Sunteți pe pagina 1din 78

Addis Ababa University Faculty of Technology Department of Civil Engineering

CHAPTER ONE

STRESS DISTRIBUTION IN SOIL

1.1 Introduction

Stress in soil is caused by one or both of the following:


a) Self weight of soil
b) External loading (unloading) = structural loading

The vertical stress in soil owing to self weight is given by:


σ'v = γ z,
where σ'v = the vertical stress in the soil at depth z below the surface
γ = unit weight of soil

Foundations are designed with an adequate factor of safety against shear failure of soil;
thus it is safe to presume that the operating stresses in soil are small enough to assume
stress-strain proportionality. Fortunately, the order of magnitudes of stress transmitted in
to soil from structural loading are also small and hence the application of elastic theory
for the determination of stress distribution in soil gives reasonably valid results. Hence
stresses due to applied structural loadings are determined by using elastic theory.

The resultant stress can be obtained by superposition of structural load stress and
overburden pressure.

1.2 Stress Distribution for a Point Load

Bossinesq (1885) has given the solution for the stresses caused by the application of a
point load at the surface of a homogenous, elastic, isotropic and semi-infinite medium.
Q

x σz τzx
r X τxz
θ y τ zy
σx
R τ yz τ xy
z τyx
Y A (x,y,z) σy
z
Q r = x2 + y2

r R = r 2 + z2 = x2 + y2 + z2
θ r z
σR sin θ = ; cos θ =
σ r τ rz R R
v = poisson' s ratio
σθ Z
Fig. 1.1 Point Load on the surface

With reference to Fig. 1.1, the expressions for the increase in stress at point A due to a
point load Q at its surface are

1
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

In rectangular coordinates
3Q z 3
σz = ⋅ (1.1)
2π R 5
3Q ⎧⎪ x 2 z 1 − 2v ⎡ 1 (2 R + z ) x 2 z ⎤ ⎫⎪
σx = ⋅⎨ 5 + ⎢ − − ⎥⎬ (1.2)
2π ⎪⎩ R 3 ⎣ R ( R + z ) R 3 ( R + z ) 2 R 3 ⎦ ⎪⎭
3Q ⎧⎪ y 2 z 1 − 2v ⎡ 1 (2 R + z ) y 2 z ⎤ ⎫⎪
σy = ⋅⎨ 5 + ⎢ − 3 − ⎥⎬ (1.3)
2π ⎪⎩ R 3 ⎣ R( R + z ) R ( R + z ) 2 R 3 ⎦ ⎪⎭
3Q ⎡ xyz 1 − 2v (2 R + z ) xy ⎤
τ xy = ⋅⎢ − ⎥ (1.4)
2π ⎣ R 5 3 R 3 (R + z) 2 ⎦
3Q xz 2
τ xz =⋅ (1.5)
2π R 5
3Q yz 2
τ yz = ⋅ (1.6)
2π R 5
In cylindrical coordinates
3Q z 3
σz = ⋅ (1.7)
2π R 5
Q ⎡ 3zr 2 1 − 2v ⎤
σr = ⋅⎢ 5 − ⎥ (1.8)
2π ⎣ R R( R + z ) ⎦
Q ⎡ 1 z ⎤
σθ = (1 − 2v) ⎢ − 3⎥ (1.9)
2π ⎣ R( R + z ) R ⎦
3Q rz 2
τ rz =
⋅ (1.10)
2π R 5
In most foundation problems it is very necessary to be acquainted with the increase in
vertical stresses (for settlements) and the increase in shear stresses (for shear strength
analysis).
The equation for the vertical stress can be rewritten as follows
52
3Q z 3 3Q z3 3Q ⎡ 1 ⎤
σz = ⋅ 5 = ⋅ 2 = ⋅⎢ 2 ⎥
2π R 2π (r + z )
2 52
2π z 2 ⎣1 + (r z ) ⎦
Q
= Iz where I z = non dim ensional inf luence factor (1.11)
z2
52
3 ⎡ 1 ⎤
Iz = ⋅⎢ 2 ⎥
2π ⎣1 + (r z ) ⎦
Values of Iz for different values of r/z ratio can be tabulated (Table 1.1) or plotted as Iz
versus r/z ratio (Fig 1.2) and hence can be used for routine stress calculation. Note that
the maximum vertical stress is observed directly below the load (r=0).

2
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
Table 1.1
r/z Iz 0.5
0 0.4775 0.45
0.2 0.4329 0.4
0.4 0.3295 0.35
0.6 0.2214
0.3
0.8 0.1386

Iz
1.0 0.0844 0.25
1.2 0.0513 0.2
1.4 0.0317 0.15
1.6 0.0200 0.1
1.8 0.0129 0.05
2.0 0.0085
0
2.2 0.0058
2.4 0.004 0 0.5 1 1.5 2 2.5 3 3.5
2.6 0.0028 r/z
2.8 0.0021
3.0 0.0015 Fig 1.2 Non dimensional influence factor
The equation for the shear stress can be rewritten as follows
3Q rz 2 r
τ rz = ⋅ 5 =σz ⋅
2π R z
r Q
⇒ τ rz = I z ⋅ 2 where I z = Influence factor given above (1.11)
zz

Pressure distributions
Graphical vertical stress distribution on a horizontal plane at any depth z below the ground
surface can be drawn as shown here under.
The vertical stress on a horizontal plane at a depth z is given by
Q
σ z = Iz z being a specified depth
2
z
For several assumed values of r, r/z is calculated and Iz is found for each, the value of σz
is then computed. As an example consider the following cases
Case 1: Let z =c (where c= constant number) and r be varied as 0, 0.25c, 0.5c,
0.75c, c, 1.25c, 1.5c, 1.75c, 2c, etc
Case 2: Let z=2c and r be varied as case 1
Case 3: Let z=4c and r be varied as case 1
Now the stresses are calculated (tabulated) and plotted together below.

3
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
Case 1 (z=c) Case 2 (z=2c) Case 3 (z=4c)
r r/z Iz (σzc2)/Q r r/z Iz (σzc2)/Q r r/z Iz (σzc2)/Q
0 0 0.4775 0.4775 0 0 0.4775 0.1194 0 0 0.4775 0.0298
0.25c 0.25 0.4103 0.4103 0.25c 0.125 0.4593 0.1148 0.25c 0.0625 0.4728 0.0296
0.5c 0.50 0.2733 0.2733 0.5c 0.250 0.4103 0.1026 0.5c 0.125 0.4593 0.0287
0.75c 0.75 0.1565 0.1565 0.75c 0.375 0.3436 0.0859 0.75c 0.1875 0.4380 0.0274
c 1.00 0.0844 0.0844 c 0.500 0.2733 0.0683 c 0.25 0.4103 0.0256
1.25c 1.25 0.0454 0.0454 1.25c 0.625 0.2094 0.0523 1.25c 0.3125 0.3782 0.0236
1.5c 1.50 0.0251 0.0251 1.5c 0.750 0.1565 0.0391 1.5c 0.375 0.3436 0.0215
1.75c 1.75 0.0144 0.0144 1.75c 0.875 0.1153 0.0288 1.75c 0.4375 0.3082 0.0193
2c 2.00 0.0085 0.0085 2c 1.000 0.0844 0.0211 2c 0.5 0.2733 0.0171

0.50
0.45 z=c
0.40
0.35 z=2c
0.30 z=4c
σzc /Q

0.25
2

0.20
0.15
0.10
0.05
0.00
-3 -2 -1 0 1 2 3
r/c

From the above curves one can see that the stress diminishes as we move down from the
ground surface and also as we move away from the point of load application.

Similarly vertical stress distribution on a vertical plane at any radial distance r from the load
can be drawn as shown here under.
For several assumed values of z, r/z is calculated and Iz is found for each, the value of σz
is then computed. As an example consider the following cases
Case 1: Let r=a (where a= constant number) and z be varied as 0, 0.5a, a, 2a,
5a, 10a, etc
Case 2: Let r=2a and z be varied as case 1
Case 3: Let r=4a and z be varied as case 1
Now the stresses are calculated and plotted below.
Case 1 (r=a) Case 2 (r=2a) Case 3 (r=4a)
z r/z Iz (σza2)/Q z r/z Iz (σza2)/Q z r/z Iz (σza2)/Q
0 - - indet. 0 - - indet. 0 - - indet.
0.5a 2.00 0.0085 0.0342 0.5a 4.0 0.0004 0.0016 0.5a 8 0.0000 0.0001
a 1.00 0.0844 0.0844 a 2.0 0.0085 0.0085 a 4 0.0004 0.0004
2a 0.50 0.2733 0.0683 2a 1.0 0.0844 0.0211 2a 2 0.0085 0.0021
5a 0.20 0.4329 0.0173 5a 0.4 0.3295 0.0132 5a 0.8 0.1386 0.0055
10a 0.10 0.4657 0.0047 10a 0.2 0.4329 0.0043 10a 0.4 0.3295 0.0033

σza /Q
2

0.00 0.02 0.04 0.06 0.08 0.10


0
r=a
2
r=2a
4 r=4a
z/a 6

10

12

4
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Stress isobar or Pressure Bulb


An ‘isobar’ is a stress contour or a line which connects all points below the ground
surface at which the vertical pressure is the same. In fact, an isobar is a spatial curved
surface and resembles a bulb in shape; this is because the vertical pressure at all points in
a horizontal plane at equal radial distances from the load is the same. Thus, the stress
isobar is also called the ‘bulb of pressure’ or simply the ‘pressure bulb’. The vertical
pressure at each point on the pressure bulb is the same. An isobar diagram, consisting of a
system of isobars appears as shown in Fig 1.3.
Q

Isobars

Z
Fig 1.3 Isobar diagram
The procedure for plotting an isobar is as follows. Let it be required to plot an isobar for
which σz=0.2Q per unit area (20% isobar)
2 2
σ z 0.2 Q z
From σ = I Q
⇒ Iz = z ⇒ Iz = = 0.2 z 2
z z 2
z Q Q
Assuming various values for z, the corresponding Iz- values are computed; for these
values of Iz, the corresponding r/z-values are obtained; and, for the assumed values of z,
r-values are got. It is obvious that, for the same value of r on any side of the z-axis, or
line of action of the point load, the value of σz is the same; hence the isobar is
symmetrical with respect to this axis.
When r=0, Iz=0.4775; the isobar crosses the line of action of the load at a depth
of: I z = 0.2 z 2 ⇒ z 2 = I z / 0.2 ⇒ z = I z / 0.2 = 0.4775 / 0.2 = 1.545 units
The calculations are best performed in the form of a table as given below. The plot is
shown too.
z Iz r/z r (unit) σz
r
-1.0 -0.5 0.0 0.5 1.0
0.5 0.0500 1.211 0.605 0.2Q
0
1.0 0.2000 0.645 0.645 0.2Q
1.5 0.4500 0.155 0.232 0.2Q 0.4
1.5452 0.4775 0.0 0.000 0.2Q
0.8
z

1.2

1.6

Westergaard (1938) has obtained an elastic solution for stress distribution in soil under a
point load. He has assumed the soil to be laterally reinforced by numerous, closely spaced
horizontal sheets of negligible thickness but of infinite rigidity, which prevent it from

5
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

undergoing lateral strain. The vertical stress σz caused by a point load, as obtained by
Westergaard, is given by (all notations as above): HL\CH1.doc
1 1 − 2v
Q 2π 2 − 2v
σz = ⋅ (1.12)
2 ⎡ 32
1 − 2v ⎞ ⎛ r ⎞ ⎤
2
⎢⎛⎜
z
⎟ +⎜ ⎟ ⎥
⎢⎝ 2 − 2v ⎠ ⎝ z ⎠ ⎥
⎣ ⎦
1.3 Stress Due to a Line Load
a) Infinite Length: The increases in stresses at point A (Fig 1.4) due to an infinite line load
of q/unit length are:
2q z 3
σz = (1.13)
π ( z 2 + x 2 )2
q/unit length
x 2q x 2 z
σx = (1.14)
π ( z 2 + x2 )2
2q z 2 x
σz τ xz = (1.15)
σx
A π ( z 2 + x2 )2
Z
Fig 1.4: Line load of infinite length
b) Finite Length: The increases in stresses at point A (Fig 1.4-a) due to finite line load of
q/unit length are: 3

⎛ ⎞
3⎤
3q z ⎢ y 1⎜ y ⎟ ⎥
σz = − ⎜ (1.13a)
B 22+ x 2 )2 ⎢ 2 2 2 3 ⎜ 2 2 2



(z
2ππ( ⎢ x +y +z ⎝ x +y +z ⎠ ⎥
q/unit length ⎣⎢ ⎦⎥
Rearranging and substituti ng m = x/z and n = y/z equation (1.13a) becomes
⎡ 3⎤
o y σz =
q ⎢ 3n


⎜ n ⎞ ⎥
⎟ (1.13b)
2 ⎢ 2 2 ⎜ 2 2 ⎟ ⎥
y x 2 π z ⎛⎜ m 2 + 1⎞⎟ ⎢ n + m +1 ⎝ n + m +1 ⎠ ⎥
⎣ ⎦
⎝ ⎠
q
x z or σz =
z
I (1.13c)
σz
σy σx 1

⎢ 3n ⎛ n
3⎤
⎞ ⎥
where, I= ⎢ 2 −⎜ ⎟ (1.13d)
A 2 2 ⎜ 2 2 ⎟ ⎥
-z 2 π⎜⎛ m 2 + 1⎟⎞ ⎢ n + m +1 ⎝ n + m +1 ⎠ ⎥
⎣ ⎦
⎝ ⎠
Fig 1.4a: Line load of finite length
1.4 Stress Due to Uniform load on an Infinite Strip
Let a uniform load of intensity q/unit area be acting on a strip of infinite length and a
constant width B (=2b) as shown in Fig 1.5. The increase in stress at point A due to the
strip load q are given with reference to Fig 1.5 as
B=2b
q/unit area
x σz =
q
[α + sin α cos(α + 2β )] (1.16)
π
βz
σx =
q
[α − sin α cos(α + 2β )] (1.17)
α π
σz q
τ xz = sin α sin(α + 2β ) (1.18)
σx A π
Fig 1.5: Uniform load
The vertical stresses at different depths
below the center of a uniform load of intensity q and width B are as follows:
Depth z 0.1B 0.2B 0.5B B 2B 5B 10B
σz 0.997q 0.977q 0.818q 0.550q 0.306q 0.126q 0.064q

6
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

A few typical pressure bulbs for this case of strip loading are shown in Fig 1.6.
B=2b
q/unit area
σz/q=1/8
σz/q=1/2

σz/q=1/4

Fig 1.6: Pressure bubs for strip load of infinite length

1.5 Stress Due to Triangular load on an Infinite Strip


When the applied load increases linearly across the width of the strip to give a triangular
distribution as in Fig 1.7, the increases in stress at point A are given by
q⎡x 1 ⎤
B=2b σz = ⎢ α − sin 2β ⎥ (1.19)
q/unit area π ⎣B 2 ⎦
⎡ 2 ⎤
z R 1
σ x = ⎢ α − ln 1 + sin 2β ⎥
q x
(1.20)
R1 R2 π ⎢B B R 2 2 ⎥
⎣⎢ 2 ⎦⎥
x β z
q ⎡ 2z ⎤
τ xz = ⎢1 + cos 2 β − B α ⎥ (1.21)
σx
α
σz 2π ⎣ ⎦
A
Fig 1.7: Triangular load
Results of section 1.4 and 1.5 above can be superimposed in order to estimate the stress
changes that result from the construction of embankments or the formation of cuttings in
a soil mass.

1.6 Stress Due to Circular loaded Area


Let a uniform load of intensity q/unit area be acting on a circular area of radius R. The
increase in vertical stress at a depth z below the center of a flexible circular area (Fig 1.8
(a)) is given by
⎧ 3/ 2⎫
⎪ ⎡ 1 ⎤ ⎪
σ z = q ⎨1 − ⎢ ⎥ ⎬ (1.22)
2
⎪ ⎢⎣1 + ( R / z ) ⎥⎦ ⎪
⎩ ⎭
q/unit area
2R 2R
q/unit area q/unit area
R R
z
σz r
z
σz
A
A
(a) (b)

Fig 1.8: Circular load


However, for points other than those under the center, the solutions have an extremely
complex form and are generally presented in the form of charts or tables. With reference
to point A in Fig 1.8 (b), the increase in total vertical stress is given as
σ z = qI (1.23)
σ
7
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

where the influence factor Iσ depends on R, z and r. Values of Iσ as a function of the


parameters z/R and r/R can be obtained from Fig 1.9. or Table 1.2

Fig 1.9: Values of influence factor Iσ.


Table 1.2: Influence coefficients Iσ for vertical stress due to uniform load on a circular area
z/R r/R
0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0
0.5 0.911 0.840 0.418 0.060 0.010 0.003 0.000 0.000 0.000
1.0 0.646 0.560 0.335 0.125 0.043 0.016 0.007 0.003 0.000
1.5 0.424 0.374 0.256 0.137 0.064 0.029 0.013 0.007 0.002
2.0 0.284 0.258 0.194 0.127 0.073 0.041 0.022 0.012 0.006
2.5 0.200 0.186 0.150 0.109 0.073 0.044 0.028 0.017 0.011
3.0 0.146 0.137 0.117 0.091 0.066 0.045 0.031 0.022 0.015
4.0 0.087 0.083 0.076 0.061 0.052 0.041 0.031 0.024 0.018
5.0 0.057 0.056 0.052 0.045 0.039 0.033 0.027 0.022 0.018
10.0 0.015 0.014 0.014 0.013 0.013 0.013 0.012 0.012 0.011

Pressure bulbs or isobar patterns for vertical stress is shown in Fig 1.10
B=2R
q/unit area
0
0.9q
0.8q
R 0.7q
0.6q
0.5q
0.4q
0.3q
2R
0.2q

0.15q
3R

4R 0.1q
0.05q

Fig 1.10: Pressure bubs for uniform circular load

8
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

1.7 Stress Due to Uniformly Loaded Rectangular Area


The more common shape of a loaded area in foundation engineering practice is a
rectangle, especially in the case of buildings. Based on Bossinesq’s theory, Newmark
(1935) has given an expression for the vertical stress at a point below the corner of a
uniformly loaded rectangular area (Fig 1.11) as

area
nit
q/u
L m=B/z
z B
n=L/z
A
Fig 1.11: Vertical stress at the corner of a uniformly loaded rectangular area

⎡ 2 2 2 2 ⎛ 2 2 ⎞⎤
σz =
q ⎢ (2mn m + n + 1) ⋅ (m + n + 2) + tan − 1⎜ (2mn m + n + 1) ⎟⎥ (1.24)
⎢ m 2 + n 2 + 1 + m 2n 2 (m 2 + n 2 + 1) ⎜ 2 ⎟
4π ⎜ m + n 2 + 1 − m 2n 2 ⎟ ⎥
⎢⎣ ⎝ ⎠⎥⎦
Since this equation is symmetrical in m and n, the values of m and n are interchangeable.
The above equation can be rewritten in the form:
σ z = qI (1.25)
σ
Where Iσ = Influence value
⎡ 2 2 2 2 ⎛ 2 2 ⎞⎤
=
1 ⎢ (2mn m + n + 1) ⋅ (m + n + 2) + tan − 1⎜ (2mn m + n + 1) ⎟⎥
⎢ m 2 + n 2 + 1 + m 2n 2 (m 2 + n 2 + 1) ⎜ 2 ⎟
4π ⎜ m + n 2 + 1 − m 2n 2 ⎟⎥
⎢⎣ ⎝ ⎠⎦⎥
Based on this equation, Fadum (1941) has prepared a chart for the influence values for
sets of values for m and n, as shown in Fig 1.12.

