Soil Mechanics: Polytechnic University of The Philippines

Download as pptx, pdf, or txt
Download as pptx, pdf, or txt
You are on page 1of 102

POLYTECHNIC UNIVERSITY OF THE PHILIPPINES

LOPEZ, QUEZON BRANCH

SOIL MECHANICS
CIEN
 CHAPTER 1 WEIGHT - VOLUME RELATIONSHIPS
o 1.2 SOIL MASS AS A THREE-PHASE SYSTEM

 The notations used in the diagram are defined below:


V = total volume of the soil mass
Vs = volume of solid particles in the soil
Vv = volume of voids in the soil
Vw = volume of water present in the voids
Va = volume of air present in the voids
W = total mass of the soil
Ws = mass of the solid particles
Ww = mass of water present in the voids
The mass of air present in the voids is negligible.
Thus, Vv = Va + Vw
and, V = Vs + Vv
or, V = Vs + Va + Vw
o 1.3 BASIC DEFINITIONS
The fundamental physical properties which govern the engineering performance of a soil are defined below:
Void ratio (e) : void ratio of a solid is defined as the ratio of volume of voids to the volume of solids.
Vv
e
Vs
Porosity (n) : defined as the ratio of the volume of voids to the total volume of the soil mass. It is generally expressed as a percentage.

Vv
n x100%
The void ratio of a soil may be greater orV less than 1. However, as the volume of voids is always less than the total volume of a
soil mass, its porosity is always less than 100%
Water Content (w) : defined as the ratio of the mass of water to the mass of solids. It is always expressed as percentage.

Degree of Saturation (s) : defined as the ration


Wwof volume of water to the volume of voids. It is always expressed as a percentage.
w x100%
Ws

The value of s may vary from 0% (for dry soils) to 100% (for fully saturated soils).
Vw
Specific gravity of solids (Gs or G) : defineds asVvthe
x100 % of the mass of a given volume of
ratio solid grains to the mass of an equal volume
of water, measured at same temperature.

Mw = mass of water of volume V


Ms where, Ms = mass of any volume V of solid
G
Mw grains
If this volume V is arbitrarily taken as unity, then in the C.G.S. system M s and Mw become numerically equal to the density of
solid grains (‫ץ‬s) and density of water (‫ץ‬w) respectively. Thus,
s = mass of unit vol of solid/mass of
G
w water
s  G  w

Mass specific gravity (Gm) : defined as the ratio of the mass of a given volume of soil to the mass of an equal volume of water.
M 
Gm  
Mw w
where, ‫ = ץ‬unit weight of the soil mass
Bulk density or unit weight (‫ )ץ‬: defined as the ratio of the total mass of a soil to its total volume. Its unit is gm/cc or t/m3 or kN/m3

W
Unit weight of solids (‫ץ‬s) : defined as the mass ofsoil

Vsolids per unit volume of solids.

Dry density (‫ץ‬d) : defined as the mass of soil solids


s per unit of the total volume of the soil
Ws mass.
Vs

Saturated unit weight (‫ץ‬sat) : When a soil mass is fullyWssaturated, its bulk density is termed as the saturated unit weight of the soil.
d 
V
Submerged density (‫ץ‬sub) : The submerged density of a soil mass is defined as the submerged weight of the soil per unit of its total
volume.
Or,

 1.4 FUNCTIONAL RELATIONSHIPS


The functional relationships have an important role to play in soil mechanics. The most important relationships are established below:
Relation between e and n :

Vv
e V  Vv  Vs Vs  V  Vv but,
By definition, or,
Vs
 e Vv Vv / V Vv / V n  Vv 
  
V  Vv (V  Vv ) / V 1  Vv 1  n  n  V 
V
n
e
1 n

Vv
n Again, by definition,
V
Vv Vv / Vs Vv / Vs e  Vv 
n   
Vs  Vv (Vs  Vv ) / Vs 1  Vv / Vs 1  e  e  Vs 

e
n
1 e
Alternative proof: The same relationships may also be deduced considering the schematic diagram of a soil mass as shown in fig. 1.2 (a) and
(b).
Vv We know
e  Vv  e  Vs
Vs that,

 Vs  1 Let us consider a soil mass having unit volume of solids.


Vv e Vv
Vv  e 1  e n  n Vv  n  V 0r Now, Again, , or
V 1 e V
V  1  n  n, Vs  V  Vv a1soil
 n mass having a total volume V=1, or,
Considering
Vv n
e 
Vs 1  n

Relation between e, G, w, and s :


Ww Vw  w
w  With reference to Fig. 1.1,
Ws Vs  s
s
G or, s  G  w But,
w
Vw  w Vw Vw / Vv Vw / Vv s se
w     
Vs  Gw Vs  G (Vs  Vv )  G G G/e G
Vv / Vs
OR ,
s e  wG

Relation between ‫ץ‬, G, s and e :


The bulk density of a three-phase soil system is given by:
W Ww  Ws Vw  w  Vs  s Vw  w  Vs  Gw Vw  G Vs
       w
V Vv  Vs Vv  Vs Vv  Vs Vv  Vs
Dividing the numerator and denominator by Vv , we get,
Vw / Vv  G  Vs / Vv sG/e
  w   w
1  Vs / Vv 1  1/ e
(s  e  G) / e G  se
 w   w
(1  e) / e 1 e
G  se
OR,   w
1 e

Expression for ‫ץ‬sat :


W Ww  Ws Vw  w  Vs  s
sat    By definition,
V Vv  Vs Vv  Vs

For a saturated soil, Vw=Vv


Vv  w  Vs  Gw Vv  G  Vs
sat    w
Vv  Vs Vv  Vs
(Vv  GVs ) / Vv
  w
(Vv  Vs ) / Vv

1  G  (1 / e) (G  e ) / e Ge
  w   w   w
1 1/ e (1  e) / e 1 e
OR, Ge
sat   w
1 e

Expression for ‫ץ‬d :


Ws Vs  s Vs  Gw By definition,
d   
V Vv  Vs Vv  Vs
G  Vs / Vv G/e
  w   w
(Vv  Vs ) / Vv 1 1/ e

OR, Gw
d 
1 e

Eqns. (1.14) and (1.15) may also be derived from eqn. (1.13) as follows:
For a saturated soil s = 1,
G  1 e Ge
sat   w   w From eqn. (1.13) we get,
1 e 1 e
For a dry soil, s = 0
G  0e Gw From eqn. (1.13) we get,
d   w 
1 e 1 e

Relation between ‫ ץ‬and ‫ץ‬d :


W Ww  Ws
  We know that,
V V
Ww  Ws
OR, V

Ws Ws
AGAIN , d  OR, V
V d
From (i) and (ii) we get,
Ww  Ws Ws

 d

Ww  Ws  Ww 
OR,   Ws
 d  1    d  (1  w)  d
 Ws 


OR, d 
1 w
Relation between ‫ץ‬sub and ‫ץ‬sat : A soil is said to be submerged when it lies below the ground water table. Such a soil is fully saturated. Now,
according to Archimedes’ principle, when an object is submerged in a liquid, it undergoes an apparent reduction in mass, the amount of such
reduction being equal to the mass of the liquid displaced by the object.
Consider a soil mass, having a volume V and mass W, which is fully submerged in water.
Volume of water displaced by the soil = V
 Mass of displaced water = V  w
W  W  V  w  V  sat  V  w
1 Apparent mass of the soil,
 V (sat  w)

The apparent density of submerged density of the soil is given by,


W 1 V (sat  w)
sub  
V V
OR, sub  sat  w

EXAMPLES :
1. A soil sample has a unit weight of 1.9 gm/cc and a water content of 12%. If the specific gravity of solids be 2.65, determine the dry
density, degree of saturation, void ratio and porosity of the soil.
SOLUTION : From the consideration of degree of saturation, a soil sample may be:
 Completely dry (s = 0)
 Fully saturated (s = 1)
 Partially saturated (0 < s < 1)
Unless otherwise mentioned in the problem, a soil sample should always be taken to be partially saturated.
 METHOD 1: Given: ‫ץ‬, w, G Required: ‫ץ‬d, s, e, n

As e and n are mutually dependent on each other, effectively three unknown parameters have to be determined from the given data. Select
the appropriate equations which may serve this purpose.
The value of ‫ץ‬d can be determined from :

d 
1 w
Here, ‫ = ץ‬unit weight of the soil = 1.9 gm/cc
…(i)
W = water content = 12% = 0.12
1 .9
d   1.696 gm / cc
1  0.12
In order to solve, for the other two unknowns, s and e , two equations are required. Evidently, the following equations will serve the
purpose: ,
wG  se se  (0.12)( 2.65)  0.318
or
G  se
AGAIN ,   w
1 e

 2.65  0.318 
OR, 1.9   (1.0)
 1 e 

1  e  1.56 e  0.56 ,
OR, or,
The expression of ‫ץ‬d may also be used.
G  w
d 
1 e
OR, (2.65)(1)
1.696 
1 e

OR , 1.696  1.696e  2.65


0.954
OR, e  0.56
1.696
FROM , 0.318
s  0.568  56.8%
0.56
e 0.56
n   0.36  36%
1  e 1  0.56
Answer: Dry density = 1.696 gm/cc, void ratio = 0.56
Degree of saturation = 56.8%, Porosity = 36%
EXERCISES :
1.A soil sample has a porosity of 35%. The soil is 75% saturated and the specific gravity of solid is 2.68. Determine its void ratio, dry
density, bulk density and moisture content.
2.The mass specific gravity of a soil is 1.95, while the specific gravity of soil solid is 2.7. If the moisture content of the soil be 22%,
determine the following: (I) Void ratio (ii) Porosity (iii) degree of saturation (iv) dry density (v) saturated density

CHAPTER 2 : INDEX PROPERTIES AND SOIL CLASSIFICATION


Various physical and engineering properties with the help of which a soil can be properly identified and classified are called the
INDEX PROPERTIES. Such properties can be broadly divided into the following two categories:
 Soil grain properties: Properties pertaining to individual solid grains and remain unaffected by the state in which a particular soil exists
in nature. The most important soil grain properties are the specific gravity and the particle size distribution.
 Soil aggregate properties: Control the behaviour of the soil in actual field. The most important aggregate properties are:
 For cohesionless soils: the relative density
 For cohesive soils: the consistency, which depends on the moisture content and which can be measured by either the Atterberg limits or
the unconfined compressive strength.
o SPECIFIC GRAVITY
The specific gravity of a soil can be determined by a pycnometer
Fig. 2.1 gives a schematic representation of the process.

