0% found this document useful (0 votes)
20 views216 pages

GT NOTES BCV502 notes (1)

The document outlines the course content for Geotechnical Engineering-I, covering topics such as soil formation, classification, permeability, compaction, consolidation, shear strength, and soil stabilization methods. It includes detailed modules on soil mechanics, stress conditions, and various testing methods, alongside course outcomes and reference materials. The course aims to equip students with the ability to classify soils, solve practical problems related to soil properties, and apply stabilization techniques.

Uploaded by

pawarlolke
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
20 views216 pages

GT NOTES BCV502 notes (1)

The document outlines the course content for Geotechnical Engineering-I, covering topics such as soil formation, classification, permeability, compaction, consolidation, shear strength, and soil stabilization methods. It includes detailed modules on soil mechanics, stress conditions, and various testing methods, alongside course outcomes and reference materials. The course aims to equip students with the ability to classify soils, solve practical problems related to soil properties, and apply stabilization techniques.

Uploaded by

pawarlolke
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as DOCX, PDF, TXT or read online on Scribd
You are on page 1/ 216

Lecture Note on GTE-I

Geotechnical Engineering-I

Course Content

Module-I

Introduction: Origin of soils, formation of soils, clay mineralogy and soil structure, basic
terminology and their relations, index properties of soils. Soil classification: Particle size
distribution, use of particle size distribution curve, Particle size classification, textural
classification, HRB classification, Unified classification system, Indian standard soil
classification system, Field identification of soils. capillary tension, capillary siphoning. Stress
conditions in soil: Total stress, pore pressure and effective stress

Module-II

Permeability: Darcy’s law, permeability, factors affecting permeability, determination of


permeability (laboratory and field methods), permeability of stratified soil deposits. Estimation
of yield from wells.

Seepage analysis: Seepage pressure, quick condition, Laplace equation for two –dimensional
flow, flow net, properties and methods of construction of flow net, application of flow net,
seepage through anisotropic soil and non-homogenous soil, seepage through earth dam.Inverted
filter and design of inverted filter.

Module-III

Soil compaction: Compaction mechanism, factors affecting compaction, effect of compaction on


soil properties, density moisture content relationship in compaction test, standard and modified
proctor compaction tests, field compaction methods, relative compaction, compaction control.

Soil consolidation: Introduction, spring analogy, one dimensional consolidation, Terzaghi’s


theory of one dimensional consolidation, consolidation test, determination of coefficient of
consolidation

Module-IV

Shear strength of soils: Mohr’s stress circle, theory of failure for soils, determination of shear
strength (direct shear test, tri-axial compression test, unconfined compression test, van shear
test), shear characteristics of cohesionless soils and cohesive soils.

Module-V

Stabilization of soil: Introduction, mechanical stabilization, cement stabilization, lime


stabilization, bituminous stabilization, chemical stabilization, thermal stabilization, electrical

1
Lecture Note on GTE-I

stabilization, Introduction to modern methods of stabilization

Reference Books:

Text Book:Geotechnical Engineering, C. Venkatramaiah, New Age International publishers.

Reference Books:

1. Geotechnical Engineering, T.N. Ramamurthy & T.G. Sitharam, S. Chand & Co.
2. Soil Mechanics, T.W. Lambe& Whiteman, Wiley Eastern Ltd, Nw Delhi.

Course Outcomes:

1 Classify soil and solve three phase soil system.

2 Solve any practical problems related to soil stresses estimation, permeability and seepage
including flow net diagram.

3 Formulate practical problems related to consolidation settlement and time rate of settlement.

4 Validate problem related to compaction in the field.

5 Use stabilization techniques for soft and expansive soil by using various methods.

PO1 PO2 PO3 PO4 PO5 PO6 PO7 PO8 PO9 PO10 PO11 PO12
CO1 2 1 1 1 1
CO2 3 2 1 1 1
CO3 2 1 3
CO4 3 2 1 2 1 1
CO5 3 2 1 2 1 1 1 1 2 2 2 1

2
Lecture Note on GTE-I

1.0 Introduction and Phase Relation


The term "soil" can have different meanings, depending upon the field in which it is considered.
To a geologist, it is the material in the relative thin zone of the Earth's surface within which roots
occur, and which are formed as the products of past surface processes. The rest of the crust is
grouped under the term "rock".
To an engineer, it is a material that can be:
 built on: foundations of buildings, bridges
 built in: basements, culverts, tunnels
 built with: embankments, roads, dams
 supported: retaining walls
Soil Mechanics is a discipline of Civil Engineering involving the study of soil, its behaviour and
application as an engineering material.
Soil Mechanics is the application of laws of mechanics and hydraulics to engineering problems
dealing with sediments and other unconsolidated accumulations of solid particles, which are
produced by the mechanical and chemical disintegration of rocks, regardless of whether or not
they contain an admixture of organic constituents.
Soil consists of a multiphase aggregation of solid particles, water, and air. This fundamental
composition gives rise to unique engineering properties, and the description of its mechanical
behavior requires some of the most classic principles of engineering mechanics.
Engineers are concerned with soil's mechanical properties: permeability, stiffness, and strength.
These depend primarily on the nature of the soil grains, the current stress, the water content and
unit weight.
In the Earth's surface, rocks extend up to as much as 20 km depth. The major rock types are
categorized as igneous, sedimentary, and metamorphic.
 Igneous rocks: formed from crystalline bodies of cooled magma.
 Sedimentary rocks: formed from layers of cemented sediments.
 Metamorphic rocks: formed by the alteration of existing rocks due to heat from igneous
intrusions or pressure due to crustal movement.
Soils are formed from materials that have resulted from the disintegration of rocks by various
processes of physical and chemical weathering. The nature and structure of a given soil depends
on the processes and conditions that formed it:
 Breakdown of parent rock: weathering, decomposition, erosion.
 Transportation to site of final deposition: gravity, flowing water, ice, wind.
 Environment of final deposition: flood plain, river terrace, glacial moraine, lacustrine or
marine.
 Subsequent conditions of loading and drainage: little or no surcharge, heavy surcharge
due to ice or overlying deposits, change from saline to freshwater, leaching,
contamination.
All soils originate, directly or indirectly, from different rock types.

3
Lecture Note on GTE-I

1.1 Physical weathering reduces the size of the parent rock material, without any change in
the original composition of the parent rock. Physical or mechanical processes taking place on the
earth's surface include the actions of water, frost, temperature changes, wind and ice. They cause
disintegration and the products are mainly coarse soils.
The main processes involved are exfoliation, unloading, erosion, freezing, and thawing. The
principal cause is climatic change. In exfoliation, the outer shell separates from the main rock.
Heavy rain and wind cause erosion of the rock surface. Adverse temperature changes produce
fragments due to different thermal coefficients of rock minerals. The effect is more for freeze-
thaw cycles.
1.2 Chemical weathering
Chemical weather not only breaks up the material into smaller particles but alters the nature of
the original parent rock itself. The main processes responsible are hydration, oxidation, and
carbonation. New compounds are formed due to the chemical alterations.
Rain water that comes in contact with the rock surface reacts to form hydrated oxides, carbonates
and sulphates. If there is a volume increase, the disintegration continues. Due to leaching, water-
soluble materials are washed away and rocks lose their cementing properties.
Chemical weathering occurs in wet and warm conditions and consists of degradation by
decomposition and/or alteration. The results of chemical weathering are generally fine soils with
altered mineral grains.
The effects of weathering and transportation mainly determine the basic nature of the soil (size,
shape, composition and distribution of the particles).
The environment into which deposition takes place, and the subsequent geological events that
take place there, determine the state of the soil (density, moisture content) and the structure or
fabric of the soil (bedding, stratification, occurrence of joints or fissures)
Transportation agencies can be combinations of gravity, flowing water or air, and moving ice. In
water or air, the grains become sub-rounded or rounded, and the grain sizes get sorted so as to
form poorly-graded deposits. In moving ice, grinding and crushing occur, size distribution
becomes wider forming well-graded deposits.
In running water, soil can be transported in the form of suspended particles, or by rolling and
sliding along the bottom. Coarser particles settle when a decrease in velocity occurs, whereas
finer particles are deposited further downstream. In still water, horizontal layers of successive
sediments are formed, which may change with time, even seasonally or daily.
Wind can erode, transport and deposit fine-grained soils. Wind-blown soil is generally
uniformly-graded. A glacier moves slowly but scours the bedrock surface over which it passes.
Gravity transports materials along slopes without causing much alteration.
1.3 BASICS OF SOIL CLASSIFICATION

Soils as they are found in different regions can be classified into two broad categories:
(1) Residual soils
(2) Transported soils

4
Lecture Note on GTE-I

1.3.1 Residual Soils


Residual soils are found at the same location where they have been formed. Generally, the depth of
residual soils varies from 5 to 20 m.
Chemical weathering rate is greater in warm, humid regions than in cold, dry regions causing a faster
breakdown of rocks. Accumulation of residual soils takes place as the rate of rock decomposition
exceeds the rate of erosion or transportation of the weathered material. In humid regions, the
presence of surface vegetation reduces the possibility of soil transportation.
As leaching action due to percolating surface water decreases with depth, there is a corresponding
decrease in the degree of chemical weathering from the ground surface downwards. This results
in a gradual reduction of residual soil formation with depth, until unaltered rock is found.
Residual soils comprise of a wide range of particle sizes, shapes and composition.
1.3.2 Transported Soils
Weathered rock materials can be moved from their original site to new locations by one or more of the
transportation agencies to form transported soils. Transported soils are classified based on the
mode of transportation and the final deposition environment.
(a) Soils that are carried and deposited by rivers are called alluvial deposits.
(b) Soils that are deposited by flowing water or surface runoff while entering a lake are
called lacustrine deposits. Alternate layers are formed in different seasons depending on flow
rate.
(c) If the deposits are made by rivers in sea water, they are called marine deposits. Marine
deposits contain both particulate material brought from the shore as well as organic remnants of
marine life forms.

(d) Melting of a glacier causes the deposition of all the materials scoured by it leading to
formation of glacial deposits.

(e) Soil particles carried by wind and subsequently deposited are known as aeolian deposits.
Soil is not a coherent solid material like steel and concrete, but is a particulate material. Soils, as they
exist in nature, consist of solid particles (mineral grains, rock fragments) with water and air in
the voids between the particles. The water and air contents are readily changed by changes in
ambient conditions and location.
As the relative proportions of the three phases vary in any soil deposit, it is useful to consider a soil
model which will represent these phases distinctly and properly quantify the amount of each
phase. A schematic diagram of the three-phase system is shown in terms of weight and volume
symbols respectively for soil solids, water, and air. The weight of air can be neglected.

5
Lecture Note on GTE-I

1.4 PHASE RELATIONSHIP OF SOILS

Fig.1: Three phase diagram of soil mass


The soil model is given dimensional values for the solid, water and air components.
Total volume, V = Vs + Vw + Vv
Soils can be partially saturated (with both air and water present), or be fully saturated (no air
content) or be perfectly dry (no water content).
In a saturated soil or a dry soil, the three-phase system thus reduces to two phases only, as
shown.

Fig.2: Soil mass showing three phase


For the purpose of engineering analysis and design, it is necessary to express relations between
the weights and the volumes of the three phases.
The various relations can be grouped into:
 Volume relations
 Weight relations
 Inter-relations
 As the amounts of both water and air are variable, the volume of solids is taken as the
reference quantity. Thus, several relational volumetric quantities may be defined. The
following are the basic volume relations:
1. Void ratio (e) is the ratio of the volume of voids (Vv) to the volume of soil solids (Vs),

𝑒 = ---------------------------
𝑉𝑣
and is expressed as a decimal.
Eq.1
𝑉𝑠

6
Lecture Note on GTE-I

2. Porosity (n) is the ratio of the volume of voids to the total volume of soil (V ), and

𝑛 = -------------------------------
𝑉𝑣
is expressed as a percentage.