9
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 1.12 Values of influence factor Iσ for calculating vertical stress at the corner

By superposition we can determine stress at any point within or out side the loaded area.
For instance the stresses at point A and point B are

T U B ⎛ ⎞
At po int A : σ z = q⎜ I +I +I +I ⎟
S ⎜ σ σ σ σ ⎟
I IV R
V ⎝ I II III IV ⎠
A
⎛ ⎞
At po int B : σ z = q⎜ I −I −I +I ⎟
II ⎜ σ σ σ σ ⎟
III ⎝ I II III IV ⎠

P Q W
PWBT:I QWBU:III
SVBT:II RVBU:IV

10
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

1.8 Stress Due to Uniformly Loaded Irregular Areas- Newmark’s Chart


Newmark (1942) devised a simple, graphical procedure for computing the vertical stress
in the interior of a soil medium (semi-infinite, homogenous, isotropic and elastic), loaded
by uniformly distributed, vertical load at the surface of a flexible area. The chart devised
is called an ‘influence Chart’.
The vertical stress underneath the center of a uniformly loaded circular area has been
⎧ 3/ 2⎫
shown to be ⎪ ⎡ 1 ⎤ ⎪ .
σ z = q ⎨1 − ⎢ ⎥ ⎬ ( See Eq. 1.22)
⎪ ⎢⎣1 + ( R / z ) 2 ⎥⎦ ⎪
⎩ ⎭
This equation can be rewritten in the form:
−2 3
R ⎛ σ ⎞
= ⎜1 − z ⎟ −1 (1.26)
z ⎜ q ⎟
⎝ ⎠
If a series of values is assigned for the ratio σz/q, such as 0, 0.1, 0.2, …., 0.9 and 1.0, a
corresponding set of values for the relative radii, R/z, may be obtained. If a particular
depth is specified, then a series of concentric circles can be drawn. Since the first has zero
radius and the 11th has infinite radius, in practice only 10 circles are drawn. Each ring or
annular space causes a stress of q/10 at a point beneath the center at the specified depth z,
since the number of annular spaces (m) is 10.
The relative radii can be tabulated as shown below:
S. No of circle σz/q Relative Radii (R/z)
1 0 0.000
2 0.1 0.270
3 0.2 0.400
4 0.3 0.518
5 0.4 0.637
6 0.5 0.766
7 0.6 0.918
8 0.7 1.110
9 0.8 1.387
Now let’s assume that a set of equally spaced rays, say n in number, is drawn emanating
from the center of the circles, thus dividing each annular area in to n sectors, and the total
area in to m*n sectors. If the usual value of 20 is adopted for n, the total number of
sectors in this case will be 10x20=200. Each sector will cause a vertical stress of 1/200th
of the total value at the center at the specified depth and is referred to as a ‘mesh’ or an
‘influence unit’. The value 1/200 or 0.005 is said to be the ‘influence value’ (or ‘influence
factor’) for the chart.
Construction of the Newmark’s chart
1. For the specified depth (say 20m), calculate the radii of the circles as:
R/z 0 0.270 0.4 0.518 0.637 0.766 0.918 1.11 1.387 1.908 ∞
R (m) 0 5.40 8.00 10.36 12.74 15.32 18.36 22.20 27.74 38.16 ∞
2. Draw circles to a convenient scale (say 1:100)
3. Draw a suitable number of uniformly spaced rays usually 20 emanating from the
center of the circles.
4. The resulting diagram will appear as shown in Fig 1.13; on it draw a vertical line
AB, representing the depth z to the scale used in drawing the circles. Thus here
AB=20cm.
11
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
One influence unit
or mesh

20cm
Influence value =0.005 B

Fig 1.13: Newmark’s influence chart


This diagram can be used for other values of z by considering the scale to which it is
drawn alters. So if z is to be 10m, line AB now represents z in scale of 1:50.

Use of the Newmark’s chart


The chart can be used for any uniformly loaded foundation of any shape. The procedure
is as follows
1. Draw the loaded area on a tracing paper, using the same scale to which the
distance AB on the chart represents the specified depth.
2. Place the plan drawn above in such a way that the point under which the vertical
stress σz is required coincides with the center of the circles on the chart.
3. Count the number of elements on the chart enclosed by loaded area including
fractional units, if any; let this total equivalent number be N.
4. Determine the vertical stress σz at the specified depth as σz =INq, where I=
influence value of the chart.

1.9 Approximate Methods


Approximate methods are used to determine the stress distribution in soil under the
influence of the complex loadings and/or shapes of loaded areas, saving time and labour
without sacrificing accuracy to any significant degree. The following are some.
a) Equivalent Point Load Method
Q2 Q3

Q1
Q4
R3
R4 R2
B
R1 z

A
Fig 1.14: Equivalent Point Load Method
Referring to Fig 1.14, if the influence values are Iσ1, Iσ2, Iσ3, …. for the point loads Q1,
Q2, Q3,…, then σz is found as:
σz = (Q1I1+ Q2I2+ Q3I3+…..) (1.27)
This method gives a good result if z/B> 3.
12
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

b) Two is to One Method

Fig 1.15: Two to One Method


The average vertical stress at depth z is obtained as:
q.B.L
σz = ( 1.27)
( B + z )( L + z )
c) The 30o or 60o Method

Fig 1.16: 30o or 60o Method


The average vertical stress at depth z is obtained as:
q.B.L
σz = ( 1.28)
( B + 1.5 z )( L + 1.5 z )

1.10 Initial Settlement or Elastic Compression


This is also referred to as the 'immediate or distortion or contact settlement' and it is
usually taken to occur immediately on application of the foundation load (within about 7
days).
Immediate settlement computation
The settlement of the corner of a rectangular base (flexible) of dimensions B' X L'
on the surface of an elastic half-space can be computed from an equation from the
Theory of Elasticity [e.g., Timoshenko and Goodier (1951)] as follows:
⎛1− v2 ⎞
S i = qo B' ⎜⎜ ⎟⎟ I S I F
⎝ Es ⎠
qo = intensity of contact pressure in units of Es
B' = least lateral dimension of contributing base area in units of S.
Es, v = elastic soil parameters
Ii = influence factors, which depend on L'/B', thickness of stratum H, Poisson's ratio v,
13
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

and base embedment depth D. The influence factor Is (see Figure 3.1 for identification of
terms) can be computed using equations given by Steinbrenner (1934) as follows:
1 − 2v
I S = I1 + I2 with I1 and I 2 as follows:
1− v
1⎡ ⎛ (1 + M 2 + 1)( M 2 + N 2 ) ⎞ ⎛ 2 ⎞⎤
⎟ + ln⎜ (M + M + 1)( 1 + N ) ⎟⎥
2
I1 = ⎢M ln⎜
π⎢ ⎜ M (1 + M 2 + N 2 + 1) ⎟ ⎜ M + M 2 + N 2 + 1 ⎟⎥
⎣ ⎝ ⎠ ⎝ ⎠⎦
N ⎡ M ⎤
I2 = tan −1 ⎢ ⎥ tan −1 in radians
2π ⎣ N M + N +1 ⎦
2 2

L' H
where; M = and N=
B' B'
B L
B' = for center and B' = B for corner I i ; L' = for center and L' = L for corner I i
2 2
IF = influence factor from the Fox (1948b) equations, which suggest that the settlement
is reduced when it is placed at some depth in the ground, depending on Poisson’s ratio
and L/B. Figure 3.1 can be used to approximate IF.
Note: if your base is "rigid" you should reduce the Is factor by about 7 percent (that is, Is,
rigid = O.931 Is, flexible)

Figure 1.17: Influence factor IF for footing at a depth D. Use actual footing width and
depth dimension for this D/B ratio.

14
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

CHAPTER TWO

SHEAR STRENGTH OF SOILS

2.1 Introduction HL\CH2.doc


The shear strength of a soil is its maximum resistance to shear stresses and a consequent
tendency for shear deformation. Its value determines such factors as the stability of slopes
(cuts, earth dams), the allowable bearing capacity of a foundation and the thrust of a soil
against retaining walls. Knowledge of the shear strength is thus an essential prerequisite
to any analysis concerned with the stability of a soil mass.
Basically, a soil drives its shearing strength from the following:
1) Resistance due to the interlocking of particles.
2) Frictional resistance between individual soil grains, which may be sliding
friction, rolling friction, or both.
3) Adhesion between soil particles or ‘cohesion’
Granular soils or sands may derive their shear strength from the first of the two sources,
while cohesive soils or clays may derive their shear strength from the second and third
source. Highly plastic clays, however, may exhibit the third source alone for their
shearing strength. Most natural soil deposits are partly cohesive and partly granular and
as such, may fail in to the second of the three categories just mentioned, from the point of
view of shearing strength.

2.2 Basic Concepts of Shearing Resistance


a) Friction between Solid Bodies
Consider a prismatic block resting on the surface MN (Fig 2.1). The block is subjected to
two forces: force Pn acting at right angle to surface MN and force Fa acting tangentially to
the plane MN.
Let Pn remain constant, Fa increases
gradually until sliding starts.
• If Fa is relatively small, the block
will remain at rest and the
resisting force can be written as
Fr=Pn tanδ. This resisting force is
due to surface roughness between
block and plane.
Fig 2.1 Friction between bodies
• As Fa increases, Fr also increases such that Fa=Fr. The block will start sliding
when angle of obliquity δ reaches a maximum value, δm.
o If the block and surface MN are the same material, δm= φ (φ=angle of
internal friction) and thus tan φ =μ (μ =coefficient of friction)
o If the block and surface MN are different materials, then δm= angle of skin
(wall) friction.

b) Internal Friction within Soil Masses


In granular or cohesionless soil masses, the resistance to sliding on any plane through the
point with in the mass involves the movement of one particle relative to another.
15
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

The components giving rise to movement include:


1) Sliding or rolling
2) Removal of interlocking between particles

The angle of internal friction, which is a limiting angle of obliquity and hence the primary
criterion for slip or failure to occur in a certain plane, varies appreciably for given sand
with the density, since the degree of interlocking is known to be directly dependent upon
the density. This angle also varies somewhat with the normal stress. However, the angle
of internal friction is mostly considered constant, sine it is almost so for a given sand at a
given density.

Since failure or slip within a soil mass cannot be restricted to any specific plane, it is
necessary to understand the relationships that exist between the stresses on different
planes passing through a point, as a prerequisite for further consideration of shearing
strength of soils.

2.3 Principal Planes and principal Stresses-Mohr’s Circle


Consider a solid mass acted on by a series of forces F1, F2, …, Fn in a 2D Plane (Fig 2.2
(a)). These forces produce stresses at a point with in the body. And every plane passing
through that point will be subjected, in general, to a normal or direct stress and a shearing
stress. In geotechnical engineering, compressive normal stress and shear producing
counter-clockwise couple/moment on the element or clockwise moment about a point just
out side the element face are usually considered positive (Fig 2.2 (b)).

Fig 2.2: a) Solid mass acted by forces b) Sign convention

Now let’s consider an element of soil whose sides are chosen as the principal planes
(planes which do not have shear stresses), the major and the minor, as shown in Fig 2.3
(a). Let it be required to determine the stress conditions on a plane inclined at angle θ
from the major principal plane measured CCW (Fig 2.3(b)). Considering the equilibrium
of this element redrawn as Fig 2.3 (c), we have
σ θ = σ1 cos2 θ + σ 3 sin 2 θ = σ 3 + (σ1 − σ 3 ) cos2 θ
(σ − σ ) (σ − σ )
= 1 3 + 1 3 cos 2θ (2.1)
2 2
(σ − σ )
τ = 1 3 sin 2θ (2.2)
θ 2
16
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 2.3 Stresses on a plane


If we square and add equations 2.1 and 2.2, we have
2 2
⎛σ +σ ⎞ ⎛ σ + σ ⎞⎛ σ − σ ⎞ ⎛σ −σ ⎞
σ 2θ + τ 2θ = ⎜ 1 3 ⎟ + 2⎜ 1 3 ⎟⎜ 1 3 ⎟ cos 2θ + ⎜ 1 3 ⎟
⎜ 2 ⎟ ⎜ 2 ⎟⎜ 2 ⎟ ⎜ 2 ⎟
⎝ ⎠ ⎝ ⎠⎝ ⎠ ⎝ ⎠
2 2
⎛σ +σ3 ⎞ ⎛ σ + σ 3 ⎞⎛ σ 1 − σ 3 ⎞ ⎛σ −σ3 ⎞
σ 2θ − ⎜ 1 ⎟ − 2⎜ 1 ⎟⎜ ⎟ cos 2θ + τ 2θ = ⎜ 1 ⎟
⎜ 2 ⎟ ⎜ 2 ⎟⎜ 2 ⎟ ⎜ 2 ⎟
⎝ ⎠ ⎝ ⎠⎝ ⎠ ⎝ ⎠
2 2
⎡ ⎛ σ + σ 3 ⎞⎤ ⎛σ −σ3 ⎞
⎢σ − ⎜ 1 ⎟⎥ + τ 2θ = ⎜ 1 ⎟ (2.3)
θ ⎜ 2 ⎟⎥ ⎜ ⎟
⎣⎢ ⎝ ⎠⎦ ⎝ 2 ⎠

This is an equation of the circle with radius σ −σ


1 3 and center ( σ1 + σ 3 , 0). This circle is
2 2

drawn in τ,δ space (Fig 2.4) and is known as Mohr’s circle of stress.

Fig 2.4 Mohr’s Circle of stress


Once this circle is drawn, we can determine normal shear stresses at any plane through a
point. The procedure is based on a unique point called ‘pole’ or ‘origin of planes’. Once
the pole is known the stress on any plane can readily be found by drawing a line from the
pole parallel to the plane where we want stresses. The coordinates of the point of
intersection with the Mohr’s circle determine the stresses on that pale.
2.4 Strength Theories for Soils
a) Coulomb Failure Equation
In 1976 Coulomb observed that if the thrust of a soil against a retaining wall caused the
wall to move forward slightly, an essentially straight slip plane formed in the retained
soil. He postulated that the maximum resistance to shear, τf, on the failure plane is given
by
τ Coulomb’s Envelope
τf = c + σf tanφ (2.4)
τf = c + σf tanφ
Where: σ is the total stress normal to the failure plane
φ is the angle of internal friction φ
c is the cohesion of the soil c σ

17
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

The use of Coulomb’s equation didn’t always result in successful design of soil
structures. This is due to the fact that water cannot sustain shear stress; the shear
resistance of a soil must result solely from the frictional resistance arising at the particle
contact points, the magnitude of which depends solely on the magnitude of the effective
stress carried by the soil skeleton. Hence, expressing Coulomb’s equation in terms of
effective stress:
τf = c' + σ'f tanφ' (2.5)
in which the parameters c' and φ' are properties of the soil skeleton, referred to
respectively as the effective cohesion and the effective angle of friction.

b) Mohr-Coulomb Theory of Shear strength


The Mohr- Coulomb theory of shearing strength of soil, first postulated by Coulomb
(1976) and later generalized by Mohr, is the most commonly used concept. Mohr stated
that when shear stress on the failure plane reaches a unique function of the normal stress
on the failure plane, failure takes place. Thus if we conduct several tests and obtain
principal stresses at failure, the Mohr’s circles at failure can be constructed as in Fig 2.5.
A line drawn tangent to the failure circles is called an envelope (Mohr failure envelope).
Therefore, it can be said that the Mohr’s circle of stress relating to a given stress
condition would represent, incipient failure condition if it just touches or is tangent to the
strength or failure envelope (circles C, D, E); otherwise, it would wholly lie below the
envelope as shown in circle A. circles lying above the Mohr failure envelope (circle B)
cannot exist.

2.5 Determination of Shearing Strength of Soils


Determination of shearing strength of a soil involves the plotting of failure envelopes and
evaluation of the shear strength parameters for the necessary conditions. The following
tests are available for this purpose:
Laboratory tests:
Field tests:
1. Direct Shear Test
1. Vane Shear Test
2. Triaxial Compression Test
2. Penetration Tests
3. Unconfined Compression Test
4. Laboratory Vane Shear Test

1) Direct Shear Test


The direct shear device, also called ‘the shear box apparatus’, essentially consists of a
brass box, split horizontally at mid height of the soil specimen, as schematically shown in
Fig 2.6. The soil sample is gripped in perforated metal grills, behind which a porous discs

18
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

can be placed if required to allow the specimen to drain. For undrained tests, metal plates
and solid metal grilles may be used.

Fig 2.6: Schematic representation of Direct Shear Test


After the sample to be tested is placed in the apparatus or shear box, a normal load which
is vertical is applied to the top of the sample by means of a loading yoke and weights.
Since the shear plane is predetermined as the horizontal plane, this becomes the normal
stress on the failure plane, which is kept constant through out the test. A shearing force is
applied to the upper-half of the box, which is zero initially and is increased until the
specimen fails.
Two types of application of shear are possible- one in which the shear stress is controlled
and the other in which the shear strain is controlled. In the stress controlled type, the
shear stress which is controlled variable, may be applied at a constant rate or more
commonly in equal increments by means of calibrated weights hung from a hanger
attached to a wire passing over a pulley. Each increments of shearing force is applied and
held constant, until the shearing deformation ceases. The shear displacement is measured
with the aid of a dial gauge attached to the side of the box. In the strain-controlled type,
the shear displacement is applied at a constant rate by means of a screw operated
manually or by motor. With this type of test the shearing force necessary to overcome the
resistance within the soil is automatically developed. The shearing force is measured with
the aid of a proving ring-a steel ring that has been carefully machined, balanced and
calibrated. The shear displacement is measured with the aid of another dial gauge
attached to the side of the box.
In both cases, a dial gauge attached to the plunger, through which the normal load is
applied, will enable one to determine the changes in the thickness of the soil sample
which will help in the computation of volume changes of the sample, if any. HL\CH2.doc

For the readings taken, we can determine the following

1) Normal stress: σi = Pi/A 2) Shear stress: τi = Fi/A


3) Vertical Strain (Volume change) εn = Δh/h 3) Shear Strain εh = ΔA/A

And the following plots can be produced (Example Fig 2.7 for overconsolidated clay)
1) Normal stress vs. shear stress (σ vs. τ) 2) Shear strain vs. shear stress (εh vs. τ)

19
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

3) Normal stress vs. volume change (σ vs. εh) 4) Mohr circle at failure

Fig 2.7 Different plots


The direct shear test is relatively simple and quick drainage is possible as the thickness of
the sample is small. However, it has some draw backs like
1. The failure plane is predetermined and this may not be the weakest plane.*
2. Drainage conditions cannot be controlled
3. The normal and shear stresses in a plane are not uniform
4. The area of the sliding surface is not constant – better to account it
5. The effect of the lateral restraints by the side walls of the shear box is likely to
affect the results.

2) Triaxial Compression Test


Apparatus and procedure Triaxial HL\CH2.doc
Fig 2.8 shows the essential details of the triaxial cell in which a soil sample is sealed with
in water tight rubber membrane and enclosed in a cell filled with water through which a
confining pressure is applied to the sample. Drainage facility is provided by strips of filter
paper placed vertically around the sample. These connect with a pours disc in the top
platen, take-off being through a nylon tube which passes out of the cell through its base.
The pore water pressure in the sample is measured through a saturated porous disc sealed
flush in the base pedestal and connected via a water-filled duct to an electrical pressure
transducer. The sample is sheared by a vertical piston load applied through the top platen.
To minimize the friction forces at the top and bottom of the sample and allow an
unrestricted lateral deformation during shear, greased rubber discs are placed between the
sample and the end caps. The form of the test may either be strain controlled, in which
case the vertical piston load is applied by a motorized loading frame greased to deform
the sample vertically at a constant rate of strain, or stress controlled, in which case the
sample is allowed to strain freely under a vertical piston load applied by dead weights.

20
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 2.8 Triaxial cell


By eliminating the shear stresses on the top and bottom boundaries of the sample,
and since no shear stress can act between the cell water and the vertical surface of the
sample, it follows that the axial stress and the ambient cell pressure are both principal
stresses. Under the conditions of the triaxial test, the major principal stress σ1 is the axial
stress and the intermediate and minor principal stresses, σ2 and σ3, are both equal to the
all-round cell pressure. This of course acts not only on the vertical surface of the sample
but also on the end faces. Thus, if at some stage of the shear test the vertical piston load is
P and the current cross-sectional area of the sample is A then
P P
σ1 = + σ 3 ⇒ = σ1 − σ 3
A A
The applied piston stress P/A is therefore in actual fact equal to the difference between
the major and minor principal total stresses, σ1-σ3, which is termed the deviator stress
and denoted as Δσ.
The application of the all-round cell pressure and application of the deviator stress
form two separate stages of the test. Whether drainage of the sample is permitted during
each stage depends upon the soil type and nature of the field problem being investigated.
A soil element in the field (notably an element of clay soil) may fail either undrained,
partially drained or drained. We thus have three basic types of tests: Undrained,
consolidated-undrained and drained.

Area correction for the Determination of Deviator Stress


During the application of the axial load, the specimen undergoes axial compression and
horizontal expansion to some extent. Little error is introduced if the volume (or area) is
assumed constant, although the area of the x-section varies as axial strain increases. This
assumption is perfectly valid if the test is conducted under undrained conditions, but, for
drained conditions, the exact relationship is somewhat different and given below.
If Ao, ho, and Vo are the initial area of cross section, height and volume of the soil
specimen respectively, and if A, h, V are the corresponding values at any stage of the test,
the corresponding changes in the values being designated ΔA, Δh, and ΔV, then

21
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
⎛ ΔV ⎞ ⎛ ΔV ⎞
V=A (ho - Δh) = Vo - ΔV ⇒ V ⎜1 −
⎜ V ⎟
⎟ ⎜1 −
⎜ V ⎟

Since axial strain,
V − ΔV o
⎝ o⎠ ⎝ o⎠
A= o = =A
h − Δh ⎛ Δh ⎞ o 1− ε εa=Δh/h
h ⎜1 − ⎟
o a
o⎜ h ⎟
⎝ o⎠
A 1
For undrained tests, ΔV = 0 ⇒ A = o . And the ratio is called the correction
1− ε 1− ε
a a
factor. Once the corrected area is determined, the additional axial stress or the deviator
stress, Δσ, is obtained as
Axial Load ( from proving ring reading)
Δσ = σ − σ =
1 3 corrected area

Mohr’s Circle for Triaxial Test (Relationship between principal


stresses and shear strength Parameters)
The stress condition in a triaxial test may be represented by a Mohr’s circle, at any stage
of the test, as well as at failure as shown in Fig 2.8 (a). The cell pressure, σc, which is also
the minor principal stress −σ3 is constant and σ11, σ22, σ13, …, σ1f are the major principal
stresses at different stages of loading and at failure. The Mohr circle at failure will be
tangential to the Mohr-Coulomb strength envelope, while those at intermediate stages
will be lying wholly below it. The Mohr’s circle at failure will be as shown in Fig 2.8 (d).