W1 = empty weight of pycnometer


W2 = weight of pycnometer and dry soil
W3 = weight of pycnometer, soil and water
W4 = weight of pycnometer filled with water
Now, weight of soil
W 2 W 1
(W 4  W 1)  (W 3  W 2) And, solids
weight= of an unequal volume of
water =
W 4 W1 W 3 W 2
G
W 2 W1
o PARTICLE SIZE DISTRIBUTION:
This is determined in the laboratory by the mechanical analysis, which consists of:
(a.) Dry mechanical analysis or sieve analysis. In this method the example is sieved through a set of sieves of gradually diminishing
opening sizes. The percent finer corresponding to each sieve size is determined and the results are plotted on a semilog graph paper to obtain
the particle size distribution curve. However, this method is applicable only to the coarser fractions of soils and not to the silt and clay
frictions as sieves having open sizes less than 0.075 mm are practically impossible to manufacture.
(b.) Wet mechanical analysis or hydrometer analysis: The percentage of finer fractions (i.e., silt and clay) in a soil can be analyzed
indirectly using a hydrometer. The method is based on Stoke’s law which states that the terminal velocity of a falling sphere in a liquid is
given by
s  w 2
v D
18
Where, ‫ץ‬s and ‫ץ‬w are the unit weights of the sphere and the liquid
respectively
= absolute viscosity of the
 D = diameter of the sphere
liquid
Fig 2.2 shows the sketch of a hydrometer. After immersing the hydrometer in the measuring cylinder containing the soil-water suspension,
readings are taken at 1/2, 1, 2, 4, 8, 15, 30, 60, 120, and 1440 minutes. Let r 1 be the reading of hydrometer at time t. The particle size and
the corresponding value of percent finer are obtained from the ff equations:

1800   Zr
D 
s  w t
AND ,
s V
N   w( r1  Cm  rw) 100%
s  w Ws
Where, D = particle size
in m
‫ץ‬s = unit weight of soil  Gs  w ‫ץ‬w = unit weight of distilled water at the room
solids temperature
t = time interval in sec
 = viscosity of water at room temperature in gm-sec/cm
r1 = reading
2
of hydrometer in suspension at time t
Zr = distance from the surface of suspension to the centre of gravity of hydrometer bulb at time t, which can be determined
from :
1 Vh 
Zr  H 1   h  cm
2 A
Where,
Vh = volume of hydrometer in cc
A = area of cross-section of measuring cylinder in cm2
H1 = distance between the surface of suspension and the neck of bulb in cm
h = length of bulb in cm
The distance H1 may be measured by a scale. However, a better proposition is to determine H1 from the ff equation :
(rd  1)  r1
H1  L
rd
Where, rd = difference between the maximum and minimum calibration marks on the stem of
hydrometer  length of
L = length of calibration ( stem)

In eqn. (2.4)
N = percent finer
V = volume of suspension in cc
Ws = weight of dry soil taken in gm
rw = reading of hydrometer in distilled water at room temperature
Cm = meniscus correction
Ws
N1  N 
W
o PARTICLE SIZE DISTRIBUTION CURVE: Fig 2.3 shows typical particle size distribution curves for various types of soils. Curves
A,B and C represent a uniform soil, a well graded soil and a gap graded soil respectively.
With reference to the particular size distribution curve of a given soil, the following two factors are helpful for defining the gradation
of the soil:
Uniformity Co-efficient: D 60
Cu 
D10

( D 30) 2
Co-efficient of Curvature: Cc 
D10  D 60

Where, D10, D30 and D60 represent the particle sizes in mm, corresponding to 10%, 30%, and 60% finer respectively.
When: Cu < 5, the soil is uniform
Cu = 5 to 15, the soil is medium graded
Cu > 15, the soil is well graded.
For a well graded soil, value of Cc should lie between 1 and 3
o RELATIVE DENSITY: It is a measure of the degree of compactness of a cohesionless soil in the state in which it exists in the field. It is
defined as,
e max  e
RD 
e max  e min
Where, emax = void ratio of the soil in its loosest state
emin = void ratio at the densest state
e = natural void ratio in the field
The relative density of a soil may also be determined from:
d max d  d min
RD  
d d max  d min
Where, = maximum dry density of
d max
the soil
= minimum dry density of
d min = in-situ dry density of
the soil
d
the soil

o ATTERBERG LIMITS: If the water content of a thick soil-water mixture is gradually reduced, the mixture passes from a
liquid state to a plastic state, then to a semi-solid state and finally to a solid state. The water contents corresponding to the transition from
one state to another are called Atterberg limits or consistency limits. These limits are determined by arbitrary but standardised tests.
To classify fine-grained soils on the basis of their consistency limits, the ff indices are used:

Ip  wl  wp Plasticity index,
w n  w p w n  wp
Il   Liquidity index,
Ip w l  wp

w l  w n wl  wn
Ic   Consistency index,
Ip w l  wp
Where, wl, wp, and wn stand for the liquid limit, plastic limit and the natural water content of the soil
Flow index (If) : It is defined as the slope of the w vs. Log10N curve obtained from the liquid limit test
w1  w2
If 
log 10 N 2 / N 1

Where, N1 and N2 are the number of blows corresponding to the water contents w1 and w2
Ip
It  Toughness index,
If

plasticity index
A Activity number,
Percentfin erthan 0.002mm

Soils can be classified according to various indices as ff:


Classification according to Plasticity index
Plasticity index Degree of Plasticity Type of soil
0 Non- plastic Sand
<7 Low plastic Silt
7-17 Medium plastic Silty clay or clayey silt
>17 Highly plastic Clay

Classification according to the liquidity index

Il Consistency
0.0 - 0.25 Stiff
0.25 - 0.50 Medium to
soft
0.50 - 0.75 Soft
0.75 - 1.00 Very soft
Classification according to the activity number
Activity Type of soil
number
<0.75 Inactive
0.75 - 1.25 Normal
>1.25 Active
o DETERMINATION OF SHRINKAGE LIMIT: The shrinkage limit of a soil is defined as the water content below which a
reduction in the water content does not result in a decrease in total volume of the soil.
In order to determine the shrinkage limit, a sample of soil having a high moisture content is filled up in a mould of known volume. The
mould containing the sample is then kept in the oven at 105℃ for 24 hours. After taking it out from the oven, the weight of the dry soil pat is
taken and its volume is measured by the mercury displacement method.
Fig. 2.4(a) and 2.4(c) represents the schematic diagrams of the initial and final states pf the sample while 2.4(b) represents that corresponding
to the shrinkage limit.

The shrinkage limit can be determined by the ff two methods:


 METHOD 1: When G is unknown
Let V0 and Vd be the initial and final volumes of the sample and W0 and Wd be its corresponding weights. Let Ww be the weight of water at
this stage. The shrinkage limit is given by:
Ww
ws 
Wd
At the initial stage, weight of
 W 0  Wd
waterof water evaporated upto shrinkage
Weight
 (V 0  Vd )w
limit
Ww  (W 0  Wd )  (V 0  Vd )w
(W 0  Wd )  (V 0  Vd )w
Ws 
Wd
METHOD 2: When G is known:
Let Vs = volume of
Wd Wd solids
Vs  
s Gw
 Wd 
 Ww  (Vd  Vs )w  Vd  w
 Gw 
Wd
  Vd  w 
G
Vd  w  Wd / G
ws 
Wd
OR,
Vd  w 1
ws  
Wd G
OR , w 1
ws 
o s G ON PARTICLE SIZE:
CLASSIFICATION BASED Soils are classified as clay, silt, sand and gravel on the basis of their particle
sizes. IS: 1498-1970 recommend the ff classification:
Soil type Particle size (mm)
Clay <0.002
Silt 0.002 to 0.075
Sand:  
(I) Fine sand 0.075 to 0.425
(II) Medium sand 0.425 to 2.0
(III) Coarse sand 2.0 to 4.75
Gravel 4.75 to 80
 

o TEXTURAL CLASSIFICATION SYSTEM: Any soil, in its natural state, consists of particles of various sizes. On the basis
of the percentage of particle sizes, and following certain definite principles, broad classification of such mixed soil is possible.

o PLASTICITY CHART: This chart is useful for identifying and classifying fine-grained soils. In this chart the ordinate and
abscissa represent the values of plasticity index
Ip  0.73( wL  and
20) liquid limit respectively. A straight line called A-line, represented by the equation
is drawn and the area under the chart is divided into a number of segments. On the chart any fine-grained soil can be represented by a single
point if its consistency limit is known. The segment in which this point lies determines the name of the soil.

