Chapter 1
Chapter 1
COMPRESSIBILITY OF SOIL
2.1 Introduction
2.2 Elastic Settlement
2.3 Primary Consolidation Settlement
2.4 Secondary Consolidation Settlement
2.5 Precompression
3.1 Introduction
3.2 Typical Response of Soil to Shearing Forces
3.3 Four Models to Interpret shear strength
3.4 Practical Implications of Failure Criteria
3.5 Laboratory Tests for Shear Strength Parameters
3.6 Sensitivity and Thixotropy
3.7 Strength Anisotropy of Clay
3.8 Field Test
3.9 Practical applications of shear strength
4.1 Introduction
4.2 Earth Pressure At-Rest
4.3 Rankine’s Lateral Earth Pressure
4.4 Coulomb’s Earth Pressure Theory
4.5 Earth Pressure Using Theory of Plasticity
4.6 Common Types of Retaining Walls
4.7 Braced Excavation
SLOPE STABILITY
5.1 Introduction
5.2 Types of Slope Failure
5.3 Causes of Slope Failure
5.4 Stability of Infinite Slopes
5.5 Finite Slopes
6.1 Introduction
6.2 Soil Response to a Loaded Footing
6.3 Conventional Failure Surface Under a Footing
6.4 Collapse Load Using the Limit Equilibrium Method
6.5 Bearing Capacity Equations
6.6 Which Equations to Use
6.7 Concentric Loadings
6.8 Eccentric Loadings
6.9 Inclined Loadings
6.10 Effect of Water Table on Bearing Capacity
6.11 Bearing Capacity from SPT
6.12 Bearing Capacity of Foundation with Uplift Forces
7.1 Introduction
7.2 Types of Piles and Installation
7.3 Vertical Load Capacity of Single Pile for Driven Piles
7.4 Vertical Load Capacity of Single Pile for Drilled Shafts
7.5 Uplift Capacity of Single Piles
7.6 Negative Skin Friction of Single Piles
7.7 Laterally Loaded of Single Piles
7.8 Pile Groups
7.9 Pile Settlements
Let us assume that under these forces the cube compressed by ∆𝑥, ∆𝑦, and ∆𝑧 in the X, Y, and Z
directions.
∆z ∆x ∆y
ε z= ε x= ε y=
z x y
1.1.2 Volumetric Strain
ε p=ε x +ε y + ε z
Let us consider, for simplicity, the XZ plane and apply a force F that causes the square to distort
into a parallelogram, as shown in figure below. Simple shear strain is a measure of the angular
distortion of a body by shearing forces.
F ∆x ∆x
τ= ; y zx =tan−1 ≈
xy z z
Stresses and strains for a linear, isotropic, elastic soil are related through Hooke’s law.
{} [ ]{ }
εx 1 −v −v 0 0 0 σx
εy −v 1 −v 0 0 0 σy
εz 1 −v −v 1 0 0 0 σz
=
γ xy E 0 0 0 2(1+ v) 0 0 τ xy
γ yz 0 0 0 0 2(1+v ) 0 τ yz
γ zx 0 0 0 0 0 2(1+ v) τ zx
The matrix on the right-hand side of equation below is called the stiffness matrix.
{ } [ ]{ }
∆σ1 1−v v v ε1
E
∆σ2 = v 1−v v ε2
(1+ v )(1−2 v )
∆σ3 v v 1−v ε 3
∆ z=∫ ε z dz
P z=∫ ∆ σ z dA
1+ v 1+ v
ε 1=
E
[ ( 1−v ) σ 1−v σ 3 ] ; ε 3=ε 1=
E
[ ( 1−v ) σ 3−v σ 1 ]
σ 2=v (σ 1 +σ 3 )
{} ε 1 1+ v 1−v
ε3
=
E v [ v ∆ σ1
1−v ∆ σ 3 ]{ }
{ }
∆σ1
=
E 1−v
∆ σ 3 ( 1+ v )(1−2 v ) v
v
[ ε1
1−v ε 3 ]{ }
Note:
Compression stresses or strain is positive
Tension stress or strain is negative
{} [
ε 1 1 1 −2 v ∆ σ 1
=
ε 3 E −v 1−v ∆ σ 3 ]{ }
{ }
∆σ1
=
E
[
1−v 2 v ε 1
∆ σ 3 (1+ v )(1−2 v ) v 1 ε3 ]{ }
Anisotropic materials have different elastic parameters in different directions. Anisotropy in soils
results from essentially two causes.
1. The manner in which the soil is deposited. This is called structural anisotropy and it is
the result of the kind of soil fabric that is formed during deposition. You should recall
that the soil fabric produced is related to the history of the environment in which the soil
is formed. A special form of structural anisotropy occurs when the horizontal plane is a
plane of isotropy. We call this form of structural anisotropy transverse anisotropy.
[ ]{ }
1 −2 v rz
{ }
∆ εz
∆ εr
=
Ez
−v zr
Er ∆σz
1−v rr ∆ σ r
Ez Er
v rz E r
=
v zr E z
Let us consider an element of a saturated soil subjected to a normal stress, 𝝈, applied on the
horizontal boundary, as shown in figure below. The stress 𝝈 is called the total stress, and for
equilibrium the stresses in the soil must be equal to 𝝈. The resistance or reaction to 𝝈 is provided
by a combination of the stresses from the solids, called effective stress (𝝈′), and from water in the
pores, called porewater pressure (u).
σ =H soil γ soil + H w γ w
Principle of effective stress and was first recognized by Terzaghi in the mid-1920s
during his research into soil consolidation. The principle of effective stress is the most
important principle in soil mechanics.
Deformations of soils are a function of effective stresses, not total stresses. The principle
of effective stresses applies only to normal stresses and not to shear stresses.
The effective stress is not the contact stress between the soil solids. Rather, it is the
average stress on a plane through the soil mass.
'
σ =σ +u
1.7.2 Effects of Seepage
As water flows through soil, it exerts a frictional drag on the soil particles, resulting in head losses.
The frictional drag is called seepage force in soil mechanics. It is often convenient to define
seepage as the seepage force per unit volume (it has units similar to unit weight), which we will
denote by j s. If the head loss ∆ h over a flow distance, L, the seepage force j s is
∆ hγw
j s= =i γ w
L
If seepage occurs downward, then the seepage stresses are in the same direction as the
gravitational effective stresses.
u z=γ w z−i γ w z
If seepage occurs upward, then the seepage stresses are in the opposite direction to the
gravitational effective stresses.
u z=γ w z +i γ w z
Let us consider the upward flow of water through a soil sample as shown in figure below.
Such a boiling condition will become imminent if the upward water force just equals the
weight of the material acting downward.
h G−1
i= =
L 1+ e
In silts and fine sands, the soil above the groundwater can be saturated by capillary
action. You would have encountered capillary action in your physics course when you
studied meniscus. We can get an understanding of capillarity in soils by idealizing the
continuous void spaces as capillary tubes.
Consider a single idealized tube, the height at which water will rise in the tube can be
found from statics. Summing forces vertically, we get
( π4 d ) h γ =πdTcos α
2
c w
4 Tcos α 4T
h c γ w= h c=
d γw dγ w
Where:
T = Surface Tension
The value of surface tension (𝑻) for water varies with temperature. At ordinary or room
temperature, 𝑻 is nearly 7.3 dynes/mm or 73 × 10 -6 N/mm and 𝛾w may be taken as 9.81 × 10 -6
N/mm3
30
hc=
d
where d is the diameter of the glass capillary in mm, and h c is the capillary rise of water
in the glass tube in mm.
1.8 STRESSES IN SOIL FROM SURFACE LOADS
∆ σ x=
{
P 2 x2 z
2 π L5
−(1−2 v) 2
[
x 2− y 2
+
y2 z
3 2
Lr ( L+ z ) L r ]}
∆ σ y=
{
P 2 y2 z
2 π L5
−(1−2 v)
[
y 2−x 2
+
x2 z
Lr 2 ( L+ z ) L3 r 2 ]}
3 3
3P z z
∆ σz= ∙ 5=
2 π L 2 π ( x + y 2 + z 2 )5/ 2
2
2 q z3
∆ σz= 2
π (x +z )
2 2
2
2q x z
∆ σ x= 2
π (x +z )
2 2
2 q xz 2
∆ σ y= 2
π(x +z )
2 2
The fundamental equation for the vertical stress increase at a point in a soil mass as the
result of a line load can be used to determine the vertical stress at a point caused by a flexible strip
load of width B.
3
2 ( q dr ) z
d σ z=
2 2
π [ ( x −r ) + z ]
2
{[ }
+B /2
∫ ( 2πq )
3
z
∆ σ z =∫ d σ z= 2
dr
( x−r ) + z 2 ]
2
−B /2
Valid for x in Positive Direction Only
For x greater than B/2
{ [ ( )]
}
2 2 B2
Bz x −z −
[ ( )] [ ( ) ] [
q −1 z −1 z 4
∆ σ z = tan −tan −
( )] + B z
π B B 2
2
x− x+ 2 B 2 2 2
2 2 x +z −
4
{ [ ( )]
}
2
B 2 2
Bz x −z −
[ ( )] [
q π z 4
∆ σz= −tan−1 −
( )] + B z
π 2 B 2
2
x+ 2 B 2 2 2
2 x +z −
4
For x less than B/2
{ [ ( )]
}
2
B
Bz x 2−z 2−
[ ( ) ] [ ( )] [
q z z 4
∆ σz= π +tan −1 −tan −1 −
( )] + B z
π B B 2
2
x− x+ 2 B 2 2 2
2 2 x +z −
4
Figure below shows the cross section of an embankment of height H. For this two-
dimensional loading condition, the vertical stress increase may be expressed as:
q0
∆ σz= ¿
π
α 1=tan
−1
( B1 + B 2
B2 )
−¿ α 2 ¿
α 2=tan
−1
( Bz ) 1
1.8.6 Vertical Stress Below the Center of a Uniformly Loaded Circular Area
Using Boussinesq’s solution for vertical stress ∆𝜎z caused by a point load, one also can
develop an expression for the vertical stress below the center of a uniformly loaded flexible
circular area
3( qr dr dα ) z3
d σ z=
2π ( r 2+ z 2 )
5 /2
α =2 π r= R 3
3q z
∆ σ z =∫ dd σ z= ∫ ∫ dr dα
α =0 r=0 2 π ( r 2 + z 2 )5 /2
{ [( ) ] }
1
∆ σ z =q 1− 3
R 2 2
+1
z
1.8.7 Vertical Stress at Any point Below a Uniformly Loaded Circular Area
A detailed tabulation for calculation of vertical stress below a uniformly loaded flexible
circular area was given by Ahlvin and Ulery (1962). Referring to figure below, we find that ∆𝜎z at
any point A located at a depth z at any distance r from the center of the loaded area can be given as
∆ σ z =q (A ' + B' )
Boussinesq’s solution also can be used to calculate the vertical stress increase below a
flexible rectangular loaded area, as shown in figure below. The loaded area is located at
the ground surface and has length L and width B.