Fig 2.8: Mohr’s Circle at Failure


The Mohr’s circles at failure for one particular cell pressure are shown for the three
typical cases of a φ-soil, a c-soil and a general c-φ soil in Figs. 2.8 (b), (c), and (d)
respectively.
With reference to Fig 2.8 (d), the relation ship between the major and the minor principal
stresses at failure may be established from the geometry of the Mohr’s circle, as follows:
From Δ AOF, 2θ = 90o +φ ⇒ θ = 45o+φ/2 ⇒ the failure plane is inclined at an
angle of 45o+φ/2 to the major principal plane.
Again from Δ AOF, sinφ = FO/AO= FO/(AB+ BO)

22
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
(σ − σ ) / 2 (σ − σ )
1 3 1 3
⇒ sinφ = =
c cotφ + (σ + σ ) / 2 2c cotφ + (σ + σ )
1 3 1 3
⇒ (σ1 − σ 3 ) = 2c cosφ + (σ1 + σ 3 ) sinφ ⇒ σ1 (1 − sinφ ) = σ 3 (1 + sinφ ) + 2c cosφ
σ 3 (1 + sinφ ) 2c cosφ
⇒σ1 = +
(1 − sinφ ) (1 − sinφ )
⇒ σ1 = σ 3 tan2 (45ο + φ / 2) + 2c tan(45ο + φ / 2) = σ 3 tan2 θ + 2c tanθ (2.6)
This is also written as σ1 = σ 3 Nφ + 2c Nφ where Nφ = tan θ = tan (45 + φ /2)
2 2 o

These state of stress at failure is defined as ‘plastic equilibrium condition’, when failure is
imminent.
From one test, a set of σ1 and σ3 is known; but it can be seen from equation 2.6 that at
least two such sets are necessary to evaluate the parameters c and φ. Conventionally, three
or more such sets are used from a corresponding number of tests. However, the usual
procedure is to plot the Mohr’s circles for a number of tests and take the best common
tangent to the circles as the strength envelope. A small curvature occurs in the strength
envelope of most soils, but since the effect is slight, the envelope for all practical
purposes, may be taken as a straight line. The intercept of the strength envelope on the τ-
axis gives cohesion and the angle of the slope of this line with the σ-axis gives the angle
of internal friction.
In derivation of equation 2.6, we have (σ 1 − σ 3 ) = 2c cos φ + (σ 1 + σ 3 ) sin φ
⇒ (σ 1 − σ 3 ) / 2 = c cos φ + ((σ 1 + σ 3 ) / 2) sin φ
⇒ q = c cos φ + p sin φ
which plots as a straight line as q against p (Fig 2.9), the slope of the line defining sinφ
and the intercept ccosφ, and thus provide an alternate method for determining shear
strength parameters from the results of triaxial tests.

Fig 2.9: Alternative Procedure of Evaluating shear strength parameters


The application of the all-round cell pressure and application of the deviator stress form
two separate stages of the test. Whether drainage of the sample is permitted during each
stage depends upon the soil type and nature of the field problem being investigated.
i) Undrained Test (UU test) HL\CH2.doc
Drainage is not permitted at any stage of the test, that is, either before the test during the
application of the normal stress or during the test when the shear stress is applied. Hence
no time is allowed for dissipation of pore water pressure and consequent consolidation of
the soil; also, no significant volume changes are expected. Usually, 5 to 10 minutes may
be adequate for the whole test (called ‘quick test’ or Q or Qu test), because of the
shortness of drainage path. However, undrained tests are often performed only on soils of
low permeability.

23
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

This test is applicable to the short term stability analysis of works formed on or in clay
deposits where it is considered that insufficient time has elapsed for any dissipation of
excess porewater pressure to have occurred by the end of construction. Such works
generally include small embankments, cuttings, retaining walls and foundation of
buildings.
Test on Saturated Clay
If the clay sample is saturated, the increase in cell pressure is carried entirely by the
porewater as an excess porewater pressure with no change in effective stress in the
sample and therefore no change in shear strength. Thus, the deviator stress required to
fail the sample is independent of the cell pressure at which the test is run. Fig 2.10 shows
the corresponding Mohr stress circles, the common tangent to which defines the failure
envelope for the soil, which in this case is horizontal giving φu=0; the intercept on the
vertical shear stress axis defines the undrained cohesion Cu. The undrained shear
strength, τu, in terms of the total stress is thus τu= Cu + σtanφu, with φu =0,
τfu = Cu. (2.7)
The undrained cohesion thus defines the undrained shear strength. And Cu is more
generally referred to as the undrained shear strength.

Fig 2.10: Failure envelope for undrained tests on Saturated Clay


Also, Cu = (σ1 - σ3)/2 = Δσf /2 , where the deviator stress at failure, Δσf, defines
the compression strength of the sample.
Here it is clear that the deviator stress, (σ1 - σ3) =constant for any values of cell-pressure,
σ3, used. Thus the parameters Cu and φu are unique and independent of the test procedure
used or the stress path followed. This is particularly important result for it means that soil
elements in the field where the stress path conditions to failure are more complex with σ1
and σ3 both varying, will all have the same value of Cu and φu=0. This would suggest that
where the stress change likely to cause failure occurs under undrained conditions, the
stability analysis may be carried out in terms of total stresses. This particular form of
analysis, known as the φu=0 analysis, leads to a very simple and rapid assessment of
stability since the undrained shear strength at any point is defined solely by the undrained
cohesion Cu.
If the porewater pressure is measured in a series of undrained tests on saturated samples it
is found that the same effective stress failure circle is obtained in each case, and hence the
effective stress parameters C' and φ' cannot be determined from such tests. Recourse is
made to consolidated-undrained or drained tests for this purpose.

24
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Unconfined Compression Test


This is a special case of triaxial test with σ3=0 or σc=0. The test is carried out in saturated
cohesive soil sample which can stand without lateral support (rubber membrane). This
test may be conducted on undisturbed or remolded cohesive soils. It cannot be conducted
on coarse grained soils such as sands and gravels as theses cannot stand with out lateral
support.
From above,
τfu = Cu = (σ1 - σ3)/2 = Δσf /2 ⇒ τfu = Cu = σ1 /2
Usually σ1 is designated in this case as qu and is referred as unconfined compressive
strength. Thus,
τfu = Cu = qu /2 (2.8)
The Mohr’s circle for this case is shown in Fig 2.10 and designated as circle A.
The value of the undrained shear strength of clay may be used to indicate its consistency
state (Table 2.1).
Table: 2.1 Consistency of Clays
Undrained shear strength Cu(Kpa)
Consistency BS5930:1981 Terzaghi and Peck: 1967
Very soft <20 <12
Soft 20-40 12-25
Firm 40-75 25-50
Medium 40-75 25-50
Stiff 75-150 50-100
Very stiff >150 100-200
Hard >200 correlationHL\CH2.doc

Test on Partially saturated Clay


Undrained tests may also be carried out on partially saturated soils, for example,
samples of a rolled clay fill or laboratory-compacted samples of earth fill. In this case the
increase in cell pressure is carried partly by the pore water and partly by the soil skeleton.
Thus, the greater σ3 the greater the increase in the effective confining pressure, and
therefore the greater the increase in shear strength and consequently the greater the
increase in deviator stress required to cause failure. The increase in the deviator stress
becomes progressively smaller and finally ceases when the applied stresses are large
enough to force all the air in the voids in to solution in the pore water, when the sample
behaves as a saturated soil with φu=0. The failure envelope with respect to the total stress
is curved (Fig 2.11(a)) and values of Cu and φu vary with the magnitude of the normal
stress on the failure plane. If the pore water pressure is measured during each test, Mohr
circles of effective stress can be drawn and it is found that the failure envelope in terms of
the effective stress is linear over a wide range of σ’(Fig 2.11(b)).
Curved failure envelope

Fig 2.11 Failure envelopes for undrained tests on partially saturated clay w.r.t (a) total stress (b) effective stress

25
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

ii)Consolidated Undrained Test (CU test)


In the previous section we have considered the relevance of the undrained test to model a
field situation where a clay deposit is subjected to a stress change which is rapid in
relation to the time for the dissipation of excess porewater pressure so that a potential
failure would occur under undrained conditions. If, however, the construction period
extends over several seasons (for example, in the case of an earth dam) it is reasonable to
assume that some consolidation of the soil mass will have occurred by the end of
construction. If at this stage the shear stresses induced in the soil are of sufficient
magnitude to cause failure, this would occur quickly under conditions of no further
drainage. This behavior is modeled in the consolidated-undrained test in which a soil
sample is fully consolidated under the cell pressure and then sheared by the deviator
stress under undrained conditions. Fig 2.12 shows a typical deviator stress-axial strain
and porewater pressure-axial strain curves obtained for normally consolidated and over
consolidated clays. For a normally consolidated soil the pore water pressure rises to
failure, reflecting the volume decrease that would occur if drainage were to be allowed.
For a heavily over consolidated clay the porewater pressure decreases during shear,
reflecting the dilatancy that would occur if the sample were free to undergo volume
change.

Fig 2.12: Stress-strain-pore pressure relationships for consolidated-undrained tests on clay


The greater the pressure to which a sample is consolidated, the greater the deviator stress
required to cause failure. Fig 2.13(a) shows typical Mohr circles of total stress. The
intercept and slope of the failure envelope define the total stress shear strength parameters
for the soil, which for a consolidated-undrained test are denoted by Ccu and φcu
respectively. If the porewater pressure is measured during the test, as is the usual practice,
then the corresponding Mohr circles of effective stress may be drawn (Fig 2.13(b)), the
failure envelope now defining the effective stress shear strength parameters C’ and φ'.
The effective stress circles may lie either to the left or right of their respective total stress
circles, depending on whether the porewater pressure is positive or negative.

Fig 2.12: Failure envelopes for consolidated-undrained tests on clay w. r. t (a) total stress, (b) effective stress.

26
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

For normally consolidated clays, the failure envelopes with respect to total and effective
stress pass through the origin giving Ccu and C' equal to zero; for overconsolidated clays
Ccu and C' usually range between 5 and 30 KPa. The value of φ' is not influenced
appreciably by overconsolidation and ranges from about 30o to 20o, decreasing with
increasing plasticity index. The value of φcu varies similarly and its relation to φ' is
determined by the magnitude of the porewater pressure at failure.
In the standard triaxial test the sample is failed in a conventional manner by holding σ3
constant and increasing σ1. In a field situation the stress changes leading to a potential
failure of an element are more complex than this and in general range from σ3 constant
and σ1 increasing to σ1 constant and σ3 decreasing. It is important to consider the possible
influence of this on the measured values of the shear strength parameters. This is
illustrated in Fig 2.14 for the case of normally consolidated clay. In test 1 a soil sample is
fully consolidated under a cell pressure σ3.

Fig 2.14: Influence of test procedure on φcu and φ’


The drainage valve is then closed and the sample sheared to failure under undrained
conditions by increasing σ1. The line OX then defines the value of φcu. In test 1 an
identical soil sample is consolidated to the same cell pressure σ3. The drainage valve is
then closed and the sample failed by decreasing σ3, at the same time maintaining σ1
constant and equal to the initial cell consolidation pressure. Since the pressures to which
the samples have been consolidated are the same, then the deviator stress at failure will be
the same as in test 1. The line OY then defines the value of φcu. It is thus seen that the
value of φcu is not unique and therefore of limited application in practice. Considering
effective stress conditions, in test 1 the porewater pressure at failure is positive therefore
the effective stress circle lies to the left of the total stress circle. In test 2, failure by
unloading the sample results in a negative porewater pressure and consequently the
effective stress circle lies to the right of the total stress circle. It is found that the two
effective stress circles coincide (circle 3). The value of φ’ is therefore unique and
consequently has a much wider application in practice. Thus, if any drainage of a soil
mass has occurred by the end of construction, stability analysis should be carried out in terms of
effective stress. This is logical since it is the effective stress in a soil that controls its shear
strength.

iii) Drained Test (CU test)


Where construction in the field is on sand or gravel deposit, a potential failure would
occur under drained conditions. Cuttings in clay may also fail many years after
construction when the initial (negative) excess porewater pressures have fully dissipated.
These conditions are simulated by the drained test in which a sample is fully consolidated
under the cell pressure and then sheared under drained conditions at a rate compatible
with no build-up in excess porewater pressure, so that u remains equal to zero throughout
27
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

the shearing stage. The form of the stress-strain-volume change relationship for sands and
clays are similar to those obtained from direct shear tests and are given in Fig 2.15.

(a) Sand

(b) Clay
Fig 2.15: Stress-strain-volume change relationships for drained tests (a) Sands (b) clay

Since u=o through out the shearing process, σ' 3=σ3, σ' 1=σ1 and the Mohr circles of
effective stress and total stress coincide (Fig 2.16). The failure envelope defines the
effective stress parameters C' and φ'. These are often denoted Cd and φd respectively.
Generally there is little difference between the effective stress parameters obtained from
drained and consolidated-undrained tests. However, for sands and heavily
overconsolidated clays φd is slightly higher than φ', because of the work done by the
sample as it expands against the confining pressure during shear.

Fig 2.16: Failure envelope for drained tests on clay

iv) Pore Pressure Parameters


In addition to determining the shear strength parameters of a soil, the triaxial test is also
used to furnish data for predicting the initial excess porewater pressure set up at a point
in a soil mass by a change in the total stress conditions. Such predictions are required in
conjunction with an effective stress stability analysis and are made by means of
experimentally determined pore pressure parameters. Pore parametersHL\CH2.doc

28
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

To introduce the idea of pore pressure parameters, consider the soil element shown in Fig
2.17 wherein the element is subjected to a triaxial loading in which Δσ2=Δσ3. We may
consider this stress system to be composed of an isotropic change of stress Δσ3 plus a
uniaxial change of deviator stress Δσ1-Δσ3. This, of course, is the stress system imposed
in the triaxial test and it is clear that the excess porewater pressure generated within the
element will result firstly from the change in the all-round stress and secondly from the
change in the deviator stress.

Fig 2.17: Soil element


If ΔUa denotes the excess porewater pressure induced in the element by the application of
Δσ3 and ΔUd that by Δσ1-Δσ3, then assuming these pore pressures to be a simple
proportion of the applied stresses, we may write
ΔUa = B Δσ3 and ΔUd = A (Δσ3-Δσ1)
Where B and A are experimentally determined pore water parameters. If ΔU denotes the
total excess porewater pressure in the element, then assuming the principle of
superposition, ΔU = ΔUa + ΔUd and hence
ΔU = BΔσ3+A(Δσ1-Δσ3) (2.9)
This problem was considered by Skempton (1954) who gives
ΔU = B [Δσ3+A (Δσ1-Δσ3)] (2.10)
This is of the same form as equation 2.9 where AB replaces A.
The pore pressure parameters are obtained from a triaxial compression test. The
parameter B is determined by measuring the increase in porewater pressure resulting from
the increase in cell pressure and varies from 1 for a fully saturated soil to 0 for a dry soil.
The parameter A, and hence A, is determined by measuring the porewater pressure
induced in the sample by the application of the deviator stress. The value of A depends
upon the initial consolidation stress system (whether isotropic or anisotropic), the stress
history (as reflected by the overconsolidation ratio), the proportion of the failure stress
applied (i.e, on the strain of the sample) and the type of stress change (whether loading or
unloading). Typical failure values (Af) are given in Table 2.2.
Table 2.2: Values of the pore pressure parameter A at failure (Af)
Soil type Af
Sensitive clay 1.5-2.5
Normally consolidated clay 0.7-1.3
Lightly consolidated clay 0.3-0.7
Heavily consolidated clay -0.5-0
With B=1 for a saturated soil, A=A.

29
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

v) P-q Diagrams and Stress Paths


In discussing about Mohr’s circle for triaxial tests, we saw how to represent and plot the
state of stress as stress point whose coordinates are (p, q), defined as
σ1 +σ σ1 −σ
p = 3
q = 3

2 2
Actually the p and q values represent the coordinates of a point on Mohr’s circle, with p
representing the center of the circle (always located on the abscissa, or σ axis) and q
representing the maximum shear stress, equal to the radius of the circle. The locus of p-q
points for a test series is known as a stress path. Such a graphical representation is known
as the p-q diagram. Fig 2.18 shows the Mohr envelope (φ line) developed from tangents
to the circles at points 1, 2, and 3 and line Kf (Kf line), which passes through p-q points
A, B, and C- points of maximum shear for the respective circles. Thus, the Kf line
represents a limiting state of stress at impending failure. The following relationship
between the φ line and the Kf line exist as illustrated in Fig 2.9.
d = c cosφ , sinφ = tan β (2.11)
where d is the intercept on the q-axis and β is the angle made by the Kf line with the p-
axis.

Fig 2.18: Stress paths for a triaxial test series


We may vary σ1 and σ3 in various ways to obtain any number of stress paths. As an
example it is shown for different cases below. For all cases assume σh=σv

Points Δσv Δσh Remark


A Δσ Δσ
B Δσ 0.5Δσ
C increses 0 Triaxial compression
D Δσ −Δσ
E 0 increases Triaxial extension
F decreases increases

Stress path HL\CH2.doc

ƒ Point A: initial conditions Po=(σh+σv)/2=σv ; qo=(σh-σv)/2=0


final conditions Pf =(σhf+σvf)/2=σv +Δσ; qf =0
ƒ Point B: initial conditions Po=(σh+σv)/2=σv ; qo=(σh-σv)/2=0
final conditions Pf=(σhf+σvf)/2=σv +3/4Δσ; qf=(σhf-σvf)/2=1/4Δσ
Note: stress paths can be developed for either total or effective stresses giving total stress
path or the effective stress path. HL\Examples CH2.doc

30
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Shear Strength of Cohesionless Soils


The shear strength of sands is given by (C = 0),
τ = σ' sinφ =(σ-u) sinφ (2.12)
Factors affecting shear strength HL\CH2.doc
Generally the value of φ (hence shear strength of sand) is influenced by
1. Void ratio or relative density or state of compaction
2. Roughness , shape and angularity of the grains
3. Grain size distribution
Ultimate values of φ may range from 29 o to 35o and peak values from 32 o to 45o for
sands. The values of φ selected for use in practical problems should be related to soil
strains expected. If soil deformation is limited, using the peak value for φ would be
justified. If the deformation is relatively large, ultimate value of φ should be used.

Stress-strain Behavior
The stress-strain behavior of sands is dependant to a large extent on initial density of
packing, as characterized by the density index. This is represented in Fig 2.15 (a) above.
There it can be seen that, the shear stress builds up gradually for loose sand, while for an
initially dense sand, it reaches a peak value and decreases at grater values of shear/axial
strain to an ultimate value compatible to that for an initially loose specimen. The
behavior of medium-dense sand is intermediate to that of loose sand and dense sand. The
hatched portion represents the additional strength due to the phenomenon of interlocking
in the case of dense sands.
Critical Void Ratio
Volume change of a soil sample during testing depends on particle size, shape, grain-size
distribution, principal stress difference, previous stress history and most significantly
relative density or void ratio. At large strains initially loose sand and initially dense sand
attain same void ratio, where further strain will not bring any volume change. Such a void
ratio is called a critical void ratio, ecr (Fig 2.19).

Fig 2.19 Critical void ratio


ƒ ecr depends on σ3 in the triaxial test and σv in the direct shear test. The larger
σ3 and σv, the smaller ecr.
Saturated Cohesionless Fine-Grain Soils
The angle of internal friction of saturated sands and some inorganic silts is only slightly
less than that of the soil in a dry state and of the same relative density (as the drainage is
instantaneous). However the shear strength might be altered significantly by a change in
the pore pressures (equation 2.12). Quite apparently, when the porewater pressure
approaches σ, the shear strength approaches zero. When that happens, we may approach
impending instability or perhaps motion (e.g., slope failures, boiling). Fluctuation in the
water table is a common cause of significant variations in the pore stress and, thereby, in
the shear strength of the soil.
31
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Liquefaction of Fine Sand and Inorganic Silts

If a saturated and/or inorganic silt is totally saturated and under hydrostatic neutral stress
such that it is not subjected to any effective stress, the mass is in a state of liquefaction.
Under such circumstances, the pore water pressure u equals the total normal stress σ,
there by reducing the shear strength to zero, and then the soil is in a quick or boiling or
flowing condition (conditions that result in impending upward movement of soil and
water).
If submerged fine sand undergoes a sudden decrease in the void ratio, an increase in the
porewater pressure, u, may result such that the pore pressure may equal or exceed the
value for σ. For example, pile driving, earthquakes, blasts, or other forms of vibration or
shock may cause a sudden decrease in the volume, thereby increasing the pore pressure u
as a result of a surge in hydrostatic excess pressure. Should the value of u reach sufficient
magnitude, say u > σ, the shear strength of the soil may be totally lost, resulting in what
is known as spontaneous liquefaction. Loose, fine silty sands are most vulnerable to such
effects from shock or dynamic loads or sudden fluctuations in the water table.
Compacting loose sand stratum is frequently a viable option to decrease the possibility of
liquefaction.

Shear Strength of Cohesive Soils HL\CH2.doc

Shear strength behavior of clays is influenced by the fact whether the clay is normally
consolidated or overconsolidated, by the fact whether it is undisturbed or remolded, by
the drainage conditions during testing, consistency of the clay, by certain structural
effects, by the type of test and by the type and rate of strain.