𝑉
Eq.2

𝑛
Void ratio and porosity are inter-related to each other as follows:
𝑒 =
1−𝑛
𝑛=
𝑒

1+𝑒
And
3. The volume of water (Vw) in a soil can vary between zero (i.e. a dry soil) and the volume

𝑆
=
𝑉𝑤
of voids. This can be expressed as the degree of saturation (S) in percentage.

𝑉𝑣
- Eq.3
For a dry soil, S = 0%, and for a fully saturated soil, S = 100%.
4. Air content (ac) is the ratio of the volume of air (Va) to the volume of voids.
𝑎
𝑎
------------------------------- Eq.4
𝑐 𝑉𝑣

n = × 100 = n ×
Va
5. Percentage air voids (na) is the ratio of the volume of air to the total volume.

a c
a
-Eq.5
V
Density is a measure of the quantity of mass in a unit volume of material. Unit weight is a
measure of the weight of a unit volume of material. Both can be used interchangeably. The units
of density are ton/m³, kg/m³ or g/cm³. The following are the basic weight relations:
1. ` The ratio of the mass of water present to the mass of solid particles is called the water

𝜔=
𝑊𝜔
content (w), or sometimes the moisture content.
----------------------- Eq.6
𝑊𝑠
Its value is 0% for dry soil and its magnitude can exceed 100%.
2. The mass of solid particles is usually expressed in terms of their particle unit

𝛾 = = 𝐺 𝛾 ----------------
𝑊𝑠
weight (𝛾𝑠) or specific gravity (Gs) of the soil grain solids.
𝑠 𝜔
Eq.7
𝑠
𝑉𝑠
where, 𝛾𝜔= unit weight of water
For most inorganic soils, the value of Gs lies between 2.60 and 2.80. The presence of organic
material reduces the value of Gs.
3. Dry unit weight (γd) is a measure of the amount of solid particles per unit volume.
𝛾--------------------
𝑠

𝑉
Eq.8
𝑠
4. Bulk unit weight (𝜸𝒐𝒓 𝜸𝒕) is a measure of the amount of solid particles plus water
per unit volume.
𝛾
= 𝑠𝛾 =𝑤
𝑤
𝑊 +𝑊
𝑉𝑠+𝑉
𝑡

7
Lecture Note on GTE-I

- ---------- Eq.9
5. Saturated unit weight (𝛾𝑠𝑎𝑡) is equal to the bulk density when the total voids is filled up
with water.

8
Lecture Note on GTE-I

6. Buoyant unit weight or submerged unit weight (γ′) is the effective mass per unit volume

𝛾; = 𝛾𝑠𝑎𝑡 − 𝛾𝑤-----------------------
when the soil is submerged below standing water or below the ground water table.
Eq.10
It is important to quantify the state of a soil immediately after receiving in the laboratory and prior to
commencing other tests. The water content and unit weight are particularly important, since they
may change during transportation and storage.
Some physical state properties are calculated following the practical measurement of others. For
example, dry unit weight can be determined from bulk unit weight and water content. The
following are some inter-relations:
1.5 Relationships Involving Porosity, Void Ratio, Degree of Saturation, Water
Content, Percent Air Voids and Air Content

𝑛 = , as a fraction
𝑉𝑣
Fig.3: Three phase diagram used for establishment of inter relationship

𝑉
= =1−
𝑉−𝑉𝑠
𝑉
--------------------------
𝑠 𝑊𝑠
=1-
𝑉 𝑉 𝐺𝛾𝜔𝑉

𝑛=1−
Eq.11
𝑊𝑑 -------------------------------------- Eq.12
𝐺𝛾𝜔𝑉

𝑒= = = −1= − 1------------------
𝑉𝑣 𝑉−𝑉𝑠 𝑉 𝑉𝐺𝛾𝜔
This may provide a practical approach to determine n
Eq.13
𝑉𝑠 𝑉𝑠 � 𝑊𝑠
𝑠

𝑒 = − 1---------------------
𝑉𝐺𝛾𝜔
Eq.14
𝑊𝑑

𝑛
= and 𝑒 =
𝑉𝑣 𝑉𝑣
The expression provides an approach to determine value of void ratio e

𝑉 𝑉
𝑉𝑠+𝑉𝑣 𝑠
= = = + = +1=
1 𝑉 𝑉𝑠 𝑉𝑣 1 1+𝑒
---------------------- Eq.15
𝑛 𝑉 𝑉𝑣 𝑉𝑣 𝑒 𝑒
𝑉𝑣
𝑒𝑣
So 𝑛 =
1+𝑒
------------------------------ Eq.16
a =
V
a
Vv
c
9
Lecture Note on GTE-I

and 𝑛 =
𝑉𝑣

∴ = = ------------------------------
𝑉𝑎

𝑛
Eq.17

𝑛𝑎𝑐 𝑉 �

By definition

10
Lecture Note on GTE-I

𝜔=
𝑊𝑤
--------------------- Eq.18
𝑊𝑠
𝑆=
𝑉𝑤
--------------------- Eq.19
𝑉𝑣
And 𝑒 =
𝑉𝑣
------------------- Eq.20
𝑉𝑠
Now , 𝑆𝑒 =
𝑉𝑤
------------------ Eq.21
𝑉𝑠
Again, 𝜔 = = = = = 𝑆𝑒/𝐺-----------
𝑊 𝑤 𝑉𝑤𝛾𝑤 𝑉𝑤𝛾𝑤 𝑉𝑤
Eq.22
𝑊𝑠 𝑉𝑠 𝑉 𝑠𝐺 𝑉𝑠𝐺
𝛾𝑠 𝛾𝑤
Hence 𝜔𝐺 = 𝑆𝑒---------------------------- Eq.23

1.6 Relationships Involving Unit Weights, Grain Specific Gravity, Void Ratio, and
Degree of Saturation:

Ww⁄
Ws(1+
Ws)
γ= = =
W Ws+Ww Eq.24
Vv
V Vs(1 ⁄
--------------------
+ )
Vs+Vv Vs

= 𝜔, water content and void ratio 𝑒 =


But 𝑊𝑤 𝑉𝑣

𝑊𝑠 𝑉𝑠
Again 𝛾 = = 𝐺𝛾 ----------------------------------
𝑊𝑠
Eq.25
𝑠 𝑤
Hence , 𝛾 = 𝐺𝛾
𝑠 𝑉 1+𝜔

𝑤 1+𝑒
---------------------------------- Eq.26

Or 𝛾 = (𝐺+𝐺𝜔)--------------------------------------
𝛾𝑤
(1+𝑒)
Eq.27

We know that , 𝜔𝐺 = 𝑒𝑆-------------------------- Eq.28

𝛾 = 𝛾(where S is taken as a fraction)--------------------


(𝐺+𝑒𝑠)
Substituting , we get
(1+𝑒) 𝑤
Eq.29

This is a general expression from which unit weight corresponding to the saturated and dry states


of soil mass may be found out by substituting S=1 and S=0 respectively.
𝛾
=
(𝐺+𝑒) --------------------
𝛾
Eq.1.30
𝑠𝑎𝑡 (1+𝑒) 𝑤

𝛾 ----------------------------
= (1+𝑒
𝐺
𝛾𝑑
And Eq.1.31
) 𝑤

𝛾′ (𝐺+𝑒)
= 𝛾𝑠𝑎𝑡 − 𝛾𝑤-------------------------
The submerged unit weight may be written as

= 𝛾 − 𝛾 ----------------------
Eq.1.32
𝑤 𝑤
Eq.1.33
(1+𝑒)

= [(𝐺+𝑒) − 1]---------------------------
𝛾𝑤 (1+𝑒)
Eq.1.34

So 𝛾′ = 𝛾 ---------------------------
(𝐺−1)

(1+𝑒) 𝑤
Eq.1.35

11
Lecture Note on GTE-I

Example 1: A soil has void ratio = 0.72, moisture content = 12% and Gs= 2.72. Determine its
(a) Dry unit weight

12
Lecture Note on GTE-I

(b) Moist unit weight, and the


(c) Amount of water to be added per m3 to make it saturated.

𝛾 = =
𝐺𝑠𝛾𝜔 2.72×9.81
Solution:
𝑑
a) =15.51 kN/m3
1+0.72

1+𝑒

𝛾 = 𝛾𝑑(1 + 𝜔)
b)

= × 2.12 × 9.81 = 17.28kN/m3


1+0.12

1+0.72

= × 9.81=19.62kN/m3
2.72+0.72

= ×

𝐺𝑠+𝑒𝑆
c)
𝛾

𝑠𝑎𝑡 𝜔 1+0.72
1+
𝑒

𝛾𝑠𝑎𝑡 − 𝛾
 Water to be added per m3 to make the soil saturated

= 19.62 – 17.38 = 2.24 kN

Example 2: The dry density of sand with porosity of 0.387 is 1600 kg/m3. Find the void ratio of
= 1000
𝑘𝑔

𝛾𝜔
the soil and the specific gravity of the soil solids. [Take ]
𝑚3

n = 0.387
= 1600 kg/m3

Solution:
𝑛 0.387
𝑒 = = = 0.631
1−𝑛 1 − 0.387
𝐺𝑠𝛾𝜔
𝛾𝑑 =
1+𝑒
𝐺 = ×𝛾 = × 1600 = 2.61
1+𝑒 1+0.631

𝑠 𝛾𝜔 1000
𝑑

Example 3

A sample of saturated soil has a water content of 35%. The specific gravity of soil solids is 2.65.
Determine its voids ratio, porosity, saturated unit weight and dry unit weight.

13
Lecture Note on GTE-I

Example 4
A sample of clay taken from a natural stratum was found to be partially saturated and when
tested in the laboratory gave the following results. Specific gravity of soil particle =2.6, wet
weight of sample =2.50 N, dry weight of sample =2.10 N and volume of sample =150 cm 3.
Determine the degree of saturation.
Sol:
Data given

14
Lecture Note on GTE-I

Alternatively, this can be found from three phase diagram

15
Lecture Note on GTE-I

Problem No 1 A soil sample has a porosity of 40 %. The specific gravity of soil solids is 2.70.
Calculate a) Voids ratio, b) dry density, c) unit weight if the soil is 56% saturated and d) unit
weight if the soil is completely saturated

2.0 INDEX PROPERTIES OF SOILS


The properties of soil which are used for identification and classification are called as Index
Properties of Soil
INTRODUCTION
Basically in soil mechanics properties of soil are broadly classified as
1 Index Properties and
2 Engineering Properties
The properties of soil which are used for identification and classification are called as Index
Properties of Soil.