The uniformly distributed load per unit area is equal to q. To determine the increase in the
vertical stress (∆𝜎z) at point A, which is located at depth z below the corner of the
rectangular area, we need to consider a small elemental area dx dy of the rectangle. The
load on this elemental area can be given by.
3
3 q dx dy z
dq=q dx dy ; ∆ σ z = 5
2 π( x + y +z )
2 2 2 2
B L 3
∆ σ z =∫ d σ z= ∫ ∫ 32qπz dx dz
5
=q I 3
y=0 x=0
(x + y + z )
2 2 2 2
At the corner
[ (
1 2 mn √ m + n +1 m + n +2
)
−1 2mn √ m +n +1
( )]
2 2 2 2 2 2
I 3= 2 2 2 2 2 2
+ tan 2 2 2 2
4 π m +n +m n + 1 m + n +1 m +n −m n +1
B L
m= n= ∆ σ z=q I 3
z z
−1
If tan () become negative add π to the equation
At the center
2
I 4= ¿
π
L z B
m 1= n1= b= ∆ σ z=q I 4
B b 2
1.8.8 Vertical Stress Caused by a Rectangular Loaded Area using Boston Rule (2:1
Method)
For uniform footing (B x L) we can estimate the change in vertical stress with depth
using the Boston Rule. Assumes stress at depth is constant below foundation influence area
q(B × L)
∆ σz=
( B+ z)( L+ z )
2.1 INTRODUCTION
A stress increase caused by the construction of foundations or other loads compresses soil layers.
The compression is caused by (a) deformation of soil particles, (b) relocations of soil particles,
and (c) expulsion of water or air from the void spaces. In general, the soil settlement caused by
loads may be divided into three broad categories:
The settlement profile and contact pressure distribution described are true for soils in which the
modulus of elasticity is fairly constant with depth.
P=P s + Pw
Ps = Load on Spring
Pw = Load on Water
2.2 ELASTIC SETTLEMENT
Elastic settlement takes place in a short period of time after the application of load and is
due to distortion of soil. As the settlement is experience in a short period of time, there will be not
enough time for the soil mass to change its water content.
2
' 1−v
Se =∆ σ ( α β ) Is If
Es
Where:
∆ σ = Net applied pressure on the foundation
v = Poison's ratio of soil
E s = Average modulus of elasticity of the soil under soil the foundation measured
from z = 0 to about z = 4B
B' = B/2 for center of foundation
B' = B for corner of foundation
B = Least Dimension of Foundation
1−2 v
I s= Shape Factor (Steinbrenner, 1934) = F 1+ F
1−v 2
'
1
F 1= ( A 0 + A1 ) and F 2= n ( tan −1 A 2 )
π 2π
(m + √m +1) √ 1+ n
' '2 '2
A1=ln
m' + √ m'2 +n '2 +1
'
m
A 2=
n √ m +n + 1
' '2 '2
I f = depth factor (Fox, 1948)
I f =f ( DB , v ,∧LB )
f
The soil specimen is placed inside a metal ring with two porous stones, one at the top of
the specimen and another at the bottom. The specimens are usually 64 mm (2.5 in.) in
diameter and 25 mm (1 in.) thick.
The load on the specimen is applied through a lever arm, and compression is measured by a
micrometer dial gauge.
The specimen is kept under water during the test. Each load usually is kept for 24 hours.
After that, the load usually is doubled, which doubles the pressure on the specimen, and
the compression measurement is continued.
At the end of the test, the dry weight of the test specimen is determined.
2.3.2 Void Ratio – Pressure Plot
After the time – deformation plots for various loadings are obtained in the laboratory, it is
necessary to study the change in the void ratio of the specimen with pressure.
V
e= −1
Vs
1+ e ∆ e
=
H ∆H
∆ e=e0 −e 1
A soil in the field at some depth has been subjected to a certain maximum effective past
pressure in its geologic history. This maximum effective past pressure may be equal to or less
than the existing effective overburden pressure at the time of sampling.
The reduction of effective pressure in the field may be caused by natural geologic processes or
human processes. During the soil sampling, the existing effective overburden pressure is also
released, which results in some expansion.
When this specimen is subjected to a consolidation test, a small amount of compression (that is,
a small change in void ratio) will occur when the effective pressure applied is less than the
maximum effective overburden pressure in the field to which the soil has been subjected in the
past.
When the effective pressure on the specimen becomes greater than the maximum effective past
pressure, the change in the void ratio is much larger, and the e-log a relationship is practically
linear with a steeper slope.
Normally Consolidated – whose present effective overburden pressure is the maximum pressure
that the soil was subjected to in the past.
Overconsolidated – whose present effective overburden pressure is less than that which the soil
experienced in the past. The maximum effective past pressure is called the preconsolidation
pressure.
'
σC
OCR= '
σ
Where:
σ 'C = preconsolidation pressure of a specimen
'
σ = present effective vertical pressure
log σ 'C =
1.112− ( ) e0
eL
'
0.0463 σ 0
0.188
Where:
e 0 = in situ void ratio
e L = void ratio at liquid limit =
[ ]
¿ (% )
100
Gs
Gs = specific gravity of soil solids
'
σ 0 = in situ effective overburden pressure
Hansbo
'
σ C =α VST C u (VST )
Where:
222
α VST = an empirical coefficient =
¿ ( %)
C u (VST ) = undrained shear strength obtained from vane shear test
In any case, these above relationships may change from soil to soil. They may be taken as an
initial approximation
2.3.4 Calculation of One-Dimensional Primary Consolidation
According to Karl von Terzaghi "consolidation is any process which involves a decrease in water
content of saturated soil without replacement of water by air." In general, it is the process in which
reduction in volume takes place by expulsion of water under long term static loads.
Let us consider a saturated clay layer of thickness H and cross-sectional area A under an existing
'
average effective overburden pressure, σ 0. Because of an increase of effective pressure, ∆ σ ' , let
the primary settlement be Sc.
SC =
Cc H
1+e 0
log
(
σ '0 + ∆ σ '
σ0
' )
Overconsolidated Soil
( )
' '
CS H σ 0 +∆ σ
SC = log
1+ e0 '
σ0
( ) ( )
' ' '
C H σ C H σ +∆ σ
SC = S log C' + C log 0 '
1+ e0 σ 0 1+ e 0 σC
Where:
C C = Compression Index
C S = Swell Index
Compression Index
The compression index for the calculation of field settlement caused by consolidation can be
determined by graphic construction after one obtains the laboratory test results for void ratio and
pressure.
Skempton
C C =0.009(¿−10) – For Undisturbed Clay
C C =0.007(¿−7) – For Remolded Clay
Rendon – Herrero
( )
2.38
1.2 1+ e0
C C =0.141 G s
GS
Swell Index
The swell index is appreciably smaller in magnitude than the index and generally can be
determined from laboratory tests. In most cases
1 1
C S= ¿ C C
5 10
Nagaraj and Murty
C S=0.0463 [ ]
¿ (%)
100
GS
∆e ∆e
C a= =
log t 2 −log t 1 t2
log
t1
Where:
C a = secondary compression index
∆ e = change of void ratio
t 2 = time after the completion of SC
t 1 = time for completion of SC
SS =C 'a H log
()t2
t1
' Ca
C a=
1+ e p
Where:
e p = void ratio at the end of the primary consolidation
'
C a = modified secondary compression index
H = thickness of clay layer
2.4.1 Time Rate of Consolidation
The total settlement caused by primary consolidation resulting from an increase in the stress
on a soil layer can be calculated using one of the three equations given. However, they do not
provide any information regarding the rate of primary consolidation. Terzaghi (1925) proposed
the first theory to consider the rate of one-dimensional consolidation for saturated clay soils. The
mathematical derivations are based on the following six assumptions.
Where:
π
M = ( 2 m+1 )
2
u0 = initial excess pore water pressure
H dr = length of maximum drainage path
T v = time factor
C v = coefficient of consolidation
[(
]
( 2 m+1 ) ) z
π
( )
2
∞ π Cv t
2u 0 2 − ( 2 m +1)
u=∑
2
2 H dr
sin e
π H dr
m=0
( 2 m+ 1 )
2
Time Factor
Cvt k
T v= 2
; C v=
H dr
γ w mv
Approximation
0 ≤ U ≤ 60 %
( )
2
π U%
T v=
4 100
0 ≤ U ≤ 60 %
Degree of consolidation
u0−u z uz
U z= =1−
u0 u0
Sd ( t )
U=
SC
Where:
u z = excess pore water pressure at time t
Sd (t ) = settlement of layer at time t
Sd (t ) = ultimate settlement of the later from primary consolidation
U = average degree of consolidation
U z = degree of consolidation at a distance z at any time t
2.4.2 Coefficient of Consolidation
For a given load increment on a specimen, two graphical methods commonly are used for
determining cv from laboratory one-dimensional consolidation tests. The first is the logarithm of-
time method proposed by Casagrande and Fadum (1940), and the other is the square-root-of-time
method given by Taylor (1942). More recently, at least two other methods were proposed. They
are the hyperbola method (Sridharan and Prakash, 1985) and the early-stage log-t method
(Robinson and Allam, 1996).