Fig. 2.6 shows a plasticity chart. The meaning of the symbols used in the chart are as follows:
 M : Silty soils
 C : Clayey soils
 O : Organic soils
 L : Low plasticity
 I : Medium or Intermediate plasticity
 H : High plasticity
Main groups of fine-grained soils are: ML, MI, MH → Silty soils CL, CI, CH → Clayey soils
OL, OI, OH → Organic soils
 EXAMPLE : Distilled water was added to 60gm of dry soil to prepare a suspension of 1 liter. What will be the reading of a hydrometer
in the suspension at t=0 sec, if the hydrometer could be immersed at that time? Assume density of water = 1gm/cc and specific gravity
of soilds = 2.70.
Solution: At t=0 sec, the solid grains have not started to settle. The suspension, therefore, is homogeneous, having constant density at any
point in it.
As G=2.70,
 s = 2.70 gm/cc
Total volume of solids in the suspension
60
 22.22cc =
2.70
22.22
Vs   0.0222cc
Volume of solids in unit volume of suspension,
1000

Vw  1  0.0222 Volume
0.9778cc
of water in unit volume of suspension
Ws  (0.0222)( 2.70 )  0.0599
Weight gm in unit volume of suspension
of solids
Ww  (0.9778)(1) Weight
 0.9778
ofgm
water in unit volume of suspension
W  Ws  Ww  0.0599  0.weight
Total 9778  1of.0377
unit gm
volume of suspension
 1.0377 gm / cc  1.038 gm / cc Density of the suspension
Therefore, reading of the hydrometer = 1038
 EXERCISE :
1. The following data were obtained from a specific gravity test performed in the laboratory:
Weight of empty pycnometer = 201.25gm
Weight of pycnometer and dry soil = 298.76 gm
Weight of pycnometer, soil and water = 758.92 gm
Weight of pycnometer full of water = 698.15 gm
Determine the specific gravity of the soil.
CHAPTER 3 Capillary and Permeability
Capillary: the interconnected pore spaces in a soil mass may be assume to form innumerable capillary tubes. At any given site, the natural
ground water table normally exists at a certain depth below the ground level. Due to space tension, water gradually rises from this level
through the capillary tubes. This cause the soil above the ground water table to be partially or even fully saturated.
Pressure Due to Capillary Water: The capillary water rises against gravity and is held by the surface tension. Therefore, the capillarity water
exerts a pressure due to its own self weight, which is always compressive.
Total,  Effective and Neutral Stresses: When an external load is applied on a saturated soil mass, the pressure is immediately transferred to
the pore water. At this point, the soil skeleton does not share any load. But with passage of time, the pore water gradually escapes due to the
pore water pressure induced and a part of the external stresses is transferred to the solid grains. The total stress is therefore divided into the
following components.
 (i) Effective stress or intergranular pressure
 (ii) Pore water pressure or neutral stress
Or
Distribution of Vertical Stresses in Various Soil-water Systems
(i) Free water: In free water, the hydrostatic pressure distribution is linear. At any depth below the water level, the vertical pressure is
given by:

(ii) Dry soil: In a dry soil mass, the distribution of vertical stresses is similar to a hydrostatic pressure distribution. At any depth below the
water level, the vertical pressure is given by,
(iii) Submerged
  soil: Fig. 3.5 shows a soil mass submerged in water with free water standing upto height . If H be the height of the soil, the
total pressure at the bottom of it is given by,

Or

or

Pore water pressure,

Effective stress
Or,
  
(iv) Saturated soil with capillary water: In Fig. 3.6, the soil mass is saturated up to a height above the water level, due to capillary rise of
water. The total stresses, pore water pressures and the effective stresses at various levels are worked out below:
Pore  Pressure in Seepage Water: the shear strength of a soil is governed by the effective stress. When no flow of water takes place through a
soil, the effective stress at a given point remains constant. However, seepage of water causes the effective stress to change, and effects the
stability of any structure built over the soil mass.
The effect of seepage of water on the effective stress can be analyzed with the following laboratory experiment.
Two containers and are interconnected through a U-tube tube. The container contains a soil mass of height with free water standing to a
height above it. The container is filled up with a water and may be raised or lowered as and when required. The water levels in both and are
maintained at constant levels with the help of inlet and outlet pipes.
Case I: When no flow of water take place: This condition occurs when the water levels in both containers are at the same level, as shown in
Fig. 3.7 (a).

At any depth z below the top of the soil mass (i.e., sec. X-X)
  

Thus, at any depth z, the effective stress depends only on the submerged density of the soil.
Case II: Downward Flow: This condition occurs when the water level in is at a higher level than that in (Fig. 3.7 c). At the section X-X

 Where,
i = hydraulic gradient =
A comparison between equations (3.11) and (3.12) clearly shows that a downward flow causes the effective stress to increase.
Case III: upward flow: this condition occurs when the water level in is at higher level than that in (Fig. 3.7b)
  

Thus an upward water flow of water causes the effective stress to decrease
Quicksand Condition: Eqn. (3.13) suggests that the reduction in effective stress at any depth due to upward flow of water depends on the
existing hydraulic gradient, . If at any site, the hydraulic gradient reaches a certain critical value (i.e., ) the seepage pressure may be become
equal to the pressure due to the self-weight of the soil. In such cases, the effective stress will be zero. In other words, a soil grains will not
carry any load any more, the entire load is transmitted to the pore water. The entire soil mass will then behave as if it were a liquid, and any
external load placed on the soil will settle immediately. As this stage the soil loses its shear strength and does not have any bearing power.
Such a condition is known as the quicksand condition. The corresponding hydraulic gradient is called the critical hydraulic gradient.
From eqn. (3.13) we get,

or

or
Darey’s
   Law: This law states that, the velocity of flow of water through soil mass is proportional to the hydraulic gradient.
i.e.,
or,
Where k= constant of proportionally, termed as the coefficient of permeability of soil.
The coefficient of permeability is measure of the resistance of the soil against flow of water through its pores.
From eqn. (3.15) we have, when , then
Thus, the coefficient of permeability of the soil is defined as the average velocity of flow which will occur under unit hydraulic gradient. It
has the unit of velocity, i.e., cm/sec, or, m/day, etc.
TABLE 3.1

Eqn. (3.15) may also be written as

Where,
q = unit discharged, i.e., the quantity of water flowing through a cross sectional area A in unit time. 
Allen Hazen’s Formula: Allen Hazen found experimentally that for loose filter sands,
 Where
 
k = coefficient of permeability in cm/sec
C = a constant, being approximately equal to 100
­= Particle size corresponding to 10% finer, in cm.
 
Laboratory Determination of K: the coefficient of permeability of a soil can be determined in the laboratory using permeameters, which are
of the following two types:
(a) constant head permeameter
(b) Falling head permeameter
The
 test
  arrangements for these two types of permeameters are shown in Fig. 3.8 (a) and (b) respectively.
Constant head permeameter: In this type of permeameters arrangement are made to keep the water levels at the top and bottom of the soil
sample constant. Water flowing through the soil from the top to bottom is collected in a graduated glass cylinder and its volume is
measured.
Let,
Q = quantity of discharge in time
L = length of the sample
h = difference in head of water at top and bottom.
Now, discharged per unit time,
We have from Darey’s law
Here,

or

Falling Head Permeameter: in this case, a stand-pipe containing water is attached to the top of the soil mass. As water percolates through
the soil from top to the bottom, the water level in the standpipe gradually falls down. Instead of measuring the discharge quantity, the fall
of water level in the stand pipe aver a certain time interval is measured.
Let,  
L = the length of the soil sample
A = cross-sectional area of the sample
a = cross-sectional area of the stand-pipe
= head of water causing flow at a time
= head of water causing flow at a time
Let, in any small interval of time , the change in head is given by –dh (the negative side indicates that the head decreases)

Or
Where,
 
The constant head permeameter is suitable for coarse-grained soils while falling head is suitable for coarse-grained soils while the falling
head permeameter is suitable for fine-grained ones.
Find Determination of k: In the field, the coefficient of permeability of a stratified or heterogeneous deposit can be determined by either
pumping-out test. The pumping-out tests for unconfined as well as confined aquifers are described below:
(a) Unconfined aquifer: Fig. 3.9 illustrates a test well fully penetrating an unconfined aquifer. As water is pumped out from the well, water
percolates from all sides into it. When the discharged q equals the rate of percolation, the water level in the well become steady.
Consider a point P on the drawdown curve at a radial distance r from the centre of the well. The hydraulic gradient at this point is given by.

Again, if h be the head of water at P then the rate of radial flow of water through a cylinder of radius r and height h is given by,

or

Integrating between proper limits,


Where,
  
and represent the radial distances of two observation wells
And and represents the height of water levels in them.

Or

 
Alternatively, when observation wells are now used,

 
Where,
a = radius of test well
R = radius of influence
ThE value of R may be determined from

Where,
s = drawdown in the test well, m
k = coefficient of permeability, m/sec
  

(b) Confined aquifer: Fig 3.10 illustrates a test well fully penetrating into a confined aquifer of thickness z,
From Darey’s law,

Or,

Or

Integrating, we get
Permeability
  of Stratified Deposits: Natural soil deposits generally are not homogenous, but consist of a number of layers. The thickness
and coefficient of permeability of the layers may vary to a large extent. In such cases, it is required to compute the equivalent coefficient of
permeability of the entire soil deposit.
Equivalent permeability parallel to the bedding planes: Fig 3.11 shows a stratified soil deposit consisting of n layers let be the thickness of
the layers while be their coefficient of permeability.

The difference in water levels on the left and right hand side of the deposit is h. This head difference causes a horizontal flow of water.
Since at any depth below G.L the head difference is constant and equals h, the hydraulic gradient is the same for each and every layer.
Let be the discharge through the individual layers and q be the total discharge through the entire deposit.

Or

Again, if be the equivalent coefficient of permeability of the entire deposit of thickness z in the direction of flow, then
From
   (i) and (ii)

Equivalent permeability perpendicular to the bedding planes: For flow in vertical direction (Fig. 3.12), the discharge velocities in each
layer must be the same.

Using Darey’s law

Now, total head loss = head loss in layer 1 + head loss in layer 2 + …... + head loss in layer n
But we have,
or head loss
From eqn. (iv)
  

EXAMPLE :
Problem #1
The void of a given soil A is twice that of another soil B, while the effective size of particles of soil A is one third that of soil B. The
height of capillary rise of water in soil A on a certain day is found to be 40 cm. Determine the corresponding height of capillary rise in soil
B.  
Solution: We have,
Let and be the heights of capillary rise in soil A and B respectively. Also, let and the respective void ratios and and be the respective
effective sizes.
From the question,

Now,
CHAPTER 4 : Seepage and Flownets
Equations of continuity:
Laplace’s equations of continuity, as applicable to two-dimensional flow problems, is given below:

Where Kx and Ky are the coefficients of permeability in the x and y directions respectively.
For an isotropic soil, Kx = Ky . therefore,

Equations (4.2) is satisfied by the potential function φ(x,y) and the stream function ψ (x,y). The properties of these functions are as follows:

Properties of a flownet :
A flownet has the following properties:
All flow line and equipotential lines are smooth curves.
A flow line and an equipotential line should intersect each other orthogonally.
No two flow line can intersect each other.
No two equipotential line can intersect each other.
Construction of a flownet:
In order to construct a flownet, the boundary conditions, the location of the two extreme flow lines and the two extreme equipotential lines,
have to be identified first, For example, Fig. 4.1 shows a flownet of sheet-pile wall. Here the boundary conditions are:
 AB is the equipotential line having the maximum peizometric head ( h=h1).
 EF is the equipotential line having the minimum piezometric head (h=h2).
 BCDE (e.i., the surface of the sheet-pile) is the shortest flow line.
 GH (e.i., the imperious boundary) is the longest flow line.
Once the boundary conditions are identified, the flownet can be drawn by trial and error. The process is tedious and each line has to be
draw, erased and redrawn a number of times.
Uses of a flownet:
A flownet enables one to determine the following:
Quantity of seepage: Fig. 4.2 shows a portion of a flownet. Let ∆q2 and ∆q1 be the quantity of seepage in unit time through two
consecutive flow channels. Let b1 and I1 be the width and length respectively of the flow element ABCD and ∆h be the head drop
between two consecutive equipotential lines.
From Darcy’s law we have,
Considering unit thickness of the soil mass, cross-sectional area of the element ABCD =b1 x 1 =b1
Hydraulic gradient,

Similarly ,

The discharge quantity through all flow channel will be equal if,

However, if the elements are made orthogonally squared(e.i., b=1) then,

Again, if H be the initial difference of head and Nd be the number of equal head drops, then,
Hydrostatic pressure: The hydrostatic pressure at any point within the soil mass is given by,

Where: hw= piezometric head at the point underconsiderations. 