2.1 PROPERTIES OF SOIL


Divided into Two Categories namely
1 Index Properties
a. Specific Gravity
b. Water Content
c. Particle Size
Distribution
d. In-Situ Density
e. Consistency
Limits
f. Relative Density

16
Lecture Note on GTE-I

2 Engineering Properties
a. Permeability
b. Shear Strength
c. Compressibility
2.1.1 SPECIFIC GRAVITY
Definition:
The ratio of weight of soil solids to the weight of equal volume of water content at 40C standard
temperature
Laboratory Determination:
a. Pycnometer Method

Fig.4: Schematic diagram for determination of Specific gravity


Specific gravity 𝐺 =
(𝑊2−𝑊1 )
------------------- Eq.2.1
(𝑊2−𝑊1)−(𝑊3−𝑊4)

2.1.2 DETERMINATION WATER CONTENT


1. Oven Dry Method
2. Sand Bath Method
3. Pycnometer Method
4. Rapid Moisture Meter
1 OVEN DRY METHOD
1. Take the empty weight of container (W1)
2. Take the weight of container along with moist soil sample (W2)
3. Place the container in the oven for 24 hours at 1050 C temperature (W3)
4. Note down the weight of container along with dried soil sample (W4)
Water content 𝑤 =
(𝑊2−𝑊3)
--------------------------- Eq.2.2
(𝑊4−𝑊1)

17
Lecture Note on GTE-I

2 SAND BATH METHOD

1. It is a rapid (Quick) and approximate method


2. Sand is collected in a container and is literally heated with fire
3. Take the empty weight of container (W1)
4. Take the weight of container along with moist soil sample (W2)
5. Place the container in the sand bath for 15 to 20 minutes. (W3)
6. Note down the weight of container along with dried soil sample.(w4)
content 𝑤 =
Water (𝑊2−𝑊3)

(𝑊4−𝑊1)

3 RAPID MOISTURE METER METHOD

It is also called as Calcium Carbide Method


This test is based on gas pressure developed by the reaction of calcium carbide with moisture of
soil.
Wet soil is placed in the cup
Holding the moisture meter horizontally, cup and chamber is brought together.
The meter is shaken for about 15 to 20 seconds.
Calcium carbide reacts with available moisture and releases acetylene gas.
From the calibrated scale of the pressure gauge the percentage of water on total mass of wet
soil is obtained.
2.1.3 PARTICLE SIZE DISTRIBUTION
Coarse Soils (Size>0.75mm)
a. Dry Sieve b. Wet Sieve
Analysis Analysis
Fine Soils (Size<75mm)
a. Sedimentation b. Hydrometer analysis

IS Sieves

18
Lecture Note on GTE-I

SIEVE ANALYSIS
Objective of sieve analysis is to classify the soil and to determine gradation.
A set of IS Sieves are taken
A suitable amount of pulverized dry soil (500g) is taken and sieved through different sieves with
the help of sieve shaker for 10 – 15 minutes.
Weight of soil retained on each sieve is determined.
Percentage weight of soil retained on each sieve is
determined. Percentage finer is determined.
Plot of Percentage finer v/s Size of the sieve on semi log sheet will give the type of gradation,
Effective size, Coefficient of uniformity and Coefficient of curvature,
TABULATION FOR SIEVE ANALYSIS
IS Sieve Weight of Soil Percentage Cumulative Percentage Finer
Size Retained(gm) Retained Percentage (100% - Cumulative Retained)
4.75mm
2mm
1mm
600μ
425μ
212μ
150μ
75μ
Pan

19
Lecture Note on GTE-I

2.1.4 TYPICAL PARTICLE SIZE DISTRIBUTION GRAPH

Fig. 5: Grain size distribution curve


Gradation and Coefficients
1. Well Graded
2. Uniformly Graded
3. Gap Graded or Step Graded

1.
Coefficient of Uniformity (Cu)
𝐶
60

D10
----------------------- Eq.2.3
𝑢
2.
Coefficient of Curvature (Cc)
𝐷2
𝐶𝐶 30

=
𝐷60×𝐷
------------------------- Eq.2.4
10
Well graded soils have high Cu values and poorly graded soils have low Cu, values. If all the
particles of soil mass are of the same size Cu is unity.
Cc lies between 1 to 3 for well graded soil.
Uniform soil: Cu = 1
Poorly graded soil: 1< Cu<4
Well graded soil: Cu>4
2.1.5 HYDROMETER ANALYSIS
1. The analysis is carried out in two stages
• Calibration of hydrometer
• Sedimentation analysis
CALIBRATION OF HYDROMETER
A = Area of stem or graduate
20
Lecture Note on GTE-I

Rh = Hydrometer Reading
He = Effective Height
H = Depth of stem up to hydrometer reading

𝐻 = (𝐻 + + ℎ) − ℎ----------------------
ℎ 𝑉 𝑉
Vh = Volume of hydrometer in cc
𝑒
Eq.2.5
2𝐴 𝐴
2
Hydrometer is immersed in the soil suspension, hydrometer readings are noted at different time
intervals
Effective height values are calculated for corresponding ‘H’ values.
A plot of He v/s Rh is used as the calibration chart.
Note that calibration of hydrometer is done only once. The same chart can be used for any ‘n’
number of trials.

Fig. 6: Calibration chart for Hydrometer

21
Lecture Note on GTE-I

Fig.7: Calibrated Hydometer

TEST PROCEDURE

1. 100 ml Soil suspension is prepared by mixing about 12 to 30 gm of soil with dispersing


agent such as sodium hexameta-phosphate (33 gm) and sodium carbonate (7 gm) and
distilled water
2. 100 ml of soil suspension is added with 900 ml of water and is poured in jar.
3. Suspension is mixed thoroughly and the hydrometer is immersed in the jar.
4. Hydrometer readings are noted at different interval of time.
5. He is taken from the calibration chart
6. Diameter of the particles is determined by following equation

×
He
𝐷=
18𝜂 √ t
------------------------- Eq.2.6
(ϒ𝑠−ϒ𝑤
)

D - Diameter of soil particle


Η - Viscosity of
water He - Effective
height t - Time in
(seconds)
γs - Unit weight of solids
γw - Unit weight of water
% finer is obtained from the following equation
𝑁=(
𝐺

22
Lecture Note on GTE-I

𝐺−1
) × 100-----------------------
𝑅𝑐
Eq.2.7
𝑊𝑠

23
Lecture Note on GTE-I

Where,
N - % Finer
G - Specific Gravity of Solids
Rc - Corrected reading of hydrometer
Ws -Weight of solids

CORRECTIONS FOR HYDROMETER READING

1. Correction for meniscus (cm)


cm = mu– ml
Where, mu = Upper meniscus
ml = Lower meniscus
Cm is always positive because upper meniscus shows lesser values
than actual

2. Correction for temperature (ct)


More the temperature than standard temperature, (27C), less will be viscosity, hydrometer
reading will be less than actual and hence correction will be positive
Lesser the temperature than standard temperature, (27 0C), more will be viscosity, hydrometer
reading will be more than actual and hence correction will be negative.

3. Correction for dispersing agent (cd)


Addition of dispersing agent increases net density of the sample.
This causes more values than the actual.
Hence, correction for dispersion agent will always be negative.

Total correction (cr)


Rc = cm ± ct – cd

2.1.6 INSITU DENSITY BY CORE CUTTER METHOD

Used for cohesive soil in which core can be easily penetrated


Take empty weight of core (Without collar) (W1)
Penetrate the cone into the earth using rammer
After chipping off extra soil, note down weight of core along with soil (W2)
Determine moisture content of the soil using oven dry method.
Bulk unit weight and dry unit weight can be determined as follows
Measure the volume of core (V)

24
Lecture Note on GTE-I

In-situ Density = γ =(𝑊2−𝑊1)-----------------------


𝑉
Eq.2.8

𝛾
𝛾𝑑
Dry Density
1+𝑤
= --------------------- Eq.2.9

2.1.7 INSITU DENSITY BY SAND REPLACEMENT METHOD

This test is done in two steps


A. Calibration of container
B. Determination of in-situ density of soil
Sand is filled in the Sand Pouring Cylinder (SPC) and the weight is taken with closed shutter.
(W1)
The SPC is placed on a smooth glass plate and the sand is allowed to run out of cone.
Weight of SPC after running sand in to cone is noted (W2)
SPC is refilled and placed on calibrating container and sand
is allowed to run out.
The weight of SPC is noted after running sand in to
calibrating container (W3).
Weight of sand in cone
Wcone =W1 –W2
Weight of sand in calibrating container
Wcc =W1 –W3 +Wcone
Volume of calibrating container is determined (Vcc)
=
𝑊𝐶𝐶
Unit weight of sand
𝑉𝐶𝐶
------------------ Eq.2.10

𝛾𝑠𝑎𝑛𝑑
SPC with sand up to W1 is taken to the field where bulk unit weight is required. The surface is
cleaned and leveled.
The tray with central hole is placed in position. A hole of approximately same size as calibrating
container is dug
The removed soil is collected on the sides of tray and its weight is noted (Wsoil).
The SPC is placed on the hole and sand is allowed to run in to hole.
Weight of SPC is noted (W4).
Weight of sand in the holeW5 =W1 –W4 +Wcone
Volume of the hole
𝑉𝑠 =
𝑊𝑆
𝛾𝑠𝑜𝑖𝑙
---------------------- Eq.2.11
𝑜𝑖𝑙

25
Lecture Note on GTE-I

In-situ Density of Soil 𝛾 =


𝑊𝑠𝑜𝑖𝑙
𝑉𝑠𝑜𝑖𝑙
-------------------- Eq.2.12

26
Lecture Note on GTE-I

2.3 ATTERBERG LIMITS

Consistency is the relative ease with which soil can be deformed.


It is applicable to fine grained soils whose consistency depends on water content. Relative
consistency can be expressed as very stiff, stiff, medium stiff, soft, very soft with increasing
water content.

Fig.8: Chart showing consistency limit


2.3.1 Liquid Limit

It is the arbitrary water content between plastic and liquid phase of soil
It is the minimum water content in the soil where it behaves as a plastic material
It is the amount of water in the soil, where the soil contains least amount of shear strength.
2.3.2 Plastic Limit
It is the minimum water content in the soil at which a thread formed by rolling on a glass plate
crumbles at 3 mm diameter.
In other words, it is the arbitrary water content between plastic and semi solid state.
2.3.3 Shrinkage Limit
It is the maximum water content at which soil behaves as a solid material.
It is the maximum water content below which, decrease in water content results in no change of
volume of soil.
In other words, it is the arbitrary water content between solid and semi solid phase of soil.
2.4 DETERMINATION OF LIQUID LIMIT
Casagrande’s Liquid Limit Apparatus
The following procedures are used

27
Lecture Note on GTE-I

1 Take 120 g of air dried soil after passing through 425μ sieve.
2 Arbitrarily mix certain amount of water and make a paste.
3 Put the soil pat in the brass cup and make a V shaped groove using grooving tool.
4 Give blows at a speed of 2 rps., till the grove closes.
5 Count the number of blows at the end of experiment.
6 Take certain amount of soil from the cup to determine water content.
7 Take at least 5 number of trials with varying number of blows from 15 to
35 The corresponding water content at 25th blow is taken as liquid limit

2.5 DETERMINATION OF PLASTIC LIMIT

Procedures to be taken
1 Take about 15 g of air dried soil after passing through 425μ sieve.
2 Arbitrarily mix certain amount of water and make soil ball.
3 Roll the soil ball on glass plate with palm.
4 Make thread till it attains a thickness of 3mm.

28
Lecture Note on GTE-I

5 If the thread crumbles at 3 mm thick, then take sample to determine water content.
6 Corresponding water content is taken as PLASTIC LIMIT.
2.6 DETERMINATION OF SHRINKAGE LIMIT
Procedure to determine Shrinkage Limit
1 Take about 30 g of air dried soil after passing through 425μ sieve in an evaporating dish.
2 Soil is mixed with sufficient quantity of water so that it may flow.
3 Fill the soil mix in dish and take the weight of dish.
4 The soil pat is allowed to dry in air till the colour of pat changes from dark to light.
5 The dish is then placed in oven at 110oc till it attains constant weight.