0.197 H 2Dr
C v=
t 50
2
0.848 H Dr
C v=
t 90
m H 2Dr
C v =0.3
D
Obtain the time t and the specimen deformation (ΔH) from the laboratory consolidation
test.
Plot the graph of t/ΔH against t
Identify the straight-line portion bc and project it back to point d.
Determine the intercept D.
Determine the slope m of the line bc.
2.4.2.4 Early-Stage log-t Method
The early-stage log-t method, an extension of the logarithm-of-time method, is
based on specimen deformation against log-of-time plot According to this method, follow
Steps 2 and 3 described for the logarithm-of-time method to determine d 0. Draw a
horizontal line DE through d0. Then draw a tangent through the point of inflection, F. The
tangent intersects line DE at point G. Determine the time t corresponding to G, which is
the time at U = 22.14%. the early-stage log-t method may provide a more realistic value
of fieldwork.
m H 2Dr
C v =0.3
D
2.5 PRECOMPRESSION
A surcharge is applied at the ground surface. This surcharge will increase the pore water
pressure in the clay. The excess pore water pressure in the clay will be dissipated by drainage-both
vertically and radially to the sand drains-which accelerates settlement of the clay layer.
Prefabricated vertical drains (PVDs), which also are referred to as wick or strip drains,
originally were developed as a substitute for the commonly used sand drain.
With the advent of materials science, these drains are manufactured from synthetic polymers
such as polypropylene and high-density polyethylene. PVDs normally are manufactured with
a corrugated or channeled synthetic core enclosed by a geotextile filter.
Installation rates reported in the literature are on the order of 0.1 to 0.3 m/s, excluding
equipment mobilization and setup time. PVDs have been used extensively in the past for
expedient consolidation of low permeability soils under surface surcharge. The main
advantage of PVDs over sand drains is that they do not require drilling and, thus, installation
is much faster.
When highly compressible, normally consolidated clayey soil layers lie at a limited depth and
large consolidation settlements are expected as a result of the construction of large buildings,
highway embankments, or earth dams, precompression of soil may be used to minimize
postconstruction settlement.
( )
'
Cc H σ +∆σ
SC =S p= log 0 ' ( p)
1+ e 0 σ0
Sc =S( p+ f ) =
Cc H
1+e 0
log
(
σ '0 + [ ∆ σ ( p )+ ∆ σ (f ) ]
σ0
' )
Where:
∆ σ ( p ) = proposed structural load
∆ σ (f ) = surcharge
Sd ( t )
U=
Sc
U=
log
( σ '0 +∆ σ (p )
σ '0 )
( σ 0 + [ ∆ σ ( p )+ ∆ σ ( f ) ]
)
'
log
σ '0
2
T v H Dr
t 2=
Cv
3.1 INTRODUCTION
3.2 TYPICAL RESPONSE OF SOIL TO SHEARING FORCE
The shear strength of a soil mass is the internal resistance per unit area that the soil mass can offer
to resist failure and sliding along any plane inside it. One must understand the nature of shearing
resistance in order to analyze soil stability problems, such as bearing capacity, slope stability, and
lateral pressure on earth-retaining structures.
We are going to describe the behavior of two groups of soils when they are subjected to
shearing forces. One group, called uncemented soils, has very weak interparticle bonds.
The other group, called cemented soils, has strong interparticle bonds through ion
exchange or substitution. The particles of cemented soils are chemically bonded or
cemented together. An example of a cemented soil is caliche, which is a mixture of clay,
sand, and gravel cemented by calcium carbonate
Let us incrementally deform two samples of soil by applying simple shear deformation
(shown if figure) to each of them. One sample, which we call Type I, represents mostly loose
sands and normally consolidated and lightly overconsolidated clays(OCR ≤ 2). The other, which
we call Type II, represents mostly dense sands and overconsolidated clays (OCR > 2).
We are going to summarize the important features of the responses of these two groups of
soils—Type I and Type II— when subjected to a constant vertical (normal) effective stress and
increasing shear strain.
Type I Soils
Show gradual increase in shear stresses as the shear strain increases (strain-hardens) until
an approximately constant shear stress, which we will call the critical state shear stress,
τ cs, is attained.
Compress, that is, they become denser until a constant void ratio, which we will call the
critical ratio, e cs, is reached.
Type II Soils
Show a rapid increase in shear stress reaching a peak value, τ p, at low shear strains
(compared to Type I soils) and then show a decrease in shear stress with increasing shear
strain(strain-softens), ultimately attaining a critical state shear stress.
The strain-softening response generally results from localized failure zones called shear
bands. These shear bands are soil pockets that have loosened and reached the critical state
shear stress. Between the shear bands are denser soils that gradually loosen as shearing
continues.
When a shear band develops in some types of overconsolidated clays, the particles become
oriented parallel to the direction of the shear band, causing the final shear stress of these clays to
decrease below the critical state shear stress. We will call this type of soil Type II-A, and the final
shear stress attained the residual shear stress, 𝜏r.
For Type I soils, the amount of compression and the magnitude of the critical state shear
stress will increase. For Type II soils, the peak shear stress tends to disappear, the critical
shear stress increases, and the change in volume expansion decreases.
An approximate straight line (OA, in figure shown) that links all the critical state shear
'
stress values of Type I and Type II soils. We will call the angle between OA and the σ n
axis the critical state friction angle, Φ cr .The line OA will be called the failure
envelope because any shear stress that lies on itis a critical state shear stress.
Curve (OBCA, in figure shown) that links all peak shear stress values for Type II soils.
We will call OBC (the curved part of OBCA) the peak shear stress envelope because
any shear stress that lies on it is a peak shear stress.
3.2.2 Effects of Overconsolidation Ratio
The initial state of the soil dictates the response of the soil to shearing forces. For
example, two overconsolidated homogeneous soils with different overconsolidation ratios but the
same mineralogical composition would exhibit different peak shear stresses and volume
expansion, as shown in figure. The higher overconsolidated soil gives a higher peak shear strength
and greater volume expansion.
3.2.3 Effects of Drainage of Excess Porewater Pressure
The rate of loading under the undrained condition is often much faster than the rate of
dissipation of the excess porewater pressure, and the volume-change tendency of the soil is
suppressed. The result of this suppression is a change in excess porewater pressure during
shearing. A soil with a tendency to compress during drained loading will exhibit an increase in
excess porewater pressure under undrained condition, resulting in a decrease in effective stress.
A soil that expands during drained loading will exhibit a decrease in excess porewater
pressure (negative excess porewater pressure) under undrained condition, resulting in an increase
ineffective stress. These changes in excess porewater pressure occur because the void ratio does
not change during undrained loading; that is, the volume of the soil remains constant.
Nearly all natural soils have some degree of cementation, wherein the soil particles are
chemically bonded. The degree of cementation can vary widely, from very weak bond
strength (soil crumbles under finger pressure) to the bond strength of weak rocks.
Cemented soils possess shear strength even when the normal effective stress is zero. They
behave much like Type II soils except that they have an initial shear strength, C cm, under
zero normal effective stress. The shear strength from cementation is mobilized at small
shear strain levels (0.001%).
In this section, we will examine four soil models to help us interpret the shear strength of
soils. A soil model is an idealized representation of the soil to allow us to understand its response
to loading and other external events. By definition, then, a soil model should not be expected to
capture all the intricacies of soil behavior. Each soil model may have a different set of assumptions
and may only represent one or more aspects of soil behavior.
Soils, in particular granular soils, are endowed by nature with slip planes. Each contact of
one soil particle with another is a potential micro slip plane. Loadings can cause a number of these
micro slip planes to align in the direction of least resistance. Thus, we can speculate that a possible
mode of soil failure is slip on a plane of least resistance.
H=μW
' −1 H −1
ϕ =tan =tan μ
W
Coulomb’s Law
τ f =(σ 'n)f tan ϕ '
Taylor (1948) used an energy method to derive a simple soil model. He assumed that the shear
strength of soil is due to sliding friction from shearing and the interlocking of soil particles.
τ dε
τ dγ=μ f σ 'z dγ ± σ 'z d ε z ; =μ f ± z
σz'
dγ
' d εz
At critical state, μf =tan ϕ cr and α = =0
dγ
( )
τ
'
σz cr
=tan ϕ 'cr
d εz
At peak shear strength =tan α p
dγ
( )
τ
'
σz p
'
=tan ϕ cr + tan α p
Unlike Coulomb failure criterion, Taylor failure criterion does not require the assumption
of any physical mechanism of failure, such as a plane of sliding. It can be applied at every
stage of loading for soils that are homogeneous and deform under plane strain conditions
similar to simple shear.
This failure criterion would not apply to soils that fail along a joint or an interface
between two soils. Taylor failure criterion gives a higher peak dilation angle than
Coulomb failure criterion.
3.3.3 Mohr Coulomb Failure Criterion
Coulomb’s frictional law for finding the shear strength of soils requires that we know the
friction angle and the normal effective stress on the slip plane. Both of these are not
readily known because soils are usually subjected to a variety of stresses. Mohr’s circle
can be used to determine the stress state within a soil mass. By combining Mohr’s circle
for finding stress states with Coulomb’s frictional law, we can develop a generalized
failure criterion.
Let us draw a Coulomb frictional failure line, as illustrated by AB in figure, and subject a
cylindrical sample of soil to principal effective stresses so that Mohr’s circle touches the
Coulomb failure line.
'
where (σ 1)f = Major Principal Effective Stress
'
(σ 3)f = Major Principal Effective Stress
' '
BOC=90 °−ϕ ; BOD=2 θ=90° + ϕ
ϕ'
θ=45+
2
Sign convention of Mohr-coulomb on a soil element
Axial stress is Compression, + σ
Shear stress is counterclockwise from the center of soil element, + τ
The Mohr–Coulomb (MC) failure criterion is a limiting stress criterion, which requires
that stresses in the soil mass cannot lie within the shaded region shown in figure. That is,
the soil cannot have stress states greater than the failure stress state.