In order to find out the piezometric head at a point, locate the flow line on which the given point lies and count the total number of head
drops in the flownet as well as the number of the head drops occurred upto the given point. The piezometric head at the given point is
obtained from,

For example, in Fig. 4.1 the piezometric head at P is,

The uplift pressure at any point below the base of a concrete dam is given by,

Exit gradient: when the percolating water comes out of the soil at the downstream end, it applies a seepage pressure on the soil which is
given by,

And a is the average dimension of the last element of a flow channel. Piping may occur if the exit gradient becomes greater than the
critical hydraulic gradient. The factor of safety against piping is given by,
In order to determine the maximum exit gradient for a given flow problem, the last element of the flow channel adjacent to the structure is
to be considered, as the value of ‘a’ is the minimum for that particular element. In Fig. 4.1 this flow element is marked by hatched lines.
Flownet in anisotropic soils:
From equations 4.1

In order to draw the flownet for anisotropic soils, a transformed sections has to be drawn fisrt by multiplying all horizontal dimensions by
but keeping the vertical dimensions unaltered. An orthogonally squared flownet is then drawn as usual for the transformed sections.
The structure, along with the flownet, is then retransformed by multiplying all horizontal dimensions by The final flownet will
consist of rectangular elements.
Multiple permeability conditions:
When the flow lines pass from one soil to another having a different permeability, they deviate from the interface of the two soils and this
deviation is similar to the refraction of light rays. This is illustrated in Fig. 4.3
The portion of the flownet lying in layer 1 is first drawn in the usual manner with square flow elements. When the flow lines as well as
equipotential lines enter layer 2, they undergo deviations according to the following equations:

Consequently, the flow elements in layer 2 are not squared any more, but become rectangle, and their width-to-height ratios are given by,

These conditions are illustrated in Fig 4.4 (a) and (b) respectively.
Unconfined flow: phreatic line:
When an impermeable structure (e.g., a sheet pile or concrete weir) retains water, all boundary conditions are known. Such a flow is known
as the confined flow or pressure flow. However, when the structure itself is previous (e.g., an earth dam) the upper boundary or the
uppermost flow line is unknown. Such a flow is termed as unconfined flow or a gravity flow. And this upper boundary is called phreatic
line.
In order to obtain the phreatic line, the basic parabola has to be drawn first then the necessary corrections at the entry and exit points have
to be made.
Construction of the basic parabola:
In Fig. 4.5, ABCD is the cross-sectional of an earth dam. In order to draw the basic parabola, proceed as follows:
 Measure the horizontal projection L of the wetted portion, ED of the upstream face.
 Locate the point P such that EP = 0.3L. The point P is first point o the basic parabola.
 With P as centre and PC radius draw an arc to intersect the extended water surface at F.
 From F draw FG perpendicular DC. The line FG is the diretrix of the basic parabola, while C is the focus.
 Locate the mid-point Q of CG.
 Let G be the origin, GF the Y-axis and GD the X-axis.
 Choose any point H on CD, such that GH = x1. With C as centre and X1 radius, draw an arc to intersect the vertical line through H at
R. The point (x1,y1) is another point on the basic parabola.
 In similar manner, locate several other points viz.,(x2,y2),(x3,y3),…., etc. join these points with a smooth curve to get the basic
parabola.
 Corrections at entry and exit points:
ED is an equipotential line and phreatic line is a flow line. These two should meet each other at right angles. This necessitates the
corrections at the entry point, which should be drawn by hand.
The phreatic line should meet the downstream face BC tangentially. This necessitates the corrections at the exit point. The basic parabola
intersects BC at M. But the phreatic line should meet BC at N. Let CN= a and NM= ∆a, The magnitude of ∆a/(a+∆a) depends on the slope
angle β of the downstream face. Its value may be obtained from Fig 4.6.

After locating N, the necessary correction should be made by hand. In Fig 4.6, the value of the slope angle β ranges from 30° to 180°. It
should be noted that for ordinary earth dams, β< 90°. However, if the dam is provided with a toe drain or a horizontal filter to arrest the
seepage water the value of β may be as high as 180°. This illustrated in Fig. 4.7 below:
It should also be noted that for an earth dam having a toe drain, chimney drain or horizontal drainage blanket, the beginning point of that
drain filter, and not the bottom corner of downstream face, should be taken into account while plotting the basic parabola.
EXAMPLE :
Problem 1: A homogeneous earth dam, 30m high, has a free board of 1.5m. flownet was constructed and the following results were noted:
No. of potential drops = 12
No. of flow channels = 3
The dam has a 18m long horizontal filter at its downstream end. Calculate the seepage loss across the dam per day if the width of the dam
be 200m and the coefficient of permeability of the soil be 3.33x10¯⁴ cm/sec.
Solution:
Total
  quantity of seepage loss per day across the entire width of the dam
=(2.185)(200) = 437mᶟ
CHAPTER 6
CONSOLIDATION- is essentially a time-dependent process. In coarse grained soils having a high co-efficient of permeability the pore water escapes very rapidly.
The time-dependent volume change of the soil mass, therefore, occurs only in less permeable fine-grained soils like clay and silts.
Definitions: the following terms are frequently used to express the compressibility characteristics of soils:
1. CO-EFFICIENT OF COMPRESSIBILITY (av)
It is defined as the change in void ratio per unit change in pressure

Where: eo and e are the void ratios of a soil under the initial and final vertical stresses 0 and respectively.
2. CO-EFFICIENT OF VOLUME CHANGE OR VOLUME COMPRESSIBBILITY (mv)
It is defined as the change in volume of a soil mass per unit of its original volume due to unit change in pressure.

Fig. 6.1 shows a soil mass having an initial void ratio eo. If the volume of solids be unity, then volume of voids is given by,
Vv = e0 x Vs = e0 x 1 =e0
V0 = Vv +Vs
= 1 + e0
If the void ratio now decreases to e due to increase in pressure, then
V1 = 1 + e
Or
 change
  in volume
V = V0 – V1
= 1 + e0 – ( 1+e)
= e0 – e
=e

From eqn 6.2 we get,

3. COMPRESSION INDEX (Cc)


It is defined as the gradient of the virgin compression curve drawn from the results of a consolidation test performed on a soil.
By definition,
Cc = gradient of AB
= tan
Consolidation
  

The value of the compression index may also be determined from the following empirical formulae:
For normally consolidated clays ( sensitivity 4),
Cc = 0.009 (w1 - 10)
Cc = 0.007 (w1 – 10)
Where: w1= liquid limit %

TERZAGHI’S THEORY OF ONE-DIMENSIONAL CONSOLIDATION:


The process of consolidation is closely related to the expulsion of pore water and dissipation of pore pressure. Terzaghi, in his theory
of one-dimensional consolidation, investigated the relationship between the rate of change of excess pore pressure and the degree of
consolidation, and deduced the following differential equation:
  

Where: u – stand for the excess pore pressure at a depth z


T – stands for the time elapsed after the application of the load.
Cv = co-efficient of consolidation, which is defined as:

Where :
k- co-efficient of permeability, cm/sec
w- unit weight of water, gm/cc
mv- coefficient of volume change, cm2/gm
the unit of Cv is cm2/sec
equation 6.6 is a second order differential equation, the solution of which may be obtained in the form,
U = f (Tv)
Where, U – degree of consolidation
Tv- time factor = Cv x t / h2
t- time required for U% consolidation, sec
h- maximum length of drainage path, cm.
in case of double drainage condition ( i.e., when a clay layer of thickness H lies between two permeable layers at top and bottom) the maximum length
of drainage path h = H/2, whereas in case of single-drainage condition (i.e., when the clay lies between a permeable and an impermeable layer), h =H.
the time
  factor Tv is a dimensionless quantity, the value of which depends on the degree of consolidation taken place at a given time, and
not on the properties of the soil. Terzaghi suggested the following equations for the determination of Tv:

6.4 LABORATORY CONSOLIDATION TEST:


In order to determine the compressibility characteristics of a clay deposit, laboratory consolidation tests are to be performed on
representative samples of the clay collected from the site. A knowledge of such characteristics is required for:
Estimating the probable consolidation settlement of a proposed structure to be constructed on this soil.
To determine the time-rate of settlement.
The sample is placed in an oedometer between two porous stones and arrangements are made to keep the sample saturated throughout the
test. The loading intensities are generally applied in the following order: 0.25, 0.5, 1.0, 2.0, 4.0, 8.0, and 16.0 kg/cm 2.
The vertical deformations of the sample under each loading intensity are measured with the help of a dial gauge. The readings are taken at
elapsed times of: 0.25, 0.5, 1, 2, 4, 8, 15, 30, 60, 120, 240, and 1440 minutes.
From the results of the test, the following three curves are drawn:
1. e vs log10p curve, to determine the value of Cc
2. Dial reading vs. log10t curve
3. Dial reading vs. t curve
In order to plot the e vs. log10p curve, the void ratio of the sample at the end of each load increment has to be determined from the
corresponding dial reading. This can be done by either of the following methods.
a. HEIGHT OF SOLIDS METHOD: after the completion of the test, the sample is taken out from the oedometer, dried in oven and its dry
weight Wd is determined.
Now, volume of solids,
  
And height of solids in the sample,

Where, A = cross-sectional area of the sample.