Apparatus required for Shrinkage Limit test


6 The shrinkage dish is weighed with the dry soil and dry weight of soil is determined.
7 The volume of dry soil pat is determined by mercury displacement
method. V1 = Volume of container in which plastic phase soil is filled initially.
W1 =Weight of wet soil sample along with container. V2 =
Volume of soil at shrinkage limit.
W2 =Weight of soil sample at shrinkage limit.
Vs=Weight of dry soil pat
γw= Unit weight of water

Shrinkage limit is nothing but water content at a particular case i.e, ws


𝑊𝑤
𝑤𝑠 = × 100
�𝑠

Weight of water at shrinkage limit is,
𝑊𝑤𝑠 = (𝑊1 − 𝑊2) − (𝑉1 − 𝑉2)𝛾𝑤

29
Lecture Note on GTE-I

=
(𝑊1−𝑊2)−(𝑉1−𝑉2)𝛾𝑤
𝑤𝑠
Hence, Shrinkage Limit
------------------------ Eq.2.13
𝑊𝑆

2.7 CONSISTENCY INDICES


2.7.1 Plasticity Index (Ip)

It is the range of water content in soil in during which soil behaves as a plastic material
Simply it is the difference in the water content at liquid limit and plastic limit
Highly plastic soils exhibit greater plasticity values

Plasticity Index 𝐼𝑃 = (𝑤𝑙 − 𝑤𝑝)-------------------- Eq.2.14

2.7.2 Shrinkage Index (Is)

It is the range of water content in soil in during which soil behaves as a semi solid material
Simply it is the difference in the water content at plastic limit and shrinkage limit

𝐼𝑠 = (𝑤𝑝 − 𝑤𝑠)------------------ Eq.2.15


Plasticity Index (Ip) Soil Description
0 Non Plastic
<7 Low plastic
7 - 17 Medium plastic
>17 Highly plastic

2.7.3 Flow Index

It is the slope of the flow curve obtained between the number of blows and water content in the
Casagrande’s test for the determination of liquid limit.
Flow index indicates the loss of shear strength of soil upon the increase of water.
Greater the slope angle less is the shear strength of soil.
Flatter slope curve indicates more shear strength in the soil.
Flow index = (𝑊1−𝑊2)------------------------
𝐼𝐹 𝑙𝑜𝑔10𝑛
Eq.2.16
2
𝑛1
2.7.4 Consistency Index (Ic)
It is the ratio of the difference between the liquid limit and natural water content of soil to its
plasticity index. This is very important to attribute the soil property
=
(𝑤𝑙−𝑤𝑛)

𝐼𝐶
Consistency Index ------------------- Eq.2.17
𝐼𝑃

30
Lecture Note on GTE-I

Condition Consistency Soil’s nature


Index
If wl = wn Ic = 0 Soil is at liquid limit.
If wl < wn Ic = Negative Soil is in liquid state.
If wl > wn Ic = Positive Soil is either in solid or semi solid
state

2.7.5 Liquidity Index (IL)

It is the ratio of the difference between the natural water content of soil and its plastic limit to its
plasticity index.
Liquidity index varies from 0 to 1
(𝑤𝑛−𝑤𝑝)
Liquidity Index, IL = ------------------ Eq.2.18
𝐼𝑃
2.7.6 Toughness Index (IT)
It is the ratio of plasticity index to the flow index.
This gives an idea of shear strength of soil at plastic limit
𝐼𝑃
Toughness Index, IT = -------------------------- Eq.2.19
𝐼𝐹
2.7.7 Density Index (ID)

It is the ratio of difference between the void ratio of a Cohesionless soil in the loosest state
and void ratio in its natural state to the difference between its void ratios in the loosest and
(𝑒𝑚𝑎𝑥−𝑒𝑛𝑎𝑡)
densest states.

(𝑒𝑚𝑎𝑥−𝑒𝑚𝑖𝑛)
Hence Density Index ID= ------------------------- Eq.2.20

The nature of soil can be predicted from Density Index value as given in table
Density Index (ID) Soil Description
<15 Very loose 15 - 35 Loose
35 - 65 Medium 65- 85 Dense
>85 Very dense 5/
<15 Very loose 15 - 35 Loose
35 - 65 Medium 65- 85 Dense
2.7.8 Shrinkage Ratio (SR)
It is the ratio of a given volume change in a soil, expressed as a percentage of the dry volume, to
the corresponding change in water content above the shrinkage limit.

31
Lecture Note on GTE-I

× 100-----------------------
𝑉1−𝑉2
SR=( 𝑉𝑑 )
𝑤1−𝑤2
Eq.2.21

Where
V1 = Volume of soil at water content w1 (%)
V2 = Volume of soil at water content w2 (%)
Vd = Volume of dry soil mass
At Shrinkage,
V2 = Vd
w2= ws
𝑉1−𝑉𝑑
(
Shrinkage Ratio (SR) = 𝑉𝑑
) × 100-------------------- Eq.2.22
𝑤1−𝑤𝑠
2.8 MISCELLANEOUS PROPERTIES
2.8.1 Sensitivity (St)
It is defined as the ratio of unconfined compressive strength of an undisturbed soil sample to the
unconfined compressive strength of the same soil after remolding at unaltered water content.
Sensitivity Soil Description
1-4 Normal
4-8 Sensitive
8 - 15 Extra sensitive
>15 Quick
2.8.2 Activity

It is the ratio of plasticity index of soil to the % finer fraction available in it.
Clayey soils exhibit more activity values.
Activity Classification Activity Classification
<0.75 Inactive
0.75 – 1.25 Normal
>1.25 Active

32
Lecture Note on GTE-I

Sol:
Data given

33
Lecture Note on GTE-I

Alternatively, this can be found from three phase diagram

34
Lecture Note on GTE-I

3.0 CLAY MINERALOGY

A soil particle may be a mineral or a rock fragment. A mineral is a chemical compound formed
in nature during a geological process, whereas a rock fragment has a combination of one or more
minerals. Based on the nature of atoms, minerals are classified as silicates, aluminates, oxides,
carbonates and phosphates.

Out of these, silicate minerals are the most important as they influence the properties of clay
soils. Different arrangements of atoms in the silicate minerals give rise to different silicate
structures.

3.1 Basic Structural Units

Soil minerals are formed from two basic structural units: tetrahedral and octahedral. Considering
the valencies of the atoms forming the units, it is clear that the units are not electrically neutral
and as such do not exist as single units.

The basic units combine to form sheets in which the oxygen or hydroxyl ions are shared among
adjacent units. Three types of sheets are thus formed, namely silica sheet, gibbsite
sheet and brucite sheet.

Isomorphous substitution is the replacement of the central atom of the tetrahedral or octahedral
unit by another atom during the formation of the sheets.

The sheets then combine to form various two-layer or three-layer sheet minerals. As the basic
units of clay minerals are sheet-like structures, the particle formed from stacking of the basic
units is also plate-like. As a result, the surface area per unit mass becomes very large.

A tetrahedral unit consists of a central silicon atom that is surrounded by four oxygen atoms
located at the corners of a tetrahedron. A combination of tetrahedrons forms a silica sheet.

35
Lecture Note on GTE-I

An octahedral unit consists of a central ion, either aluminium or magnesium, that is surrounded
by six hydroxyl ions located at the corners of an octahedron. A combination of aluminium-
hydroxyl octahedrons forms a gibbsite sheet, whereas a combination of magnesium-hydroxyl
octahedrons forms a brucite sheet.

Kaolinite and halloysite clay minerals are the most common.


3.1.1 Kaolinite Mineral
The basic kaolinite unit is a two-layer unit that is formed by stacking a gibbsite sheet on a silica
sheet. These basic units are then stacked one on top of the other to form a lattice of the mineral.
The units are held together by hydrogen bonds. The strong bonding does not permit water to
enter the lattice. Thus, kaolinite minerals are stable and do not expand under saturation.
Kaolinite is the most abundant constituent of residual clay deposits.

Fig.3.1: Kaolinite Structure


3.1.2 Halloysite Mineral
The basic unit is also a two-layer sheet similar to that of kaolinite except for the presence of
water between the sheets.

36
Lecture Note on GTE-I

Montmorillonite and illite clay minerals are the most common. A basic three-layer sheet unit
is formed by keeping one silica sheet each on the top and at the bottom of a gibbsite sheet.
These units are stacked to form a lattice as shown.
3.1.3 Montmorillonite Mineral
The bonding between the three-layer units is by vander Waals forces. This bonding is very weak
and water can enter easily. Thus, this mineral can imbibe a large quantity of water causing
swelling. During dry weather, there will be shrinkage.

Fig.3.2: Structure of Montomorillonite Mineral


3.1.4 Illite Mineral

Illite consists of the basic Montmorillonite units but are bonded by secondary valence
forces and potassium ions, as shown. There is about 20% replacement of aluminium with
silicon in the gibbsite sheet due to isomorphous substitution. This mineral is very stable and
does not swell or shrink.

Fig.3.3: Illite Mineral Structure


Natural soils are rarely the same from one point in the ground to another. The content and nature
of grains varies, but more importantly, so does the arrangement of these. The arrangement and
organisation of particles and other features within a soil mass is termed its fabric.
CLAY particles are flaky. Their thickness is very small relative to their length & breadth, in
some cases as thin as 1/100th of the length. They therefore have high specific surface values.
These surfaces carry negative electrical charge, which attracts positive ions present in the pore
water. Thus a lot of water may be held as adsorbed water within a clay mass.

37
Lecture Note on GTE-I

Fig. 3.4: Schematic diagram of Clay Structure

3.2 IMPORTANT PROPERTIES OF CLAY MINERALS


Some of the important properties that influence the behaviour of clay minerals are presented
below:
3.2.1 High surface area

Specific surface area (SSA) is defined as the surface area of soil particles per unit mass (or
volume) of dry soil. Its unit is in m2/g or m2/m3. Clay minerals are characterized by high specific
surface area (SSA) as listed in Table 2.2. High specific surface area is associated with high soil-
water-contaminant interaction, which indicates high reactivity. The reactivity increases in the
order Kaolinite< Illite< Montmorillonite. For the purpose of comparison, SSA of silt and sand
has also been added in the table. There is a broad range of SSA values of soils, the maximum
being for montmorillonite and minimum for sand. As particle size increases SSA decreases.

For smectite type minerals such as montmorillonite, the primary external surface area amounts to
50 to 120 m2/g. SSA inclusive of both primary and secondary surface area (interlayer surface
area exposed due to expanding lattice) and termed as total surface area would be closed to 800
m2/g.For kaolinite type minerals there is possibility of external surface area where in the
interlayer surface area does not contribute much.
3.2.2 Plasticity and cohesion

Clay attracts dipolar water towards its surface by adsorption. This induces plasticity in clay.
Therefore, plasticity increases with SSA. Water in clays exhibits negative pressure due to which
two particles are held close to each other. Due to this, apparent cohesion is developed in clays.
Surface charge and adsorption
Clay surface is charged due to following reasons:

Iso-morphous substitution (Mitchell and Soga 2005): During the formation of mineral, the
normally found cation is replaced by another due to its abundant availability. For example: when
Al+3 replace Si+4 there is a shortage of one positive charge, which appears as negative charge on

38
Lecture Note on GTE-I

clay surface. Such substitution is therefore the major reason for net negative charge on clay
surface.
O-2 and OH- functional groups at edges and basal surface also induce negative charge.
Dissociation of hydroxyl ions or broken bonds at the edges is also responsible for unsatisfied
negative or positive charge. Positive charge can occur on the edges of kaolinite plates due to
acceptance of H+ in the acid pH range (Berkowitz et al. 2008). It can be negatively charged under
high Ph environment.

Absence of cations from the crystal lattice also contributes to charge formation.
In general, clay particle surface are negatively charged and its edges are positively charged.
Due to the surface charge, it would adsorb or attract cations (+ve charged) and dipolar molecules
like water towards it. As a result, a layer of adsorbed water exists adjacent to clay surface.
3.2.3 Exchangeable cations and cation exchange capacity

Due to negative charge, clay surface attracts cations towards it to make the charge neutral. These
cations can be replaced by easily available ions present in the pore solution, and are termed as
exchangeable ions. The total quantity of exchangeable cations is termed as cation exchange
capacity, expressed in milli equivalents per 100 g of dry clay. Cation exchange capacity (CEC) is
defined as the unbalanced negative charge existing on the clay surface. Kaolinite exhibits very
low cation exchange capacity (CEC) as compared to montmorillonite. Determination of CEC is
done after removing all excess soluble salts from the soil. The adsorbed cations are then replaced
by a known cation species and the quantity of known cation required to saturate the exchange
sites is determined analytically.

3.3 FLOCCULATION AND DISPERSION

When two clay particles come closer to each other it experiences (a) interparticle attraction due
to weak van-der-Waal’s force (b) repulsion due to –ve charge. When particles are sufficiently
close, attraction becomes dominant active force and hence there is an edge to face configuration
for clay particles as shown in Fig. 3.5 a. Such a configuration is termed as flocculant structure.
When the separation between clay particles increase, repulsion becomes predominant and hence
the clay particles follows face to face configuration called dispersed structure. A lot of micro and
macro level behaviour of clays are associated with these arrangement of clay particles (Mitchell
and Soga 2005).