The shaded areas are called regions of impossible stress states. For dilating soils, the
bounding curve for possible stress states is the failure envelope, AEFB. For nondilating
soils, the bounding curve is the linear line AFB. The MC failure criterion derived here is
'
independent of the intermediate principal effective stress σ 2, and does not consider the
strains at which failure occurs.
Because MC is a limiting stress criterion, the failure lines AG and AH are fixed lines in
'
[τ f ,σ ] space. The line AG is the failure linefor compression, while the line AH is the failure line
n
for extension (soil elongates; the lateral effective stress is greater than the vertical effective stress).
The shear strength in compression and in extension from interpreting soil strength using the MC
failure criterion is identical. In reality, this is not so.
3.3.4 Tresca Failure Criteria
The shear strength of a fine-grained soil under undrained condition is called the
undrained shear strength, Su . We use the Tresca failure criterion—shear stress at failure is one-
half the principal stress difference—to interpret the undrained shear strength. The undrained shear
strength, Su , is the radius of the Mohr total stress circle; that is
The shear strength under undrained loading depends only on the initial void ratio or the
initial water content. An increase in initial normal effective stress, sometimes called confining
pressure, causes a decrease in initial void ratio and a larger change in excess porewater pressure
when a soil is sheared under undrained condition.
The result is that the Mohr’s circle of total stress expands and the undrained shear
strength increases. Thus, Su is not a fundamental soil property. The value of Su depends on the
magnitude of the initial confining pressure or the initial void ratio. Analyses of soil strength and
soil stability problems using Su are called total stress analyses.
With the exception of Taylor’s criterion, none of the failure criteria provide information
on the shear strains required to initiate failure. Strains (shear and volumetric) are important in the
evaluation of shear strength and deformation of soils for design of safe foundations, slopes, and
other geotechnical systems. Also, these criteria do not consider the initial state (e.g., the initial
stresses, overconsolidation ratio, and initial void ratio) of the soil. In reality, failure is influenced
by the initial state of the soil.
There are several laboratory methods now available to determine the shear strength
parameters (i.e., c , ϕ ,) of various soil specimens in the laboratory. They are as follows:
A popular apparatus to determine the shear strength parameters is the shear box. This test
is useful when a soil mass is likely to fail along a thin zone under plane strain conditions
The shear box consists of a horizontally split, open metal box. Soil is placed in the box,
and one-half of the box is moved relative to the other half. Failure is thereby constrained
along a thin zone of soil on the horizontal plane.
Horizontal forces are applied through a motor for displacement control or by weights
through a pulley system for load control. Most shear box tests are conducted using
displacement control because we can get both the peak shear force and the critical shear
force. In load control tests, you cannot get data beyond the maximum or peak shear force.
Coulomb failure criterion is used to determine the shear strength. Taylor failure criterion
may also be used, but Coulomb failure is better suited for the direct shear test.
3.5.2 Conventional Triaxial Apparatus
The triaxial shear test is one of the most reliable methods available for determining shear
strength parameters. It is used widely for research and conventional testing.
In this test, a soil specimen about 36 mm (1.4 in.) in diameter and 76 mm (3 in.) long
generally is used. The specimen is encased by a thin rubber membrane and placed inside
aplastic cylindrical chamber that usually is filled with water or glycerine. The specimen is
subjected to a confining pressure by compression of the fluid in the chamber. (Note: Air
is sometimes used as a compression medium.) To cause shear failure in the specimen, one
must apply axial stress through a vertical loading ram (sometimes called deviator stress).
The axial load applied by the loading ram corresponding to a given axial deformation is
measured by a proving ring or loadcell attached to the ram.
If the axial stress is greater than the radial stress, the soil is compressed vertically, and the
test is called triaxial compression. If the radial stress is greater than the axial stress, the
soil is compressed laterally, and the test is called triaxial extension.
The applied stresses are principal stresses and the loading condition is axisymmetric. For
compression tests, we will denote the radial stresses σ r asσ 3 and the axial stresses σ z as
σ 1. For extension tests, we will denote the radial stresses σ r asσ 1and the axial stresses σ z
asσ 3
Pz
Deviatoric Stress; q=σ 1−σ 3 =
A
∆z
Axial Strain; ϵ 1=
H0
∆r
Radial Strain; ϵ 3=
H0
∆V
Volumetric Strain; ϵ p= =ϵ 1+2 ϵ 3
V0
2
Deviatoric Strain; ϵ q = ( ϵ 1 −ϵ 3)
3
Change in Area
V V −∆ V
A= = =
0
( V ) A (1−ϵ )
∆V
=
V 0 1−
0 0 p
( H)
H H −∆ H 0 ∆H (1−ϵ ) 1
H 1− 0
0
Connections to measure drainage into or out of the specimen, or to measure pressure in the
pore water (as per the test conditions), also are provided. The following three standard types of
triaxial tests generally are conducted:
The purpose of this test is to determine the undrained shear strength of saturated clays
quickly. In the UC test, no radial stress is applied to the sample (σ 3=0)The axial
(plunger) load, P z, is increased rapidly until the soil sample fails, that is, it cannot
support any additional load. The loading is applied quickly so that the porewater cannot
drain from the soil; the sample is sheared at constant volume.
Consolidated Drained (CD) Compression Test
In the second stage, the pressure in the cell (cell pressure or confining pressure)
is kept constant, and additional axial loads or displacements are added very
slowly until the soil sample fails.
The displacement rate (or strain rate) used must be slow enough to allow the
excess porewater pressure to dissipate. Because the hydraulic conductivity of
fine-grained soils is much lower than that of coarse-grained soils, the
displacement rate for testing fine-grained soils is much lower than for coarse-
grained soils. Drainage of the excess porewater is permitted and the amount of
water expelled is measured. It is customary to perform a minimum of three tests
at different cell pressures.
Since the CD test is a drained test, a single test can take several days if the
hydraulic conductivity of the soil is low (e.g., fine-grained soils)
The Mohr–Coulomb failure criterion is used to interpret the results of a CD test.
'
The elastic moduli for drained conditions, E' and E s, are obtained from the CD
' '
test from the plot of deviatoric stress,(σ 1−σ 3 ) as ordinate and 𝜀ଵ as abscissa
The results of CD tests are used to determine the long-term stability of slopes,
foundations, retaining walls, excavations, and other earthworks.
3.5.2.3 Consolidated Undrained (CU) Compression Test
The purpose of a CU test is to determine the undrained and drained shear
strength parameters ¿ ¿). The CU test is conducted in a similar manner to the CD test
except that after isotropic consolidation, the axial load is increased under undrained
condition and the excess porewater pressure is measured.
One represents total stress condition, and the other effective stress condition. For
each test, Mohr’s circle representing the total stresses has the same size as Mohr’s circle
representing the effective stresses, but they are separated horizontally by the excess
porewater pressure. Mohr’s circle of effective stresses is shifted to the right if the excess
porewater pressure at failure is negative and to the left if the excess porewater pressure is
positive.
The CU test is the most popular triaxial test because you can obtain not only. Su
' '
but also ϕ cr and ϕ p and most tests can be completed within a few minutes after
consolidation, compared with more than a day for a CD test. Fine-grained soils with low
k values must be sheared slowly to allow the excess porewater pressure to equilibrate
throughout the test sample. The results from CU tests are used to analyze the stability of
slopes, foundations, retaining walls, excavations, and other earthworks.
3.5.2.4 Unconsolidated Undrained (UU) Compression Test
Two or more samples of the same soil and the same initial void ratio are
normally tested at different cell pressures. Each Mohr’s circle is the same size, but the
circles are translated horizontally by the difference in the magnitude of the cell pressures.
Only the total stress path is known, since the porewater pressures are not measured to
enable the calculation of the effective stresses.
The undrained shear strength, Su , and the undrained elastic moduli, Eu and
(E¿¿ u) p ¿ , are obtained from a UU test. Tresca failure criterion is used to interpret the
UU test. The UU tests, like the UC tests, are quick and inexpensive compared with CD
and CU tests. The advantage that the UU test has over the UC test is that the soil sample
is stressed in the lateral direction to simulate the field condition. Both the UU and UC
tests are useful in preliminary analyses for the design of slopes, foundations, retaining
walls, excavations, and other earthworks.
3.6 SENSITIVITY AND THIXOTROPY
For many naturally deposited clay soils, the unconfined compression strength is reduced greatly
when the soils are tested after remolding without any change in the moisture content. This
property of clay soils is called sensitivity. The degree of sensitivity may be defined as the ratio of
the unconfined compression strength in an undisturbed state to that in a remolded state
qu (undisturbed )
St =
qu (remolded )
The loss of strength of clay soils from remolding is caused primarily by the destruction of the clay
particle structure that was developed during the original process of sedimentation. If, however,
after remolding, a soil specimen is kept in an undisturbed state (that is, without any change in the
moisture content), it will continue to gain strength with time. This phenomenon is referred to as
thixotropy.
Thixotropy is a time-dependent, reversible process in which materials under constant composition
and volume soften when remolded. This loss of strength is gradually regained with time when the
materials are allowed to rest.
Time
3.7 STRENGTH AND ANISOTROPY OF CLAY AND THIXOTROPY
The unconsolidated-undrained shear strength of some saturated clays can vary, depending on the
direction of load application; this variation is referred to as anisotropy with respect to strength.
Anisotropy is caused primarily by the nature of the deposition of the cohesive soils, and
subsequent consolidation makes the clay particles orient perpendicular to the direction of the
major principal stress. Parallel orientation of the clay particles can cause the strength of clay to
vary with direction
Fairly reliable results for the undrained shear strength, cu (Φ = 0 concept), of very soft to medium
cohesive soils may be obtained directly from vane shear tests. The shear vane usually consists of
four thin, equal-sized steel plates welded to a steel torque rod. First, the vane is pushed into the
soil. Then torque is applied at the top of the torque rod to rotate the vane at a uniform speed. A
cylinder of soil of height h and diameter d will resist the torqu euntil the soil fails.