Let e be the void ratio corresponding to a height h of the sample.

Thus if the value of h is known at any time during the test, the corresponding void ratio can be determined. The value of h may be obtained
from:
H=H–RxC
Where: H = initial height of sample.
R = dial reading
C = dial gauge constant.
b. CHANGE IN VOID RATIO METHOD:
V1 = initial volume of the sample
V2 = volume of the sample at the end of compression under a loading intensity p
V = change in volume = V1-V2
  

DETERMINATION OF Cv : for a given soil, the value of Cv is not constant but depends on the magnitude of the applied stress. In order to
determine the degree of consolidation of clay layer under an external load, it is required to determine the initial and final pressures ( z and z
+ z respectively) on the soil. If, for example, the initial and final pressure before and after the application of external load be 1kg/cm 2, then
the value of Cv must be obtained from this particular range of loading in the consolidation test.
The value of Cv may be determined from either of the following methods:
a. square root of time fitting method:
B. Logarithm of time fitting method:

COMPUTATION OF SETTLEMENT: the total settlement, S, of a footing is given by,


S = S i + S c + Ss
Where, Si = immediate settlement
Sc = primary consolidation settlement
Ss = secondary consolidation settlement
The secondary consolidation settlement is of importance only in case of highly organic and peats.
IMMEDIATE SETTLEMENT: the immediate settlement due to vertical concentrated load Q at a depth z and radial distance r is given
by,

The immediate settlement due to a uniformly loaded area is given by,

Where,
q = inetensity of contact pressure
B
 = least
  lateral dimension of loaded area
= poisson’s ratio of soil
E = modulus of elasticity of soil
If = influence factor, the value of which depends on:
Type of footing (i.e., whether it is rigid or flexible)
Shape of the footing
The location of the point below which settlement is required (I.e., the centre, corner or any other point of the footing)
Length to breadth ratio of the footing

EXAMPLES:
A normally consolidated clay stratum of 3m thickness has two permeable layers at its top and bottom. The liquid limit and the initial void
ratio of the clay are 36% and 0.82 respectively, while the initial overburden pressure at the middle of clay layer is 2kg/cm 2. Due to the
construction of a new building this pressure increases by 1.5 kg/cm 2. Compute the probable consolidation settlement of the building.
  

A 3m thick layer of silty clay is sandwitched between two layers of dense sand. The effective overburden pressure at the centre of the silty
clay layer is 2kg/cm2. However, due to the construction of a raft foundation, this pressure increases to 4 kg/cm 2. Laboratory consolidation
test was performed on a 2.5 cm thick sample of the silty clay. Under applied stresses of 2 kg/cm 2 and 4 kg/cm2 the compressions of the
sample were found to be 0.26 cm and 0.38 cm respectively. Compute the probable consolidation settlement of the raft.
SOLUTIOI:
S c = mv x H 0 x p
Where;
H0 = initial thickness
= 2.5 cm for the soil sample and 300 cm for the soil in situ
= change in effective pressure
= 4-2 = 2 kg/cm2
Mv = co-efficient of volume change for the pressure range of 2kg/cm 2 to 4kg/cm2
For the laboratory test:
Initial thickness of the sample = 2.5 cm
Thickness under a pressure of 2 kg/cm2 = 2.5-0.26 = 2.24 cm
Thickness under a pressure of 4 kg/cm2 = 2.5-0.38 = 2.12 cm
Change in thickness when the pressure increases from 2 kg/cm2 to 4 kg/cm2 = 2.24-2.12= 0.12 cm
The consolidation settlement of the silty clay layer
Sc = (0.024)(300)(2)
14.4 cm
CHAPTER 7 COMPACTION
Moisture-density Relationships
While compacting a soil in the field, it is always desirable to compact the soil in such a way that its dry density is maximum. If a
given soil is compacted under a specified compactive effort, its dry density will be the maximum at a certain moisture content, known as
the optimum moisture content. Hence, before compacting a soil in the field, its optimum moisture content and the corresponding dry
density must be determined in the laboratory. The test employed for this purpose is called Standard Proctor Test.
Standard Proctor Test
In this test, samples of the given soil are prepared at various moisture contents and are compacted in a cylindrical mould, 127.3 mm
high and having an internal diameter of 100 mm. The sample is compacted in three layers of equal height, each layer being subjected to 25
blows of a compaction rammer having a self-weight of.2600 gm and a height of free fall of 310 mm.
Samples
  are compacted in the mould at increasing moisture contents. After each test, weight of the sample compacted is determined and its
bulk and dry densities are computed. A curve is then plotted to show the variation of dry density with moisture content (Fig 7.1). The curve
is usually parabolic in shape. Initially the dry density increases with increasing moisture content, until a certain peak value is reached.
Further increase in moisture content results in a decrease in the dry density. The moisture content represented by the peak of the curve is the
optimum moisture content (OMC) and the corresponding dry density is the maximum dry density of the soil under that particular
compactive effort.

For heavier field compaction, the moisture-density relationship can be investigated by the modified AASHO test. The test procedure is
similar to that of Proctor test except that a heavier rammer (weight = 4900 gm, free fall = 450 mm) is used and the soil is compacted in 5
layers.
Under heavier compaction, the moisture-density curve (Fig 7.1.) is shifted upwards and simultaneously moves to the left, resulting in a
lower OMC but a greater .
Zero Air Voids Line: Compaction is achieved by the expulsion of air from the voids. However, as the external load acts for a very short
time, it is nearly impossible to drive out all the air from the voids. Thus, during compaction, a soil is not fully saturated. If the remaining air
could be driven out, its void ratio would have been reduced and consequently, its dry density would have increased. The zero air voids line
(Fig 7.2) is a theoretical curve which represents the relationship between water content and dry density of the soil when it is 100%
saturated.
  

At any given moisture content, the dry density of a soil in the fully saturated condition can be derived as follows:
We have
AND
For a fully saturated soil, s = 1, ∴

From eqn. (7.1) it is evident that, for a given soil, an increase in moisture content will always result in a decrease in . Hence the zero air
voids line is always a steadily descending line.
California Bearing Ratio (CBR)
The California bearing ratio test is of immense importance in the field of highway engineering. The CBR value of a soil or a paving material
is a measure of its strength against probable rutting failure due to moving wheel loads. The California bearing ratio is defined as the ratio of
the force per unit area required to drive a cylindrical plunger of 50 mm diameter at the rate of 1.25 mm/min into a soil mass to that required
to drive the same plunger at the same rate into a standard sample of crushed stone.
THUS , (
 The test is performed by first compacting the given soil in the AASHO mould at the specified compactive effort as stated in Art.7.3. The sample
is compacted upto a height of. 127 mm at the particular moisture content and density at which the CBR value is required. The plunger is then
driven into the soil under a steadily increasing static load. The settlement of the plunger is measured with the help of a dial gauge while the
corresponding load is obtained from the proving ring. From the results a load-settlement curve is plotted and the test loads for 25 mm and 5.0
mm penetration are determined. The values of unit standard loads corresponding to these two penetrations are 70 kg/cm² and 105 kg/cm²
respectively. Therefore, the CBR-values at 2.5 mm and 5.0 mm penetrations can be determined.
 Generally, the CBR value at 2.5 mm penetration should be greater than that at 5.0 mm penetration. In that case, the former value is accepted as
the CBR value for design purposes. If the CBR value corresponding to 5 mm penetration exceeds that for 2.5 mm penetration, the test should be
repeated, However, if identical results are obtained once again, the CBR value for 5 mm penetration should be used.
Correction to the curve

The load-penetration curve should always be convex upwards (curve A in Fig.7.3). However, due to surface irregularities, the initial portion of the
curve is sometimes concave upwards (curve B in Fig.7.3). The curve then must be corrected in the following manner:
(i) The straight portion of curve I is projected backwards to meet the X-axis at O'.
(ii) The origin O is shifted to O'.
(iii) Subsequently, all penetrations are measured from the new origin O'. 
Thus, the points corresponding to 2.5 mm and 5.0 mm penetration should be shifted towards the right by an amount equal to the shift of origin.
In order to simulate the worst possible field conditions, the CBR test is sometimes performed on soaked samples. After compacting the sample in the
mould, the sample is kept submerged in water for a period of 4 days, after which the sample becomes almost saturated. The CBR test is then
performed on this soaked sample. 
EXAMPLE
   :
Problem . The optimum moisture content of a soil is 16.5% and its maximum dry density is 1.57 gm/cc. The specific gravity of solids is 2.65.
Determine:
(i) the degree of saturation and percentage of air voids of the soil at OMC.
(ii) the theoretical dry density at OMC corresponding to zero air voids. 
Solution: (i) When the soil is at OMC, it has a moisturc content of 16.57% and a dry density of 1.57 gm/cc.
Now we have,

OR
OR
AGAIN , s = wG, or, s =
∴s=
Hence,the required degree of saturation is 63.5% and the percentage of air void is (100 - 63.5)% = 36.5%
(ii) At zero air void the soil is fully saturated i.,e .,s = 1.
= = 0.437
γd = = 1.844 gm/cc
 
CHAPTER
   8 SHEAR STRENGTH
According to Coulomb's law, the shear strength,

Where,

The factors and re called the shear parameters of a soil.


When expressed graphically, eqn. (8.1) can be represented by a straight line called the failure envelope; The general form of failure envelope
for a cohesionless, a cohesive and a - soil are shown in Fig. 8.1 (a), (b) and (c) respectively.