39
Lecture Note on GTE-I

Fig. 3.5: Different arrangement of clay particle


3.4.1 Swelling and shrinkage
Some clay minerals when exposed to moisture are subjected to excessive swelling and during
drying undergo excessive shrinkage. A lot of engineering properties of soil is affected by this
behaviour and the stability of structures founded on such soils become detrimental. The swelling
of clay minerals decreases in the order montmorillonite> illite> kaolinite.
3.5 MINERALS OTHER THAN SILICA AND CLAY

Other than silica and clay, subsurface contains a variety of minerals such as oxides and
carbonates that governs the reactivity of soil and its interaction with the environment. Some of
the abundant metal oxide minerals present are iron oxides (hematite, magnetite, goethite etc.) and
aluminium oxides (gibbsite, boehmite). Other oxide minerals (such as manganese oxide, titanium
oxide) are far less than Fe and Al oxides, but because of small size and large surface area, they
would affect very significantly the geochemical properties of subsurface. These oxides are
mostly present in residual soils of tropical regions. Other major components include soluble
calcium carbonate and calcium sulphate, which has relatively high surface area. In most soils,
quartz is the most abundant mineral, with small amount of feldspar and mica present. Carbonate
minerals such as calcite and dolomite are found in some soils in the form of bulky particles,
precipitates etc. Sulphate minerals mainly gypsum are found in semiarid and arid regions.

3.6 SOIL MINERALOGY CHARACTERIZATION

One of the very well established methods for mineralogy characterization of fine-grained soils is
by using X-ray diffraction (XRD) analysis. Majority of the soil minerals are crystalline in nature
and their structure is defined by a unique geometry. XRD identifies minerals based on this
unique crystal structure. In XRD, characteristic X-rays of particular wave length are passed
through a crystallographic specimen. When X-ray interacts with crystalline specimen it gives a
particular diffraction pattern, which is unique for a mineral with a particular crystal structure.
The diffraction pattern of the soil specimen (according to its crystal structure), which is based on
powder diffraction or polycrystalline diffraction, is then analyzed for the qualitative and
quantitative (not always) assessment of minerals. Sample preparation method for XRD should be
done with great care as the XRD reaches only a small layer (nearly 50 µm) from the surface of
the sample. Hence, homogeneity is very important. Soil sample is initially dried and sieved
through 2 mm sieve. Sieved sample is homogenized in a tumbler mixer for 30 min. A control
mix of 30 g was taken and ground in lots of 15 g in a gyratory pulverizer. 15 percent by weight

40
Lecture Note on GTE-I

of KIO4 (internal standard) was added to 5 g of specimen and again homogenized in a mixer..
X-ray wave of monochromatic radiation (K α) is commonly obtained from copper radiation,
which is commonly known as Cu- Kα. A typical XRD output is represented by Fig. 2.5. It can be
noted from the figure that ordinate represent relative intensity of X-ray diffraction and abscissa
represents twice of angle at which a striking X-ray beam of wave length λ makes with parallel
atomic planes. Based on this diffraction pattern, the minerals can be identified by matching the
peak with the data provided by International Centre Diffraction Data (ICDD) formerly known as
Joint Committee on Powder Diffraction Standards (JCPDS).

It is understood that the area under the peak of diffraction pattern gives the quantity of each
phase present in the specimen. However, quantitative determination of mineral composition in
soils based on simple comparison of diffraction peak heights or area under peak is complex and
uncertain because of different factors such as mineral crystallinity, hydration, surface texture of
the specimen, sample preparation, non-homogeneity of soil samples, particle orientation etc. The
method of quantification will be more precise for those soils with less number of minerals. Al-
Rawas et al. (2001) have discussed about constant mineral standard method and constant clay
method for quantification of clay minerals

Fig.3.6: A typical XRD pattern with mineral identification for two different soils (modified from
Sreedeep 2006)
In the first method, increasing quantity of clay are added to the fixed mass of known standard
and the difference in X-ray diffraction intensity when the specimen changes from 100 percent
standard to 100 percent clay is noted. The peak area ratio for each component is then plotted
against percentage of clay, based on which regression equation is determined. This regression
equation is further used for mineral quantification. In the second method, known weight of pure
standard mineral is added to clay containing the same components, and the change in the

41
Lecture Note on GTE-I

reflection peak-area intensity of each component is measured to estimate the weight proportion
of that component.

3.7 INFLUENCE OF WATER ON CLAY MINERAL BEHAVIOUR


In dry clay, adsorbed cations are bound tightly to the soil surface. The cations in excess, and their
associated anions are present as a salt precipitate. When this dry clay is placed in water, the salts
go into solution and the adsorbed cations are hydrated. The negatively charge mineral surface
and the distributed charge in the adjacent phase are together known as the diffuse (or electrical)
double layer. Factors which affect the nature and extent of the double layer include: (i)
composition and concentration of the electrolyte solution, (ii) density of charge on the surface of
the particle, (iii) valence of the cations, (iv) dielectric constant of the medium, and (v)
temperature. Mathematical theories concerning the double layer which include these factors are
well summarised in Lee (1968), and Mitchell (1977).
Previously it has been mentioned that clay particles can be of colloid size. In solution, the
particles will move in a random manner. If the attraction forces between particles exceed forces
of repulsion, the particles will increase in size - up to a stage where gravitational forces will
predominate and the particles will sediment out, i.e. the system flocculates. The forces
controlling attraction and repulsion can be classified as: (a) Independent of a system: eg.
electrostatic or electromagnetic (Van der Waals) or (b) System Dependent: eg. electrostatic
(electron or ion clouds) or ion hydration. Electrostatic and electromagnetic forces both contribute
to attraction, whilst electron or ion clouds and ion hydration are forces of repulsion, with ion
hydration being the major repulsive force.
Thus the behaviour of clay minerals in suspension can be controlled by changing the
environment. For the Cardinia Creek reservoir turbidity of the water was reduced by introducing
gypsum - the thickness of the double layer of the clay particles in suspension was reduced,
leading to forces of attraction exceeding repulsion forces. For this case the enlarged particles
settled out. (Grant et al, 1976). Any factor which changes the “thickness” of the double layer will
alter the characteristics displayed by a clay soil. Two of the more important system variables are:
(a) Cation Valence: an increase in valence suppresses the concentration of the solution between
adjacent particles, and the potential between particles, leading to a decrease in inter-particle
repulsion. (b) Ion Size: the larger the ion size the thicker the layer required to accommodate the
necessary number of cations, hence the greater the repulsion. Associated with both of these
factors is the phenomenon that multi-valent ions are adsorbed preferentially onto the mineral
surface. For a given number of exchange sites, the higher the valence the lower the number of
ions required to satisfy charge neutrality and hence the double layer is affected accordingly. This
feature can be detected through very simple soil tests.
Na Kaolinite
wL = 53%, wp= 32%
Ca Kaolinite

42
Lecture Note on GTE-I

wL = 38%, wp= 27%

Since the difference between wL and wp is indicative of the compressibility of the soil we can
argue that a Na Kaolinite would be more compressible than a Ca Kaolinite. When the double
layer and mineral particle size are considered together it can be argued that the smaller the
particle (and hence the higher the specific surface) the larger the amount of water that (with
respect to the mass of dry soil) will be associated with the double layer. This feature is of great
importance when the swell of a soil, when wetted from a dry condition, is considered. Of
particular significance is the clay mineral montmorillonite: the bonding between successive
layers is weak and when an ion associated with this interlayer position hydrates, the clay mineral
stack will expand. Expansion associated with moisture can cause structural failure in some
lightly loaded structures.
3.8 STRUCTURE OF SOILS
The ‘structure’ of a soil may be defined as the manner of arrangement and state of aggregation of
soil grains. In a broader sense, consideration of mineralogical composition, electrical properties,
orientation and shape of soil grains, nature and properties of soil water and the interaction of soil
water and soil grains, also may be included in the study of soil structure, which is typical for
transported or sediments soils. Structural composition of sedimented soils influences, many of
their important engineering properties such as permeability, compressibility and shear strength.
Hence, a study of the structure of soils is important.
The following types of structure are commonly studied:
(a) Single-grained structure
(b) Honey-comb structure
(c) Flocculent structure
3.8.1 Single-grained Structure

Single-grained structure is characteristic of coarse-grained soils, with a particle size greater than
0.02mm. Gravitational forces predominate the surface forces and hence grain to grain contact
results. The deposition may occur in a loose state, with large voids or in a sense state, with less
of voids

43
Lecture Note on GTE-I

3.8.2 Honey-comb Structure


This structure can occur only in fine-grained soils, especially in silt and rock flour. Due to the
relatively smaller size of grains, besides gravitational forces, inter-particle surface forces also
play an important role in the process of settling down. Miniature arches are formed, which bridge
over relatively large void spaces. This results in the formation of a honeycomb structure, each
cell of a honey-comb being made up of numerous individual soil grains. The structure has a large
void space and may carry high loads without a significant volume change. The structure can be
broken down by external disturbances
.

3.8.3 Flocculent Structure


This structure is characteristic of fine-grained soils such as clays. Inter-particle forces play
predominant role in the deposition. Mutual repulsion of the particles may be eliminated by
means of an appropriate chemical; this will result in grains coming closer together to form a
‘floc’. Formation of flocs is ‘flocculation’. But the flocs tend to settle in a honeycomb structure,
in which in place of each grain, a floc occurs. Thus, grains grouping around void spaces larger
than the grain-size are flocs and flocs grouping around void spaces larger than even the flocs
result in the formation of a ‘flocculent’ structure. Very fine particles or particles of colloidal size
(< 0.001 mm) may be in a flocculated or dispersed state. The flaky particles are oriented edge-to-
edge or edge-to-face with respect to one another in the case of a flocculated structure. Flaky
particles of clay minerals tend to from a card house structure (Lambe, 1953), when flocculated.
When inter-particle repulsive forces are brought back into play either by remoulding or by the
transportation process, a more parallel arrangement or reorientation of the particles occurs, as
shown in Fig. This means more face-to-face contacts occur for the flaky particles when these are
in a dispersed state. In practice, mixed structures occur, especially in typical marine soils.

41
Lecture Note on GTE-I

4.0 SOIL CLASSIFICATION SYSTEM


Classification systems are used to group soils according to their order of performance under
given set of physical conditions. Soils that are grouped in order of performance for one set of
physical conditions will not necessarily have the same order of performance under some other
physical conditions.
Classification systems are used to group soils according to their order of performance under
given set of physical conditions. Soils that are grouped in order of performance for one set of
physical conditions will not necessarily have the same order of performance under some other
physical conditions.
4.1 Different Classification of Soils for Engineering Purpose
1. Classification based on grain size
2. Textural classification
3. AASHTO classification system
4. Unified soil classification system
4.1.1 Grain Size Classification System for Soils
Grain size classification systems were based on grain size. In this system the terms clay, silt,
sand and gravel are used to indicate only particle size and not to signify nature of soil type.
There are several classification systems fin use, but commonly used systems are shown here.
2.00 0.6 0.2 0.06 0.02 0.006 0.002
MIT Gravel C M F C M F Clay
Sand Silt
4.1.2 Textural Classification of Soil
The classification of soil exclusively based on particle size and their percentage distribution is
known as textural classification system. This system specifically names the soil depending on the
percentage of sand, silt and clay. The triangular charts are used to classify soil by this system.

Fig-1: Textural Classification of U.S. Public Roads Administration

42
Lecture Note on GTE-I

4.1 .3 AASHTO classification system of Soil


80 20 4.75 0.075 0.002
ISI C F C M F Silt Clay
Gravel Sand
AASHTO classification is otherwise known as PRA classification system. It was originally
developed in 1920 by the U.S. Bureau of Public Roads for the classification of soil for highway
sub-grade use.
This system is developed based on particle size and plasticity characteristics of soil mass. After
some revision, this system was adopted by the AASHTO in 1945.
In this system the soils are divided into seven major groups. Some of the major groups further
divided into subgroups. A soil is classified by proceeding from left to right on the classification
chart to find first the group into which the soil test data will fill.
Soil having fine fractions are further classified based on their group index. The group index is
defined by the following equation.
Group index = (F – 35)[0.2 + 0.005 (LL – 40)] + 0.01(F – 15)(PI – 10)
F – Percentage passing 0.075mm size
LL – Liquid limit
PI – Plasticity index
When the group index value is higher, the quantity of the material is poorer.