Normally Consolidated Clays

When the sample is extracted from the ground, the overburden pressure is removed and
the pore pressure altered significantly, that is, negative pore pressures are developed. In
order to simulate a somewhat realistic in-situ state of stress, the characteristics of
saturated, normally consolidated clays extracted from a given stratum are commonly
investigated via a CU triaxial test. A confining pressure, σ3, and a deviator stress, Δσ, are
applied for undrained conditions. A confining pressure, say σo, of the in-situ value may
be estimated as the overburden pressure for the depth of the stratum from which the
sample was extracted. If several such tests are run for varying confining cell pressures, a
Mohr envelope may be obtained as indicated in Fig 2.20 (a). If the confining pressure is
less than the in-situ value σo, the Mohr envelope depicts a range of preconsolidation of
the soil; that is, relative to the confining pressures, the soil specimen appears
overconsolidated. The shear strength of the clay specimen tested in this range is higher
than that indicated by a straight line through the origin. The relationship between the
shear strength and the normal stress in this range is designated by line portion ab, which
is slightly curved, but frequently interpreted as a straight line. On the other hand, if the
tests are run under confining cell pressures larger than σo, the envelope to the rupture
circles is approximately a straight line, represented by segment bc. Although the effect of
porewater pressure is present for all ranges of normal stress within the Mohr envelope,

32
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

the pore pressure is larger for the case where the confining pressures are smaller than σo.
Generally, there is an increased drainage of water with increasing confining pressure.
However, the effective stress is more significant in the region where the confining
pressures are larger than σo.

Fig 2.20 τ−σ relationships from CU and drained tests (a) CU test (b) CD test
Figure 2.20 (b) shows the relationship between the shear strength and the effective stress
for a case of CD tests. These tests are appreciably more time consuming than the CU test,
and are therefore not as common during investigations. The specimen is subjected to a
confining pressure, and then the deviator stress is applied at a very slow rate and drainage
is permitted at each end of the sample. A series of such tests run under varying confining
cell pressures provides a strength envelope as shown in Fig 2.20 (b). The strength
envelope for such test is again somewhat curved for the range of confining pressures less
than σo; for confining pressures greater than σo, the Mohr envelope is approximately a
straight line. Briefly, a comparison of the two cases shows the following
1. The corresponding slopes of straight-line segments of the Mohr envelopes (Fig 2.20)
are significantly different.
2. For the drained condition the neutral (pore-water) stress is virtually negligible.
3. The effective friction is significantly larger for the drained case than for the undrained
case.
The relationship between angles φ and φcu may be illustrated by means of Fig 2.20 (c).
The shear strengths for the CD and CU tests are τ and τ1 as shown. The actual strength
may lie somewhere between that for a CD and that for an undrained condition since total
drainage is unlikely; that, the actual strength of the sample may range between values of τ
and τ1. HL\CH2.doc

Overconsolidated Clays HL\0.doc


A soil being evaluated or tested may have been subjected to a great deal of
precompression (i.e. overconsolidation pressure, Pc) induced by loads, which since then
may have disappeared. Fig 2.21 shows the Mohr strength envelope for an
overconsolidated clay. One notes that for σ < Pc, the Mohr envelope deviates from a
straight line, that is, the shear strength is larger than that given by the dotted straight line.
Generally preconsolidation causes or results in a smaller void ratio at failure than would
otherwise exist, even though the specimen tends to expand as a result of extraction from
an in-situ condition. Cohesion and general capillary forces tend to resist the volume
increase, thereby resulting in somewhat greater shear strength, as indicated by the curved
portion of the envelope. Beyond the preconsolidation pressure Pc, the effective normal
stress and shear strength relationship is given by a reasonably straight line.

33
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 2.21 Shear strength vs. effective normal stress for OCC
The shear strength of clays that have fine discontinuities, hairline cracks, and the like,
generally referred to as fissures, may be appreciably different from that of the typical
overconsolidated clay described above. Depending on the magnitude and orientation of
these fissures, test results may be particularly misleading in the overall evaluation. For
example, the results from a direct shear test on a sample where fissures are parallel to the
shearing force may be appreciably smaller than for one where the orientation of fissures
is 90o to the shear force. On the other hand, the triaxial test yields somewhat more reliable
results, improved perhaps by the lateral restraint of the confining pressure.
Sensitivity of Clay
If the strength of undisturbed sample of clay is measured and its strength is again
measured after remolding at the same water content to the same dry density, a reduction
in strength is often observed. This is an important phenomenon which is quantitatively
characterized by ‘sensitivity’, defined as follows:
Unconfined compression strength, undisturbed
Sensitivity S f =
Unconfined compression strength, remolded
A comparison of stress-strain curves for sensitive clay in the undisturbed and remolded
sates is shown in the figure below. Sensitivity classification is given in the table below.
Table: Sensitivity classification of clays
Sensitivity, Sf Classification
1 Insensitive
1-2 Low
2-4 Medium
4-8 Sensitive
8-16 Extra-sensitive
>16 Quick (Sf can be up to 150)
Overconsolidated clays are rarely sensitive, although some quick clays have been found
to be overconsolidated.

3) In-Situ Evaluation of Shear Strength


As mentioned previously, the extraction of a soil sample and the subsequent changes
induced by the extraction process, the handling, and the testing procedures may greatly
alter the characteristics of the specimen and, therefore, the test results. Furthermore, it is
not always feasible and practical to duplicate the in-situ conditions. Frequently, it is both
practical and desirable to test the soil in the in-situ condition.
A number of field tests are used to estimate the shear strength of the soil. Some of
these give results that fit in to the theoretically based expressions used to designate the
shear strength of the soil. Others are empirical in nature and greatly dependent on the
engineer’s judgment and experience.

34
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

The Vane shear Test

Is used for the determination of shear strength of soft clays (clays which may be disturbed
during the extraction and testing process). The test is performed at any given depth by
first augering to the prescribed depth, cleaning the bottom of the borings, and then
carefully pushing the vane instrument (Fig 2.22) in to the stratum to be tested. The torque
is then applied gradually and the peak value noted. The shear strength of the soil can then
be estimated by using the formulae derived below.

Fig 2.22 Vane Instrument


The torque is resisted by T1 and T2.
If both ends of the vane are ‘submerged’ in the soil stratum, and if the maximum shear
stress is Cu for all shear surfaces, then
Resisting moment = cylindrical surface resistance + two circular end face resistance
T = 2πr L (Cu r) + 2[πr2 Cu (2/3r)] =2πr2Cu (L+2/3r)
⇒ Cu = T ⇒ Cu = 3T 3 if L = 4 r (commonly used ratio)
2π r2 (L + 23 r) 28 π r
If one end of the vane is ‘submerged’ in the soil stratum,
Resisting moment = cylindrical surface resistance + one circular end face resistance
T = 2πr L (Cu r) + πr2 Cu (2/3r) =2πr2Cu (L+1/3r)
⇒ Cu = T ⇒ Cu = 3T 3 if L = 4 r
2 π r 2 (L + 13 r) 26 π r
This test is also made in laboratories using small vane instrument.

Standard Penetration Test (SPT)

The number of blows required to drive the split-spoon sampler is still another means used
to estimate the soil’s resistance to shear (shear strength of the soil). Here a split-spoon
sampler (Fig 2.23 (a)) is lowered to the bottom of the bore hole by attaching it to the drill
rod and then driven by forcing it in to the soil by blows from a hammer (64Kg) falling
from a height of 76cm. The sampler is initially driven 15cm below the bottom of the bore
hole to exclude the disturbed soil while boring. It is then further driven 30cm in two
stages (each 15cm). The number of blows required to penetrate the last 30cm is termed as
the standard penetration value, N. Corrections have to be made for overburden pressure,
dilatancy etc. Then corrected N value is used for correlation. HL\6.doc

35
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 2.23 Standard penetration test (SPT)


The test is especially suited for cohesionless soils as a correlation has been established
between the SPT value and the angle of internal friction of the soil.
Terzaghi and Peck also gave the following correlation between SPT value, φ and Dr.
Table: Correlation between N, φ, and Dr for Sands
Condition N φ (degree ) Dr(%)
Very loose 0-4 <20 0-15
Loose 2-4 28-30 15-35
Medium 10-30 30-36 35-65
Dense 30-50 36-42 65-85
Very dense >50 >42 >85
Table: Correlation between N and qu for Clays
Consistency N qu(KPa)
Very soft 0-2 <25
Soft 2-4 25-50
Medium 4-8 50-100
Stiff 8-15 100-200
Very stiff 15-30 200-400
Hard >30 >400
The correlation for clays is unreliable. Hence, vane shear test is recommended for more
reliable information.
Usually SPT is conducted at every 2m depth or at the change of stratum. If refusal is
noticed at any stage, it should be recorded.

Other Penetrometers

Penetrometers are some times used to test the shear strength of the soil at the surface-
lateral or vertical. Their use is primarily applicable to fine-grained soil; coarse and
gravelly strata tend to give erroneous results. The procedure for using the penetrometer
consist of first cleaning the surface of any loose material, pushing the pentrometer in to
the stratum to the calibration mark on the head of the penetrometer, and recording the
maximum reading on the penetrometer scale. This reading represents the pressure in force
per unit area necessary to push the penetrometer to the designated mark. The reliability of
the results must be interpreted with a view to the condition present at the time of testing
(like water content).

36
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

CHAPTER THREE

LATERAL EARTH PRESSURE AND RETAINING WALLS

1. Introduction
Soil is neither a solid nor a liquid, but it exhibits some of the characteristics of both. One
of the characteristics similar to that of a liquid is that of its tendency to exert a lateral
pressure (earth pressure) against any object in contact. Thus structures which retain or
support soil like retaining walls, abutments, sheet pile walls, basement walls and under
ground conduits need estimation of the lateral pressure for their design.
Various types of retaining walls are shown in Fig 3.1 and are widely employed in civil
engineering works ranging from their use in road and rail construction to support cuts and
fills where space is limited to prevent the formation of appropriate side slopes, to the
construction of marine structures such as docks, harbours and jetties.

Fig 3.1 Types of Earth retaining structures


The design of a retaining wall requires the determination of the pressures which act on it.
These are influenced by:
• The physical property of the soil (γ, φ and C)
• The time dependent nature of the soil strength (for clays⇒drainage)
• The interaction between the soil and the retaining structure at the interface (δ)
• The type of wall and the degree and mode of wall movement
• The imposed loading (e.g. height of back fill, surcharge loads)

2. Lateral Earth Pressures


Let us consider a retaining wall which holds back a mass of soil. The soil exerts a push
against the wall by virtue of its tendency to slip laterally and seek its natural slope or
angle of repose, thus making the wall to move slightly away from the backfilled soil
mass. This kind of pressure is known as the active earth pressure of the soil. The soil
being the actuating element is considered to be active and hence the name active earth
pressure. Next, let us imagine that in some manner the retaining wall is caused to move
toward the soil. In such a case the retaining wall or is the actuating element and the soil
37
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

provides the resistance for maintaining stability. The pressure or resistance which soil
develops in response to movement of the structure toward it is called the passive earth
pressure which may be very much greater than the active earth pressure. The surface over
which the sheared-off soil wedge tends to slide is referred to as the surface of sliding or
rupture.
Fig 3.2 depicts the relationship between the earth pressure and the wall movement. Po
represents the magnitude of pressure when no movement of the retaining wall takes place;
it is commonly referred to as earth pressure at rest. As the wall moves toward the backfill,
the pressure increases, reaching a maximum value of Pp (passive earth pressure) at point
C. On the other hand, if the wall moves away from the backfill, the force decreases,
reaching a minimum value of Pa (active earth pressure) at point B.

Fig 3.2 Relationship between earth pressure and wall movement


Very little movement (about 0.5% horizontal strain) is required to mobilize the active
pressure; however, relatively much larger movement (about 2% of horizontal strain for
dense sands as high as 15% for loose sands) may be required to mobilize full passive
resistance (Lambe and Whitman, 1969).About 50% of the passive resistance may be
mobilized at a moment comparable to that required for the active case. HL\30-1.doc

The relative magnitude of the active and passive earth pressures may perhaps be
better illustrated with the aid of Fig 3.3. For the sake of simplicity, several specific
assumptions are made. Deviations from these will be discussed in the next sections.
1. Frictional forces between backfill and retaining wall are assumed negligible
2. The wall is vertical, and the surface of the backfill is horizontal
3. The backfill is homogenous, granular material
4. The failure surface is assumed to be a plane

Fig 3.3 Relationship between earth pressure and wall movement for cohesionless soil
The magnitude of the active and passive forces Pa and Pp could be derived from the basic
condition of static equilibrium as follows.

38
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Case of Active Pressure: From Fig 3.3(a),


1 1
P a = Wtan (β - φ ) but W = γ H ( H cot β ) = γ H 2 cot β
2 2
1
⇒ P a = γ H 2 cot β tan (β - φ )
2

The maximum value of Pa may be obtained when cPa /cβ=0. Thus,


∂Pa 1 2 ⎡
= γH ⎢cot β sec 2 (β − φ ) + tan (β − φ )⎛⎜ − cosec 2β ⎞⎟ ⎤⎥ = 0
∂β 2 ⎣ ⎝ ⎠⎦
φ
⇒ β = β cr = 45 + substituti ng this back in the equation for Pa, we have
2
1 ⎛ φ⎞
Pa = γH 2 tan 2 ⎜ 45 − ⎟ (3.1)
2 ⎝ 2⎠
1 2 φ
Pa = γH K a where : K a = tan 2 ⎛⎜ 45 − ⎞⎟
2 ⎝ 2⎠
Case of Passive Pressure: From Fig 3.3(b),
1 1
P p = Wtan (β + φ ) but W = γ H ( H cot β ) = γ H 2 cot β
2 2
1
⇒ P p = γ H 2 cot β tan (β + φ )
2

The maximum value of Pp may be obtained when cPp /cβ=0. Thus,


∂ Pp 1
= γH 2 ⎡⎢cot β sec 2 (φ + β ) + tan (φ + β )⎛⎜ − cos ec 2 β ⎞⎟ ⎤⎥ = 0
∂β 2 ⎣ ⎝ ⎠⎦
φ
⇒ β = β cr = 45 − substituti ng this back in the equation for P p , we have :
2
1 φ
Pa = γH 2 tan 2 ⎛⎜ 45 + ⎞⎟ (3.2)
2 ⎝ 2⎠
1 ⎛ φ⎞
Pp = γH 2 K p where : K p = tan 2 ⎜ 45 + ⎟
2 ⎝ 2⎠
Ka and Kp are generally referred to as coefficients for active and passive pressure,
respectively. They are constants for any given soil where φ = constant. It is clear that the
value of Kp is significantly larger than Ka.

3. Earth Pressure at Rest


Earth pressure at rest may be obtained theoretically from the theory of elasticity applied
to an element of soil, remembering that the lateral strain of the element is zero. Referring
to Fig 3.4 (a), the principal stresses acting on an element of soil situated at a depth z from
the surface in semi-infinite, elastic, homogenous and isotropic soil mass are σv, σh and σh
as shown; σv and σh denoting the stresses in the vertical and horizontal directions
respectively.

Fig 3.4: Stress conditions relating to earth pressure at rest


39
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

The soil deforms vertically under its self weight but is prevented from deforming laterally
because of an infinite extent in all lateral directions. Let E and ν be the modulus of
elasticity and Poisson’s ratio of the soil respectively.
σ ⎛σ σ ⎞
Lateral strain, ε h = h − v⎜ v + h ⎟ = 0
E ⎝ E E ⎠
σh v
⇒ = = Ko where K o is known as the coefficient of earth pressure at rest.
σ v 1− v
Ko is the ratio of the intensity of the lateral earth pressure at rest to the vertical stress at a
specified depth.
As the vertical pressure at any depth is σv=γz, then σh=Ko σv = Ko γz. The distribution of
the earth pressure at rest with depth is linear (hydrostatic nature) for constant properties
such as E,ν, and γ, as shown in Fig 3.4 (b)
If a structure such as a retaining wall of height H is interposed from the surface and
imagined to be held without yield, the total thrust on the wall per unit length Po is given
H H
by 1 . This is considered to act at H/3
Po = ∫σ dz = ∫ .γ z . dz = .γ . H 2
h K o K o
0 0
2
above the base of the wall.
Choosing an appropriate value for the Poisson’s ratio,ν, is by no means easy; this is the
limitation in arriving at Ko from the equation Ko=[ν/(1-ν)]. However, various researchers
proposed empirical relationships for Ko, some of which are given below:
Ko = (1-sinφ') [Jaky, 1944]
Ko =0.9 (1-sinφ') [Fraser, 1957]
Ko = 0.19 + 0.233logIp [Kenney, 1959]
Ko = [1+ (2/3) sinφ'] [(1- sinφ)/(1+sinφ)] [Kezdi,1962]
Ko = (0.95-sinφ') [Booker and Ireland, 1965]
φ' in these equations represents the effective angle of friction of the soil and Ip the
plasticity index. HL\30-1.doc 30-2
4. Earth Pressure Theories
The magnitude of earth pressure is evaluated by the application of one or the other of the
so-called lateral earth pressure theories or simply earth pressure theories. Several
investigators have proposed many theories of earth pressure after a lot of experimental
and theoretical work. Coulomb’s and Rankine’s are perhaps the two best-known theories
and are frequently referred to as the classical earth pressure theories.
Rankine’s Theory
Rankine (1857) developed his theory of lateral earth pressure when the backfill consists
of dry, cohesionless soil. He has made the following important assumptions.
1) The soil mass is semi-infinite, homogenous, dry and cohesionless
2) The ground surface is plane which may be horizontal or inclined
3) The face of the wall in contact with the backfill is vertical and smooth. HL\30-1.doc 30-3
a) Cohesionless Backfill and Level Surface
The basic concept behind Rankine’s theory can be depicted via Mohr’s circle. Consider
the element shown in Fig 3.5 (a) subjected to the geostatic stresses shown. The value of
σ1 could be approximated as σ1 = γ h. the horizontal principal stress (σ3 or σh) or the
lateral earth pressure at rest in this case is σ3=σh=Koγh. If the wall were to move to the
left, thereby creating the case of active stress, the value of σ1 would become the major
40
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

principal stress. The corresponding Mohr circle for this case is depicted by circle I in Fig
3.5 (b). On the other hand, if the wall were to push against the backfill, a case of passive
pressure would be developed. The vertical stress then becomes the minor principal stress,
and the lateral stress would thus become the major principal stress. The Mohr circle for
this condition is depicted by circle II in Fig 3.5 (b).

Fig 3.5 The orientation of slip planes in granular soil mass with a level surface
Considering circle I (active case), we have
(σ − σ ) / 2 (σ − σ )
sin φ = 1 3
= 1 3

(σ 1 + σ 3 ) / 2 (σ 1 + σ 3 )
Rearranging terms and solving for σ3,
1 − sin φ
σ 3 = σ 1 = σ 1 t an 2 (45 - φ /2 ) = γ ht an 2
(45 - φ /2 ) (3.3)
1 + sin φ
Again considering circle II (passive case), we have
(σ ' 3 − σ 1 )/ 2 (σ '
− σ )
sin φ = = 3 1

(σ '
3 + σ 1 ) / 2 (σ '
3 + σ 1 )

Rearranging terms and solving for σ'3,


1 + sin φ
σ '3 = σ = σ 1 t an 2
(45 + φ /2 )= γ ht an 2
(45 + φ /2 ) (3.4)
1 − sin φ
1

We note that tan2 (45-φ /2) and tan2 (45+φ /2) in Eqs. 3.3 and 3.4 are constants for
constant values of φ. Hence, the corresponding pressures against the retaining wall vary
linearly with depth, as indicated by Fig 3.6. The corresponding resultant pressures, active
and passive, can be calculated for a unit length of retaining wall as
1 1
Pa = γ H 2 tan 2
(45 − φ /2 γ H 2K )= a ( 3 .5 )
2 2
1
Pp = γ H 2 tan 2
(45 + φ /2 ) = 1 γ H 2 K p ( 3 .6 )
2 2
which correspond to equations 3.1 and 3.2 respectively.

b) Cohesionless Backfill and Inclined Surface


Let’s consider a cohesionless mass with a sloping surface behind a smooth vertical
retaining wall (δ =0). Considering an element on Fig 3.7(a), the lateral stresses acting on
the vertical faces of the element (i.e. the faces parallel to the wall) are parallel to the
inclined surface. Thus, any such planes experience not only normal but also shear stresses
and are no longer principal planes as was the case for horizontal surfaces.
The weight of the soil column above the element is γz; length of side is 1/cosβ. Hence
σv=γz cosβ. The corresponding resultant pressure on the wall could be determined with
the aid of Mohr’s circle.