The shear parameters of any soil depend not only on the nature of the soil but also on such factors like moisture content and loading
conditions. At very low moisture content a cohesive soil may develop a certain amount of internal friction. Likewise, at high moisture
contents a cohesionless soil may show the signs of having an apparent cohesion.
 Mohr's
  circle of stress:
This is a graphical representation of the stress conditions in a soil mass which enables one to find out the stresses developed on any plane
within the soil due to an external loading system.
In a stressed material, a plane which is subjected to only a normal stress, but no shear stress, is called a principal plane. Through any point
in the material, two such planes exist. These planes are called the major and the minor principal planes and are orthogonal to each other. If
the principal stresses and are known, the normal stress and shear stress on a plane inclined at an angle to the major principal plane is
given by,

Pole:
The concept of the pole, or the origin of the planes, is very useful in such problems where the locations of the principal planes are not
known. Consider the soil element subjected to a system of external stresses as shown in Fig. 8.3. It is required to determine the normal and
shear stresses acting on the plane AA, inclined at an angle o the horizontal.
Considering the free body diagram of the element it can be proved that the element can be in equilibrium only if,
.
Thus, the pole may be defined as a particular point on the Mohr's circle such that, if a line is drawn from this point making it parallel to any
given plane within the soil mass, then, the co-ordinates of the point of intersection of this line with the circle will represent the stresses
acting on that plane
  

Sign convention:
The following sign conventions are normally followed for plotting the stress co-ordinates :
Normal stress: Compressive stresses are taken as positive and tensile stresses as negative.
Location of the failure plane :
Fig. 8.4 represents a soil sample subjected toa major principal stress and a minor principal stress . s the sample is on the verge of failure, the
Mohr circle has touched the failure envelope at P. Evidently, the pole of the Mohr circle is at A. The highest point on the circumference of
the Mohr circle is the crown R. The line AR is inclined to the -axis at 45°. The corresponding plane in the soil is MN, which is the plane
subjected to the maximum shear stress However, the potential failure plane in the soil is not MN, but the plane represented by the point P,
because the stress co-ordinates given by P are such that coulomb's equation is satisfied as the point P lies on the failure envelope. In order to
determine the location of this plane, join PA and PC.
Now,
As AC = PC,

Again, since ,
In
 ,  
Or
Or ,
In Fig. 8.4, the plane BB, drawn at (45+ ) to the major principal plane, represents the failure plane.
It can be proved that, at failure the relationship between the two principal stresses is given by,
)
Or ,
Where ,

Determination of Shear Strength:


The following tests are employed for the evaluation of the shear strength of a soil :
A. Laboratory tests :
 Direct Shear Test
 Triaxial Compression Test
 Unconfined Compression Test.
B. Field Test :
1. Vane Shear Test
For a detailed description of the test procedures, the reader is referred to any standard textbook of Soil Mechanics. Only the essential points regarding the
computation of shear strength will be highlighted here.
 
Direct
  Shear Test:
In this test, soil samples compacted at known densities and moisture contents in a shear box of 6 cm x 6 cm size, which can be split
into two halves, is sheared by applying a gradually increasing lateral load. Three identical samples of a soil are tested under different
vertical compressive stresses and the corresponding shear stresses at failure are determined. A graph is then plotted between normal stress
and shear stress. Results of each test are represented by a single point. Three points obtained from the three tests are joined by a straight
line which is the failure envelope for the given soil. The slope of this line gives the angle of internal friction, while the intercept from the
-axis gives the value of cohesion of the soil.
Triaxial Compression Test:
In this test, cylindrical soil specimens of 3.8 cm diameter and 7.6 cm height, enclosed in an impermeable rubber membrane, are
replaced inside the triaxial cell. An all-round cell pressure, is applied on the sample. Simultaneously, a gradually increasing vertical stress
is applied until either the sample fails, or its axial stain exceeds 20%. Stress vs strain curves are plotted to determine the normal stress at
failure. This stress is called the deviator stress, . The major principal stress , is obtained from the following relation (refer Fig. 8.5) :
Three  samples of a soil are tested under different cell pressures. From the results, three Mohr circles are constructed, and a common tangent
is drawn to them. This is the failure envelope.
The normal stress at any point during the test is determined by dividing the normal load obtained from the reading of the proving ring by the
cross-sectional area of the sample. Due to the bulging of the sample during shear, the cross-sectional area should be modified using the
following equation :

Where, = coffected area


= initial area
= axial strain =
Where , = axial compression
= initial length
In the drained triaxial tests, the volume of the sample may change during the test due to expulsion or absorption of water. In that case, the
corrected area should be determined from :

Where ,

Unconfined Compression Test : This is a special case of triaxial test in which = 0


We have
  eqn. = (45° + /2) + 2 tan(45°+ )
As , = 0 , for an unconfined compression test
= 2c tan(45° + /2)
A number of tests on identical specimens will give the same value of . Thus, only one equation is available while two unknowns, viz., c and
, are involved. Hence, eqn. (8.11) cannot be solved without having a prior knowledge of any one of the unknowns. Due to this reason, the
unconfined compression test is employed to determine the shear parameters of purely cohesive soils only. For such soils, S = and hence, =
2 c tan 45° = 2c
The vertical stress at failure, known as the unconfined compressive strength and denoted by is obtained by dividing the normal load at
failure by the corrected area, as given by eqn.
Thus,
Or,
Vane Shear Test.
This is a field test used for the direct determination of the shear strength of a soil. Generally this test is conducted in soft clay situated
at a great depth – samples of which are difficult to obtain.
The apparatus consists of four metal blades, called vanes, mounted on a steel rod, as shown in Fig. 8.6. The device is
pushed slowly up to the desired depth and is rotated at a uniform speed by applying a torque through the torque rod. The amount of torque
applied is recorded on a dial fitted to the rod. Failure occurs when the vane can be rotated without any further increase in the torque.
For a cohesive soil, = 0. Hence coulomb's equation reduces to :
s=c
   Thus, for a cohesive soil, the shear strength is equal to its cohesion. In a vane shear test, the cohesion, and hence the shear strength can
be determined from:

Where,

Sensitivity:
When the shear stresses developed in a soil exceeds its shear strength, the soil fails by shear and loses its strength. However, if the soil
is left in that state for some time, it regains some of its original strength. The sensitivity of a soil is a measure of its capability of regaining
strength after a disturbance has been caused in the soil. It is expressed as,
(8.14)
On the basis of the sensitivity, clayey soils are divided in the following categories:

Sensitivity Nature of Clay


1 Insensitive
1-2 Low Sensitive
2-4 Medium
Sensitive
4-8 Sensitive
8-16 Extra Sensitive
>16 Quick Clay
EXAMPLE :
Problem 2. Stresses acting on a soil element are shown in Fig. 8.8 (a).

(i) Determine the magnitude and direction of the principal stresses.


(ii) Find out the stresses acting on the plane XX.
(iii) If the soil has a cohesion of 5 kN/m² and an angle of internal friction of 25°, find out whether a shear failure is likely to occur along the plane
XX.
Solution:
The graphical solution of the problem is presented in Fig. 8.8
(b). The procedure is as follows:
1. Two orthogonal co-ordinate axes and an appropriate vector scale (1 cm = 5 kN/m²) are chosen.
2. The points M (20, -10) and N (40, 10) are chosen to represent the stresses on the planes AB and BC respectively.
3. M and N are joined and the mid-point O of MN is located.
4. With O as centre and MN as diameter, the Mohr circle is drawn.
5. The point M represents the stresses on the plane AB. From M, a straight line MP is drawn parallel to AB, to intersect the circle at P. P is the pole.
6. From P, PQ
 
  

(i) The points of intersection, R and S between the circle and the -axis give the principal stresses Here,
= 48 kN/m²
and = 16.2kN/m²
In order to locate the directions of the principal planes, the points R and S are joined to the pole P. Through any point Z in the soil element, Z – 1
PS and Z - 3 PR are drawn.
The planes Z - 1 and Z - 3 give the directions of the major and minor principal planes respectively.
(ii) The stresses on XX are given by the co-ordinates of Q. From the figure we obtain.
= 16.5 kN/m² and = 3.6kN/m²
(iii) The normal stress on XX is 16.6 kN/m². From coulomb's equation, the shear strength of a soil is given by,
s = c + tan
Here, c = 5 kN/m², = 16.6 kN/m²,
s = 5 + (16.6)(tan25°)
= 12.74kN/m° > 3.6 kN/m²
As < s, failure along XX is not possible.
CHAPTER 9 – EARTH PRESSURE
Depending on the conditions prevailing at the site, the lateral earth pressure may be divided into the following three categories:
(i) Earth pressure at rest.
(ii) Active earth pressure.
(iii) Passive earth pressure.
Earth Pressure at Rest: Fig. 9.1 (a) shows a retaining wall, embedded below the ground level up to a depth D, and retaining earth up to a
height H. If the wall is perfectly rigid, no lateral movement of the wall can occur. And hence, no deformation of the soil can take place. The
lateral pressure exerted by the soil is then called the earth pressure at rest