2T
S u=
πd ( )
3 h
+
d 3
1
Results from SPT have been correlated to several soil parameters. Most of these correlations are
weak. Typical correlation among N values, relative density, and ϕ ' are given in Tables below.
You should be cautious in using the correlation in Table below. SPTs are not recommended for
fine-grained soils, so the correlation shown in Table should be used only to provide an assessment
of the relative shear strength of fine-grained soils.
3.8.3 Cone Penetrometer test (CPT)
The cone resistance q c is normally correlated with the undrained shear strength. Several
adjustments are made to q c . One correlation equation is
q c −σ z
Su =
Nk
where N k is a cone factor that depends on the geometry of the cone and the rate of penetration.
Average values of N k as a function of plasticity index can be estimated from
PI−10
N k =19− ; PI >10
5
Results of cone penetrometer tests have been correlated with the peak friction angle. A number of
correlations exist. Based on published data for sand (Robertson and Campanella, 1983), you can
'
estimate ϕ p using
Retaining structures such as retaining walls, basement walls, and bulkheads commonly are
encountered in foundation engineering as they support slopes of earth masses. Proper design and
construction of these structures require a thorough knowledge of the lateral forces that act
between the retaining structures and the soil masses being retained. These lateral forces are caused
by lateral earth pressure. This chapter is devoted to the study of the various earth pressure theories
Consider a mass of soil, the mass is bounded by a frictionless wall of height AB. A soil element
'
located at a depth z is subjected to a vertical effective pressure σ 0 and a horizontal effective
'
pressure σ h. There are no shear stresses on the vertical and horizontal planes of the soil element.
' '
Let us define the ratio of σ h to σ 0 as a nondimensional quantity K, or
'
σh
K= '
σ0
'
σh
K= K 0= '
σ0
' '
σh σa
K= K a= '
= '
σ0 σ0
( σ '1 )f −( σ '3 )f
'
sin Φ = '
( σ 1 )f + ( σ '3 )f
( σ '3 )f 1−sin Φ'
( )
' '
σh σa Φ
'
= '= ' = =tan 2 45− =K a
σ 0 σ 0 ( σ 1 )f 1+ sinΦ
' '
2
' '
σh σp
K= K a= '
= '
σ0 σ0
( σ '1 )f −( σ '3 )f
'
sin Φ = '
( σ 1 )f + ( σ '3 )f
( σ '1 )f 1+ sinΦ '
( )
' ' '
σh σp Φ
= '= ' = =tan 2 45+ =K p
σ 0 σ 0 ( σ 3 )f 1−sin Φ
' '
2
4.2 EARTH PRESSURE AT – REST
For coarse-grained soils (loose sand), the coefficient of earth pressure at rest can be estimated by
using the empirical relationship
'
K 0 =1−sin Φ
K 0 =1−sin Φ' +
[ γd
γ d ( min ) ]
−1 5.5
K 0 =( 1−sin Φ ) ( OCR )
' sin Φ
K 0 =0.44+0.42 [ PI ( % )
100 ]
Where: PI = Plastic Index
A soil mass that is bounded by a frictionless wall, AB, that extends to an infinite depth.
The stress condition in the soil element can be represented by the Mohr's circle a.
However, if the wall AB is allowed to move away from the soil mass gradually, the horizontal
principal stress will decrease. Ultimately a state will be reached when the stress condition in the
soil element can be represented by the Mohr's circle b, the state of plastic equilibrium and failure
of the soil will occur.
This situation represents Rankine's active state, and the effective pressure a on the vertical plane
(which is a principal plane) is Rankine's active earth pressure.
( ) ( )
' '
' 2 Φ ' Φ
σ a=γz tan 45− −2c tan 45−
2 2
Ka=
σ 'a
σ '0
=tan 2 45− ( Φ'
2 )
4.3.2 Rankine's Theory of Passive Pressure
AB is a frictionless wall that extends to an infinite depth. The initial stress condition on a soil
element is represented by the Mohr's circle a. If the wall gradually is pushed into the soil mass, the
'
effective principal stress σ h will increase. Ultimately, the wall will reach a situation where the
stress condition for the soil element can be expressed by the Mohr's circle b.
At this time, failure of the soil will occur. This situation is referred to as Rankine's passive state.
'
The lateral earth pressure σ p which is the major principal stress, is called Rankine's passive
earth pressure.
( ) ( )
' '
' 2 Φ ' Φ
σ p=γz tan 45+ + 2c tan 45+
2 2
( )
' '
σp Φ
K p= '
=tan 2 45+
σ0 2
4.3.3 Generalized Case for Rankine Active and Passive Pressure-Granular Backfill
Section 4.3.1 and 4.3.2 can be extended to general cases of frictionless wall with inclined backfill
(granular soil).
Ka=
cos2 θ(cos α ¿ + √ sin2 Φ' −sin2 α )¿
ψ a=sin
−1
( sin α
sin Φ' )
−α +2 θ
'
The pressure σ a will be inclined at an angle ß with the plane drawn at right angle to the
back-face of the wall
β=tan
−1
( sin Φ' sinψ a
1−sin Φ cos ψ a
' )
Failure wedge will be inclined at an angle
π Φ' α 1
η= + + − sin−1
4 2 2 2
sin α
sin Φ
' ( )
4.3.3.2 Generalized Case for Rankine Passive
Ka=
cos2 θ(cos α ¿− √ sin2 Φ' −sin2 α)¿
ψ p=sin
−1
( sinsinΦα )+α −2 θ
'
'
The pressure σ p will be inclined at an angle ß with the plane drawn at right angle to the
back-face of the wall
( )
'
−1 sinΦ sin ψ p
β=tan
1+sin Φ' cos ψ p
( )
'
π Φ α 1 −1 sin α
η= − + + sin
4 2 2 2 sinΦ '
Active Case - A retaining wall with cohesionless soil backfill that has a horizontal ground surface.
Passive Case
Active Case - A retaining wall with cohesive soil backfill that has a horizontal ground surface.
( ) ( )
' '
' 2 Φ ' Φ
σ a=γz tan 45− −2c tan 45−
2 2
( )
'
2 Φ
K a =tan 45−
2
2c'
z=
γ √Ka
Passive Case
( ) ( )
' '
Φ Φ
σ 'p=γz tan 2 45+ + 2c ' tan 45+
2 2
( )
'
Φ
K p =tan2 45+
2
σ p=γz K p −2 c ' √ K p
'
4.4 COULOMB'S EARTH PRESSURE THEORY
Coulomb (1776) proposed that a condition of limit equilibrium exists through which a soil mass
behind a vertical retaining wall will slip along a plane inclined an angle θ to the horizontal.
Let AB be the back face of a retaining wall supporting a granular soil; the surface of which is
constantly sloping at an angle α with the horizontal. BC is a trial failure surface. In the stability
consideration of the probable failure wedge ABC.
W Pa
=
sin ( 90+θ−δ −β+Φ ) sin ( β−Φ' )
' '
W sin ( β −Φ )
'
Pa=
sin ( 90+ θ−δ −β +Φ )
' '
1 2
Pa = γ H ¿
2
d Pa 1 2
=0 ; P a= γ H K a
dβ 2
cos 2 ( Φ ' −θ )
K a ( C )=
[√ ]
2
sin ( δ + Φ ) sin ( Φ −α )
' ' '
2 '
cos θ cos( δ +θ) 1+
cos ( δ ' + θ ) cos ( θ−α )
4.4.2 Generalized Case for Coulomb Passive
cos ( Φ +θ )
2 '
K a ( C)=
[ √ ]
2
sin ( δ +Φ ) sin ( Φ + α )
' ' '
cos2 θ cos(δ ' −θ) 1−
cos ( δ ' −θ ) cos ( θ−α )
4.5 EARTH PRESSURE USING THEORY OF PLASTICITY
Rosenfarb and Chen (1972) developed a closed-form earth pressure solution using plasticity
theory that can be used for both active and passive earth pressure computations. The closed-form
solution requires a computer program with an iteration routine, which is not particularly difficult.
Figure shows the curved failure surface in the granular backfill of a retaining wall of height H. The
curved lower portion BC of the failure wedge is an arc of a logarithmic spiral.
'
θtan Φ
r =r 0 e
In practice, the common types of retaining walls constructed can be divided into two major
categories: rigid retaining walls and flexible retaining walls.
1. Gravity retaining walls are constructed with plain concrete or stone masonry. They depend
on their own weight and any soil resting on the masonry for stability. This type of
construction is not economical for high walls
2. Semi gravity retaining walls in many cases, a small amount of steel may be used for the
construction of gravity walls, thereby minimizing the size of wall sections. Such walls
generally are referred to as semigravity walls.
3. Cantilever retaining walls are made of reinforced concrete that consists of a thin stem and a
base slab. This type of wall is economical to a height of about 8 m.
4. Counterfort retaining walls are similar to cantilever walls. At regular intervals, however,
they have thin, vertical concrete slabs known as counterforts that tie the wall and the base slab
together. The purpose of the counterforts is to reduce the shear and the bending moments.
4.6.1.1 Modes of failure for rigid retaining wall
1. Translation - A rigid retaining wall must have adequate resistance against translation.
That is, the sliding resistance of the base of the wall must be greater than the resultant
lateral force pushing against the wall.
T
( FS )T =
P ax
2. Rotation - A rigid retaining wall must have adequate resistance against rotation. The
rotation of the wall about its toe is satisfied if the resultant vertical force lies within the
middle third of the base.
B
e≤
6
where; e = eccentricity of the vertical resultant
B = Length of base
3. Bearing Capacity - A rigid retaining wall must have a sufficient margin of safety
against soil bearing capacity failure. The maximum pressure imposed on the soil at the
base of the wall must not exceed the allowable soil bearing capacity.
( σ Z ) Max ≤ qu
where; ( σ Z ) Max = Maximum vertical stress imposed
qu=Allowable Soil Bearing Capacity
4. Deep-Seated Failure - A rigid retaining wall must not fail by deep-seated failure,
whereby a slip surface encompasses the wall and the soil adjacent to it.