41
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 3.7: Lateral pressure and slip planes in granular sloped backfill (active case)
Fig 3.7(c) symbolizes an active state of stress. The magnitude of the vertical stress is
depicted by distance OC; the lateral stress, acting parallel to the sloped surface, is
represented by the distance OA. Hence, from Fig 3.7 (c) we have
DE
sinφ = DE = OD sin φ = AD
OD
(σ + σ 3 ) (σ 1 − σ 3 )
OD = 1 AD =
2 2
OB = ODcos β BD = ODsin β
AB = (AD) 2 − ( BD ) 2 = (ODsin φ ) 2 − (OD sin β ) 2 = OD sin 2 φ − sin 2 β = BC
[
σ h = OA = OB − AB = ODcos β − OD sin 2 φ − sin 2 β = OD cos β − sin 2 φ − sin 2 β ]
σ v = OC = OB + BC = ODcos β + OD sin φ − sin 2 2
β = OD [cos β + sin φ − sin
2 2
β]
σh [
OD cos β − sin φ − sin β
2 2
] [
cos β − sin φ − sin β 2 2
]
⇒ Ka = =
[
σ v OD cos β + sin 2 φ − sin 2 β
=
] [
cos β + sin 2 φ − sin 2 β ]
σh [
cos β − cos 2 β − cos 2 φ ]
Ka =
σv
=
[
cos β + cos 2 β − cos 2 φ ] ( 3 .7 )

Thus σ h = σ v K a = γ z cos β K a
For a given slopped surface and uniform soil properties Ka becomes a constant. Thus, the
intensity of load, or stress, varies linearly with depth. Hence, as before the total resultant
active force may be given by
1
Pa = γH 2 cosβ K a where Ka is given in equation 3.7. The direction of the
2
resultant is parallel to the sloped surface.
Now suppose there is inclined backfill and the wall face is also
inclined. To calculate the total lateral force acting on the wall,
combine the lateral force acting on an imaginary vertical surface
passing through the heal of the wall with the weight of the wedge
of soil between the imaginary surface and back of wall to get the
resultant.

Similarly for the passive state


σh [
cos β + cos 2 β − cos 2 φ ]
Kp =
σv
=
[
cos β − cos 2 β − cos 2 φ]
(3 .8 )

1
Pp = γH 2 cosβ K p where Kp is given in equation 3.8. The direction of the
2
resultant is parallel to the sloped surface.
For the case of level surface, equations 3.7 and 3.8 reduce to equations 3.5 and 3.6
respectively.

42
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Coulomb’s Theory
In 1776 Coulomb introduced an expression for determining the thrust on retaining walls.
Basic assumptions were
1. The soil is isotropic and homogenous and posses both cohesion and internal friction
2. The failure surface is plane. The back fill surface is planner.
3. Frictional forces are distributed uniformly along the plane failure surface
4. The failure wedge is a rigid body undergoing translation
5. There is a wall friction which develops as the sliding wedge moves along the back of
the wall.
For the active case derivation was made considering Fig 3.8 as follows.

Fig 3.8: Coulomb’s theory


From Fig 3.8 (a) :
H H
AB = H/sin α ; BD = ABsin[180 − (α + θ)] = sin[180 − (α + θ)] = sin(α + θ)
sin α sin α
AE AB H sin( α + β )
using sin law : = ⇒ AE =
sin( α + β ) sin(θ − β ) sin α sin(θ − β )
1 1 H sin(α + β) H 1 H2 sin(α + β)
Area (ΔABE) = AE * BD = * sin(α + θ) = sin(α + θ)
2 2 sinα sin(θ − β) sinα 2 sin α 2 sin(θ − β)
1γH ⎡ 2 sin(α + β) ⎤
Weight W = γ * area (ΔABE) = ⎢sin(α + θ) ⎥
2
2 sin α ⎢⎣ sin(θ − β) ⎥⎦
From Fig 3.8 (b) :
P W
a =
sin(θ − φ ) sin[180 - α - θ + φ + δ )
Wsin( θ − φ ) ⎡1 γ H2 ⎡ sin(α + β) ⎤ ⎤⎥ ⎡ sin(θ − φ ) ⎤
⇒ Pa = =⎢ ⎢ sin(α + θ) ⎥ ⎢ ⎥ (3 . 9 )
sin[180 - α - θ + φ + δ ) ⎢ 2 2 ⎥
sin(θ − β) ⎦ ⎦ ⎣⎢ sin[180 - α - θ + φ + δ ) ⎦⎥
⎣ sin α ⎣
To determine the orientation of the failure plane (θ cr) that produces a maximum Pa, set
∂Pa
= 0 . After determining θ cr and substituting it in equation 3.9 we have,
∂θ
γ H2 sin 2 (α + φ ) γ H2
Pa = 2
= Ka (3 .10 )
2 ⎡ sin( φ + δ )sin( φ - β ) ⎤ 2
(sin α)sin(α − δ ) ⎢1 +
2

⎣ sin( α − δ )sin( α + β ) ⎦
sin 2 (α + φ )
where K a = 2
⎡ sin( φ + δ )sin( φ - β ) ⎤
(sin α)sin(α − δ ) ⎢1 +
2

⎣ sin( α − δ )sin( α + β ) ⎦

The corresponding passive trust can similarly be derived and is given as follows
γ H2 sin 2 (α − φ ) γ H2
Pp = 2
= Kp (3 .11)
2 ⎡ sin( φ + δ )sin( φ + β ) ⎤ 2
(sin α)sin(α + δ ) ⎢1 −
2

⎣ sin( α + δ )sin( α + β ) ⎦

43
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
sin (α − φ )
2
where K p = 2
⎡ sin( φ + δ )sin( φ + β ) ⎤
(sin 2 α)sin(α + δ ) ⎢1 − ⎥
⎣ sin( α + δ )sin( α + β ) ⎦

For a vertical smooth back face and horizontal back fill surface (that is, α=90, β=0 and
δ=0) equations 3.10 and 3.11 reduce to equations derived based on the Rankine theory
(3.7 &3.8).
Coulomb arbitrarily placed the resultant thrust Pa (Pp) at the third point from the bottom.
Correspondingly, he assumed the pressure distribution to vary linearly with depth.
Although this assumption appears to give results acceptable for very rigid walls and
granular backfill, it is not valid for relatively flexible bulk heads, for cohesive backfills,
or where the retaining wall rotates about points not close to the bottom.
5. Lateral Earth Pressures In cohesive (C-φ) soils
The Mohr’s circle may be used to determine the lateral thrust on retaining walls
supporting cohesive soils.
Let’s assume a vertical and smooth retaining wall and a C-φ soil backfill with a horizontal
surface. At a given depth z on the element shown in Fig 3.9 (a), the vertical stress σ1=γ z.
The lateral stress is σ3. Both of these are principal stresses. The corresponding Mohr’s
circle for this case is shown in Fig 3.9 (b).

Fig 3.9: Cohesive (c-φ) backfill


For the active state, from Fig 3.9 (b) we note that
(σ1 − σ 3 ) /2 (σ1 − σ 3 )
Sin φ = =
(σ1 + σ 3 )/2 + c cot φ (σ1 + σ 3 ) + 2 c cot φ
Rearrangin g terms,
⎛ 1 − sin φ ⎞ ⎛ cos φ ⎞
σ 3 = σ 1 ⎜⎜ ⎟⎟ − 2c ⎜⎜ ⎟⎟
⎝ 1 + sin φ ⎠ ⎝ 1 + sin φ ⎠
cos φ 1 - sin 2 φ (1 - sin φ ) (1 + sin φ ) 1 - sin φ
But = = =
1 + sin φ 1 + sin φ (1 + sin φ ) (1 + sin φ ) 1 + sin φ
⎛ 1 − sin φ ⎞ 1 - sin φ
⇒ σ 3 = σ 1 ⎜⎜ ⎟⎟ − 2c
⎝ 1 + sin φ ⎠ 1 + sin φ
1 - sin φ ⎛ φ⎞
But σ = γ h and = tan 2 ⎜ 45 - ⎟. Hence,
1 1 + sin φ ⎝ 2⎠
⎛ φ⎞ ⎛ φ⎞ ⎛ φ⎞
⇒ σ 3 = γh tan 2 ⎜ 45 − ⎟ − 2ctan ⎜ 45 − ⎟ = γh K − 2c K where K = tan 2 ⎜ 45 - ⎟
⎝ 2⎠ ⎝ 2⎠ a a a ⎝ 2⎠

Fig 3.10: Pressure distribution in c-φ soil


44
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

The maximum horizontal stress (or pressure) for the active case occurs when h=H. The
pressure distribution is shown in Fig 3.10(b). The corresponding resultant Pa is
1
P = γH 2 K − 2cH K (3 .12 )
a 2 a a
For the passive state, from Fig 3.9 (c) we note that
(σ 3 − σ 1 ) /2 (σ 3 − σ 1 )
Sin φ = =
(σ1 + σ 3 )/2 + c cot φ (σ1 + σ 3 ) + 2 c cot φ
Rearrangin g terms,
⎛ 1 + sin φ ⎞ ⎛ cos φ ⎞
σ 3 = σ 1 ⎜⎜ ⎟⎟ + 2c ⎜⎜ 1 − sin φ ⎟⎟
⎝ 1 − sin φ ⎠ ⎝ ⎠
cos φ 1 - sin 2 φ (1 - sin φ ) (1 + sin φ ) 1 + sin φ
But = = =
1 − sin φ 1 − sin φ (1 − sin φ ) (1 − sin φ ) 1 − sin φ
⎛ 1 + sin φ ⎞ 1 + sin φ
⇒ σ 3 = σ 1 ⎜⎜ ⎟⎟ + 2c
⎝ 1 − sin φ ⎠ 1 − sin φ
1 + sin φ ⎛ φ⎞
But σ = γ h and = tan 2 ⎜ 45 + ⎟. Hence,
1 1 − sin φ ⎝ 2⎠
⎛ φ⎞ ⎛ φ⎞ ⎛ φ⎞
⇒ σ 3 = γh tan 2 ⎜ 45 + ⎟ + 2ctan ⎜ 45 + ⎟ = γh K + 2c K where K = tan 2 ⎜ 45 + ⎟
⎝ 2⎠ ⎝ 2⎠ p p p ⎝ 2⎠

The maximum horizontal stress (or pressure) for the passive case occurs when h=H. The
pressure distribution is shown in Fig 3.10(c). The corresponding resultant Pp is
1
P = γH 2 K + 2cH K (3 .13 )
p 2 p p
6. Unsupported Cuts in Cohesive (C-φ ) Soils
Unsupported excavations would theoretically be possible in c-φ soils if the lateral
pressure (σ3 for the active case) would not exceed the strength of the soil. From above we
have
σ = γh K − 2c K (See eqn. 3.12)
3 a a
At ground surface, h=0. Thus, σ 3 = − 2c K (tension)
a
This implies the formation of a crack as depicted in Fig 3.11(a). The corresponding
pressure distribution based on eqn. 3.12 is shown in Fig 3.11(b).

Fig 3.11:Tension crack in c-φ soil and corresponding pressure distribution (active state)
The theoretical depth of crack ht can be determined by recognizing that, at the bottom of
the crack, σ3=0. Thus from equation 3.12,
0 = γh K − 2c K
a a
2c
⇒h = (3 .14 )
t γ K
a
The theoretical maximum depth Hc of unsupported excavation may be calculated as the
point where the tension forces equal the cohesive strength. Hence from Fig 3.11(b),
Hc=2ht. This could also be obtained from eqn. 3.12 if σ = 2c K , when h=Hc:
3 a
4c
2c K = γh K − 2c K ⇒H = = 2h (3.15)
a a a c γ K t
a
45
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Though eqn 3.15 provides a theoretical depth to which an excavation may be made with
out lateral support, it should be used cautiously. HL\30-1.doc30-4

7. Effect of Uniform Surcharge


The extra loading carried by a retaining structure is known as surcharge. If the uniform
surcharge is of intensity q per unit area, then the vertical stress at every elevation in the
backfill is considered to increase by q. As such, the lateral pressure has to increase by
Ka. q. Thus, at any depth z, σ h= γh Ka + Ka. q.
Fig 3.12 (b) and (c) show two different ways in which the pressure distribution may be
shown. In Fig 3.12 (c), the uniform surcharge is also considered to have been converted
into an equivalent height He, of backfill, which is easily established, as shown.

Fig 3.12: Effect of uniform surcharge on lateral pressure


8. Effect of Submergence
When the backfill is fully saturated /submerged, the lateral pressure will be due to two
components: 1) lateral earth pressure due to submerged unit weight of the backfill soil
2) lateral pressure due to pore water
This is shown in Fig 3.13(a)

Fig 3.13: Effect of submergence on lateral earth pressure


At any depth z below the surface, the lateral pressure, σh, is given by:
σ h= γsub z Ka + γw z.
The pressure at the base is obtained by substituting H for z, i.e. (σ h= γsub H Ka + γw H)
In case water stands to the full height of the retaining wall on the other side of the
submerged backfill, as shown in Fig 3.13(b), the net lateral pressure from the submerged
back fill will be only from the submerged unit weight of the backfill soil, as the water
acting on both sides will get cancelled.

46
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

9. Effect of Soil Layering


When the backfill is layered as shown in Fig 3.14 the lateral pressure will be determined
as indicated on the figure.

Fig 3.14: Effect of soil layering on lateral earth pressure


10. Graphical Method of Determination of Lateral Earth Pressure
Coulomb’s theory is very adaptable to graphical solution and the effects of wall friction
and batter are automatically allowed for. Poncelet (1840), Culmann (1866), Rebahann
(1871) are the notable figures who contributed to further development of Coulomb’s
theory.
10.1. Rebhann’s Condition and Graphical Method
Rebhann is credited with having presented the criterion for the direct location of the
failure plane assumed in the Coulomb’s theory. His presentation is somewhat as follows:
Fig 3.15 (a) represents a retaining wall retaining a cohesionless backfill inclined at β to
the horizontal. Let BC be the failure plane, the position of which is to be determined.

Fig 3.15: Rebhann’s condition for Coulomb’s wedge theory- Location of failure plane for
the active case
Let BD be a line inclined at φ to the horizontal through B, the heel of the wall, D being
the intersection of this φ-line with the surface of the backfill. The value of Pa depends
upon the angle θ relating to the location of the failure plane. Pa will be zero when θ=φ (as
no wall required to retain the soil mass at angle φ), and increases with an increase in θ up
to a limit, beyond which it decreases and reaches zero again when θ=180-α(here the
failure wedge has no mass). Thus, the failure plane will lie between the φ-line and the
back of the wall.
Let AE be drawn at an angle (φ+δ) to the wall face AB to meet the φ-line in E. Let CG be
drawn parallel to AE to meet the φ-line in G. Let the distances be denoted as follows: AE
= a, BG = c, CG = x, BD = b, BE = d.
Now it is required to determine the criterion for which Pa is the maximum, which is supposed to give the correct
location of the failure surface.
Weight of the soil in the sliding wedge is

47
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

W= γ ∗(area Δ ABC)*1=γ∗(area Δ ABD- area Δ BCD)*1= ½[γb(a-x)sinψ]


Δ BCG is similar to triangle of forces (AA similarity). Hence, W/c =Pa/x ⇒Pa=W x /c
γ bx
∴ Pa = .(a − x ) sin ψ
2c
If DG = k (CG ) = kx, then c = b - kx
γ bx
⇒ Pa = .(a − x)sin ψ
2(b − kx)
The only value which var ies with the orientatio n of the failure plane is x. Thus for the value
∂ Pa
of Pa to be a maximum, =0
∂x
∂P
⇒ a = (b − kx )(a − 2x ) + kx (a − x ) = 0 ⇒ b (a − x ) = x (b − kx )
∂x
⇒ b (a − x ) = cx
1
Multiplyin g both sides by sin ψ we have,
2
1 1 1 1 1
b (a − x )sin ψ = cx sin ψ ⇒ basin ψ − bxsin ψ = cx sin ψ
2 2 2 2 2
⇒ area( Δ ABD) - area( Δ BCD) = area( Δ BCG)
Thus we see that, area( Δ A BC) = area( Δ BCG)
This equation signifies that for BC to be the failure plane the requirement is that the area
of the failure wedge ABC be equal to the area of the triangle BCG (Rebhann’s condition).
The triangles ABC and BCG have a common base BC, so altitudes on BC should be
equal. Thus, AJ [sin (<AJB)] = CG [sin (<BCG)]. But sin<AJB = sin<BCG as CG is
parallel to AJ. This leads to CG=AJ=x; and JE= a- x.
Triangles DAE and DCG are similar (AA similarity). Hence b − d . x = a (i)
b−c
Also triangles BCG and BJE are similar and consequently d
.x = a - x (ii)
c
From (i) and (ii) we get,
⎡b − d d⎤
x⎢ − ⎥ =x ⇒ c 2 = bd or c= bd ..... ( Poncelet Rule )
⎣b − c c⎦
Thus if c is known the position of G and hence that of the most dangerous rupture plane
(surface BC) can be determined and the weight of the sliding wedge, W, and the active
thrust, Pa, can be calculated.
The value of x can now be determined as c x = b(a − x) ab Summary
⇒ x=
b+c ψ=α-δ
Substituting c = b (a − x) in the equation of Pa, one gets c = bd
x
1
x=bd/ (b+c)
Pa = γx 2 sin ψ (3.16)
2
This is a semi-analytical procedure for the computation of the active thrust by Coulomb’s
wedge theory.
However, elegant graphical methods have been devised and preferred to the analytical
approach, in view of their versatility, coupled with simplicity.
A graphical method is given by Poncelet and sometimes known as the Rebhann’s
graphical method, since it is based on Rebhann’s condition. The steps involved are as
follows with reference to Fig 3.16.

48
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 3.16: Poncelet graphical construction for active thrust


1. Let AB represent the back face of the wall and AD the backfill surface.
2. Draw BD at inclination of φ ⇒ BD = φ -line.
3. Draw BK inclined at ψ (=α−δ) with BD, which is the ψ-line
4. Through A, draw AE Parallel to the ψ-line to meet BD at E.
5. Describe a semi-circle with BD as a diameter
6. Erect a perpendicular line to BD at E to meet the semi-circle in F
7. With B as center and BF as radius draw an arc to meet BD at G
8. Through G draw a line parallel to ψ-line or AE to meet AD at C
9. With G as a center and GC as radius draw an arc to cut BD at L
10. Join CL with a straight line and also draw a perpendicular line CM from C on to LG.
Then BC is the rapture surface.
The criterion may be checked as follows: Since Δ ABC=ΔBCG, and BC is their common
base, their altitudes on BC must be equal; or AN sin (<ANB)=NG sin(<GNC) that is to
say AN=NG, since (<ANB)= (<GNC). Thus, if AN and NG are measured and found to
be equal the construction is correct.
Let CM = n, CG=LG= x ⇒ n = x sinψ
1 1
Then Pa = γx 2sin ψ = γ x n HL\30.doc 4&5
2 2
10.2. Culmann’s Graphical Method
Karl Culmann (1866) gave his own graphical method to evaluate the earth pressure from
Coulomb’s theory. Culmann’s method permits one to determine graphically the
magnitude of the earth pressure and to locate the most dangerous rupture surface
according to Coulomb’s wedge theory. This method has more general application than
Poncelet’s and is, in fact, a simplified version of the more general trial wedge method. It
may be conveniently used for ground surface of any shape, for different types of
surcharge loads, and for layered backfill with different unit weights for different layers.
a) Active Earth Resistance -Cohesionless soil
The steps in the construction may be set as follows (Refer Fig 3.17):

Fig 3.17: Culmann’s


graphical method for
active thrust
49
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

1. With a suitable scale draw the ground line and the retaining wall. This should include
height and slope of the retaining wall, surface configuration of the backfill, location
and magnitude of concentrated (line) surcharge loads, uniformly distributed
surcharge, and so on.
2. From point B draw line BD making an angle φ with the horizontal (φ-line)
3. Draw line BK at angle ψ (=α−δ) with line (φ-line)
4. Draw rays BC1, BC2, BC3, and so on, that is , assumed failure surfaces
5. Determine the weight of each wedge, accounting for variations in densities if the
backfill is layered system, for variable moisture content, and so on.
6. To a convenient scale, plot these weights along line BD (φ-line). For example, the
distance from B to W1 along line BD equals W1; similarly, the distance from W1 to
W2 along line BD equals W2, and so on.
7. From each of the points located on line BD, draw lines parallel to line BK to intersect
the corresponding assumed failure surfaces; that is, the line from W1 will intersect
line BC1 at F1; the one from W2 will intersect line BC2 at F2, and so on.
8. Connect these points of intersection (F1, F2, F3…) with a smooth line, Culmann’s
curve.
9. Parallel to BD, draw a tangent line to Culmann’s curve. In Fig 3.17 point F represents
such a tangent point. More than one tangent is possible if the Culmann line is
irregular.
10. From the point of tangency, draw line FW parallel to line BK. The magnitude of FW,
based on the used scale, represents the active pressure Pa. If several tangents to the
curve are possible, the largest of such values becomes the value of Pa. The failure
surface passes through F and B, as shown in the figure.
b) Passive Earth Resistance -Cohesionless soil
Fig 3.18 illustrates the procedure for determining the passive resistance via Culmann’s
method. The approach is similar to that for the active pressure, with some notable
differences: (1) line BD makes an angle of φ below rather than above the horizontal; (2)
the reference line makes an angle of ψ with line BD, with ψ measured as indicated. For
the assumed sliding wedges, the weights W1, W2, and so on, are plotted along line BD.
From these points, lines are drawn parallel to the ψ-line to intersect the corresponding
rays. The Culmann line represents a smooth curve connecting such points of intersection.
A tangent to the Culmann curve parallel to line BD is drawn, with the resultant earth
pressure being the scaled value of line FW.