The conjugate relationship between the lateral earth pressure and he vertical overburden pressure is given by:
σh = K0 ∙ σv, or σh = K0 ∙ γ z
where:
K₀ = co-efficient of earth pressure at rest.
γ = unit weight of soil
z = depth at which lateral pressure is measured.
The
 value
  of K₀ depends on the properties of the soil and its stress history, and is given by:
K₀ =
where, µ= Poisson's ratio of the soil.
9.3. Active and Passive Earth Pressures: In reality, a retaining wall is not rigid, but flexible, i.e., it is free to rotate about its base. In Fig. 9.1
(a), let P₀ and P₀’ be the at-rest lateral thrusts acting on the back and front faces of the wall respectively. Due to the difference in elevation
levels, P₀ ˃ P₀’. Hence, a flexible wall will yield away from the backfill. The soil wedge ABC will then tend to slide down along the potential
sliding surface BC. This condition is illustrated in Fig.9.1(b). The frictional resistance F R against such movement will act upward along BC. Its
horizontal component FH will act in the opposite direction to that of P0. Thus the net pressure on the wall will decrease. Such a state is called
the active state of plastic equilibrium and the lateral pressure is called the active earth pressure.
The equation governing the relationship between the major and minor principal stresses, acting on a soil element, is given by,
σ 1 = σ 3 Nϕ +
where, Nϕ = (45° + ϕ/2)
ϕ = angle of internal friction
c = cohesion.
Let us consider an infinitesimally small soil element at a depth Z below the ground level, adjacent to a retaining wall, as shown in Fig.
9.2.
σ
 v = vertical
  overburden pressure on the element.
σh = lateral earth pressure on the element.
According to the fourth assumption stated above, a conjugate relationship exists between σ v and σh. The relationship is similar to the
one expressed by eqn. (9.3). However, the exact form of the equation depends on the prevailing conditions, i.e., whether the backfill is in an
active state or in a passive state.
(i)Active state:
In this case, σ1 = σv and σ3 = σh.
But, σv = γz
and, σh = active pressure intensity = Pɑ
Eqn. (9.3) gives,
γ z = Pɑ ∙ Nϕ
Pɑ = -
(ii) Passive state:
Here, σ1 = σh and σ3 = σv
But, σv = γ z
and, σh = active pressure intensity = Pp
Eqn. (9.3) gives,
Pp = γ z N ϕ
   Computation of Earth Pressure Using Rankine's Theory:
(A) Active Earth Pressure:
(a) Cohesionless soils:
For a cohesionless soil, c = 0.
Eqn. (9.4) reduces to
Pɑ = = =
or, Pɑ = Kɑ γz
Eqn. (9.6) and (9.7) can be used to compute the active earth pressure for various backfill conditions, as discussed below:
Dry or Moist Backfill with Horizontal Ground Surface:
Fig. 9.3 (a) shows a retaining wall supporting a homogeneous backfill of dry or moist soil, upto a height H.
At any depth z below the top of the wall.
Pɑ = Kɑ γz
At the top of the wall (z = 0), Pɑ = 0
At the base of the wall (z = H), Pɑ = Kɑ Γh
Fig.
 9.3  (b) shows the distribution of active Pressure intensity. The magnitude of resultant thrust per unit length of wall may be obtained
by multiplying the average pressure intensity by the height of the wall. Average pressure intensity, P ɑv = = Kɑ γH
Resultant thrust, PA = Kɑ γH ∙ H = Kɑ γH²
It is evident from eqn. (9.8) that the resultant thrust is given by the area of the pressure distribution diagram. This thrust acts through
the centroid of the triangle ABC, i.e., is applied at a height of H/3 above the base of the wall.
Fully Submerged Backfill:
This condition is shown in Fig. 9.4 (a). As the soil is fully submerged, its effective unit weight is,
γ' = γsat – γw
At any depth z below the top of the wall, the total active pressure is the sum of pressures exerted by the soil and water. According to
Pascal's law, a fluid exerts equal pressure in all directions at any given depth.
Hence, at a depth z,
Pɑ = Kɑ γ’z + γɑ z
The corresponding pressure distribution diagram is shown in Fig. 9.4 (b)
(iii) Partially Submerge Backfill:
Backfill having similar properties above and below water table:
   In Fig. 9.5 (a), the retaining wall has to retain earth upto a height H. The ground water table is located at a depth H 1 below ground level.
The active pressure intensities are given by:
Above ground water table: Pɑ = Kɑ γz (0 ≤ z ≤ h1)
Below ground water table: Pɑ = Kɑ γ h1 + Kɑ γ’z + γwz (0 ≤ z ≤ h2) z being measured from G.W.T.)
Fig. 9.5 (b) shows the corresponding pressure distribution diagram. The resultant active thrust per unit run of the wall is given by the
entire area of this diagram. It is easier to determine the area by dividing it into a number of triangle and rectangles. In Fig. 9.5 (b).
P1 = Δ ABC, P2 = Δ area of BCED
P3 = Δ DEF, P4 = Δ DFG.
Resultant active thrust,

The point of application of PA can be determined by taking moments of individual pressure areas about the base of the wall. Thus,
PA ∙ ӯ = P1 γ1 + P2 γ2 + P3 γ3 + P4 γ4

Eqns. (9.10) and (9.11) may be used to determine the resultant thrust and its point of application corresponding to any pressure distribution
diagram.
(b) Backfill having different properties above and below water table:
Fig. 9.6 (a) and (b) illustrate this backfill condition and the corresponding pressure distribution diagram.

(iv) Backfill with Uniform Surcharge:


Fig.9.7 (a) illustrates a retaining wall supporting a backfill loaded with a uniform surcharge q. The corresponding pressure distribution
diagram is shown in Fig. 9.7 (b). From the figure it is evident that the effect of the surcharge is identical to that of an imaginary backfill
having a height Zs, placed above G.L., where,
  

(v) Backfill with a Sloping Surface.


The condition is shown in Fig. 9.8 (a). The active earth pressure at any depth z below the top of the wall acts in a direction parallel to the
surface of the backfill and is given by:
Pɑ = Kɑ γH
where, Kɑ = cos β ∙ )
(vi) Wall Having an Incline Backface:
In order to determine the active earth pressure in this case using Rankine's theory, the following steps should be followed (Ref. Fig. 9.9)
 Draw the wall section and the ground line.
 Draw a vertical line through the base of the wall to intersect the ground line at c.
 Compute the length BC from:
  

 Determine the active pressure on this imaginary plane BC, using eqn. (e.13).
 For designing the wall, compute the self-weight of the soil wedge ABC and consider is effect on the stability of the wall separately.

(b) Cohesive frictional Soils:


From eqn. (9.4), the active earth pressure at a depth z is given by.
Pɑ = -
At z = 0, Pɑ = -
At z = H, Pɑ = -
Let Hc be the depth at which pressure intensity is zero.
- = 0, or, -
or, Hc =
  

Fig. 9.10 (b) shows the distribution of active pressure. The negative side of this diagram (i.e., Δ abc) indicates the development of tension
upto a depth Hc. Since soils cannot take tension, cracks will be formed in this zone. The depth H c is, therefore, called the zone of tension
crack. The resultant lateral thrust is obtained by computing the area of the positive side of the diagram (i.e. Δ cde).
(B) Passive Earth Pressure.
(a) Cohesionless soils:
For a cohesionless soil, eqn. (9.5) reduces to:
Pp = γz Nϕ
or, Pp = Kp γz
where, Kp = co-efficient of passive earth pressure

Fig. 9.11 (a) and (b) shows a retaining wall subjected to a passive state' and the corresponding passive pressure distribution diagram.
  

(b) Cohesive-frictional Soils:


From eqn. (9.5), we have,
P p = γ z Nϕ  
For the retaining wall shown in Fig. 9.11(a),
At z = 0, Pp
At z = H, Pp = γH Nϕ
The corresponding pressure distribution diagram is shown inFig.9.ll(c).
9.5. Coulomb's Earth Pressure Theory: Instead of analysing the stresses on a soil element, coulomb considered the equilibrium of the failure soil
wedge as a whole. The major assumptions in Coulomb's theory are:
(i) The soil is dry, homogeneous and isotropic.
(ii) The failure surface formed due to the yielding of the wail is a plane surface.
(iii) The failure wedge is a rigid body.
(iv) The backface of the wall is rough.
The resultant thrust acts on the backface of the wall at one-third height and is inclined to the normal on the wall at this point at an angle δ , where,
δ = angle of wall friction.
9.5.l Wall friction: The concept of wall friction is illustrated in Fig. 9.12 (a) and (b).
In the active state, the wall moves away from the backfill and the failure wedge ABC tends to move downwards. As it slides down'
frictional resistances act upward along the backface of the wall (soil-wall friction),and the failure plane (soil-to-soil friction). In absence of the
frictional force FR1, the active thrust P would have been acting normally on the backface. But now the resultant PA of P and FR1, is inclined at an
angle δ to the normal on the backface. Due to similar reasons, the soil reaction PA will also be inclined at an angle ϕ to the normal on the failure
surface. The same arguments lead us to the conclusion that in a passive state also, Pp and Rp will be inclined at angles δ and ϕ respectively to the
normals on AB and BC. However, in the active state, the lines of action of Pp and RA lie below the respective normals, whereas in the passive
state, the lines of action of Pp and Rp lie above them.
A number of graphical and analytical methods for the determination of lateral earth pressure have been proposed or the basis of Coulomb's theory.
The most important methods arc:
Graphical method:
(i) Culmann's method
(ii) Rebhann's construction
Analytical method:
Trial wedge method.
For detailed description of these methods, the reader may refer to any standard text-book of Soil Mechanics. However the application of these
methods have been illustrated in this chapter by a number of worked-out problems.
Some of the special techniques required to enable us to solve more complex problems involving external loads, or irregularities in the shape of
the wall or the ground surface have also been dealt with.
EXAMPLE :
A gravity retaining wall with a rough backface having a positive batter angle of 10, has to retain a dry, cohesionless backfill upto 4.5 m above G.L. The
properties of the backfill are:
γ = 17 kN/m³, ϕ = 25
The top of the backfill is sloped upwards at 20 °to the horizontal. The angle of wall friction may be taken as 15°. Determine the total active thrust on the
wall by Rebhann's construction.
Solution: This problem cannot be solved by the conventional Rebhann's method, as the ground line and ϕ-line will meet at a great distance (β is nearly
equal to ϕ). However, certain modifications over Rebhann's method will enable us to solve the problem. The solution is presented in Fig. 9.21, while the
procedure is explained below:

 The backface of the wall, A8, is drawn to e scale of 1: 80

 The ground-line AC ϕ-line BD and ψ-line BX are drawn.


 Here, ψ = 90 - (10° + 15°) = 65°
 An arbitrary point E is taken on BD
 A semi-circle is drawn with BE as diameter.
 EF ǁAC is drawn. lt intersects A-B at F.
 FG ǁBX is drawn. It intersects BD at G.

  ւ BD is drawn. It intersects the circle at H.
GH
 With B as centre and BH radius, an arc HI is drawn to intersect BD at I.
 FI is joined.
 AJ ǁFI is drawn. AJ intersects AD at J.
 From J as centre and JK ǁBX is drawn to intersect AC at K.
 With J as cente and JK as radius, an arc KL is drawn to intersect BD at L.
 KL is joined.
 From K, KM ւ BD is drawn.
Now, Pa = weight of the soil wedge JKL = ∙ KM ∙ LJ ∙ γ
= (0.5) (0.3) (3.3) (17) = 84.2 kN/m.

CHAPTER 10 STABILITY OF SLOPES


The soil mass bounded by a slope has a tendency to slide down. The principal factor causing such a sliding failure is the self-weight of the soil.
However, the failure may be aggravated due to seepage of water or seismic forces. Every man-made slope has to be properly designed to ascertain the
safety of the slope against sliding failure.
Various methods are available for analysing the stability of slopes. Generally these methods are based on the following assumptions:
1. Any slope stability problem is a two-dimensional one.
2. The shear parameters of the soil are constant along any possible slip surface.
3. In problems involving seepage of water, the flownet can be constructed and the seepage forces can be determined.
 