5. Seepage - A rigid retaining wall must have adequate protection from groundwater
seepage. The porewater pressures and the maximum hydraulic gradient developed under
seepage must not cause any of the four stability criteria stated above to be violated and
static liquefaction must not occur.
i cr
( FS )S = ≥3
i Max
Consists of long, slender members of either steel or concrete or wood or plastic and relies on
passive soil resistance and anchors for stability.
1. Cantilever - commonly used to support soils to a height of less than 3 m. Cantilever sheet pile
walls rely on the passive soil resistance for their stability.
Cantilever sheet pile walls are analyzed by assuming that rotation occurs at some point, O,
just above the base of the wall. The consequence of assuming rotation above the base is that,
below the point of rotation, the lateral pressure is passive behind the wall and active in front
of the wall.
To simplify the analysis, a force R is used at the point of rotation to approximate the net
passive resistance below it (the point of rotation). By taking moments about O, the unknown
force R is eliminated and we then obtain one equation with one unknown, that is, the
unknown depth, do.
2. Anchored or tie-back - commonly used to support deep excavations and as waterfront
retaining structures. anchored sheet pile walls rely on a combination of anchors and passive
soil resistance for their stability.
3. Propped
There are two methods used to analyze anchored sheet pile walls. One is the free earth method, the
other is the fixed earth method. We will be discussing the free earth method, because it is
frequently used in design practice. In the free earth method, it is assumed that (1) the depth of
embedment of the wall is insufficient to provide fixity at the bottom end of the wall, and (2)
rotation takes place about the point of attachment of the anchor, O.
Mechanical stabilized earth (MSE) walls are used for a variety of retaining structures. Metal
strips, geotextiles, or geogrids reinforce the soil mass. Other names used are geosynthetics and
geo composites.
You should recall from Chapter 1 that if a load is applied to a soil mass under axisymmetric
undrained condition, the lateral strain (ε3) is one-half the axial strain (ε₁).
If the undrained restriction is lifted, then you can expect lateral strains greater than one-half
the vertical strains. If we were to install strips of metal in the lateral directions of the soil
mass, then the friction at the interfaces of the metal strips and the soil would restrain lateral
displacements.
The net effect is the imposition of a lateral resistance on the soil mass that causes Mohr's
circle to move away from the failure line. The lateral force imposed on the soil depends on the
interface friction value between the reinforcing element and the soil mass and the vertical effective
stress. For a constant interface friction value, the lateral frictional force would increase with depth.
The reinforcing material will fail if the lateral stress exceeds its tensile strength.
There are two sets of stability criteria to be satisfied for MSE walls. One is the internal
stability; the other is the external stability. The external stability of an MSE wall is determined by
analogy to a gravity retaining wall with a vertical face. The internal stability depends on the tensile
strength of the reinforcing material and the slip at the interface of the reinforcing material and the
soil.
Tensile failure of the reinforcing material at any depth leads to progressive collapse of the wall,
while slip at the interface of the reinforcing material and the soil mass leads to redistribution of
stresses and progressive deformation of the wall.
Internal stability
The frictional resistance develops outside the active slip or failure zone.
Pr
=( σ 'z +q s ) tan Φi
2 w Le
T
=( σ 'z +q s ) K a ( R )
Sz
L=LR+ ¿
LR =( H 0 −z)tan 45−
Φ 'cs
2 ( )
External stability
Pax
Lb =
Sw
where; Pax = lateral active earth force, Coulomb
Lb= length of reinforcement at the base
Su
Sw = adhesion stress =
2
For long-term loading.
'
T =γ H 0 Lb tan Φ b
'
where; Φ b = effective interfacial friction angle between the reinforcement and the soil
at the base
Lb = length of reinforcement at the base
1. Gabion baskets. Gabion basket walls consist of prefabricated steel wire or polypropylene or
polyethylene or nylon baskets filled with rocks and stacked horizontal and vertically.
2. Bin walls. Bin walls are gravity walls in which earth fill is placed in a bin made from metal
(steel) or timber or concrete.
3. Precast modular concrete walls. These walls are constructed by stacking precast blocks
made from concrete and other materials. There is a large variety of blocks, the major
differences being the material used to make the blocks, size, and interlocking mechanism.
1. Soil Nailing. These walls utilize reinforcing elements to form a composite with the soil. The
reinforcing elements are nails or small-diameter cast-in-place concrete piles or small diameter
steel pipe piles. One popular in situ reinforced wall is a soil nail wall.
Closely spaced nails are installed by drilling inclined holes into the soil and grouting the holes.
Shear stresses from the soil are transferred to the nails and are resisted by tensile forces in the
nails. The faces of soil nail walls are shotcrete or precast concrete panels or cast-in- place
concrete.
2. Chemically Stabilized Earth Walls (CSE) - CSE walls are in situ soils mixed with chemical
grouts such as lime or lime cement mixtures to form columns of overlapping soils. Sometimes
reinforcements are added to the soil-grout mixtures before they harden. CSE walls can retain
soils up to great depths. They are also used in seepage control.
The critical design elements in a braced excavation are the loads on the struts, which are usually
different because of different lateral loads at different depths, the time between excavations, and
the installation procedure.
Failure of a single strut can be catastrophic because it can lead to the collapse of the whole
system. The analysis for the. forces and deflection in braced excavation should ideally consider
the construction sequence, and numerical methods such as the finite element method are
preferred.
Semi-empirical methods are often used for shallow braced excavations and in the preliminary
design of deep braced excavations. The finite element method is beyond the scope of this
material. We will only discuss a semi-empirical method.
Braced wall AB of height H that deforms by rotating about its top. The wall is assumed to
be rough, with the angle of wall friction equal to δ'. The point of application of the active thrust
(that is, η aH) is assumed to be known. The curve of sliding is assumed to be an arc of a
logarithmic spiral.
4.7.2 Semi-empirical method (Peck's pressure diagram)
Lateral stress distributions used in the semi-empirical method are approximations from
field measurements of strut loads in different types of soil. These lateral stress distributions are not
real but average approximate stress distributions to estimate the maximum strut load. The real
lateral stress distributions are strongly affected by arching action, as we will discuss in Chapter 5.
Geotechnical engineers must pay particular attention to geology, surface drainage, groundwater,
and the shear strength of soils in assessing slope stability. However, we are handicapped by the
geological variability of soils and methods for obtaining reliable values of shear strength. The
analyses of slope stability are based on simplifying assumptions, and the design of a stable slope
relies heavily on experience and careful site investigation. The failed soil mass can move very
quickly over large distances. Your job is to prevent such failure.
Slope failures depend on the soil type, soil stratification, groundwater, seepage, and the slope
geometry.
Failure of a slope along a weak zone of soil is called a translational slide. The sliding mass can
travel long distances before coming to rest. Translational slides are common in coarse-grained
soils.
A common type of failure in homogeneous fine-grained soils is a rotational slide that has its point
of rotation on an imaginary axis parallel to the slope. Three types of rotational failure often occur.
One type, called a base slide, occurs by an arc engulfing the whole slope. A soft soil layer resting
on a stiff layer of soil is prone to base failure.
The second type of rotational failure is the toe slide, whereby the failure surface passes through
the toe of the slope.
The third type of rotational failure is the slope slide, whereby the failure surface passes through
the slope.
A flow slide occurs when internal and external conditions force a soil to behave like a viscous
fluid and flow down even shallow slopes, spreading out in several directions. The failure surface is
ill defined in flow slides. Multiple failure surfaces usually occur and change continuously as flow
proceeds. Flow slides can occur in dry and wet soils.
Block or wedge slides occur when a soil mass is shattered along joints, seams, fissures, and weak
zones by forces emanating from adjacent soils. The shattered mass moves as blocks and wedges
down the slope.
Rainfall - Long periods of rainfall saturate, soften, and erode soils. Water enters into existing
cracks and may weaken underlying soil layers, leading to failure, for example, mud slides.
Earthquakes - Earthquakes induce dynamic forces especially dynamic shear forces that reduce
the shear strength and stiffness of the soil. Porewater pressures in saturated coarse-grained soils
could rise to a value equal to the total mean stress and cause these soils to behave like viscous
fluids—a phenomenon known as dynamic liquefaction. Structures founded on these soils would
collapse; structures buried within them would rise.
Geological Features - Many failures commonly result from unidentified geological features. A
thin seam of silt (a few millimeters thick) under a thick deposit of stiff clay can easily be
overlooked in drilling operations, or one may be careless in assessing borehole logs only to find
later that the presence of the silt caused a catastrophic failure. Sloping, stratified soils are prone to
translational slide along weak layers. You must pay particular attention to geological features in
assessing slope stability.
External Loading - Loads placed on the crest of a slope (the top of the slope) add to the
gravitational load and may cause slope failure. A load placed at the toe, called a berm, will
increase the stability of the slope. Berms are often used to remediate problem slopes.
Construction Activities - Construction activities near the toe ofan existing slope can cause failure
because lateral resistance is removed. We can conveniently divide slope failures due to
construction activities into two cases. The first case is excavated slope and the second case is fill
slope.
Excavated slope - When excavation occurs, the total stresses are reduced and negative porewater
pressures are generated in the soil. With time the negative porewater pressures dissipate, causing a
decrease in effective stresses and consequently lowering the shear strength of the soil. If slope
failures were to occur, they would most likely take place after construction is completed.
Fill slope - are common in embankment construction. Fill (soil)is placed at the site and compacted
to specifications, usually greater than 95% Proctor maximum dry unit weight. The soil is
invariably unsaturated, and negative porewater pressures develop. The soil on which the fill is
placed, which we will call the foundation soil, may or may not be saturated. If the foundation soil
is saturated, then positive porewater pressures will be generated from the weight of the fill and the
compaction process. The effective stresses decrease, and consequently the shear strength
decreases. With time the positive pore water pressures dissipate, the effective stresses increase,
and so does the shear strength of the soil. Thus, slope failures in fill slopes are most likely to occur
during or immediately after construction.