Fig 3.18: Culmann’s graphical method for passive earth pressure (cohesionless soil)

50
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

c) Active Earth Resistance -Cohesive soil


The lateral earth pressure of cohesive soil may be obtained from Coulomb’s wedge
theory; however, one should consider the tension zone near the surface of cohesive
backfill and consequent loss of contact and loss of adhesion and friction at the back of the
wall and along the plane of rupture, so as to avoid getting erroneous results. The trial
wedge method may be applied to this case as illustrated in Fig 3.19. The following five
forces act on a trial wedge:
1. Weight of the wedge including the tension zone, W.
2. Cohesion along the wall face or adhesion between the wall and the fill, Ca.
3. Cohesion along the rupture plane, C.
4. Reaction on the plane of failure, R, acting at φ to the normal to the plane of failure.
5. Active thrust, Pa, acting at δ to the normal to the face of the wall.
The total adhesion force, Ca, is given by Ca=ca.BF , where ca is the unit adhesion
between the wall and the fill, which cannot be greater than the unit cohesion, c, of the
soil. ca may be obtained from tests; however, in the absence of data, ca may be taken as
equal to c for soils with c up to 50KPa, ca may be limited to 50Kpa for soils with greater
than this value (Smith, 1974)

Fig 3.19: Active earth pressure of


cohesive soil- Culmann’s method
d) Passive Earth Resistance -Cohesive soil
The procedure adopted to determine the active earth pressure of cohesive soil may also be
used to determine the passive earth resistance of cohesive soil. The only differences are
that the signs of friction angles, φ and δ, will be reversed and the directions of Ca and C
also get reversed.
Comparison of Coulomb and Rankine Theory
The following are the important points of comparison:
1. Coulomb considers a retaining wall and the backfill as a system; he takes into account
the friction between the wall and the backfill, while Rankine does not.
2.The backfill surface may be plane or curved in Coulomb’s theory, but Rankine’s
allows only for a plane surface.
3. In Coulomb’s theory, the total earth thrust is first obtained and its position and
direction of the earth pressure are assumed to be known; linear variation of pressure
with depth is tactically assumed and the direction is automatically obtained from the
concept of wall friction, In Rankine’s theory, plastic equilibrium inside a semi-infinite
soil mass is considered, pressures evaluated, a retaining wall is imagined to be
interposed later, and the location and magnitude of the total earth thrust are
established mathematically.
51
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

4. Coulomb’s theory is more versatile than Rankine’s in that it take into account any
shape of the back fill surface, break in the wall face or in the surface of the fill, effect
of stratification of the backfill, effect of various kinds of surcharge on earth pressure,
and the effects of cohesion, adhesion and wall friction. It leads itself to elegant
graphical solutions and gives more reliable results, especially in the determination of
the passive earth resistance; this in spite of the fact that static equilibrium condition
does not appear to be satisfied in the analysis.
5. Rankine’s theory is relatively simple and hence is more commonly used, while
Coulomb’s theory is more rational and versatile although cumbersome at times;
therefore, the use of the latter is called for in important situations or problems.

52
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

CHAPTER FOUR

STABILITY OF SLOPES

4.1 Introduction
The term slope as used here refers to any earth mass, natural or man-made, whose
surface forms an angle with the horizontal. Hills and mountains, river banks and coastal
formations are common examples of sloped earth masses formed by nature perhaps as a
result of glacial movements, weathering, erosion, deposit buildup and sedimentation, or
other factors. Examples of man-made slopes may include fills such as embankments,
earth dams, levees; examples associated with cuts may include highway and rail way cuts,
canal banks, foundation excavations, and trenches.
Virtually every slope experiences gravitational forces and subsequent changes.
Indeed, the combining effects of gravity and water are primary and direct influences on
changes for most of the slopes (failure of an earth slope). Some slopes may possibly also
be influenced by such natural phenomena as chemical actions, earth quakes, glacial
forces, or wind. In virtually all instances, the effect is a general flattening of the slope,
either suddenly or slowly and cumulatively. Predicting the change with any degree of
accuracy may be a difficult task; preventing such change may be an even greater task for
the soil engineer.
Determination of the potential failure surface and the forces tending to cause slip
and those tending to restore or stabilize the mass of the earth are essential steps in the
stability analysis (functional) of earth slopes and available margin of safety. The soil
mass is assumed to be homogenous. It is also assumed that it is possible to compute the
seepage forces from the flow net and the shearing strength of the soil from the Mohr-
Coulomb theory. The same principles of limiting equilibrium mechanics are used to
analyse the stability of slopes (unretained soil masses) as for retained soil masses. For
purposes of analysis, slopes may be classified either as an infinite or finite. In practice, a
slope may be considered to be infinite where the soil properties at corresponding depths
are the same and where the depth D to a hard stratum is constant and small compared
with the overall length of the slope, as shown, for example, in Fig 4.1(a). For this
geometry, any mass movement or failure of the slope usually involves the displacement
of a soil mass having a length many times greater than its depth, as indicated by the
potential slip path [Fig 4.1(a)]. Neglecting end effects, a typical vertical slice through the
soil of width Δb can be taken as representative of the whole failure mass, in which case a
consideration of the simple statics governing equilibrium of the typical slice can be used
to derive stability relationship for the slope.

Fig 4.1: Slopes


A slope of finite extent (bounded by the top surface) is illustrated in Fig 4.1(b). This
represents the more general problem and in such slopes the statics of the whole failure
mass must be considered. Because of this, the analysis of finite slopes is usually more
53
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

complex than that for infinite slopes. This category includes cuttings for roads, railways,
canals, etc., and embankments for roads and earth dams.
4.2 Infinite slopes
Let’s consider an infinite slope inclined at angle β to the horizontal in a general C-φ soil
as shown in Fig 4.2.

Fig 4.2 Forces on element of infinite slope


The material above the slip plane is assumed homogenous and the slip plane is also
assumed parallel to the surface of the slope. A unit thickness of the element is assumed in
the direction normal to the page. Forces F1 and F2 are assumed equal and opposite, and
therefore ignored in the analysis. Thus, by decomposing forces [Fig 4.2 (b)] in the
perpendicular and parallel directions to the slope, we obtain the resulting normal and
shear stresses on the failure surface (acting stresses) as shown below.
Normal load N Wcos β Wcos β W (4.1)
Normal stress, σ = = = = = cos 2 β = γ Hcos 2 β
Area l*l l*l b/cos β b
Effective normal stress is
σ' = σ-u = γ H cos2β -u where u = pore water pressure (4.2)
Shear force T Wsin β Wsin β W (4.3)
Shear stress, τ = = = = = sin β cos β = γ Hsin β cos β
Area l*l l*l b/cos β b
The shear strength of the soil is, σ = c' + σtanφ' = c' + (γ Hcos 2 β - u)tan φ ' (4.4)
Shear strength c' + (γ Hcos β - u)tan φ '
2
Factor of safety against slippage, F = =
shear stress γ Hsin β cos β At
c' + (γ Hcos 2 β - u)tan φ '
⇒ F = ( 4 .5 )
γ Hsin β cos β
the verge of failure the shear strength is equal to the acting shear stress, i.e. F=1. If we
make F=1 in equation 4.5, we can find that the slope will be stable to a maximum depth
Hc, called the critical depth of the stratum. Thus
c' - u tan φ '
⇒ Hc = ( 4 .6 )
γ sin β cos β - γ cos 2 βtanφ'
From equations (4.5) and (4.6) different cases can be seen as:
Case (i) Cohesionless soil - No seepage
c' + (γ Hcos 2 β - u)tan φ ' tan φ '
F= =
γ Hsin β cos β tan β
c' - u tan φ '
Hc = =0
γ sin β cos β - γ cos 2 βtanφ'
⇒ γ sin β cos β - γ cos 2 βtanφ' = 0 ⇒ tan β = tan φ ' ⇒ β cr = φ ' is called the critical slope
Case (ii) Cohesionless soil –Seepage
c' + (γ Hcos 2 β - u)tan φ ' (γ Hcos 2 β - u)tan φ '
F= =
γ Hsin β cos β γ Hsin β cos β
c' - u tan φ ' u tan φ '
Hc = =
γ sin β cos β - γ cos 2 βtanφ' γ sin β cos β - γ cos 2 βtanφ'
54
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Case (iii) Cohesive soil and No seepage

c' + (γ Hcos 2 β - u)tan φ ' c' + γ Hcos 2 βtan φ '


F= =
γ Hsin β cos β γ Hsin β cos β
c' - u tan φ ' c' c' ⎛ sec 2 β ⎞
Hc = = = ⎜⎜ ⎟⎟
γ sin β cos β - γ cos βtan φ ' γ sin β cos β - γ cos βtan φ ' γ ⎝ t an β - tan φ ' ⎠
2 2

Rapid draw down in a slope


When the water level in a river or reservoir recedes, say after floods or after a drawdown,
the water in the slope of the embankment may fall as rapidly as that in the reservoir,
depending upon the permeability of the soil. This gives rise to a condition commonly
known as “sudden or rapid drawdown”. The effect of this is that seepage occurs from the
high water level in the slope to the lower water level of the river. A flow net can be drawn
for this condition and the excess hydrostatic head at any point within the slope can be
determined. If seepage coupled with a water table at the surface of the ground
is assumed as shown in Fig 4.3 and the excess pore water pressure at the center of the
base of the element is represented by a height hw, the excess pore water pressure u at the
base of the element is determined as follows.
AB = H cosβ
h = ABcosβ = Hcos2β
w
Thus, u = γ h = γ Hcos2β
w w w

Fig 4.3: Slope with seepage and water table at surface


If we substitute this value of u in equation 4.5 we have,
c' + (γ Hcos 2 β - u)tan φ ' c' + (γ Hcos 2 β - γ w Hcos 2 β )tan φ '
F= =
γ Hsin β cos β γ Hsin β cos β
c' + H( γ - γ w )cos 2 βtan φ '
⇒F= ( 4 .7 )
γ Hsin β cos β
Normally γ = γsat ⇒ γ - γw = γsat- γw = γsub (or γb)
c' + H γ b cos 2 βtan φ '
⇒F= ( 4 .8 )
γ Hsin β cos β
The critical depth Hc obtained from eqn. 4.8 when F=1 is,
c'
Hc = ( 4 .9 )
⎡ γ ⎤
γcos 2β ⎢ tan β - b tan φ '⎥
⎣ γ ⎦
c' ⎡ γ ⎤
= cos 2β ⎢ tan β - b tan φ '⎥ = stability number (dimension less)
γH c ⎣ γ ⎦
Example: A relatively cohesive soil at a constant infinite slope has a negligible seepage
and negligible pore water pressure. If γ=18 KN/m3, c' = 36Kpa, φ =140, H=3m, and
β=220, find the factor of safety against sliding.
Solution: From eqn 4.7 F = c' + H(γ - γ w )cos βtan φ ' = 36 + 3 * (18 − 0)(cos0 22 )tan
2 2 0
14 0
= 2.54
γ Hsin β cosβ 0
18 * 3 * sin 22 cos 22
4.3 Finite slopes
A finite slope is one with a base and a top surface, the height being limited. The inclined
faces of earth dams, embankments, excavations and the like are all finite slopes.

55
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Investigation of the stability of finite slopes involves the following steps according to the
commonly adopted procedure:
a) Assuming a possible slip surface,
b) Studying the equilibrium of forces acting on this surface (activating & resisting)
c) Repeating the process until the worst slip surface, that is, the one with minimum
margin of safety is found.
Methods of analysis can be by:
1) Considering the whole sliding mass as a free body which include,
ƒ Culmann method
ƒ The Circular arc analysis
2) Dividing the sliding mass in to a number of vertical slices and considering the
equilibrium of each slice
ƒ Swedish method of slices- (Fellenius and Peterson)
ƒ Bishop Method
ƒ Taylor’s method
4.3.1 Culmann Method
This method of analysis is based on the assumption that failure occurs on a slip plane
through or above the toe of the slope. Fig 4.4 shows the forces acting on the sliding mass.

Fig 4.4: Culmann’s Method


1
Weight of wedge ABC = hL γ * 1
2
H h Hsin( β − θ)
AB = = ⇒ h =
sin β sin( β − θ) sin β
1 sin( β − θ)
W = HL γ
2 sin β
The resisting force to sliding along AC
F R = cL + Wcos θ tan φ
The force acting to cause sliding
F A = Wsin θ
1 sin( β − θ)
c+ Hγ cos θ tan φ
F cL + Wcos θ tan φ 2 sin β
Factor of safety, FS = R = =
FA Wsin θ 1 sin( β − θ)
Hγ sin θ
2 sin β
A value of FS ranging from 1.25 to 1.5 is considered to be satisfactory for an earth slope. For
economic reasons, a value greater than 1.5 is not desired. Hence, FS=1.5 may be considered
to be necessary as well as sufficient.
Referring to the force polygon shown in Fig 4.4(c)
cL W 1/2 γ LHsin( β - θ ) c 1 sin( β - θ )sin( θ - φ )
= = ⇒ = ( 4 .11) w
sin( θ - φ ) cos φ sin β cos φ γ H 2 sin β cos φ
here the quantity c/γH is known as the stability number. The condition for impending slip
occurs when c/γH is a maximum. Thus, from eqn 4.11,

56
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
∂ (c / γ H ) 1
= 0 ⇒ θ cr = (β + φ ) ( 4 .12 )
∂θ 2
The method gives good results for very steep or vertical slopes; it does not provide
satisfactory results for relatively flat slopes.
Example: Find the factor of safety of the slope if γ=17KN/m3, c=26KPa, φ=180, β=480
and H=15m.
Solution: The critical failure plane orientation from eqn 4.12 is θ cr = 1 (β + φ ) = 33 0 and
2
0
L=H/sinθ =15/sin33 =27.54m.Thus factor of safety from eqn 4.10 is then
1 sin( β − θ) 1 sin(48 − 3 3 )
c + Hγ cos θ tan φ 26 + * 15 * 17 cos 3 3tan 1 8
2 sin β 2 sin48
FS = = = 1 . 58
1 sin( β − θ) 1 sin(48 − 3 3 )
Hγ sin θ * 15 * 17 sin 3 3
2 sin β 2 sin48

4.3.2 The Circular Arc Analysis


This method of analysis is based on the assumption that failure occurs on a circular slip
surface. Fig 4.5 shows the forces acting on the sliding mass.

Fig 4.5 Forces acting on mass circular failure surface


Generally, the pore water pressure at any point along the arc may be estimated via the
flow net construction. Thus, we may divide the arc length AB in to n equal lengths and
compute the average pressure head for each length. The average force per unit thickness
for any segments i become . The vector
hi (AB/n) γw summation of these forces gives
the pore pressure resultant U; it acts through the center of the circle. The total cohesion
force is r θ where θ is in radian.
AB = c(AB).
For an assumed center of failure arc, the actuating or overturning moment is Mo= We.
The lines of action of W pass through the centroid of wedge’s cross section. The centroid
may be estimated by cutting a piece of cardboard to the shape of the wedge and balancing
the cutout on a pin or sharp point (then e can be determined). Thus, for a shear strength
τ =c(AB)the resisting moment is Mr = r τ. The pore pressure on the arc
+ tanφΣσ',
does not contribute to the moment since U passes through the circle center O. The
corresponding safety factor is FS= Mr /Mo, or
⎡ n

r ⎢ c r θ + tan φ ∑ σ' i ⎥
FS = ⎣ ⎦ (4.13)
i =1

We
If there is a tension crack, the cracked part will not have shear resistance and hence the
total cohesion resistance will be c(AB'). And AB' =r θ' where θ' is in radian

If the backfill is purely cohesive φ =0 [or if analysis is on total stresses (φ=0 analysis)],

57
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

c r 2θ
FS =
We
In eqn 4.13 the term σ'i varies over the arc and is not symmetrical. Its value is influenced
mainly by the weight of the soil above a point i. This approach leads to the method of
slices discussed below.
4.3.3 The Swedish Method of Slices
In this method of analysis the area within the slip circle is divided in to a number of
vertical slices generally of equal width [Fig 4.6 (a)]. The forces on each slice are
evaluated from the limiting equilibrium of slices. These forces are shown in Fig 4.6 (b).
The equilibrium of the entire mass is determined by summation of the forces on each slice
as forces activating failure and forces resisting failure.

Fig 4.6: Swedish method of slices.


The reactions R1i and R2i at the sides of the slice are assumed equal. And forces on the
vertical faces of the slice F1i and F2i are neglected.
Here N i = W i cos α i and T i = W i sin α i
Ni W i cos α i Ti W i sin α i
⇒ Normal stress, σ i = = and Shear stress, τ i = =
Li *1 Li Li *1 Li
Wi
Shear strength : τ = c + σtan φ = c + cos α i tan φ
Li
Shearing resistance force on a slice : T fi = cL i + W i cos α i tan φ
n n n
Shearing resistance force of entire soil wedge is : ∑ τL
i =1
i = ∑ cL
i =1
i + ∑ W i cos α i tan φ
i =1

Shearing force acting on the slice to cause failure is : τ i L i = W i sin α i


n
Total Shearing force acting on the wedge to cause failure is : ∑ W sin α
i =1
i i

n n

Force resisting failure


∑ cL
i =1
i + ∑ W i cos α i tan φ
i =1
Thus , f actor of safety, FS = = n
( 4 . 14 )
Force causing failure
∑ W sin α
i =1
i i

Note : The tangentia l components of few slices may cause restoring moments instead of
activating moments. The shearing resistance with in the tension cracked zone is zero.
The above procedure can be tabulated as follows:
Slice No Ai (m2) Wi (KN) αi (deg) Li (m) CLi (KN) Wi cosαi (KN) Wi sinαi(KN)
1
2
etc
Sum
In this method a number of slip circles are analyzed and the minimum factor of safety is
finally obtained. The circle with this minimum factor of safety is the failure circle.
58
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

The Effective Stress Analysis


Total stress (φ =0) analysis is applicable for the analysis of stability of a slope soon after
construction under undrained conditions. If the pore water pressure exists in an
embankment under certain conditions of drainage or seepage, an analysis in terms of
effective stress is considered appropriate. Thus we have,
W i cos α i W i sin α i
σ 'i = − ui and τi =
Li Li
n n

∑ c' L + ∑ (W cos α
i =1
i
i =1
i i − u i L i )(tan φ ' )
FS = n
( 4 . 15 )
∑ W sin α
i =1
i i

Here ui is obtained from flow nets or other methods.

Estimation of Pore Water Pressure


The pore water can be determined from flow nets. The procedure is to draw flow lines
and the equipotential lines up to the top of the flow net. Consider the slice abcd
The height to which water will rise in a stand
pipe inserted at the point is hw. This height
for each slice will be the vertical distance b/n
base of slice (mid point) and the intersection
of the equipotential line through the middle
point of the sliding surface of slice with the

Fig 4.7a: determination of pore water pressure from flow net.


The pore water pressure for each slice is then: ui =γw hw
The pore pressure ratio, ru, is defined as γ h
ru = w w
γ z
The effect of pore water pressure is to reduce the effective stress and thereby reduce the
stability because the shear strength mobilized would be decreased.
The factor of safety is calculated from,
n n

∑ c' L i + ∑ (W i cos α i − u i L i )(tan φ ' ) (4.16)


i =1 i =1
FS = n

∑ W sin α
i =1
i i

In the absence of a flow net, the approximate value of FS may be obtained from:
n n

∑ c' L i + ∑ (W i cos α i )(tan φ ' ) (4.17)


i =1 i =1
FS = n


i =1
W i sin α i

Here the normal component Ni (=Wi cosαi) of the weights of the slices have to obtained
using effective or buoyant unit weight γsub (γb or γ’) and the tangential components Ti
(=Wisinαi) using the saturated unit weight.

59
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Steady Seepage- Long term stability


The case when steady seepage is occurring at the maximum possible rate through an earth
dam or an embankment is considered as the critical condition for the stability of the
downstream slope. The seeping water below the phreatic line, exerts a pore water pressure
on the soil mass which lies below the phreatic line. Hence if the slices of the critical circle
arc, happen to include the submerged soil (Fig 4.7 b)
The factor of safety is calculated from,
n n

∑ c' L + ∑ (W cos α
i =1
i
i =1
i i − u i L i )(tan φ ' )
FS = n

∑ W sin α
i =1
i i

In the absence of a flow net, the approximate


value of FS may be obtained from eqn. 4.17

Fig 4.7a: determination of pore water pressure from flow net.


The pore water pressure at a point is represented by the piezometric head at that point, as
explained earlier.
Rapid Drawdown
For the upstream slope of an embankment or an earth dam, the case of rapid or sudden
drawdown represents the critical condition since seepage force in that condition adds to
the sliding moment while it reduces the shear resistance mobilized by decreasing the
effective stress. The effect of rapid drawdown on slope stability depends very much on
the opportunity for drainage at the base. If the base material is pervious the flow pattern
tends to be down wards, which is conducive to stability; otherwise, the seepage forces
may create more unfavorable conditions with respect to stability. The pore water pressure
along the slip surface can be determined from the flow net constructed after the
drawdown or alternatively the flow net before the drawdown can be employed to estimate
the porewater pressure as described here under (Fig 4.8).
ƒ Pore water pressure before drawdown, uo is
uo=γw hw=γw(z +h1- h2)
ƒ Pressure due to weight of water on a slice
(before draw down), σw =γw h1
ƒ Pore water pressure after draw down, ui,
will be ui= uo- σw = γw(z +h1- h2)-γw h1
⇒ ui= γw(z - h2)

Fig 4.8: Up stream slope subjected to rapid drawdown


The factor of safety is then calculated from,
n n

∑ c' L i + ∑ (W i cos α i − u i L i )(tan φ ' ) (4.18)


i =1 i =1
FS = n


i =1
W i sin α i

In the absence of a flow net, the approximate value of FS may be obtained from eqn 4.17
given previously.