 
2 Stability
  of Infinite slopes: In Fig. 10.1, X-X represent an infinite slope which is inclined to the horizontal at an angle β. On any plane YY
(YY ǁ XX) at a depth z below the ground lever the soil properties and the overburden pressure are constant. Hence, failure may occur along a
plane parallel lo the slope at some depth. The conditions for such a failure may be analysed by considering the equilibrium of the soil prism
ABCD of width b.
Considering unit thickness, volume of the prism V = z b cos β and, weight of the prism, W = γz b cos β
Vertical stress on YY due to the self-weight.
σz = = γz cos β
This vertical stress can be resolved into the following two components:
σ = σz cos β = γz cos² β
and, τ = σz sin β = γz cos β sin β
Failure will occur if the shear stress τ exceeds the shear strength τ f of the soil. The factor of safety against such failure is given by,

F=
cphensionless soils: We have from Coulomb’s equation,
τf = c + σ tan ϕ
For a cohensionless soil, c = 0,
τf = σ  tan ϕ
Substituting in eqn. (10.4)
F=
Again, substituting the expressions for σ and τ.
F= =
When ϕ = β, F = 1. Thus a slope in a cohesionless soil is stable till β ≤ ϕ, provided that no external force is present.
c – ϕ soils: In this case, the factor of safety against slope failure is given by,
F=
F=
Let Hc be the critical height of the slope for which F = 1 (i.e, τf = τ)
γ Hc cos β sin β = c + γ Hc cos² β tan ϕ
or, Hc =
or, Hc =
Eqn. (10.7) may also be written as:

Or,
where,
   sn is a dimensionless quantity known as the stability number and is given by:
sn =
If a factor of safety Fc is applied to the cohesion such that the mobilized cohesion at a depth H is,
Cm =
Then, sn = =
From eqns. (10.10) and (10.12), we get,
=
or, Fc = = FH
Hence, the factor of safety against cohesion, Fc is the same as the factor of safety with respect to height, FH.
10.3 Stability of Finite slopes: In case of slopes of limited extent, three types of failure may occur. These are: face failure, toe failure and
base failure (Fig. 10.2 a, b and c respectively).
Various methods of analysing the failure of finite slopes are discussed below.
Swedish Circle Method: In this method, the surface of sliding is assumed to be an arc of a circle.
 Purely cohesive soils: Let AB represent the slope whose stability has to be investigated. A trial slip circle AS 1 C is drawn with O as centre
and OA = OC = R as radius.

 Let W be the weight of the soil mass AS 1 CB acting vertically downwards through the centre of gravity and c be the unit cohesion of the
soil. The self-weight tends to cause the sliding while the shear resistance along the plane AS 1 C counteracts lt.
Now, arc length AS1 C = R ∙ θ
where, θ = ∟AOC (expressed in radius)
Total shear resistance along the plane AS 1 C = R θ c
Restoring moment = shear resistance x lever arm
or, MR = R θ c x R = r² θ c
Considering unit thickness of the soil mass,
W=A∙1 ∙γ = Aγ
where,
γ = unit weight of the soil
A = cross-sectional area of the sector AS 1 CB
 The
  area A can be determined either by using a planimeter or by drawing the figure to a proper scale on a graph paper and counting the
number of divisions of the graph paper covered by the area. Now, disturbing moment, Md = W . d
where, d = lever arm of W with respect to O.
The distance d may be determined by dividing the area into an arbitrary number of segments of small width, and taking moments of all
these segments about O.
Thus, the factor of safety against slope failure,
F= =
(b) Cohesive frictional soils: With reference to Fig. 1O.4, a trial slip circle AS1 C is taken and the sector AS1 CB is divided into a number
of vertical slices, preferably of equal width. The forces acting on each slice are:
(i) Self-weight, W, of the slice, acting vertically downwards through the centre of gravity. Considering unit thickness of the slice,
W = γ x ba x ιa
Where, ba and ιa represent the average height and length of the slice respectively.
(ii) The cohesive force, C, acting along the arc in a direction opposing the probable motion of the sliding soil.
   C = c ∙ ιa
Where, c = unit cohesion
ιa = average length of slice
(iii) Lateral thrust from adjacent slices, EL and ER. In simplified analysis it is assumed that, EL and ER. Hence the effects of these two forces
are neglected
(iv) Soil across the arc: According to the laws of friction, when the soil is about to slide, R will be inclined to the normal at an angle ϕ.
(v) The vertical stresses, VL and VR, which are equal and opposite to each other and hence need not be considered.
The weight W is resolved into a normal component N and a tangential component T. For some of the slices T will enhance the failure, for
the others it will resist the failure. The algebraic sum of the normal and tangential components are obtained from:
Σ T = Σ (W sin α)
and, Σ N = Σ (W cos α)
now, driving moment, MD = R Σ T
and, restoring moment, MR = R [c Σ Δ ι + Σ N tan ϕ]
but Σ Δ ι = are length of arc AS1 C = R θ
MR = R [c R θ + Σ N tan ϕ]
Factor of safety,
F = = OR F =
1O.5 Method of Locating the Centre of the Trial slip circle: The number of trials required to find out the critical slip circle can be
minimised by an empirical method proposed by Fellenius. According to him, the centre of the critical slip circle is located on a straight line
PQ, which can be obtained as follows:
 Draw the given slope AB and determine the slope angle β.
 Determine the values of the angles α1 and α2 (Fig. 10.5) from Table 10.1.
 From A, draw AP at angle of α1 to AB.

 From B, draw BP making it inclined to the horizontal at α. BP and AP intersect at P, which is a point on the desired line PQ.
 The other point Q is located at a depth H below the toe of the slope and it a horizontal distance of 4.5 H away from it. Locate this
point and join PQ. The centre of the critical slip circle will be located on PQ.
10.8 Friction Circle Method: This method is based on the assumption that the resultant force R on the rupture surface is tangential to a circle
of radius γ = R sin ϕ which is concentric with the trial slip circle. Various steps involved ate given below:
1. Draw the given slope to a chosen scale.
2. Select a trial slip circle of radius R, the centre of which is located at O (Fig. 10.6 a)
3. Compute r (= R sin ϕ) and draw another circle of radius r, with O as centre.
4. Now consider the equilibrium of the sliding soil mass under the following forces:
(i) Self-weight W of the sector ABCD.
ii) The cohesive force C along the plane ADC, the magnitude and direction of which can be computed as follows:
Let c be the unit cohesion. The arc ADC is divided into a number of small elements. Let C 1, C2,….., Cn be the mobilised cohesive forces
along them.
The resultant C of these forces can be determined by drawing a force polygon.
Now, the mobilised unit cohesion, Cm’, is given by:
where,
  Fc = factor of safety with respect to cohesion.
The cohesive force is given by
C = Cm’ Lc =
But, summing up the moments of all forces about o and equating to zero, we get,
C ∙ La ∙ R = C ∙ La ∙ α
where, α = perpendicular distance of line of action of C from the centre of the slip circle.
α=
The other force is the soil reaction F, which is assumed to be tangential to the friction circle.
5. Draw the triangle of forces in the following manner:
(i) Draw a vertical line ab to represent W (Fig. 10.6 (b)).
(ii) From a draw ac, making it parallel to the line of action of Fr.
(iii) From b drop a perpendicular bd on ac. The line bd now represents' in magnitude and direction, the cohesive force CR required to
maintain the equilibrium of the soil mass ABCD along the chosen slip circle.
6. Determine the unit cohesion required for stability from:
Cr =
7. The factor of safety w.r.t cohesion is now obtained from:
Fc = =
8. The factor of safety w.r.t shear strength can be obtained as follows:
 (i)
  Assume a certain factor of safety with respect to the angle of internal friction. Let it be Fϕ. The mobilized angle of internal friction
is then given by:
tan ϕm =
(ii) Draw a new friction circle with O as centre and r’ as radius, where,
r' = R sin ϕm
(iii) The factor of safety w.r.t. cohesion Fc is then obtained by forming another triangle of forces. Compare F c and Fϕ. If they are different,
go for another trial.
(iv) In this manner, adjust the radius of the circle until Fϕ and Fc become equal to each other. This value is then accepted as the factor of
safety for shear strength of the soil w.r.t. the given trial slip circle.
10.9 Taylor's Stability Number: Taylor carried out stability analysis of a large number of slopes having various heights, slope angles and
soil properties. On the basis of the results, he proposed a simple method by which the factor of safety of a given finite slope can be easily
determined with reasonable accuracy. Taylor introduced a dimensionless parameter, called Taylor's Stability Number, which is given by,
Sn =
The value of Sn may be obtained from Fig. 10.7.
 The
  stability numbers are obtained for factor of safety w.r.t cohesion, while then factor of safety w.r.t friction, F ϕ is initially taken as
unity.
 The values of Sn obtained from Fig.10.7 are applicable for slip circles passing through the toe. However for slopes made in cohesive
soils of limited depth and underlain by a hard stratum, the critical slip circle passes below the toe. In such cases, the value of S n should
be obtained from Fig.10.8. In this figure, the depth factor plotted along the x-axis is defined as:
nd =

 where, D = Depth of hard stratum below toe


 H = Height of slope above toe.
 Fig. 10.8 consists of a family of curves for various slope angles'-Each curve consists of two parts. The portions drawn with firm lines are
applicable to field conditions illustrated in Fig. 10.9 (a), while the portions drawn with broken lines are meant for the conditions shown
in Fig. 10.9(b).
 The figure also consists of a third set of curves, shown with broken lines, for various values of n where, n represents the distance x of the
rupture circle from the toe, as illustrated in Fig. 10.9(a), and is given by,
n=
  

EXAMPLE :
 Slope of infinite extent is made in a dense sand layer at an angle of 30° to the horizontal. Determine the factor of safety of the slope
against shear failure if the angle of internal friction of the soil be 36°.
Solution: With reference to Fig. 10.1, XX represents the given slope, while YY is a plane parallel to it at a depth z.
Vertical stress on YY due to overburden,
σz = γz
where, γ = unit weight of the soil
Normal stress on YY, σ = σz cos² β (β = slope angle)
Shear stress on YY, τ = σz cos β sin β .
Shear strength of the soil on the plane YY,
τf = σ tan ϕ = σz cos² β tan ϕ.
But, factor of safety against shear failure,
Fs = = =
= = 1.258
THANK YOU 

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy