding is the shear force due to the shear strength f i mobilized on the base of the slice i.

289

Fig. 10.13
Si = f i l i = ( i tan i + ci ) li = G i cos i tan i + ci li ci, i represent the shear strength parameters corresponding to the soil layer in which is located the base of the slice i. The equilibrium of the slice i can be expressed considering the moment of the forces Ti and Si about 0. The factor of safety FS can be defined as the ratio between the stability moment MS of forces opposing to the sliding and the overturning moment M0 of forces producing the sliding.
FS = M S R (G i cos i tan i + ci li ) (G i cos i tan i + ci li ) = = (10.24) R G i sin i G i sin i M0

A particular case occurs when the vertical passing through 0 divides the sliding mass (ABCD) in two parts. As one can see, for the slices located at the left of the vertical line, the tangential component of the weight W is directed against the sliding, therefore it must be included among the stability forces. The expression of the factor of safety becomes:
FS = M S (G i cos i tan i + G i sin i l + ci li ) = G i sin i M0 r

(10.25)

Where sin i l corresponds to the slices located at the left of the vertical and sin i r to the slices located at the right of the vertical. In relations (10.24) and (10.25) the factor of safety is expressed in terms of the total stresses.

290

The Swedish method underestimates the factor of safety; the error, compared with more accurate methods of analysis, is usually within the range 5-20%. An approximative method to obtain the minimum factor of safety FS min is the following one: It is assumed that the centres of circular failure surfaces susceptible to induce low factors of safety are located on a line O1M (fig. 10.14). The point O1 is found at the intersection of lines making angles 1 and 2 with the slope, respectively the horizontal passing through point E. Values of 1 and 2 in function of the slope are given in the table 10.4.

Fig. 10.14

tan 1 2

1.75.1 60o 29o 40o

1:1 45o 28o 37o

1:1.5 33o45o 26o 35o

1:2 26o34o 25o 35o

1:3 18o25o 25o 35o

Table 10.4 1:5 11o19o 25o 37o

The point M has the coordinates (4.5H) and (H), in respect with the point B. On the line O1M are chosen several points as centres of circular failure surfaces. Using the slices method, the factor of safety is computed for each surface and is drawn at a convenient scale normal to the line O1M, taken as a reference. A curve of variation of FS is obtained. The tangential to the curve parallel to the reference line defines FS min . If FS min is larger than a preestablished allowable value FS all , the stability condition of the slope is fulfilled. If FS min < FS all , appropriate measures have to be considered in order to improve stability conditions, such as diminishing the slope (angle ), removing soil from the upper part of the slope or creating an embankment at the base of the slope (fig. 10.15).

291

Fig. 10.15

10.3.2 Translational slip in the case of a plane failure surface parallel to the surface of the slope
It is assumed that the potential failure surface is parallel to the surface of the slope and is at a depth that is small compared with the length of the slope. The slope can be considered as being of infinite length, therefore end effects can be ignued. The slope is inclined at angle to the horizontal and the depth of the failure plane, which usually is the thickness of the soil layer covering a rock layer, is z (fig. 10.16). The water table is taken to be parallel to the slope at a height of m Z (0 < m < 1) above the failure plane. Steady seepage is assumed to be taking place in a direction parallel to the slope. The forces on the side of any vertical slice are equal and opposite and the stress conditions are the same at every point in the failure plane.

Fig. 16
In terms of effective stress, the shear strength of the soil along the failure plane is:
f = ( u ) tan ' + c'

(10.26) 292

and the factor of safety is: FS =


f req

(10.27)

For a slice of width 1 and height z, the weight G is: G = [(1 m ) + m sat ] z The vertical pressure on the base is: p = [(1 m ) + m sat ]z cos

The expressions for , and u are:


= [(1 m ) + m sat ]z cos 2 = [(1 m ) + m sat ]z sin cos

u = m z w cos 2 By replacing and u in the relation (10.26) and then in the relation (10.27), the factor of safety FS is found.

Special cases If c ' = 0 and m = 0, when the soil between the surface and the failure plane is not fully saturated, then:
FS =
tan ' tan

(10.28)

The relation (10.28) is identical to the one previously established. If c ' = 0 and m = 1, when the water table coincides with the surface of the slope, then: F=
' tan ' sat tan

(10.29)

293

As already shown in the p. 10.1, when c ' = 0, in the case of non-cohesive soils, the factor of safety is independent of the depth Z. For a total stress analysis the shear strength parameters cu and u are used and u is taken zero.