60
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Earth quake
In the case of an earth quake, a horizontal seismic force is applied at the centroid of the
slice as shown in Fig 4.9 and the factor of safety is estimated as usual.
n n

∑ c' Li +∑ (Wicosαi − u iLi )(tanφ ')


FS = i =1 i =1 (4.19)
n n

∑ W sinα + ∑ C W a
i =1
i i
i =1
s i i

Cs = seismic coefficient (Cs=0.03g to 0.27g) –common 0.1g


ƒ The seismic force has no vertical component but only
increases the driving force.
Fig 4.9: Earth quake force ƒ This method is pseudostatic analysis. Dynamic analysis
shall be made for important dams

4.3.4 Bishop’s Simplified Method of Slices


As mentioned in the Swedish method of slices, the side forces were neglected. This
violates the equilibrium requirement with respect to translation. Thus, the safety factor
based on the Swedish method of slices is somewhat in error. Yet, to include these forces
complicates the approach quite appreciably. Bishop suggested an alternative approach to
determine Ni. He assumed the resultant in the vertical direction on the forces acting on the
sides of the slice to be zero. Thus, referring to Fig 4.10, we have

Fig 4.10: Bishop’s method.


N i = Wi cosα i − u i li and Ti = Wisinα i
The shear resistance of the soil introducing some factor of safety is :
c' li + N i tanφ '
Ti = (4.20)
FS
From ∑ Fy = 0, we have : (u ili + N i )cosα i − Wi + Tisinα i + F2i − F1i = 0 (4.21)
Substituting eqn 4.20 to eqn 4.21 and dividing by cos α i and rearranging, we have
c ' li
Wi − ui li cos α i − sin α i + (F1i − F2i )
Ni = FS
⎛ sinα i tan φ ' ⎞
⎜ cos α i + ⎟
⎝ FS ⎠
The factor of safety is :
c' li
Wi − ui li cos α i − sin α i + (F1i − F2i )
∑ c' l + ∑
i

FS
sinα i tan φ ' ⎞
tanφ '
⎜ cos α i + ⎟
FS =
∑ c' l + ∑ N tanφ ' =
i i ⎝ FS ⎠
∑T i ∑ Wisinαi
61
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

⇒ FS =
∑ c' l i cos α i + ∑ [W i − u i l i cos α i + (F1i − F 2i )]tan φ '
⎡ sin α i tan φ ' ⎤

W i sin α i ⎢ cos α i +
⎣ FS ⎥

Calculatio ns are simplified if ∑ (F1i − F 2i ) = 0 and the error introduced is very small ( < 1%).
Thus factor of safety wil l be

FS =
∑ c' l i cos α i + ∑ [W i − u i l i cos α i ]tan φ '
( 4 . 22 )
⎡ sin α i tan φ ' ⎤
∑ W i sin α i ⎢ cos α

i +
FS ⎥

The solution of this equation requires an iterative analysis since FS appears on both sides
of the equation. Trial and error approach consists of assuming an FS value on the right
hand side of the equation, then solving for FS on the left. If the difference between the
assumed FS and the computed FS is significant, a new FS is assumed and the procedure
repeated until a satisfactory FS is determined. Computer programs are generally preferred
in this method. If φ'=0, the eqn 4.22 may be solved directly for FS.

The following table summarizes different situations and their corresponding preferred
methods.
S. No. Situation Preferred Method
1 End of construction with partly Total stress analysis (φ=0 analysis)by
saturated soil-construction period is using Cu and φu from UU tests
short compared to consolidation time
2 Steady Seepage Effective stress analysis by using
C', φ' with the pore water pressure
obtained from flow nets
3 Rapid draw down Effective stress analysis with C', φ'
from CU tests. Pore water pressure
estimated from flow nets or other
methods
4 Earth quake Total or Effective stress analysis
depending on the data available

4.3.5 Friction Circle Method


The reaction R against the failure surface makes an angle φ to the normal to the failure
surface. So, all reactions will be tangent to the circle with O as center and rsinφ as a
radius (Fig 4.11). This circle is called the friction circle.

Fig 4.11: Concept of Friction Circle


62
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

In the friction circle method with known φ, the following quantities are known:
1. Magnitude of weight sliding wedge W and its line of action
2. Direction of R (at angle φ to the normal to the failure surface). Magnitude of R and its
distribution along the failure surface not known
3. The direction of total cohesion Cm=cm L along the rupture surface parallel to chord AB
Let’s assume a uniform distribution of cohesion along the failure surface. If we resolve
the cohesion force parallel and perpendicular to the chord AB, the components
perpendicular to chord are oppositely directed, they do not contribute any moments. If cm
is mobilized cohesion, then
Lr
c m L r = c m L' r c ⇒ rc =
L'
The line of action of W and Cm are located first. Then the triangle of forces may be
completed as shown in Fig 4.11(b) and the values of R and Cm can then be determined
and the factor of safety is calculated.
The factor of safety with respect to cohesion, assuming the friction is mobilized in full, is
given by:
c
FS c = (4.23)
cm
The factor of safety with respect to friction, assuming the cohesion is mobilized in full, is
given by:
tan φ
FS φ = (4.24)
tan φ m
Where φ and φm are the total (effective) angle and mobilized friction angle. If the factor of
safety with respect to the total shear strength FS is required, φm is to be chosen such that
FSc and FSφ are equal. Here FS is:
shear strength ( = τ )
FS = (4.25)
mobilized shear resistance ( = τ m )
Where: τ = c + σ'tanφ and τm = cm + σ' tanφm
4.3.6 Taylor’s Method
Taylor (1948) prepared two charts relating the stability number (N=c/γH) to the angle of
slope, β, based on the friction circle method and analytical approach. The first is for the
general case of a c-φ soil with the angle of slope less than 530 (Fig 4.12). The second is
for a soil with φ=0 and a layer of rock or stiff material at a depth DH below the top of the
embankment, as shown in Fig 4.13. Here, D is known as the depth factor; depending
upon its value, the slip circle will pass through the toe or will emerge at a distance nH in
front of the toe (the value of n may be obtained from the curves).

Fig 4.12: Taylor’s charts for slope stability (After Taylor, 1948)
(for φ=00 and β<530, use Fig 4.13)
63
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 4.13: Taylor’s charts for slope stability (After Taylor, 1948)

Taylor’s charts are based on the assumption of full mobilization of friction, that is, these
give the factor of safety with respect to cohesion. This is all right for a purely cohesive
soil; but, in the case of c-φ soil, where the factor of safety FS with respect to shearing
strength is desired, φm should be used for φ:
tan φ
tan φ m = (4.26)
FS
φ
(Also φm ≈ )
FS
The charts are not applicable for a purely frictional soil (c = 0). The stability then depends
only upon the slope angle, irrespective of the height of the slope.

64
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

CHAPTER FIVE

BEARING CAPACITY OF SOILS

1. Introduction

All civil engineering structures impose a loading on the underlying soil or rock. The part
of the structure, usually lying below ground level which transmits the load to the
supporting strata is referred to as the foundation. To ensure stability, foundations must
provide an adequate factor of safety against shear or bearing failure of the underlying
soil and the structure must be capable of withstanding the settlements that will result, in
particular the differential settlements. In the vast majority of cases the choice of the
foundation is governed by the second of theses two factors, thus in general the need to
limit settlements of the structure will control the design of its foundation.
Thus the criteria for the determination of the bearing capacity of a foundation are based
on the requirements for the stability of the foundation. The design value of the safe
bearing capacity would be the smaller of the two values, obtained from the two criteria:
1. Shear strength criterion HL\51.doc
2. Settlement criterion
The soil’s limiting shear resistance is referred to as the ultimate bearing capacity, qu, of
the soil. For design, one uses an allowable bearing capacity, qa, obtained by dividing the
ultimate bearing capacity by a suitable safety factor (i.e. qa=qu/FS). HL\52.doc
The bearing capacity is affected by factors like
a) Nature of the soil and its physical and engineering properties
b) Size, shape, depth, rigidity and roughness of the foundation
c) Water table conditions and initial stresses in the foundation soil
d) Total and differential settlements that the structure can withstand without
functional failure
In general, foundations constructed at a depth below the ground level which is
approximately less than or equal to 2 (i.e. Df <2B) –like footings and mat foundations are
designated as shallow foundations. In contrast, piles, piers and caissons (Df >2B) are
designated as deep foundations.
Numerous proposals have been advanced regarding considerations, criteria, and
procedures for the evaluation of the bearing capacity of soils. Some analytical methods
of estimating bearing capacity are given below.

2. Bearing Capacity Based on Rankine Wedge


Fig 5.1 (a) shows a long, narrow footing (i.e., L/B is very large) at a depth Df in to a c-φ
soil. Fig 5.1 (b) depicts the Rankine wedges used in this analysis. Wedge I is assumed to
be an active Rankine wedge, which is pushed down and slides to the right during the
failure sequence; wedge II is assumed to be a passive wedge, which is pushed to the right
and upward in the process. The lateral resistances (P) are assumed to act at the interface
of the two wedges as shown [Fig 5.1(c)]; they have the same magnitude but opposite

65
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

directions. The P associated with wedge I represents the active pressure resultant,
whereas the P for wedge II is the passive thrust.

Fig 5.1: Footing in c-φ soil


( Wedge I (active case):
1
Pa = γK a H 2 − 2 cH K a + q u K a H (a)
2
where K a = tan 2 (45 − φ /2 )
( Wedge II (passive case):
1
Pp = γK p H 2 + 2cH K p + γ D f K p H (b)
2
where K p = tan 2 (45 + φ /2 )
The two resultants at the interface are assumed to have same magnitude. Hence equating
eqn (a) and (b), we have
1 1
γK a H 2 − 2 cH K a + qu K a H = γK p H 2
+ 2cH K p + γ D fK pH
2 2
solving for q u , we have

qu =
1
2
γH
⎛ 1 ⎞
⎜⎜
K
⎟⎟ (K p − K a ) +
2c
K
K ( p + K a )+ γDfK 2
p (c )
⎝ a ⎠ a

But K p = 1/K a ; also, from Fig 5.1 (b)


B B
H = =
2tan(45 - φ /2) 2 K a

Hence, eqn (c) becomes

q u =
1
4
γB K ( p
5
2
− K p) + 2c (K + K
1
2
p
3
2
p
1
2
)+ γDfK 2
p (d)

Let N γ =
1
2
(K − K ) p
5
2
p
1
2

= 2 (K ) 3 1
N c + K p
2
p
2

1 + sin φ
N = K
2
and K = = tan 2
(45 + φ /2 )
1 − sin φ
q p p

Thus eqn (d) becomes,


1
q u = cN c + γBN γ + γ DfN q ( 5.1)
2
For a purely cohesive soil (φ=0): Kp =1, Nγ =0, Nc =4, Nq =1 and hence the ultimate
bearing capacity from eqn 5.1 is qu = 4c +γ Df . So in purely cohesive soils increase or
decrease in the width, B, of the footing has no effect on qu. But increasing Df increases qu
by a little factor.
The above derivation is base on less-than-accurate assumptions: (1) the shear at the
interface of the two wedges was neglected and (2) the failure surfaces are not straight
lines as was indicated for the two wedges.
3. Prandtl’s Theory for Ultimate Bearing Capacity
Prandtl’s theory of plastic equilibrium reflects on the penetration (deformation) effects of
hard objects in to much softer material. The assumptions made were: (1) Soil is isotropic
and homogenous (2) Soil is weightless (3) Footing is long and has a smooth base and
66
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

placed at the surface of the soil. Fig 5.2 shows three zones developed with in the soil
resulting from bearing failure.
Zone 1: Active zone
Zone 2: zone of radial shear or
plastic zone
Zone 3: passive zone

Fig 5.2: Prandtl’s theory of plastic equilibrium


The penetrating wedge I pushes sides of zone II and III. Shearing resistance mobilizes
along log spiral and straight line segment. Considering this shear resistance, the ultimate
bearing capacity of the soil based on Prandtl’s theory is given by:
[
q u = ccot φ tan 2 (45 + φ/2 )e π tan φ − 1 ] (5.2)
Reissner included a term for the overburden pressure q =γ Df as:
[ ]
q u = ccot φ tan 2 (45 + φ/2 )e π tan φ − 1 + γ D f tan 2 (45 + φ/2 )e π tan φ (5.3)
Terzaghi added further term for the weight of the soil.
4. Terzaghi’s Bearing Capacity Theory
Terzaghi improved on the wedge analysis described in the preceding section by
working with trial wedges of the type assumed by Prandtl. However, he expanded on
Prandtl’s theory to include the effects of the weight of the soil above the footer (bottom)
level, an aspect the Prandtl omitted in his work. Terzaghi assumed the general shape of
the various zones to remain unchanged, as illustrated in Fig 5.3(a). Terzaghi assumed the
angle that the wedge faced formed with the horizontal to be φ, rather than the (45+φ/2)
assumed in Prandtl’s and most other theories.
As did Prandtl, Terzaghi assumed a strip footing of infinite extent and unit width.
Unlike Prandtl, however, Terzaghi assumed a rough instead of a smooth base surface.
Further more, although he neglected the shear resistance of the soil above the base of the
footing (segment gf in Fig 5.3 (a)), he did account for the effects of the soil weight by
superimposing an equivalent surcharge load q=γ Df. Otherwise, the shape of the failure
surface is similar to that assumed by Prandtl.

Fig 5.3: Terzaghi’s bearing capacity theory


Fig 5.3 (b) shows the penetrating wedge, where the down ward load is resisted by the
forces of the inclined faces of the wedge. Assuming a unit length of the footing normal to
the page, we obtain qu as follows
∑ Fy = 0 ⇒ q u B = 2 P p + 2(bd)csin φ
But bd = (B/2)/cos φ . Thus,

67
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

q u B = 2 Pp + Bctan φ (5.4)
Terzaghi represented the value of Pp as the vector sum of the three components: (1) that
from cohesion, (2) that from surcharge, and (3) that resulting from the weight of the soil
(bdef in Fig 5.3 (b)). With the addition for the shape factors in the cohesion and base
terms, Terzaghi obtained expressions for the ultimate bearing capacity for general shear
conditions as:
1
Long footings : q u = cN c + γ D f N q + B γ N γ ( 5.5)
2
Square footings : q u = 1 . 3 cN c + γ D f N q + 0 . 4 γB N γ ( 5.6)
Circular footings : q u = 1 . 3 cN c + γ D f N q + 0 . 3 γB N γ ( 5.7)

where : N =
a 2
; N = [N ]
− 1 cot φ ;
q
2 cos 2
(45 + φ / 2 ) c q


⎜ 3π − φ ⎞
⎟ tan φ
1 ⎡ K pγ ⎤ ⎜
4 2 ⎟
N γ = tan φ ⎢ − 1⎥ with a = e ⎝ ⎠
2 ⎣ cos
2
φ ⎦
⎡ ⎛ φ + 33 ⎞ ⎤
N pγ = 3tan 2
φ ⎢ 45 + ⎜ ⎟⎥ (After S. Husain)
⎣ ⎝ 2 ⎠⎦

Table 5.1 below gives the values for the various bearing capacity factors recommended
for the above equations.
Table 5.1: Terzaghi’s N-factors
φ0 0 2 4 6 8 10 12 14 16 18 20 22 24
Nc 5.7 6.3 6.97 7.73 8.6 9.61 10.76 12.11 13.68 15.52 17.69 20.27 23.36
Nq 1 1.22 1.49 1.81 2.21 2.69 3.29 4.02 4.92 6.04 7.44 9.19 11.4
Nγ 0 0.18 0.38 0.62 0.91 1.25 1.7 2.23 2.94 3.87 4.97 6.61 8.58
φ0 26 28 30 32 34 36 38 40 42 44 46 48 50
Nc 27.09 31.61 37.16 44.04 52.64 63.53 77.5 95.67 119.67 151.95 196.2 258.29 347.52
Nq 14.21 17.81 22.46 28.52 36.51 47.16 61.55 81.27 108.75 147.74 204.2 287.86 415.16
Nγ 11.35 15.15 19.73 27.49 36.96 51.7 73.47 100.39 165.69 248.29 427 742.61 1153.2

The results obtained here are quite within acceptable limits for shallow footings (e.g.
Df/B<1) subjected to only vertical loads. But they are limited to concentrically loaded
horizontal footings; they are not suitable for footings that support eccentrically-loaded
columns or to tilted footings. Furthermore, they are regarded as somewhat overly
conservative.
Terzaghi developed his bearing-capacity equations assuming a general shear failure in a
dense soil and a local shear failure for a loose soil. For the local shear failure he proposed
reducing the cohesion and φ as:
2
c ''= c
3
⎛ 2 ⎞
φ ' ' = tan −1
⎜ tan φ ⎟
⎝ 3 ⎠

5. Meyerhof’s Bearing Capacity Equation


Meyerhof proposed a bearing capacity equation similar to that of Terzaghi but
added shape factors, s, depth factors, d, and inclination factors, i (Eqn 5.8).
1
InclinedLoad: q u = cNc s c d c ic + γ Df N q s q d q iq + Bγ N γ s γ d γ iγ (5.8)
2
where: N q = e π tanφ
tan (45 + φ / 2)
2
N c = (N q − 1)cotφ
N γ = (N q − 1) tan(1.4φ )

The N values are given in Table 5.2 (a) and (b).

68
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Table 5.2 (a): Meyerhof’s N- factors


0 0 2 4 6 8 10 12 14 16 18 20 22 24
φ
Nc 5.1 5.63 6.19 6.81 7.53 8.34 9.28 10.37 11.63 13.1 14.83 16.88 19.32
Nq 1 1.2 1.43 1.72 2.06 2.47 2.97 3.59 4.34 5.26 6.4 7.82 9.6
Nγ 0 0.01 0.04 0.11 0.21 0.37 0.6 0.92 1.37 2 2.87 4.07 5.72
0 26 28 30 32 36 38 40 42 44 46 48 50
φ 34
Nc 22.25 25.8 30.14 35.49 42.16 50.59 61.35 75.32 93.71 118.37 152.1 199.27 266.89
Nq 11.85 14.72 18.4 23.18 29.44 37.75 48.93 64.2 85.38 115.31 158.51 222.31 319.07
Nγ 8 11.19 15.67 22.02 31.15 44.43 64.08 93.69 139.32 211.41 328.74 526.47 873.89

Table 5.2 (b): Meyerhof’s factors (s, d, i)


φ Shape Depth Inclination
Any φ
2
s c = 1 + 0.2K p
B
d c = 1 + 0.2 K p
D ⎛ α ⎞
ic = iq = ⎜1 - 0 ⎟
L B ⎝ 90 ⎠
For φ =00 s q = s γ = 1.0 d q = d γ = 1.0 i γ = 1 .0

For φ >100 s = s = 1 + 0.1K B


2
D ⎛ α ⎞
q γ p d q = d γ = 1 + 0.1 K p i γ = ⎜⎜ 1 - ⎟⎟
L B ⎝ φ ⎠
⎛ φ⎞
K p = tan 2 ⎜ 45 + ⎟ where α = angle of resultant measured from vertical axis
⎝ 2⎠
V α Q
H

⎛ B⎞
When triaxial φ tr is used for plain strain, adjust φ tr to obtain φ ps = ⎜1.1 − 0.1 ⎟φ tr
⎝ L⎠
Meyerhof suggested that footing dimensions B'=B-2ey and L'= L-2ex be used in
determining the total allowable load eccentrically applied in the x and y directions,
respectively (i.e., Qu=qu B' L'), and in the corresponding terms in the ultimate bearing
capacity equations and in the various correction factors for shape and inclination.
6. Hansen’s Bearing Capacity Equation
Hansen proposed the general bearing capacity equation (eqn 5.9) which includes ground
factors and base factors to include conditions for a footing on a slope.
1
q u = cN c s c d c i c b c g c + γ D f N q s q d q i q b q g q + Bγ N γ s γ d γ i γ b γ g γ (5.9)
2
where : N q = e π tan φ tan 2 (45 + φ / 2 ); N c = (N q − 1)cot φ ; N γ = 1 .5 (N q − 1) tan φ E
N q a nd N c are same as Meyerhof' s N q & N c . [Table 5.3 (a)]
xpressions for inclination, shape, depth, base, and ground inclination expressions
proposed by Hanson are given in Table 5.3 (b).
Table 5.3 (a): Hansen’s N- factors