Rapid Drawdown - Reservoirs can be subjected to rapid drawdown. In this case the lateral force
provided by the water is removed and the excess porewater pressure does not have enough time to
dissipate. The net effect is that the slope can fail under undrained condition. If the water level in
the reservoir remains at low levels and failure did not occur under undrained condition, seepage of
groundwater would occur, and the additional seepage forces could provoke failure.
An ‘infinite slope’ is one which represents the boundary surface of a semi-infinite soil mass
inclined to the horizontal. In practice, if the height of the slope is very large, one may consider it
as an infinite one. It is assumed that the soil is homogeneous in its properties. If different strata are
present the strata boundaries are assumed to be parallel to the surface. Failure tends to occur only
along a plane parallel to the surface.
Let us consider an infinite slope in cohesionless soil, inclined at an angle β to the horizontal.
where:
The shear strength of a soil consists of two components, cohesion and friction.
c’ = cohesion
Φ’ = angle of friction
𝜎’ = normal stress on the potential surface
Therefore,
If there is steady state seepage through the soil and the groundwater table coincides with the
ground surface, as shown in Figure below, the factor of safety against sliding can be determined
as.
5.5 FINITE SLOPES
When the value of 𝐻CR approaches the height of the slope, the slope generally may be considered
finite. For simplicity, when analyzing the stability of a finite slope in a homogeneous soil, we need
to make an assumption about the general shape of the surface of potential failure. Although
considerable evidence suggests that slope failures usually occur on curved failure surfaces,
Culmann’s analysis is based on the assumption that the failure of a slope occurs along a plane
when the average shearing stress tending to cause the slip is more than the shear strength of the
soil. Also, the most critical plane is the one that has a minimum ratio of the average shearing stress
that tends to cause failure to the shear strength of soil.
Factor of Safety
Cohesion at any angle
Critical Angle
Stability number
A soil shear failure can result in excessive building distortion and even collapse. Excessive
settlements can result in structural damage to a building frame, nuisances such as sticking doors
and windows, cracks in tile and plaster, and excessive wear or equipment failure from
misalignment resulting from foundation settlements.
This chapter will be concerned with evaluation of the limiting shear resistance, or ultimate bearing
capacity qult of the soil under a foundation load. Chapter 2 will be concerned with estimation of
settlements.
Depending on the structure and soil encountered, various types of foundations are used.
a. Spread Footing
b. Mat Foundation
c. Pile Foundation
d. Drilled Shaft Foundation
Shallow Foundation
A shallow foundation is a type of building foundation that transfers building loads to the earth
very near to the surface. Generally D / B ≤ 1
Deep Foundation
A deep foundation is a type of foundation which transfers building loads to the earth farther down
from the surface than a shallow foundation does, to a subsurface layer or a range of depths,
generally Lpile / Dpile ≥ 4
Centric vertical loads are now incrementally applied on the footing. As the load increases, some
regions of the soil would yield and behave plastically (plastic flow).
If the soil were a rigid–perfectly plastic material, some regions would flow plastically while other
regions would show no deformation. We will call the soil regions that have reached the plastic
state the “plastic zones”. As more loads are added, the plastic zones increase and eventually break
free to the surface, and soil “pileup” on the sides of the footing.
The surface between the plastic zones and the nonplastic or nondeforming zones (applicable to
rigid–perfectly plastic material) is called a slip surface or limiting stress surface.
The “pileup” is influenced by the overburden pressure and the strain-hardening ability of the
material. If the footing is embedded in the soil and/or the soil has a large potential to strain-harden,
the plastic flow that causes “pileup” of soil around the edges of the footing would be restrained,
creating large lateral pressures to force the soil to move laterally.
TWO CONSEQUENCES
1. A soil that would normally show a peak shear stress because of dilatancy and then strain-
soften would be forced to behave as a strain-hardening material, pushing the plastic zone
farther into the soil mass.
2. The failure mechanism discussed might not develop. Therefore, in this situation, there
would not be any distinct collapse load but an increasing load with increasing footing
displacement until critical state is achieved. Generally, this would occur at displacements
that are intolerable.
Prandtl (1920) studied a rigid–perfectly plastic half space loaded by a stiff wedge that is subjected
to centric loads. Terzaghi (1943) applied Prandtl’s theory to a strip footing with the assumption
that the soil is a semi-infinite, homogeneous, isotropic, weightless rigid–plastic material.
6.3.1 GENERAL SHEAR FAILURE MECHANISM
One zone, ABD, is a fan with radial slip planes stopping on a logarithmic spiral slip
plane. The other zone, ADE, consists of slip planes oriented at angles of 45 + Φ/2 and 45 − Φ/2 to
the horizontal and vertical planes, respectively. Zone ADE is called the Rankine passive zone.
According to the Mohr–Coulomb criterion, slip planes form when soil is sheared to
failure. No slip plane, however, can pass through the rigid footing, so none can develop in the
soiljust below the footing. The collapse mechanism shown in previous figure is called the general
shear failure mechanism.
Conventional collapse mechanism shown in previous figure may not develop. Therefore,
calculation of a collapse load from this mechanism (general shear failure mechanism) could be
considerably inaccurate.
Other collapse mechanisms have been proposed. For example, it is assumed that for loose
soils, the slip planes, if they developed, are expected to lie within the soil layer below the base of
the footing and extend laterally. This is called local shear failure.
6.3.3 PUNCHING
For very loose soil, the slip surfaces may be confined to the surfaces of the rigid wedge.
This type of failure is termed punching shear.
Two potential failure modes, where the footing, when loaded to produce the maximum
bearing pressure qult, will do one or both of the following:
6.4.1 ROTATE
Rotate as in figure below about some center of rotation (probably along the vertical line
Oa) with shear resistance developed along the perimeter of the slip zone shown as a circle.
When the foundation pushes into the ground, stress block 1 to the left of vertical line OY
has principal stresses as shown. The push into the ground, however, displaces the soil on the right
side of the line OY laterally, resulting in the major principal stress on block 2 being horizontal as
shown.
When the two blocks are adjacent to each other at the vertical line OY, it is evident that
𝜎3,1 = 𝜎1,2 but with a principal stress rotation of 90° between blocks.
For 𝜱’ = 0
Based on passive pressure,
Based on rotation,
Average of combined Passive and Rotation with ,
6.4.2 PUNCH
Net Ultimate Bearing Capacity - If the difference between the unit weight of concrete used in the
foundation and the unit weight of soil surrounding is assumed to be negligible, then.
Allowable or Gross Allowable Bearing Capacity,
One of the early sets of bearing-capacity equations was proposed by Terzaghi (1943).
Terzaghi's equations were produced from a slightly modified bearing-capacity theory developed
by Prandtl (ca. 1920) from using the theory of plasticity to analyze the punching of a rigid base
into a softer
(soil) material.
Hansen (1970) proposed the general bearing-capacity case and N factor equations. This equation is
readily seen to be a further extension of the earlier Meyerhof (1951) work. These represent
revisions and extensions from earlier proposals in 1957 and 1961. The extensions include base
factors for situations in which the footing is tilted from the horizontal 𝑏i and for the possibility of a
slope 𝛽 of the ground supporting the footing to give ground factors 𝑔i.
6.5.4 VESIC’S BEARING CAPACITY EQUATION
The Vesic (1973, 1915b) procedure is essentially the same as the method of Hansen (1961) with
select changes. Vesic equation is somewhat easier to use than Hansen's
6.6 WHICH EQUATIONS TO USE
It is good practice to use at least two methods and compare the computed values of qu. If
the two values do not compare well, use a third method,
• Terzaghi - Very cohesive soils where D/B < 1 or for a quick estimate of quit to compare with
other methods. Do not use for footings with moments and/or horizontal forces or for tilted bases
and/or sloping ground
• Hansen, Meyerhof, Vesic - Any situation that applies, depending on user preference or
familiarity with a particular method.
• Hansen, Vesic When base is tilted; when footing is on a slope or when D/B > 1.
6.10 EFFECT OF WATER TABLE ON BEARING CAP
Equations from Terzaghi through Vesic’s give the ultimate bearing capacity, based on the
assumption that the water table is located well below the foundation. However, if the water table is
close to the foundation, some modifications of the bearing capacity equations will be necessary.
Case I. If the water table is located so that 0 ≤ 𝐷1 ≤ 𝐷f the factor 𝒒 in the bearing capacity
equations takes the form below. Also, the value of 𝜸 in the last term of the equations has to be
replaced by 𝛾’ = 𝛾sat− 𝛾w.
Case II. For a water table located so that 0 ≤𝑑 ≤ 𝐵. the factor 𝑞 in the bearing capacity equations
takes the form 𝒒=𝜸𝑫𝒇. In this case, the factor in the last term of the bearing capacity equations
must be replaced by the factor.
Case III. When the water table is located so that 𝑑 > 𝐵 the water will have no effect on the
ultimate bearing capacity.
6.11 BEARING CAPACITY FROM SPT
The SPT is widely used to obtain the bearing capacity of soils directly. One of the earliest
published relationships was that of Terzaghi and Peck (1967). This has been widely used, but an
accumulation of field observations has shown these curves to be overly conservative.
Meyerhof (1956, 1974) published equations for computing the allowable bearing capacity
for a 25-mm settlement. These could be used to produce curves similar to those of Terzaghi and
Peck and thus were also very conservative.
Considering the accumulation of field observations and the stated opinions of the authors
and others, this author adjusted the Meyerhof equations for an approximate 50percent increase in
allowable bearing capacity to obtain the following:
These equations have been in existence for quite some time and are based primarily on N
values from the early 1960s back and, thus, 𝐸s is likely on the order of 50 to 55 and not 70+ as
suggested. Since lower 𝐸s produces higher blow counts N if the preceding equations are
standardized to 𝑁’70, we must use revised values for factors 𝐹1and 𝐹2 as shown in the table of
Ffactors.
In these equations N is the statistical average value for the footing influence zone of
about 0.5B above footing base to at least 2B below.
We note in these equations that footing width is a significant parameter. Obviously if the depth of
influence is on the order of 2B a larger footing width will affect the soil to a greater depth and
strains integrated over a greater depth will produce a larger settlement.
Footings to develop tension resistance are idealized in Figure below. Balla (1961)
considered this problem. He assumed a failure surface (the dashed line ab in figure below) as
circular and developed some highly complicated mathematical expressions that were verified on
model tests in a small glass jar and by some larger tests of others.
The following equations are developed by neglecting the larger pull-out zone observed in
the tests and using an approximation of shear resistance along line ab’. Shape factors are used
together with a limiting depth ratio D/B or H/B to make the simplified equations adequate for
design use. In the general case we have for the ultimate tension. with adjustments for depth and
shape (whether perimeter is round or rectangular). This equation gives (only for footings in sands)
the following:
The lateral earth pressure coefficient 𝐾u can be taken as one of the following:
4
5
6
7
8
7.1 INTRODUCTION
A pile is a slender, structural member installed in the ground to transfer the structural
loads to soils at some significant depth below the base of the structure. Structural loads include
axial loads, lateral loads, and moments. Another term commonly used in practice for pile
foundations is deep foundations. Structures that cannot be supported economically on shallow
foundations are normally supported by pile foundations.
There are several types of concrete piles that are commonly used. These include cast-in-
place concrete piles, precast concrete piles, drilled shafts, and barrette piles.
7.2.1.2 PRECAST CONCRETE PILES usually have square or circular or octagonal cross
sections and are fabricated in a construction yard or a factory from reinforced or
prestressed concrete. They are preferred when the pile length is known in advance.
The disadvantages of precast piles are problems in transporting long piles, cutting,
and lengthening. A very popular type of precast concrete pile is the Raymond
cylindrical prestressed pile. This pile comes in sections, and lengths up to 70 m can
be obtained by stacking the sections.
7.2.1.3 MICROPILES (Also called minipiles, pin piles, needle piles, or root piles) are
small-diameter (50 mm to 340 mm) pipe piles (pushed or driven) or grouted (jet or
post or pressure) piles. They are particularly useful for (1) sites with low headroom,
(2) congested areas, (3) sites with restricted access, and (4) foundation repair or
strengthening.
Steel piles come in various shapes and sizes and include cylindrical, tapered, and H-piles.
Steel H-piles are rolled steel sections. They are non-displacement piles. Steel pipe piles are
seamless pipes that can be welded to yield lengths up to 70 m. They are usually driven with open
ends into the soil. A conical tip is used where the piles must penetrate boulders and rocks. To
increase the load capacity of steel pipe piles, the soil plug is excavated and replaced by concrete.
These piles are called concrete-filled steel piles. The soil plug may adhere to the pile surface and
move down during driving. This is called plugging.
Plastic piles comprise a variety of composite materials that include polymer composites,
PVC, and recycled materials. These piles are used in special applications such as in marine
environments and within soil zones exposed to seasonal changes.
Concrete, steel, and timber can be combined to form a composite pile. For example, the
portion of a timber pile above groundwater level that is likely to suffer from decay due to termites
or rot may be replaced by concrete. Similarly, the portion of a steel pile within a corrosive
environment can be covered with concrete or other protective materials.
Accurate estimation of pile load capacity is a rather difficult task because it is difficult, if
not impossible, to account for (a) the changes in stress and strain states from installation effects,
(b) the variability of soil types, and (c) the differences in the quality of construction practice.
Therefore, calculations of pile load capacity are approximations and rely heavily on empiricism or
semi empiricism (part mechanics, part empirical).
If the skin friction is greater than about 80% of the end bearing load capacity, the pile is deemed a
friction pile and, if the reverse, an end bearing pile. If the end bearing is neglected, the pile is
called a floating pile.
The a-method is based on a total stress analysis (TSA) and is normally used to estimate
the short-term load capacity of piles embedded in fine-grained soils. In the 𝛼-method, a
coefficient, 𝛼u, is used to relate the undrained shear strength, 𝑠u, to the adhesive stress (𝑓s) along
the pile shaft.
The end bearing capacity is found by analogy with the conventional failure mode of
shallow foundations and is expressed as;
7.3.2 β – METHOD
7.3.2.1 SKIN FRICTION
The 𝛽-method is based on an effective stress analysis and is used to determine the short-
term and long-term pile load capacities of coarse-grained soils and the long-term load capacity of
fine-grained soils. The friction along the pile shaft is found using Coulomb’s friction law, where
the frictional stress is given by where 𝜎x’ is the coefficient of friction, 𝜎x’ is
the lateral effective stress, and ϕi’ is the interfacial effective friction angle. The skin friction is
expressed as
The end bearing capacity is calculated by analogy with the bearing capacity of shallow
footings and is determined from
where the angle 𝜓p (called the angle of pastification, varies from 𝜓p ≤ 𝜋/3 for
soft, fine-grained soils to 𝜓p ≤ 0.58𝜋 for dense, coarse-grained soils and
overconsolidated fine-grained soils. Janbu recommended that for soft, compressible soils,
𝜓p should not exceed 𝜋/3, while for dense, coarse-grained soils, 𝜓p should not exceed
𝜋/2.
7.3.3 I METHOD
Vijayvergiya and Focht (1972) presented a method of obtaining the skin resistance of a
pile in overconsolidated clays and have claimed a correlation between design and load tests on the
order of ± 10 percent. The original development was based primarily on pile load tests. These
were on long piles used for offshore oil production structures and founded in clays located in or
along the U.S. coastline of the Gulf of Mexico. This method has also been used in other marine
installations with some success (e.g., North Sea oil production structures).
The end bearing capacity is found by analogy with the conventional failure mode of
shallow foundations and is expressed as;
7.4 VERTICAL CAP. OF SINGLE PILE FOR DRILLED PILES
7.4.1 a – METHOD
The load capacities of drilled shafts are calculated similarly to driven piles except that the
empirical adhesion, friction, and end bearing factors are different.
7.4.2 β – METHOD
The 𝛽 for drilled shafts in coarse-grained soils have been obtained from back calculations
from load tests on 1-mdiameter drilled shafts in cemented sand at particular locations (Texas Gulf
Coast region and Los Angeles, California). They are not related to any soil parameters. They have
to be used with careful judgment based on experience.
The soil near the top of the drilled shaft is subjected to environmental and construction
effects, while the soil just above the base may develop tensile cracking. Consequently, the upper
1.5 m of the shaft and one pile diameter above the base are ignored in calculating skin friction for
drilled shafts.
Piles may be required to resist uplift forces for example, in foundations of structures
subjected to large overturning moments such as tall chimneys, transmission towers, or jetty
structures. Methods of calculating the adhesion to resist uplift are the same as those used for
friction pile.
• The principal effect of negative skin resistance is to increase the axial load in the lower fixed
portion of the pile. It may result also in increased pile settlements due to the axial shortening
and/or additional point penetration of the pile under the increased axial load.
• Negative skin friction can produce large tension stresses when the effect is from expansive soils -
especially if no, or insufficient, gap is left between soil and pile cap and the soil expands against
both the pile and the cap.
For negative skin resistance forces to develop significantly, a portion of the pile must be fixed
against vertical movement, such as the point being on rock or the lower part being in a dense sand.
If the entire pile moves down with the consolidation effect no negative skin resistance forces
develop.
CASE 1 - CLAY FILL OVER GRANULAR OIL
For laterally loaded piles, Broms (1965) developed a simplified solution based on the
assumptions of:
a. shear failure in soil, which is the case for short piles, and
b. bending of the pile, governed by the plastic yield resistance of the pile section, which is
applicable to long piles.
The preceding sections have considered the soil aspects of single piles in some detail
together with a brief discussion of pile-driving operations. Rarely, however, is the foundation
likely to consist of a single pile. Generally, there will be a minimum of two or three piles under a
foundation element or footing to allow for misalignments and other inadvertent eccentricities.
Building codes may stipulate the minimum number of piles under a building element.
When several piles are clustered, it is reasonable to expect that the soil pressures
produced from either side friction or point bearing will overlap as idealized in the figure.
The superimposed pressure intensity will depend on both the pile load and spacing, and if
sufficiently large the soil will fail in shear, or the settlement will be excessive.
The stress intensity from overlapping stressed zones will obviously decrease with
increased pile spacing s; however, large spacing s are often impractical since a pile cap is
cast over the pile group for the column base and/or to spread the load to the several piles
in the group.
When several pile butts are attached to a common structural element termed a pile cap the
result is a pile group. A question of some concern is whether the pile group capacity is the sum of
the individual pile capacities or something different either more or less. If the capacity is the sum
of the several individual pile contributions, the group efficiency 𝐸g = 1.0. There are mixed
opinions on pile group efficiency defined as
The ASCE Committee on Deep Foundations report [CDF (1984)] recommends not using
group efficiency as a description of group action. It suggests that friction piles in
cohesionless soils at the usual spacings s of s = 2 to 3D will have a group efficiency 𝐸g >
1.0.
The reason given is that in cohesionless soil the pile displacement + driving vibrations
increase the soil density (or 𝛾s) in a zone in the vicinity of the pile, which is further
increased as other piles are driven nearby.
For friction piles in cohesive soils the block shear + point bearing of the group in plan is
used as the group capacity, but in no case is the group capacity to be considered greater
than the single pile capacity times the number of piles in the group.
If the pile material is assumed to be elastic, the deformation of the pile shaft can be evaluated, in
accordance with the fundamental principles of mechanics of materials, as
The settlement of a pile caused by the load carried at the pile point may be expressed in the form:
The settlement of a pile caused by the load carried by the pile shaft is given by a relation
Sometimes, a pile group may be embedded above a soft clay layer and transfer sufficient
load to it (soft clay) to cause consolidation settlement
to estimate the consolidation settlement, the full design load is assumed to act at a depth
of 2/3 L
then distributed in the ratio of 2:1 (vertical : horizontal). The increase in vertical stress at
a depth z in the soft clay layer shown in figure below