10.3.3 Translational slip on a predetermined failure surface taken as a polygonal surface


As in the previous case, the failure surface is predetermined and is represented by the surface of a rock layer covered by a mass of soil. Quite often this is a diluvium made primarily of cohesive soils deposited over a rock. To check the stability, the contact surface between the deluvial mass and the rock is assumed to be a polygonal surface (fig. 10.17a). Vertical planes are drawn in the points of change in slope of the contact surface, dividing the sliding soil mass in a number of blocks. Horizontal interblock forces are assumed. The analysis starts with the block at the top of the sliding mass, on which the following forces are acting and must be in equilibrium (fig. 10.17b):

Fig. 10.17 the weight G1, of known magnitude and direction; the reaction Q1 on the base of the slice, with an angle to the normal, where is the angle of internal friction of the soil; the cohesion force C1 on the base, known in magnitude and direction; C1 = cl1, where c is the cohesion of the soil; the normal force E2-1 on the side, representing the interblock force, known only as direction. Forces G1 and C1 are composed and their resultant is decomposed on the directions of Q1 and E2-1. Moving to the second block, besides the forces G2, Q2, C2 and E3-2, the normal force transmitted by the slice 1, E1-2 = E2-1 should be considered. The known forces E1-2, G2 and C2 are decomposed and their resultant is decomposed on the known directions of Q2 and E3-2. In the same manner, the thrust E transmitted by each block to the next block is found. 294

At the block at the bottom, there are Gn, Cn and En-1-n known in magnitude and direction and Qn known in direction. The polygon of the forces is constructed. There are three possibilities (fig. 10.18):

Fig. 10.18
a. the polygon closes; this situation corresponds to the limiting equilibrium; b. in order to close the polygon, is needed a force E directed in the direction of the slide; the slope is stable, has a reserve of stability proved by the force E; to define the factor of safety FS, the analysis is repeated by gradually increasing the weight of the slices until the situation a of limiting equilibrium is met; the factor of safety is then equal to the factor to which the weight of the slices was increased; c. in order to close the polygon, is needed a force E opposed to the direction of the slide; the slope is not stable and the force E, multiplied by a factor of safety FS could be used for the dimensioning of a wall placed at the bottom of the slope (fig. 10.19).

Fig. 10.19

10.4 Choice of the shear strength parameters to be used in the stability of slopes analysis
The use of the appropriate shear strength parameters, in correlation with the construction conditions, is essential for the analysis of the stability of slopes. In what follows, two examples will be considered.

295

10.4.1 Stability of an embankment on a saturated clay


Even if the slopes of the embankment are stable, the loss of stability can occur along a failure surface originating in the foundation soil (fig. 10.20).

Fig. 10.20
A point M is taken in the possible failure surface. Diagrams of variation in time of the load exerted on the foundation soil by placing successive layers of soil within the embankment, of the shear stress on a plane passing through the point M, of the shear strength f of the clay and the factor of safety FS. The initial water pressure is: uo = w h o (10.30)

where ho is the depth of the considered point below the ground water table. Due to the very low permeability of the clay and to the relatively short construction period of the embankment, it can be assumed that there is no drainage of the pore water and that no significant dissipation is likely during construction. The clay will be loaded in the undrained conditions, responding with the shear strength in undrained conditions. After the end of construction, total stresses in the point under consideration remain unchanged, while the excess pore water pressure dissipates gradually until becoming zero after a time t2. The reduction in the pore water pressure is joined by a decrease of the soil porosity and an increase of the effective stresses and of the shear strength of the foundation soil. Checks of the stability should be made for two characteristic moments of the construction:

a. end-of-construction (time t1)


A total stress analysis is performed using the parameter cu obtained from U U shear tests (unconsolidated-undrained tests) on saturated samples.

b. long-term stability (time t2) 296

An effective stress analysis is performed, using the shear strength parameters , c ' obtained in drained conditions. If one examines the diagram of variation in time of the factor of safety FS = f , (fig. 10.21) realizes that in this example the most dangerous condition corresponds to the end-of-construction time t1. In the long term, following t1, the stability conditions are improving.

Fig. 10.21

10.4.2 Stability of an excavated slope


An excavated slope is considered (fig. 10.22) in a cohesive soil mass. The excavation is joined by a lowering of the water table. A point P is taken on a potential failure surface. As in the previous example, it is assumed that the excavation take place over a short period of time; therefore undrained conditions are valid for the clay. As the excavation proceeds the shear stress in the point P increases.

297

Fig. 10.22
The reduction of the geological pressure on the point P leads to a decrease of the pore water pressure (the pore water pressure change is negative). It was shown that for an increment of isotropic stress associated with a major principal stress increment, the excess pore water pressure is:
n = B 3 + AB ( 1 3 )

(10.31)

For a saturated clay in undrained conditions B = 1


n = 3 + A ( 1 3 )

(10.32)

In the case of an excavated slope, the minimum principal stress 3 decreases more than the maximum principal stress 1 , therefore 3 is negative, ( 1 3 ) is positive and n is negative, with a value depending on A. In time, the excess pore water pressure dissipates producing a heave of the clay and a reduction of the shear strength. Checks of the stability for the two characteristic moments of the construction: a. end-of-construction (time t1) As in the previous example, a total stress analysis is performed, using the parameter cu . b. long-term stability (time t2) An effective stress analysis is performed (by knowing the pore water pressure corresponding to the final position of the ground water) using the parameters ' and c ' . One can realize that, unlike the previous example, the most dangerous conditions appears after a long time t2 because the shear strength decreases until the excess negative pore water pressure become zero (fig. 10.23).

298

Fig. 10.23
In conclusion, when choosing the shear strength parameters the following recommendation have to be followed: for the end-of-construction, condition when the construction means both the loading or the unloading of the saturated clay susceptible to loose its stability, and the construction period is short in comparison to the time required for the dissipation of the excess pore water pressure, a total stress analysis is performed using the parameter cu; similarly, but on partly saturated clay, analysis can be done either in total stresses with parameter u obtained from U U tests or in effective stresses (which implies establishing the parameter pressure) with parameters ' , c ' ; for the long term stability analysis, the effective stress are used, after establishing the pore water pressure corresponding to the final level of the ground water with the parameter ' and c ' .

299

CHAPTER 11. BEARING CAPACITY


In fig. 11.1 is shown a strip footing, which is a shallow foundation supporting a load-bearing wall. When establishing the area A of contact between the foundation and the soil, two fundamental requirements must be satisfied: - to ensure safety against the risk of shear failure of the supporting soil (fig. 11.1 a), - to limit the settlement s of the foundation to values allowable for the structure and for its normal exploitation (fig. 11.1 b).

The problem of bearing capacity, this chapter is dealing with, refers to the first of the two above outlined requirements. Bearing capacity represents the ability of a soil to carry a load. The allowable bearing capacity is defined as the maximum pressure which may be applied to the soil such that the two fundamental requirements are satisfied. The ultimate bearing capacity as defined as the least pressure which would cause shear failure of the supporting soil immediately below and adjacent to a foundation. As shown in the chapter 7, the problem of ultimate bearing capacity is a special case of limiting or plastic equilibrium in a soil mass. In the following paragraphs, the particular problem of the ultimate bearing capacity of shallow foundations will be considered.

11.1 Failure modes


Present knowledge concerning the way in which failure of the soil supporting shallow foundations takes place is based on analysis of both causes of accidents in which various structures lost stability and interpretation of experimental data. The experiments were conducted, in general, at small scale in installations allowing to visualize the trajects followed by soil particles during the process of gradual loading until the failure condition was reached. On that basis, three main modes of failure were recognized, depending in essence on the ground conditions. a. general shear failure

300

Continuous failure surfaces develop between the edges of the footing and the ground surface (fig. 11.2 a). As the pressure is increased towards the value of the ultimate bearing capacity pcr, the state of plastic equilibrium is reached initially in the soil around the edges of the footing then gradually spreads downwards and outwards. Ultimately, the state of plastic equilibrium is fully developed throughout the soil above the failure surfaces. Heaving of the ground surface occur on both sides of the footing although the final slip movement would occur only on one side, accompanied by tilting of the footing, as shown in fig. 11.1 a. The load-settlement diagram, which accompanies this mode of failure is shown in the diagram a in fig. 11.3, puts into evidence clearly the values of the ultimate bearing capacity pf for which deformations increase indefinitely. The transition from the initial, quasi-linear, part of the diagram and the point corresponding to pcr is a short one.

The general shear failure (sometimes named complete shear failure) is typical for soils of low compressibility (dense sands, stiff clays) and for rocks. b. local shear failure In this mode of failure, there is significant compression of the soil under the footing and only partial development of the state of plastic equilibrium. The failure surfaces do not reach the ground surface and tilting of the foundation is unlikely to occur. The loadsettlement diagram (b in the fig. 11.3) shows that the ultimate bearing capacity is not clearly defined and is characterized by the occurrence of relatively large settlements. This mode of failure is associated with soils of medium to high compressibility, (non-cohesive soils of medium relative density, cohesive soils of medium consistency). 301

c. punching shear failure This mode of failure occurs when there is compression of the soil under the footing, accompanied by shearing in the vertical direction around the edges of the footing. As the pressure is increased, the foundation penetrates into the soil like a piston. There is no heaving of the ground surface away from the edges and no tilting of the footing. The loadsettlement diagram (c in fig. 11.3) shows that large settlements are also characteristics to this mode of failure and the ultimate bearing capacity, like in the case b, is not well defined. Punching shear failure, is associated with soils of very high compressibility such as loose sands and soft clays. In cases of local shear and punching shear failures, the ultimate bearing capacity should be defined based on a deformation criterion. Available experimental data show that settlements of shallow foundations corresponding to a failure load are of the order of (3%...7%) B for clay soils and of (5%...15 %) B for sands where B is the width of the foundation. Hence, a settlement of 10% B could be adopted as a deformation criterion for any soil condition in order to define pf (fig. 11.4). It follows also that plate load tests on compressible soils should be conducted to settlements equal to at least 0.25 B, to be able to define the ultimate load from the load-settlement diagram.

302

Besides the nature of the soil, the mode of failure depends also on other factors such as: - the depth of the foundation; punching shear failure will occur in a soil of low compressibility, for instance dense sands, if the foundation is located at considerable depth (deep foundation); - the kind of loading; a dense sand subjected to cyclic loading will exhibit punching shear failure; - the rhythm of loading; a saturated, normally consolidated clay, exhibits a general shear failure under a sudden loading, when no volume change takes place, and a punching shear failure when the rhythm of applying the load is slow and after each load stage the time required for the consolidation of the soil is provided.

11.2 General hypothesis adopted for computing the ultimate bearing capacity
For the computation of the ultimate bearing capacity pcr the following hypothesis are adopted: - a continuous failure surface characteristic for the general shear failure mode (fig. 11.5);

- the failure condition f = tan + c is fulfilled in each point of the failure; - the shear strength of the soil between the level of the foundation and the ground surface (part CD of the failure surface) is neglected; - the friction between the soil above the level of the foundation and the lateral face of the foundation (EB) is neglected; - the friction between the soil located above and below the foundation level (on the line BC) is neglected; - the friction between the base of the foundation (AB) and the soil to which it c.. in contact, is neglected. With these hypothesis, the soil located above the foundation level is replaced by a surcharge q = D, where D is the foundation depth.

303

11.3 Ultimate bearing capacity in the case of a failure surface made by two planes
The two failure planes (fig. 11.6) have the inclinations in respect to the horizontal of (45o + ) and (45o ) , corresponding to the development in the mass of soil under the 2 2 footing of two Rankine zones on both sides of a imaginary, fictitious, perfectly smooth (frictionless) wall BD, namely the active zone on the left of the wall and the passive zone on the right of the wall.

Computing pcr is based on expressing the active earth thrust Pa behind a vertical wall BD limited by an horizontal ground surface on which a surcharge pcr is applied and the passive resistance Pp in front of the same wall, limited by an horizontal ground surface on which a surcharge q = D is applied (fig. 11.7).

1 Pa = H 2 K a + p f H K a 2 c H K a 2

(11.1 a)

304

1 Pp = H 2 K p + q H K p + 2 c H K p 2

(11.1. b)

To find pf, the condition Pa = Pp is written, considering that:


H = B tan (45o + ) = B K p 2
Ka = 1 Kp

1 1 pf 1 1 H + 2c = H Kp + q Kp + 2c Kp 2 Kp Kp Kp 2

Kp 1 1 2 p f = H K2 + q K + 2 c K K H + 2 c = p p p p 2 2 Kp 1 2 = B Kp K2 p Kp + q Kp + 2c Kp Kp + Kp = 2
5 3 1 1 1 2 +K 2) = B (K p 2 K p 2 ) + q K 2 + 2 c ( K p p p 2

(11.2)

The expression (11.2) can be put into the form:


1 p f = q Nq + c Nc + B N 2

(11.3)

where N q , N c , N , named bearing capacity factors, are depending on the angle of internal friction , and have the following expressions:
5 5 1 1 2 2 2 2 N q = K p ; N c = 2 (K p + K p ); N = (K p K p 2 )

(11.4)

11.4 Ultimate bearing capacity in the case of a curved failure surface


The problem is solved in three phases, corresponding to the following conditions: a. cohesionless, weightless soil ( 0; c = 0; = 0) b. frictionless, weightless soil ( c 0; = 0; = 0 ) c. soil with weight ( 0 ) a. In the case of a soil without cohesion and weight, a suitable failure mechanism for a strip footing is shown in fig. 11.8. The footing, of width B and infinite length, carries a 305

uniform pressure on the surface of a mass of homogeneous, isotropic soil. When the pressure becomes equal to the ultimate bearing capacity pcr the footing will have pushed downwards into the soil mass producing a state of plastic equilibrium, in the form of an active Rankine zone, below the footing, the angles ABC and BAC being ( 45o + ). The 2 downward movement of the wedge ABC forces the adjoining soil sideways, producing outward lateral forces on both sides of the wedge. Passive Rankine zones ADE and BGF develop on both sides of the wedge ABC, the angles DEA and GFB being ( 45o ). The 2 transition between the downward movement of the wedge ABC and the lateral movement of the wedges ADE and BGF takes place through zones of radial shear ACD and BCG. In his solution, Prandtl admits that the surfaces CD and CG are logarithmic spirals, to which BC and ED, or AC and FG, are tangential. The equation of the spiralis r = ro e tan where is the angle between the initial radius ro and the one corresponding to a point on the spiral; is the angle made by the radius with the normal in any point of the spiral. A state of plastic equilibrium exists above the surface EDCGF, the remainder of the soil mass being in a state of elastic equilibrium.

To find pcr, first the equilibrium of the wedges ABC and BDE, as equilibrium of forces on vertical direction, will be considered. Then, the equilibrium of the transition zone BCD, as equilibrium of moments toward the point B, will be written. On the conjugated failure planes AC and CB are acting the reactions RI, making an angle with the normal (fig. 11.9 a). The equation of projection of forces on the vertical direction:
p f AB = 2 R I cos (45o ) 2 AB = 2 ro cos (45 o + ) 2 p f 2 ro cos (45o + ) = 2 R I cos (45o ) 2 2

306

cos (45o + ) 2 = p r tan (45o ) R I = p f ro f o 2 cos (45o + ) 2

(11.5)

On the conjugated failure planes BD and DE are acting the reactions RIII, making an angle with the normal (fig. 11.9 b). The equation of projection of forces on the vertical direction:
q BE = 2 R III cos (45 o + ) 2 BE = 2 r1 cos (45 o ) 2 q 2 r1 cos (45o ) = 2 R III cos (45 o + ) 2 2

cos (45o ) 2 = q r tan (45 o + ) R III = q r1 1 2 o cos (45 + ) 2

(11.6)

The equilibrium of the transition zone II (fig. 11.9 c) is expressed in terms of the moment around the point B.

The arc of the spiral CD belongs to the failure surface, therefore the reaction RII makes an angle with the normal to the arc. Hence, the direction of RII coincides with the direction of the radius and RII produces no moment in respect to B. The moment equation becomes:

307

r r R I cos o = R III cos 1 2 2 2 2 p f ro tan (45o ) = q r1 tan (45o + ) 2 2


tan But r1 = ro e 2

2 2 tan p f ro tan (45o ) = q ro e tan (45o + ) 2 2 p f = q e tan tan 2 (45o + ) 2 By writing: e tan tan 2 (45o + ) = N q 2

(11.7)

equation (11.7) becomes: pf = q Nq (11.8)

From (11.8) follows that, in the case of a cohesionless and weightless material, there is a bearing capacity only if there is a surcharge q. To consider the effect of the cohesion, a normal stress equal to c cot is added to the normal stresses p and q. The equation (11.8) becomes:
p f + c cot = (q + c cot ) N q p f = q N q + c cot ( N q 1)

(11.9)

By writing cot ( N q 1) = N c equation (11.9) becomes: pf = q Nq + c Nc Nq and Nc are bearing capacity factors depending on . An additional term should be added to equation (11.10) to take into account the selfweight of the soil. Experimental observations showed that a wedge of soil remaining in elastic state, with faces making an angle with the horizontal, is developed below the foundation and moves downwards together the foundation, tending to produce the lateral 308

movement of the soil along the failure surfaces CDE and CFG (fig. 11.10). The passive resistance of the soil mass above the failure surfaces is mobilized. The problem consists on computing the passive resistance force Pp of a mass of soil ( 0, 0 ), limited by a horizontal ground surface, behind a wall BC with inclination and height H = B tan . 2

The failure surface CDE is made of the line DE, corresponding to the passive Rankine zone BDE and by the arc of logarithmic spiral CD. The passive resistance force Pp can be expressed:
1 1 B2 1 2 Pp = H K p = tan 2 K p = B 2 tan 2 2 2 4 8

(11.11)

The equilibrium of the elastic wedge ABC:


Q = 2 Pp cos ( )

(11.12)

The ultimate bearing capacity is:


Q 2Pp 2 B 2 tan 2 cos ( ) pf = = cos ( ) = = B B 8B 1 1 = B tan 2 cos ( ) = BN 4 2

(11.13)

The following notation was used:


1 N = tan 2 cos ( ) 2

309

Terzaghi assumed that = and obtained the value of the passive resistance force in the hypothesis of a curved failure surface with the method presented in the chapter 8. Adding the additional term bringing the effect of the self-weight of the soil, the expression of the ultimate bearing capacity pf becomes:
1 p f = q Nq + c Nc + B N 2

(11.14)

Relations of the kind of (11.14) were established by Terzaghi and other authors. Most of them differ only with respect of the third component, introducing the influence of the self-weight of the soil. These relations are theoretically incorrect for a plastic material since they are superposing terms corresponding to different failure figures such as those represented in fig. 11.8 and 11.10. However, the error implied is considered to be on the safe side and is accepted in engineering practice.

11.4 Ultimate bearing capacity in the case of a purely cohesive soil


This is a particular case of the problem previously considered. The failure mechanism show in fig. 11.8 is transformed, when = 0 , in the one shown in fig. 11.11. Equation (11.10) becomes:
p f = c Nc + D Nq = c Nc + D

(11.15)

(For = 0 , Nq = 1)

One defines as netto ultimate bearing capacity the difference between the critical pressure in the geological pressure at the level of the foundation base:
pf
netto

= p f D = c Nc

(11.16)

The problem is to find the bearing capacity factor Nc for this case ( c 0, = 0, = 0) . An approach similar to the one used for the case ( 0, c = 0, = 0) is adopted:

310

Equilibrium of forces acting on the prism I (fig. 11.12 a)


B 2 2 c Nc B = 2 R I + 2c 2 =RI 2 +cB 2 2 2 2

RI =

1 c B ( N c 1) 2

(11.17)

The normal stress acting on the faces AC and BC:


c B ( N c 1) 2 = = c ( N c 1) B 2 2 2

pI =

RI B 2 2 2

(11.18)

Equilibrium of forces acting on the prism III (fig. 11.12 b)


B 2 2 2 R III = 2c 2 2 2 2 2 R III = cB 2

(11.19)

311

The normal stress acting on the faces BD and DE:


cB 2 =c p III = B 2 2 2

(11.20)

pf is obtained by writing the condition that the moment of all forces acting on the failure prism, in respect to the point B, is zero. Normal pressures acting on the circular are CD having the direction of the radius, do not give moment toward B.
cDE BD+ c AC = BD RD + c AC BC + c ( N c 1) AC 2 2 AB DE = 2 N c AB + c DE 2 2

(11.21)

But AC = BC = BD = DE = r
AB = 2 r 2 =r 2 2

Relation (11.21) becomes:


r2 + r 2 2 r2 r2 2 r + ( N c 1) = N c + ; 2 2 r 2

2 2 r2 2 r 2 r + r + ( N c 1) = N c r + 2 2 2 N 1 1 2 + + c = Nc + 2 2 2 N 2 + 1 = c 2 2 2

N c = + 2 = 5.14 pf = 5.14 c
netto

(11.22) (11.23) (11.24)

p f = 5.14 c + D

Skempton has shown that, in fact, the netto ultimate bearing capacity increases with the depth D of the foundation until a depth D = 5B (fig. 11.13), reaching a limit value 9 for Nc. 312

For rectangular foundations B x L, for which relation:


pf D B = 5 c (1+ 0.2 ) (1+ 0.2 ) netto B L

D 2.5 , Skempton proposed the B

(11.25)

313

You might also like