69
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

φ0 0 2 4 6 8 10 12 14 16 18 20 22 24
Nc 5.1 5.63 6.19 6.81 7.53 8.34 9.28 10.37 11.63 13.1 14.83 16.88 19.32
Nq 1 1.2 1.43 1.72 2.06 2.47 2.97 3.59 4.34 5.26 6.4 7.82 9.6
Nγ 0 0.01 0.05 0.11 0.22 0.39 0.63 0.97 1.43 2.08 2.95 4.13 5.75
φ0 26 28 30 32 34 36 38 40 42 44 46 48 50
Nc 22.25 25.8 30.14 35.49 42.16 50.59 61.35 75.32 93.71 118.37 152.1 199.27 266.89
Nq 11.85 14.72 18.4 23.18 29.44 37.75 48.93 64.2 85.38 115.31 158.51 222.31 319.07
Nγ 7.94 10.94 15.07 20.79 28.77 40.05 56.18 79.54 113.96 165.58 244.65 368.68 568.59

Table 5.3 (b): Hansen’s factors (s, d, i, b, g)


Shape factors Depth factors Inclination factors
B ' (for φ=0) d = 0 .4 k (for φ=0) 1 1 Hi
s c = 0 .2 c
ic = − 1− (for φ =0)
L' 2 2 AfCa
d c = 1 . 0 + 0 .4 k (for φ>0)
sc = 1 +
N q B ' (for φ>0) 1 − iq
N c L' ic = iq − (for φ >0)
d q = 1 + 2 tan φ (1 − sin φ ) k 2
Nq −1
B'
s q = 1 .0 + sin φ d γ = 1.0 α1
L' ⎡ 0 .5 H i ⎤
D D i q = ⎢1 − ⎥
B' k = if ≤ 1 V + A f C a cot φ ⎦
s γ = 1.0 - 0.4 ≥ 0 .6 B B ⎣
L' ⎛
k = tan − 1 ⎜
D ⎞
⎟ if
D
> 1 2 ≤ α1 ≤ 5
⎝ B ⎠ B
α2
k in radians ⎡ ⎛ η ⎞ ⎤
⎢ ⎜ 0 .7 − 0
⎟H i ⎥
⎝ 450 ⎠
i γ = ⎢1 − ⎥
⎢ V + A f C a cot φ ⎥
⎢ ⎥
⎣ ⎦
2≤α2 ≤5
Ground factors (base on
slope)
β 0 (for φ=0)
gc = 0
147

β 0
(for φ>0)
g c = 1 .0 −
147 0
g q = g γ = (1 − 0 . 5 t an β )
5

Base factors (tilted base)


β+η < 900; β < φ ; D measured vertically. η 0 (for φ=0)
bc =
For L/B < 2 use φtr 0
147
For L/B>2 use φps=1.5 φtr -170 but for φtr< 340use φtr= φps η0
bc = 1− (for φ>0)
δ = friction angle between base and soil (0.5φ < δ < φ) 147 0

Af = B' L' (effective area) b q = e − 2η tan φ η in radians


ca = base adhesion (0.6c to 1.0c)
b γ = e − 2 .7η tan φ η in radians
Notes
9 Failure can take place either along the long side or along the short side and thus
shape , depth and inclination factors shall be calculated in both sides
9 Use Hi as either HB or HL for inclination factors

7. Vesic’s Bearing Capacity Equation

70
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

The Vesic procedure is essentially the same as the method of Hansen with select changes.
The Nc and Nq terms are those of Hansen but Nγ is slightly different as is given by:
N γ = 2(N q + 1) tan φ (also see Table 5.4 (a))
There are also differences in the ii, bi and gi, terms (Table 5.4 (a)).
Table 5.4 (a): Vesic’s Nγ - factors
φ0 0 5 10 15 20 25 26 28
Nγ 0 0.4 1.2 2.6 5.4 10.9 12.5 16.7
φ0 30 32 34 36 38 40 45 50
Nγ 22.4 30.2 41 56.2 77.9 109.3 271.3 761.3
HL\53.doc

Table 5.4 (b): Vesic’s factors (s, d, i, b, g)


Shape factors Depth factors Inclination factors
= 1.0 +
Nq B d c = 0 .4 k (for φ=0) mH i (for φ =0)
sc ic = 1 −
Nc L AfCaNc
B d c = 1 . 0 + 0 .4 k (for φ>0) 2 + B/L
s q = 1 . 0 + tan φ m = mB =
L Where: 1 + B/L
s γ = 1.0 - 0.4
B
≥ 0 .6 d q = 1 + 2 tan φ (1 − sin φ ) 2 k 2 + L/B
m = mL =
L 1 + L/B
d γ = 1.0
D D 1 − iq
k =
B
if
B
≤ 1
ic = iq − (for φ >0)
⎛ D ⎞ D Nq −1
k = tan − 1 ⎜ ⎟ if > 1
⎝ B ⎠ B m
⎡ Hi ⎤
k in radians i q = ⎢1 − ⎥
⎣ V + A f C a cot φ ⎦
m +1
⎡ Hi ⎤
i γ = ⎢1 − ⎥
⎣ V + A f C a cot φ ⎦

Ground factors (base on slope)


β
gc = − − β in rad (for φ=0)
5 . 14

1 − iq
g c = iq − (for φ>0)
0
5 . 14 tan φ
β+η < 90 ; β<φ ; D measured vertically. g q = g γ = (1 . 0 − t an β )
2

For L/B < 2 use φtr


Base factors (tilted base)
For L/B>2 use φps=1.5 φtr -170 but for φtr< 340use φtr= φps
b c = g c (for φ=0)
δ = friction angle between base and soil (0.5φ < δ < φ)
Af = B' L' (effective area) 2β (for φ>0)
bc = 1−
ca = base adhesion (0.6c to 1.0c) 5 . 14 tan φ
bq = bγ = (1 − η tanφ ) η in radians
2

Notes
9 Compute m=mB when Hi=HB (H parallel to B) and m=mL when Hi=HL (H // L). If
you have both HB and HL use m = mB2 + mL2 . Note use of B and L, not B',L'.

9 When φ =0 and βK0, use Nγ = -2sin(± β) in Nγ term

9 Always iq, iγ > 0. For Vesic use B' in the Nγ term even when Hi=HL

71
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Example: Compute the allowable bearing capacity via Terzaghi’s, Meyerhof’s, and
Hansen’s equations for the footing and soil parameters shown in Fig (Ex1). Use a safety
factor of 2.5 to obtain qall.

Solution:
1. Terzaghi:

1
q u = cN c + γ D f N q + Bγ N γ
2
a 2
F or φ = 24 , N = = 11 . 4 ; N = [N − 1 ] cot φ = 23 . 4 ;
2 cos (45 + φ / 2 )
q 2 γ c

⎛ 3π φ ⎞
1 ⎡ K pγ ⎤ ⎜ − ⎟ tan φ
N γ = tan φ ⎢ − 1⎥ = 8 .9 with a = e ⎝ 4 2 ⎠
= 2 . 60
2 ⎣ cos
2
φ ⎦
⎡ ⎛ φ + 33 ⎞ ⎤
N pγ = 3tan 2
φ ⎢ 45 + ⎜ ⎟ ⎥ = 34 . 19
⎣ ⎝ 2 ⎠⎦
1
q u = cN c + γ D f N q + Bγ N γ = 662.2KPa
2
q u
q all = = 264 . 88 KPa
FS
2. Meyerhof’s equation:
1
q u = cN c s c d c ic + γ D f N q s q d q iq + Bγ N γ s γ d γ iγ
2
For φ = 24 : N q = e π tan φ tan 2 (45 + φ / 2 ) = 9 .60 , N c = (N q − 1) cot φ = 19 .32 ,
N γ = (N q − 1) tan( 1 . 4φ ) = 5 .72
⎛ φ⎞
K p = tan 2 ⎜ 45 + ⎟ = 2 .37
⎝ 2⎠
⎛B⎞ ⎛B⎞
Hence, s c = 1 + 0.2K p ⎜ ⎟ = 1.02 s q = s γ = 1 + 0.1K p ⎜ ⎟ = 1.01
⎝L⎠ ⎝L⎠
⎛D⎞ ⎛D⎞
d c = 1 + 0.2 K p ⎜ ⎟ = 1.31 d q = d γ = 1 + 0.1 K p ⎜ ⎟ = 1.15
⎝B⎠ ⎝B⎠
i c = i q = i γ = 1.0
1
Thus : q u = cN c s c d c ic + γ D f N q s q d q iq + Bγ N γ s γ d γ iγ = 678.81Kpa
2
qu
q all = = 271 .52 KPa
FS
3. Hansen’s equation

72
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

1
q u = cN c s c d c i c b c g c + γ D f N q s q d q i q b q g q +
Bγ N γ s γ d γ i γ b γ g γ
2
For φ = 24 : N q = e π tanφ tan 2 (45 + φ / 2 ) = 9.60 ; N c = (N q − 1)cot φ = 19 .32 ;
N γ = 1.5(N q − 1) tan φ = 5.75
⎛ Nq ⎞⎛ B ⎞
⎟⎟⎜ ⎟ = 1.02 s q = 1 + sin φ ⎛⎜ ⎞⎟ = 1.02 s γ = 1 − 0.4⎛⎜ ⎞⎟ = 0.98
B B
Hence, s c = 1 + ⎜⎜
⎝ Nc ⎠⎝ L ⎠ ⎝L⎠ ⎝L⎠
⎛D⎞
as D/B < 1, k = D/B ⇒ d c = 1 + 0.4 k = 1.40 d q = 1 + 2 tan φ (1 − sin φ ) 2 ⎜ ⎟ = 1.31
⎝B⎠
d γ = 1.0; i c = i q = i γ = 1.0; b c = b q = b γ = 1.0; g c = g q = g γ = 1.0
1
q u = cN c s c d c i c b c g c + γ D f N q s q d q i q b q g q + Bγ N γ s γ d γ i γ b γ g γ = 729 .1Kpa
2
qu
q all = = 291 .64 KPa
FS

8. Effect of Water Table on Bearing Capacity


The soil’s unit weight used in the second and third term (the γ in Nq and Nγ terms) of the
bearing capacity equations presented in the preceding sections are the effective unit
weights. When the effective unit weights are used, they results in a decrease in the
ultimate bearing capacity of the soil. The effective unit weight γb (γ') is roughly half γsat;
consequently there will be about 50% reduction in the value of the corresponding term in
the bearing capacity formula. Indeed, a rise in the water table may result in swelling of
some fine-grain soils, possible loss of apparent cohesion, a reduction of the angle of
internal friction and a decrease in the shear strength of the soil. However, the effect of
water table on the shear strength parameters of the foundation soil is usually considered
small and hence ignored. Also the effective stress parameters, c' and φ', obtained from an
appropriate test in the laboratory, on saturated sample of the soil, are to be used when the
soil is submerged under water.
The water table location can be in one of the following cases (Fig 5.4):
1. Water Table Above the base of the Footing
Fig 5.4(a) depicts a case of the water table located between the ground surface and base
of the footing. This condition is not frequently encountered; where practical, designers
will circumvent such conditions by relocating the foundation to higher elevations since,
typically, high water tables create construction problems, require dewatering, lower
bearing capacity, and so on. However, when this condition is encountered, both the 2nd
and 3rd terms of the bearing capacity equations are affected by a lower value of γ [= γb (γ')
or γsub].

73
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Fig 5.4: Effect of Water Table


2. Water at the Base of Footing
For this case, the γ in the second term (Nq) requires no adjusting. The third term will be γb
(Fig 5.4 (b)).
3. The Water Table Below the Base of Footing but with in the wedge zone
When the water table lies with in the wedge zone [depth approximately
H=0.5Btan(45+φ/2) from base of footing], some small difficulty may be obtained in
computing the effective unit weight to use in the Nγ term [Fig 5.4 (c)]. In many cases this
term for such situation can be ignored to obtain a conservative solution. However, one
can compute effective weight (γe) for the soil within the wedge zone as
dw γ'
γ e = (2H − d w ) 2
γ+ 2
(H − d w )2
H H
Where: H=0.5B tan (45+φ/2); dw =depth of water table below base of footing
γ and γ’ (=γ-γw) are wet and submerged unit weight of the soil respectively
4. The Water Table Below the wedge zone
When the water table is below the wedge zone [depth approximately H=0.5Btan(45+φ/2)
from base of footing], the water table effects can be ignored for computing the bearing
capacity.
9. Gross and Net Bearing Capacities
The bearing capacity obtained in the above equations is the gross bearing capacity. The
net bearing capacity (qnet) is obtained by deducting the original overburden pressure or
surcharge pressure from the gross bearing capacity (qult).
1
q u = cN c + γ D fN q + Bγ N γ
2
q net = q u - γ D f
⎡ 1 ⎤
⇒ q net = ⎢ cN c + γ D f N q + Bγ N γ ⎥ − γ D f
⎣ 2 ⎦
1
q net = cN c + γ D f (N q − 1) + Bγ N γ
2
Floating foundation: If a foundation on the soil exerts a pressure of intensity less or equal
to the overburden pressure removed during excavation, there will be no need of checking
bearing capacity and such a foundation is called floating (compensating) foundation. For
example in a clay soil (γ=20KN/m3), a building producing a pressure of 80Kg/m2 with a
basement placed at 5m depth has a floating foundation. (γH=100 KPa >80KPa).
10. Bearing Capacity Based on Tolerable Settlement
As discussed previously, the bearing capacity of a foundation is based on two criteria-the
pressure that might cause shear failure of the foundation soil and the maximum allowable
pressure such that the settlements produced are not more than the tolerable values. The
first criterion has already been discussed in detail. For the second criterion, the tolerable
values of the total and differential settlements which a particular structure, on a particular

74
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

type of foundation in a given soil, can undergo without sustaining any harmful effects are
to be decided up on. These values have already been specified, basing on experience and
judgement.HL\54.doc Once the limiting values of settlement are fixed, the procedure
involves determining that pressure which causes settlements just equal to the limiting
value. This is allowable bearing capacity on the basis of the settlement criterion. It is to
be noted that there is no need to apply a further factor of safety to this pressure, since it
would have been applied even at the stage of fixing up tolerable settlement values.
The smaller pressure of the values obtained from the two criteria is termed the ‘allowable
bearing pressure’, which is used for design of the foundation.
The bearing capacity based on settlement criterion may be determined from the field load
tests or plate load tests, standard penetration tests or from the charts like those prepared
by Terzaghi and Peck, based on extensive investigation.
i. Bearing Capacity From SPT
The SPT is widely used to obtain the bearing capacity of soils directly.
9 Terzaghi and Peck (1967) made the first suggestion
9 Meyerhof come up (1974) with modification
9 Bowels (1995) made another modification which he claims to yield an average of
50% increase in allowable bearing capacity. This equation is given as:
• For Smax = 25mm
q all (KPa) = 20 N 55 K d ; B ≤ 1.2m
2
⎡ B + 0.3 ⎤
q all (KPa) = 1 2.5 N 55 ⎢ ⎥ K d; B ≥ 1.2m
⎣ B ⎦
Df
where K d =1+ ≤ 1 . 33
3B
N 55 = corrected N for E r = 55% → (use averaged value)
• For Smax > 25mm
S(mm)
q all = (q all )25mm
25mm
ii. Bearing Capacity From CPT
9 Meyerhof (1956,1965) suggested for Smax =25mm and sands
qc
q all (KPa) = ; for B ≤ 1.2m (a)
30
2
q c ⎛ B + 0 .3 ⎞
q all (KPa) = ⎜ ⎟ ; for B > 1.2m (b)
50 ⎝ B ⎠
where qc= point resistance in KPa
Meyerhof proposed doubling the result obtained from (b) for mat foundations.
9 Schmertmann (1975) gave for footings on sands

N γ = c with this value of Nγ, φ is determined followed by other factors. Then


q
80
Meyerhof’s bearing capacity equation is employed to determine qult. This approximation
should be applicable for D/B <1.5. qc is averaged over the depth interval from about B/2
above to 1.1B below the footing base.
9 For clays one may use[Schmertmann]:
Strip : q all (KPa) = 200 + 28qc
Square :q all (KPa) = 500 + 34qc q c in KPa

iii. Bearing Capacity From Field Load Tests

75
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

Obviously the most reliable method of obtaining the ultimate bearing capacity and the
settlement characteristics at a site is to perform a load test. The test is also used in the
design of highways and runways. The probable settlement of the soil for a given loading
and at a given depth can also be determined.
Round plate with standard diameter (30cm and 70cm) or square plate of side (0.3m x
0.3m and 60cmx60cm) is loaded in a pit excavated in the ground, at a depth equal to the
roughly estimated depth of the foundation for which the bearing capacity is to be
estimated. The procedure is:
9 Excavate a pit to the depth the test is to be performed. The test pit should be at least 4B
(or 4R) wide as the plate to the depth the foundation is to be placed.
9 A load is applied on the plate by increments (ΔP=qult, estimated/5), and settlements are
recorded from dial gauges (at least 3 in no.) for each load increment. Then plots of
settlement vs. time and settlement vs. applied stress are made.
9 The test is continued until a total settlement reaches 25mm or until the capacity of the
testing apparatus is reached or until the soils fails by shear(plate starts to sink rapidly)
Fig 5.5 presents the essential features of the test and typical plots obtained from the plate
load test.

Fig 5.5: Plate Load test


9 When the load vs. settlement approaches the vertical, one interpolates qult. Sometimes,
however, qult is obtained as that value corresponding to a specified displacement (as,
say, 25mm)
Determination of Bearing Capacity from Plate Load Test
Terzaghi and Peck have suggested the following relation between the settlement of the
plate Sp and the settlement of the footing SF for sandy soils:
2
⎡ b (B + 0.3) ⎤ bP
SP = SF ⎢ P ⎥ for sands and SP = SF for clays
⎢⎣ B(b p + 0.3)⎥⎦ B
where: B= width of footing (least dimension) and bp= width (diameter) of plate
The permissible settlement value, such as 25mm, should be substituted for SF in the
above equations and the SP value will be calculated. Then from the load-settlement curve,
the pressure corresponding to the computed settlement SP, is the required value of the
ultimate bearing capacity, qult, for the footing.
The coefficient of subgrade reaction, ks, can also be estimated as:

76
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering

ks =
Δσ 0.4σ max
ΔS
=
ΔS
(
KN/m3 )
The parameter is also employed in immediate settlement computation.
Limitations of the test
1. Size effects: Since the size of the test plate and the size of the prototype foundation are
very different, the results of a plate load test do not directly reflect the bearing capacity
of the foundation. The bearing capacity of footings in sands varies with the size of
footing; thus, the scale effect gives rather misleading results in this case. However, this
effect is not pronounced in cohesive soils as the bearing capacity is essentially
independent of the size of footing in such soils.
2. Consolidation settlements in cohesive soils, which may take years, cannot be
predicted, as the plate load test is essentially a short-term test. Thus, load tests don’t
have much significance in the determination of qall based on settlement criterion w.r.t
cohesive soils.
3. The load test results reflect the characteristics of the soil located only within a depth of
about 2B of plate. This zone of influence in the case of a prototype footing will be
much larger and unless the soil is essentially homogenous for such a depth and more,
the results could be terribly misleading. Thus it may be misleading if there is weak soil
and ground water with in this influence zone.

11. Bearing Capacity Based on Building Codes(Presumptive Pressure)


Building codes stipulates values of allowable soil pressure to use when designing
foundations. These values are usually based on years of experience, although in some
cases they are simply used from building code of another city. These arbitrary values of
soil pressures (Presumptive pressures) are base on a visual classification.
Table 5.5 indicates representative values of building code pressures. These values are
primarily for illustrative purposes, since it is generally agreed that in all but minor
construction projects some soil exploration should be undertaken. Major drawbacks to
the use of presumptive soil pressures are that they don’t reflect the depth of the footing,
size of footing, location of water table, or potential settlements.
Table 5.5 Presumed Design Bearing Resistances* under Vertical Static Loading (EBCS 7, 1995)
Compactness** Presumed Design
Supporting
Description or Bearing Remarks
Ground Type
Consistency*** Resistance (KPa)
Massively crystalline igneous and
Rocks metamorphic rock (granite, basalt, Hard and sound 5600 These values are
gneiss) based on the
Foliated metamorphic rock (slate, Medium hard and assumptions that
2800 the foundations are
schist) sound
carried down to un
Sedimentary rock (hard shale, Medium hard and weathered rock
2800
siltstone, sandstone, limestone) sound
Weathered or broken-rock (soft
Soft 1400
limestone)
Soft shale Soft 850

77
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C
Addis Ababa University Faculty of Technology Department of Civil Engineering
Decomposed rock to be assessed as
soil
Dense 560
Width of foundation
Gravel, sand and gravel Medium dense 420
(B) not less than 1 m
Loose 280
Non-cohesive
Soils Dense 420 Ground water level
assumed to be depth
Sand Medium dense 280 not less than B below
Loose 140 the base of the
foundation
Hard 280
Silt Stiff 200
Medium stiff 140
Soft 70
Cohesive soils
Hard 420
Stiff 280
Clay Medium stiff 140
Soft 70
Very soft Not Applicable
* The given design bearing values do not include the effect of the depth of embedment of the
foundation
** Compactness: dense: N > 30,
medium dense: N is 10 to 30
loose: N < 10, where N is standard penetration value
*** Consistency: hard: qu > 400 kPa,
stiff: qu = 100 to 200 kPa
medium stiff: qu = 50 to 100 kPa
soft: qu = 25 to 50 kPa, where qu. is unconfined compressive strength

78
Lecture Note (Instructor: Amsalu Gashaye) Academic Year: 2001E.C

S-ar putea să vă placă și