General Granular Materials(35% 0r less passing 0.075mm sieve) Silt-clay materials


classification More than 35% passing 0.075mm
Group A-1 A-3 A-2 A-4 A-5 A-6 A-7-5
classification A-1-a A-1-b A-2-4 A-2-5 A-2-6 A-2-7 A-7-6
(a) Sieve
analysis
% passing
i) 2.00 mm 50max
ii) 0.425 mm 30max 50max 51min
iii)0.075mm 15 max 25max 10max 35max 35max 35max 35max 36min 36min 36min 36min
(b) Characteris
tics passing
through 0.425
mm
i) Liquid limit 40max 41min 40max 41min 40max 41min 40max 41min
ii) Plastic limit 6 max NP 10max 10max 11min 11min 10max 10max 11min 11min*
(c)Usual type Stone Fragments Fine Silty or Clayey Gravel Sands Silty Soils Clayey Soils
of significant Gravel and Sands sands
constituent
materials
(d)General Excellent to Good Fair to Poor
rating as
Subgrade
Note: *If Plasticity index is equal to or less than (Liquid limit-30), the soil is A-7-5(PL>30%)
If Plasticity index greater than (Liquid limit-30), the soil is A-7-6(PL<30%)

43
Lecture Note on GTE-I

4.1.4 Unified Soil Classification System


Unified soil classification system was originally developed by Casagrande (1948) and was
known as airfield classification system. It was adopted with some modification by the U.S.
Bureau of Reclamation and the U.S. Corps of Engineers.
This system is based on both grain size and plasticity characteristics of soil. The same system
with minor modification was adopted by ISI for general engineering purpose (IS 1498 – 1970).
IS system divides soil into three major groups, coarse grained, fine grained and organic soils and
other miscellaneous soil materials.
Coarse grained soils are those with more than 50% of the material larger than 0.075mm size.
Coarse grained soils are further classified into gravels (G) and sands (S). The gravels and sands
are further divided into four categories according to gradation, silt or clay content.
Fine grained soils are those for which more than 50% of soil finer than 0.075 mm sieve size.
They are divided into three sub-divisions as silt (M), clay (c), and organic salts and clays (O),
based on their plasticity nature they are added with L, M and H symbol to indicate low plastic,
medium plastic and high plastic respectively.
Examples:
GW – well graded gravel. GP – poorly graded gravel, GM – Silty gravel
SW – Well graded sand, SP – Poorly graded sand, SM – Silty sand, SC – Clayey sand
CL – Clay of low plastic, CI – Clay of medium plastic, CH – Clay of higher plastic
ML – Silt of medium plastic, MI – Silt of medium plastic, MH – Silt of higher plastic
OL – Organic silt and clays of low plastic, OI – Organic silt and clays of medium plastic
OH – Organic silt and clays of high plastic.
Fine grained soils have been sub-divided into three subdivisions of low, medium and high
compressibility instead of two sub-divisions of the original Unified Soil Classification System.
Table-2: Significance of letters for group symbol in table-3.
Soil Soil Component Symbol
Coarse Grained Boulder None
Cobble None
Gravel G
Sand S
Fine Grained Silt M
Clay C
Organic Matter O
Peat Peat Pt
Applicable to Coarse grained Soils Well graded W
Poorly Graded P
Applicable to Fine grained soils Low compressibility W L <35 L
Medium compressibility I
(WL 35 to 50)

44
Lecture Note on GTE-I

The standard recommends that when a soil possesses characteristics of two groups either in
particle size distribution or in plasticity, it is designed by combination of group symbols.

Table 3: Unified Soil Classification System (based on particle passing through 75 mm sieve)

4.2 INDIAN STANDARD CLASSIFICATION SYSTEM FOR SOIL


Indian Standard Classification System (ISC) was adopted by Bureau of Indian Standards is in
many respect similar to the Unified Soil Classification (USC) system.
Soils are divided into three broad divisions:
1. Coarse grained soils, when 50% or more of the total material by weight is retained on 75
micron IS sieve.
2. For fine grained soils, when more than 50% of the total material passes through
75 micron IS sieve.
3. If the soil is highly organic and contains a large percentage of organic matter and
particles of decomposed vegetation, it is kept in a separate category marked as peat
(Pt).
In all there are 18 groups of soils: 8 groups of coarse grained, 9 groups of fine grained and one of

45
Lecture Note on GTE-I

Table 4: Basic Soil Components in ISC System


Soil Soil components Symbol Particle size range and description
Boulder None Rounded to angular, hard, rock, particle average
diameter more than 300 mm
Cobble None Rounded to angular, hard, rock, particle average
diameter smaller than 300 mm but retained on 80
mm IS sieve.
Gravel G Rounded to angular, hard, rock, particle average
diameter smaller than 80 mm but retained on 4.75
Coarse grained mm IS sieve.
Coarse : 80 to20 mm IS sieve
Fine: 20 mm to 425 micron IS sieve
Sand S Rounded to angular, hard, rock, particle passing
through 4.75 mm but retained on 75 micron IS
sieve.
Coarse: 4.75 mm to 2 mm IS sieve
Medium: 2.00 mm to 425 micron
Fine: 425 micron to 65 micron IS sieve
Fine grained Silt M Particle smaller than 75 micron , Identify behavior
that slightly plastic or non-plastic regardless of
moisture and exhibits little or no strength when air
dried.
Clay C Particle smaller than 75 micron , Identify behavior
that can exhibit
Plastic properties within certain considerable
strength when air dried
Organic matter O Organic matter in various sizes and stage of
decomposition.

46
Lecture Note on GTE-I

Table: 5 Classification of Coarse-grained Soils (ISC) System


Division Sub-division Group Typical Laboratory Criteria Remark

𝐶𝑢 > 4
Symbol Names

𝐶𝑐between 1 &3
Coarse Gravel Clean GW Well graded When
grained soil More than gravel gravel fines are
(More than half of (Fine less GP Poor graded Not meeting all between
half of material is than 5%) gravel requirement for GW
5% to
material is larger than Gravel with GM Silty gravel Atterberg Atterberg
12%
larger than the 4.75 mm fines (Fines limit limits
the 75 – IS sieve more than below A plotting boarder

𝐼𝑝 < 4 line with


micron IS 12%) line or above A line
sieve cases
GC Clayey gravel Atterberg Ip requiring
limit between use of
above A 4 & 7 are dual

𝐼𝑝 > 7 line
line or boarder
symbol
cases GW-
requiring GP.SW-
use of SC etc
dual
symbol

𝐶𝑢 > 6
GM-GC

𝐶𝑐between 1 &3
Sands Clean sand SW Well graded
More than (Fine less sands
half of than 5%) SP Poor graded Not meeting all
material is sands requirement for SW
smaller Sands with SM Silty Sands Atterberg Atterberg
than the fines (Fines limit limits
4.75 mm IS more than below A plotting
sieve
𝐼𝑝 < 4 line with
12%) line or above A

SC Clayey Sands Atterberg Ip


limit between
above A 4 & 7 are

𝐼𝑝 > 7 line
line or boarder

cases
requiring
use of
dual
symbol
SM-SC

47
Lecture Note on GTE-I

Table: 6 Classification of Fined-grained Soils (ISC) System


Division Sub-division Group Typical Names Laboratory Criteria Remark
Symbol
Fine Low ML Organic silts Atterberg limit Atterberg
grained compressibility with none to low below A line limits Black

𝐼𝑝 < 7
soil (L)Liquid limit plasticity or plotting cotton soils
(More less than 35% in India
above A line
than CL Inorganic clays Atterberg limit lies partly
half of with Ip above and
with low above A line

𝐼𝑝 > 7
material plasticity or between 4 & partly
is 7 (hatched below the
smaller zone) ML- A-Line
than the CL
75 µ IS
sieve ) OL Organic silts Atterberg limit below A line Organic
with low and
plasticity inorganic
MI Inorganic silts Atterberg limit below A line soils
Intermediate with medium plotted in
compressibility plasticity same zone
(I) CI Inorganic clays Atterberg limit above A line of plasticity
Liquid limit with medium chart are
more than 35% plasticity distinguish
but less than OI Organic silts Atterberg limit below A line ed by
50% with medium colour,
plasticity odor or
MH Inorganic silts Atterberg limit below A line liquid limit
High with High test after
compressibility plasticity oven
(H) CH Inorganic clays Atterberg limit above A line drying.
Liquid limit with High Reduction
more than 50% plasticity in value
after and
OH Organic silts Atterberg limit below A line
before
with medium to
drying by
High plasticity
three fourth
indicates
clear
organic
soils

Field Identification of Soils


In field identification of soil, the engineer concerned first determines whether the soil is coarse
grained or fine grained. To make this determination, soil sample is spread on a flat surface. If
more than half of the particles are visible to the naked eye, then it is classified as coarse grained
or otherwise it is classified as fine grained. If the soil is coarse gained, follow the procedures
outlined under the heading coarse grained soil; if the soil is fine grained, follow the procedure
mentioned under the heading of fine grained soil.
48
Lecture Note on GTE-I

Coarse Grained Soil:


Once the soil has been determined as coarse grained, further examination is required to
determine the grain size distribution, the grain shape and gradation of coarse grained soils.
Coarse grained soil is classified as cobble or sand depending on whether more than half of the
coarse fraction is of cobble size (76 mm or larger) or sand size (5 mm to 0.074 mm). Soil
particles can also be described according to a characteristic shape.
Particle shape may vary from angular to round to flat or elongated. Coarse grained soil may be
described as well graded, poorly graded or gap graded. A soil is said to well grade if it has a
good representation of all grain sizes. If the soil grains are approximately of same size, then the
sample is described as poorly graded. A soil is said to be gap graded if the intermediate grain
sizes are absent
Fine Grained Soil:
Following field tests are performed to classify fine grained soil or for the fine fraction of coarse
grained soil
(i) Dilatancy test:
Prepare a part of moist soil having a volume equivalent to a 25 mm cube by adding enough water
to make the soil soft but not-sticky. Place the pat in the open palm of one hand and shake
horizontally by striking against the other hand several times. If the reaction is positive, water
appears on the surface of the pat giving a glossy appearance. On squeezing the sample between
the fingers the water and glossiness disappear from the surface, the soil becomes stiff and cracks.
The phenomenon of appearance of water on the surface of soil on shaking and disappearance on
squeezing, followed by cracking is called as “dilatancy”. The rapidity of appearance and
disappearance of water from the surface of the soil help to identify the character of fines in the
soil. Table 3.6 shows character of fines in soil w.r.t. the positive reactions.
(ii) Dry strength test:
Prepare a part of soil to the consistency of putty by adding water. Allow the pat to dry by oven,
sun or air. The strength is tested by breaking and crumbling the dry pat between the fingers. Dry
strength of soil increases with increasing plasticity. Clays have high dry strength and silts have
slight dry strength.
(iii) Toughness test:
Take a part of soil to the consistency of putty, add water or allow drying as necessary. Roll the
soil between the palms into a thread of 3 mm diameter. Fold the thread of soil and repeat the
procedure a number of times till the thread starts crumbling when rolled into 3 mm diameter. The
crumbled pieces are lumped together and subjected to kneading until the lump crumbles. The
threads are stiffer and lumps are tougher at plastic limit for soils having higher clay contents.
(iv) Dispersion test:
Pour small quantity of soil in a jar of water. Shake the jar containing soil and water and allow the
soil to settle. The coarser particles settle first followed by finer ones. Sands settle in about 30 to
60 sees, silts settle in 30 to 60 mins and clay particles remains in suspension for at least several
hours.

49
Lecture Note on GTE-I

(v) Bite lest:


Take a pinch of soil and place between the teeth and grind lightly. Fine sand is felt gritty. Silt
have rough feeling but do not stick to the teeth, clays have smooth feeling and stick to the teeth.
(vi) Colour and Odour test:
Organic soils have darker colours like dark grey, dark brown etc. and a musty odour. The odour
can be more noticeable by heating a wet sample. Inorganic soils have clean, bright colours like
light grey, brown, red, yellow or white.

5.0 SOIL WATER AND WATER FLOW


INTRODUCTION
All soils are permeable materials, water being free to flow through the interconnected pores
between the solid particles. The pressure of the pore water is measured relative to atmospheric
pressure and the level at which the pressure is atmospheric (i.e. zero) is defined as the water table
(WT) or the phreatic surface. Below the water table the soil is assumed to be fully saturated,
although it is likely that, due to the presence of small volumes of entrapped air, the degree of
saturation will be marginally below 100%.

SOIL WATER: Water presence in the voids of soil mass is called soil water. It can be
classified in several ways:

Broad classification: 1. Free water 2. Held water


a. Structural water b. Adsorbed water c. Capillary water
Classification on phenomenological basis
1. Ground water 2. Capillary water 3. Adsorbed water 4. Infiltrated water
Classification on structural aspect
1. Pore water 2. Solvate water 3. Adsorbed water 4. Structural water
Free water: Water is free to move through a soil mass under the influence of gravity.
Held water: It is the part of water held in the soil pores by some force existing within the pores.
Such water is not free to move under gravitational force
Adsorbed water: Adsorbed water is that water which the soil particles freely adsorb from
atmosphere by physical force of attraction and held by force of adhesion. Water is the vicinity of
soil particles subjected to an attractive force basically consists of two components. i) Attraction
of bipolar water to be electrical charged soil. ii) Attraction of dipolar water to the action in the
double layer, cation in turn attract to the particles.

Structural water: It is the water chemically combined in the crystal structure of the soil mineral.
Structural water cannot be separated or removed and also not removed by oven drying at 105-

50
Lecture Note on GTE-I

110°c.It can be destroyed at higher temperature which will destroy the crystal structure.
Infiltrated water Infiltrated water is the portion of surface precipitation which soaks into ground,
moving downwards through air containing zones.

Pore Water: It is cable of moving under hydrodynamic forces unless restricted in its free
movement such as when entrapped between air bubbles or retention by capillary forces.
Gravitational and capillary water are the two types of pore water.

Solvate Water: The water which forms a hydration shell around soil grains is solvate water. it
is subjected to polar electrostatic and binding forces.

Ground water: Subsurface water that fills the voids continuously and is subjected to no force
other than gravity is known as gravitational water.

Capillary water: The minute pores of soil serve as capillary tubes through which the moisture
rise above the ground water table. Capillary water is the soil moisture located within the
interstices and voids of capillary size of the soil. Capillary water is held in the interstices of soil
due to capillary forces. Capillary action or capillarity is the phenomenon of movement of water
in the interstices of a soil due to capillary forces. The capillary forces depend upon various
factors such as surface tension of water, pressure in water in relation to atmospheric pressure,
and the size and conformation of soil pores.

The pores of soil mass may be looked upon as a series of capillary tubes, extending vertically
above water table. The rise of water in the capillary tubes, or the fine pores of the soil, is due to
the existence of surface tension which pulls the water up against the gravitational force. The
height of capillary rise, above the ground water (or free water) surface depends upon the
diameter of the capillary tube (or fineness of the pores) and the value of the surface tension.
When a capillary tube is inserted in water, the rise of water will take place up to reach the
equilibrium. At this stage the rise of water in the tube is stopped.

5.1 SOIL SUCTION


The tensile stress in the meniscus circumferences caused in water is called the capillary tension
or the capillary potential. The capillary tension or capillary potential is the pressure deficiency,

which water is retained in a soil mass. It decreases linearly from a maximum value of ℎ𝑐𝛾𝑤 at
pressure reduction or negative pressure in the pore water (or the pressure below atmospheric) by

the level of the meniscus to zero value at the free water surface. The pressure deficiency in the
held water is also termed as soil suction or suction pressure. Soil suction is measured by the
height hc in centimeters to which a water column could be drawn by suction in a soil mass free
from external stress. The common logarithm of this height (cm) or pressure (g/cm2) is known as
the pF value (Schofield, 1935): pF = log10 (hc) Thus, a pF value of 2 represents a soil suction of
100 cm of water or suction pressure and capillarity of 100 g /cm2.
51
Lecture Note on GTE-I

5.1.1 Factors affecting soil suction


1. Particle size of soil
Smaller the size of the particles, smaller will be pore size resulting greater capillary rise and
hence greater suction
2. Water content
Smaller the water content, greater will be soil suction and attends its maximum value when the
soil is dry condition.
3. Plasticity index of soil mass
Soil suction will be greater in a soil which Hs more plasticity index.
4. Soil structure
Changes in the soil structure result in the change in the size of the interstices and hence soil
suction will change
5. History of wetting and drying
For a same soil, suction is greater during drying cycle than during wetting cycle
6. Soil density
Increase in density of soil results in decrease in the size of pores of the soil and hence increase in
suction. At the loose state the pore sizes will be more resulting a decrease in soil suction.
7. Temperature
Rise in temperature results in decrease in surface tension and hence decrease in suction.
8. Angle of contact
Soil suction decreases with increase in the angle of contact and will be maximum when angle of
contact becomes zero.
9. Dissolved salts in water
Impurities such as dissolved salts etc increases the surface tension resulting in increase in soil
suction.
5.2 CAPILLARITY PRESSURE:

The magnitude of the pressure is the same at all height above the free water surface. The
capillarity pressure transferred from grain to grain called as inter angular or effective pressure.

Capillary action (or) capillarity: It is the phenomenon of movement of water in the interstices
of a soil due to capillary forces. The capillary forces depend upon various factors depend upon
various factors such as surface tension of water, pressure in water in relation to atmospheric
pressure and three size and conformation of soil pores.

52
Lecture Note on GTE-I

Capillary Rise of Water in Soils

In contrast to capillary tubes the continuous voids in soils have a variable width. They
communicate with each other in all directions and constitute an intricate network of voids. When
water rises into the network from below, the lower part of the network becomes completely
saturated. In the upper part, however, the water occupies only the narrowest voids and the wider
areas remain filled with air. Fig. shows a glass tube filled with fine sand. Sand would remain
fully saturated only up to a height h' which is considerably smaller than h c. A few large voids

hc only in the smaller voids. The zone between the depths ( ℎ� 𝑐 − ℎ′ ) will remain
may effectively stop capillary rise in certain parts. The water would rise, therefore, to a height of


partially
saturated.

Fig.5.1: Capillary rise


The height of the capillary rise is greatest for very fine grained soils materials, but the rate of rise
in such materials is slow because of their low permeability. Fig.5.1 (b) shows the relationship
between the height of capillary rise in 24 hours and the grain size of a uniform quartz powder.
This clearly shows that the rise is a maximum for materials falling in the category of silts and
fine sands.
As the effective grain size decreases, the size of the voids also decreases, and the height of
capillary rise increases. A rough estimation of the height of capillary rise can be determined from
the equation,

= 𝑒𝐷 ---------------------
𝐶 Eq.5.1
𝑐
10
in which e is the void ratio, DIQ is Hazen's effective diameter in centimeters, and C is an
empirical constant which can have a value between 0.1 and 0.5 sq. cm.
5.2.1 Capillary Siphoning
53
Lecture Note on GTE-I

The capillary phenomenon which enables raising the water in capillary tube against the force of gravity
can also be taken place in voids of dry soil. In case of the earth dam, this may create serious
problems. Capillary forces are able to raise water against the force of gravity not only into
capillary tubes or the voids in columns of dry soil, but also into narrow open channels or V-
shaped grooves. If the highest point of the groove is located below the level to which the surface
tension can lift the water, the capillary forces will pull the water into the descending part of the
groove and will slowly empty the vessel. This process is known as capillary siphoning. The same
process may also occur in the voids of soil. For example, water may flow over the crest of an
impermeable core in a dam in spite of the fact that the elevation of the free water surface is
below the crest of the core as shown in

Fig. 5.2: Capillary Siphoning

In an earth dam, the central impervious core provides to check the seepage of water through the body of
dam. When the water level in the reservoir reaches HFL, the portion of the upstream of the dam
will be saturated. The water level in u/s of the pervious shell will be practically the same as HFL.
Now due to capillarity, the water will raise through the height ‘h’. If the top of the impervious
core is situated at height y<h, above the HFL, the capillary force will pull the water to the
downstream of the dam. This effect will slowly empty the reservoir. This process is known as
Capillary Siphoning action in earth dam.

This will not only result in the water loss but also damage the downstream and create instability to the
earth dam. So the height of the impervious core is a major factor for design consideration.

Contact moisture. Water can also be held by surface tension round the point of contact of two particles
(spheres) capillary water in this form is known as contact moisture (or) contact capillary water.
6.0 EFFECTIVE STRESS CONCEPTS IN SOIL
At any plane in a soil mass, the total stress or unit pressure σ is the total load per unit area. This pressure
may be due to i) self weight of soil ii) over burden on the soil. The total pressure consists of two
distinct components: inter granular pressure or effective pressure and the neutral pressure or pore
pressure. Effective pressure σ' is the pressure transmitted from particle through their point of
contact through the soil mass above the plane. Such a pressure, also termed as inter

54
Lecture Note on GTE-I

granular pressure, is effective in decreasing the voids ratio of the soil mass and in mobilizing its
shear strength. The neutral pressure or the pore water pressure or pore pressure is the pressure
transmitted through the pore fluid. Therefore, this pressure is also called neutral pressure (u).
Since the total vertical pressure at any plane is equal to the sum of the effective pressure and pore

𝜎 = 𝜎′ + 𝑢-------------------
water pressure we have
Eq.6.1

6.1 TOTAL AND EFFECTIVE STRESS IN SOILS


When a load is applied to soil, it is carried by the solid grains and the water in the pores.
The total vertical stress acting at a point below the ground surface is due to the weight of
everything that lies above, including soil, water, and surface loading. Total stress thus increases

Vertical total stress at depth z, 𝜎𝑣 = 𝛾𝑍


with depth and with unit weight.

Fig.6.1: Stresses acting on a soil element


At any plane in a soil mass, the total stress or unit pressure σ is the total load per unit area. This
pressure may be due to i) self weight of soil ii) over burden on the soil. The total pressure
consists of two distinct components: inter granular pressure or effective pressure and the neutral
pressure or pore pressure. Effective pressure σ' is the pressure transmitted from particle through
their point of contact through the soil mass above the plane. Such a pressure, also termed as
inter granular pressure, is effective in decreasing the voids ratio of the soil mass and in
mobilizing its shear strength. The neutral pressure or the pore water pressure or pore pressure is
the pressure transmitted through the pore fluid.
Therefore, this pressure is also called neutral pressure (u). Since the total vertical pressure at any
plane is equal to the sum of the effective pressure and pore water pressure we have,
The pressure of water in the pores of the soil is called pore water pressure (u). The magnitude
of pore water pressure depends on:
 the depth below the water table.
 the conditions of seepage flow.

55
Lecture Note on GTE-I

Fig.6.2: Pore water pressure


Under hydrostatic conditions, no water flow takes place, and the pore pressure at a given point

𝑢 = 𝛾𝑤ℎ--------------------------------
is given by
Eq.6.2
Where, h = depth below water table or overlying water surface
It is convenient to think of pore water pressure as the pressure exerted by a column of water in
an imaginary standpipe inserted at the given point.
The natural level of ground water is called the water table or the phreatic surface. Under
conditions of no seepage flow, the water table is horizontal. The magnitude of the pore water
pressure at the water table is zero. Below the water table, pore water pressures are positive.
The principle of effective stress was enunciated by Karl Terzaghi in the year 1936. This

1. At any point in a soil mass, the effective stress (represented by or 𝜎′) is related to
principle is valid only for saturated soils, and consists of two parts:

𝜎′ = 𝜎 − 𝑢-----------------------------
total stress (σ) and pore water pressure (u) as
Eq.6.3
Both the total stress and pore water pressure can be measured at any point.
2. All measurable effects of a change of stress, such as compression and a change of shearing

Compression = 𝑓1(𝜎)
resistance, are exclusively due to changes in effective stress.

Shear Strength = 𝑓2(𝜎)

Fig. 6.3: Total Stress and Effective Stress acting on Soil


In a saturated soil system, as the voids are completely filled with water, the pore water
pressure acts equally in all directions.
The effective stress is not the exact contact stress between particles but the distribution of load
carried by the soil particles over the area considered. It cannot be measured and can only be
computed.
If the total stress is increased due to additional load applied to the soil, the pore water pressure
initially increases to counteract the additional stress. This increase in pressure within the pores
might cause water to drain out of the soil mass, and the load is transferred to the solid grains.
This will lead to the increase of effective stress.
Above the water table, when the soil is saturated, pore pressure will be negative (less than
atmospheric). The height above the water table to which the soil is saturated is called

56
Lecture Note on GTE-I

the capillary rise, and this depends on the grain size and the size of pores. In coarse soils, the
capillary rise is very small.
6.2 STRESSES WHEN NO FLOW TAKES PLACE THROUGH THE
SATURATED SOIL MASS

Fig.6.4: Stresses when no flow takes place


The above figures show that the container A is filled with sand to a depth z1 and water to a depth
z2 above the sand surface. A flexible tube connects the bottom of the container A to another
container B. The water levels are kept constant in these two containers.
The water surfaces in both the containers in Fig. 6.4(a) are kept at the same level. Under this
condition, no flow takes place from one container to another. Consider two points M and N as

pressure at N according to the laws of hydraulics. Therefore, the water pressure at N, 𝑢𝑧 = (𝑧


shown in the figure on a horizontal plane. The water pressure at M should be equal to the

+
𝑧2)𝛾𝑤
The pressure uz is termed as the pore water pressure acting on the grains at depth z from the

weight of the submerged soil above N. If 𝛾𝑏 is the submerged unit weight of the soil, the total
surface of the sample. However, the total pressure at point N is due to the water head plus the

𝜎𝑧 = 𝑧𝛾𝑏 + (𝑧 + 𝑧2)𝛾𝑤-----------------
pressure at N is
Eq.6.4
The inter-granular or effective pressure at the point N is the difference between the total and the

𝜎′ = 𝜎 − 𝑢 =𝑧𝛾 ---------------
pore water pressures. Therefore, the effective pressure
𝑧 𝑧 𝑧 𝑏
Eq.6.5
The above equation clearly demonstrates that the effective pressure is independent of the depth
of water z2 above the submerged soil surface.
6.3 STRESSES WHEN FLOW TAKES PLACE THROUGH THE SOIL FROM TOP
TO BOTTOM
57
Lecture Note on GTE-I

Fig. 6.5: Stresses when flow takes place in downward direction


In Fig. 6.5 (a) the water surface in container B is kept at h units below the surface in A. This difference
in head permits water to flow from container A to B. Since container B with the flexible tube can
be considered as a piezometer tube freely communicating with the bottom of container A, the
piezometric head or the pore water pressure head at the bottom of container A is (z1 + z2 - h).

𝑢𝑐 = (𝑧1 + 𝑧2 − ℎ)𝛾𝑤-------------------------
Therefore, the pore water pressure uc at the bottom level is
Eq.6.6
As per Fig.6.5(a), the pore water pressure at the bottom of container A when no flow takes place through

𝑢𝑐 = (𝑧1 + 𝑧2)𝛾𝑤-------------------
the soil sample is
Eq.6.7

𝛾𝑤when water flows through the soil sample from top to bottom. It may be understood that this
It is clear from the above expressions, that there is a decrease in pore water pressure to the extent of

decrease in pore2 water pressure is not due to velocity of the flowing water. The value of the
velocity head 𝑣 is a negligible quantity even if we take the highest velocity of flow that is
2𝑔
encountered in natural soil deposits. As shown in Fig. 6.5. (a), the total pressure 𝜎𝑐 at the bottom of the

𝜎𝑐 = 𝑧1𝛾𝑏 + (𝑧1 + 𝑧2)𝛾𝑤---------------


container in this case also remains the same. Therefore,
Eq.6.8

𝜎′ = 𝜎 − 𝑢 =𝑧 + ℎ𝛾 ----------------
The effective pressure at the bottom of the container is
𝑐 𝑐 𝑐 1 𝑤
Eq.6.9
Equation indicates that in this case there is an increase in the effective pressure by ℎ𝛾𝑤 at the bottom of
the container A as compared to the earlier case. The effective pressure at the top surface of the
sample is zero as before. Therefore, the effective pressure at any depth z can be written as

58
Lecture Note on GTE-I

𝜎′ = 𝜎′ ( ) = (𝑧 𝛾 + ℎ𝛾 )( )----------------------
𝑧 𝑧

𝑧 1 𝑤
Eq.6.10
𝑐 𝑧 𝑏
=𝑧𝛾 +
𝑧1 ℎ𝑧𝛾𝑤
1
------------------ Eq.6.11
𝑧1
𝑏
ℎ𝑧𝛾𝑤
The above equation indicates that is the increase in the effective pressure as the water flows
𝑧1
from the surface to a depth z. This increase in effective pressure due to the flow of water through
the pores of the soil is known as seepage pressure. It may be noted that h is the total loss of head
as the water flows from the top surface of the sample to a depth z1.
The corresponding loss of head at depth z is (z/z1)h. Since (z/z1) = i the hydraulic gradient, the

be′ expressed as 𝑖𝑧𝛾𝑤.The effective pressure at depth z can be written as


loss of head at depth z can be expressed as iz. Therefore the seepage pressure at any depth may

𝜎 = 𝑧𝛾 + 𝑖𝑧𝛾 ---------------------
𝑧 𝑏 𝑤
Eq.6.12
The distribution of pore water and effective pressures are shown in Fig. 6.5.(b). In normal soil
deposits when flow takes place in the direction of gravity there will be an increase in the
effective pressure.

6.4 STRESSES WHEN FLOW TAKES PLACE THROUGH THE SOIL


FROM BOTTOM TO TOP

Fig. 6.6: Stresses when flow takes place in upward direction


In Fig. 6.6 (a), the water surface in container B is kept above that of A by h units. This
arrangement permits water to flow upwards through the sample in container A. The total
piezometric or the pore water head at the bottom of the sample is given by (z1+z2+h).
Therefore, the pore water pressure uc at the bottom of the sample is
59
Lecture Note on GTE-I

𝑢𝑐 = (𝑧1 + 𝑧2 + ℎ)𝛾𝑤------------------
As before the total pressure head 𝜎𝑐 at the bottom of the sample is
Eq.6.13

𝜎𝑐 = 𝑧1𝛾𝑏 + (𝑧1 + 𝑧2 + ℎ)𝛾𝑤------ Eq.6.14

𝜎′ = 𝜎 − 𝑢 = 𝑧 𝛾 + (𝑧 + 𝑧 + ℎ)𝛾 − (𝑧 + 𝑧 + ℎ)𝛾
The effective pressure at the bottom of sample is, therefore,
𝑐 𝑐 𝑐 1 𝑏 1 𝑤 1 2 𝑤
2
= 𝑧1𝛾𝑏 − 𝑧𝛾𝑤------------- Eq.6.15
This equation indicates that there is a decrease in the effective pressure due to upward flow of
water. At any depth z, zγb is the pressure of the submerged soil acting downward and izγb is the
seepage pressure acting upward. The effective pressure reduces to zero when these two pressures

𝜎′ = 𝜎 − 𝑢 = 0 0r
balance. This happens when
𝑐 𝑐 𝑐
𝑧 𝛾 − 𝑧𝛾 =0 and hence 𝑖 = 𝑖 =𝛾𝑏-----------------
1 𝑏 𝑤
Eq.6.16
𝑐 𝛾𝑤
This indicates that the effective pressure reduces to zero when the hydraulic gradient attains a
maximum value which is equal to the ratio of the submerged unit weight of soil and the unit
weight of water. This gradient is known as the critical hydraulic gradient ic. In such cases,
cohesionless soils lose all of their shear strength and bearing capacity and a visible agitation of
soil grains is observed. This phenomenon is known as boiling or a quick sand condition.
Hence critical hydraulic gradient can be expressed as
𝑖𝑐 =
𝐺−1

1+𝑒
--------------- Eq.6.17
Quick conditions are common in excavations below the ground water table. This can be
prevented by lowering the ground water elevation by pumping before excavation. Quick
conditions occur most often in fine sands or silts and cannot occur in coarse soils. The larger the
particle size, the greater is the porosity. To maintain a critical gradient of unity, the velocity at
which water must be supplied at the point of inflow varies as the permeability. Therefore a quick
condition cannot occur in a coarse soil unless a large quantity of water can be supplied.
6.5 EFFECTIVE PRESSURE DUE TO CAPILLARY WATER RISE IN SOIL
Since the capillary pressure inside a soil mass is below atmospheric pressure, it draws the grains
of soils closer to each other at all points where the menisci touch the soil grains. Inter-granular
pressure of this type is called capillary pressure. The effective or inter-granular pressure at any

𝜎� ′ = 𝜎 − 𝑢𝑤-------------
point in a soil mass can be expressed by

where σ is the total pressure, 𝜎′


Eq.6.18
� is the effective or the inter-granular pressure and uw is the pore


water pressure. When the water is in compression uw is positive, and when it is in tension uw is
negative. Since uw is negative in the capillary zone, the inter-granular pressure is increased by uw.
The equation, therefore, can be written as

60
Lecture Note on GTE-I

𝜎 = 𝜎 + 𝑢𝑤---------------
Fig.6.7: Effect of capillary pressure uc on soil vertical stress diagram

Eq.6.19
The increase in the inter-granular pressure due to capillary pressure acting on the grains leads to
greater strength of the soil mass.

6.6 Important of Effective Stress

At any point within the soil mass, the magnitudes of both total stress and pore water pressure are
dependent on the ground water position. With a shift in the water table due to seasonal
fluctuations, there is a resulting change in the distribution in pore water pressure with depth.
Changes in water level below ground result in changes in effective stresses below the water
table. A rise increases the pore water pressure at all elevations thus causing a decrease in
effective stress. In contrast, a fall in the water table produces an increase in the effective stress.
Changes in water level above ground do not cause changes in effective stresses in the ground
below. A rise above ground surface increases both the total stress and the pore water pressure by
the same amount, and consequently effective stress is not altered.
If both total stress and pore water pressure change by the same amount, the effective stress
remains constant.
Total and effective stresses must be distinguishable in all calculations. Ground movements and
instabilities can be caused by changes in total stress, such as caused by loading by foundations
and unloading due to excavations. They can also be caused by changes in pore water pressures,
such as failure of slopes after rainfall.

61
Lecture Note on GTE-I

Example 1: For the soil deposit shown below, draw the total stress, pore water pressure and
effective stress diagrams. The water table is at ground level.

Solution:

Total stress
At - 4m, = 1.92 x 4 = 7.68
At -11m, = 7.68 + 2.1 x 7 = 22.38
Pore water pressure
At - 4 m, u = 1 x 4 = 4
At -11 m, u = 1 x 11 = 11
Effective stress
At - 4 m , = 7.68 - 4 = 3.68
At -11m , = 22.38 - 11 = 11.38

62
Lecture Note on GTE-I

Example 2: An excavation was made in a clay stratum having = 2 T/m3. When the depth was
7.5 m, the bottom of the excavation cracked and the pit was filled by a mixture of sand and
water. The thickness of the clay layer was
10.5 m, and below it was a layer of pervious water-bearing sand. How much was the artesian
pressure in the
sand layer?

Solution:
When the depth of excavation was 7.5 m, at the interface of the CLAY and SAND layers, the
effective stress was equal to zero.
Downward pressure due to weight of clay = Upward pressure due to artesian pressure
(10.5 - 7.5) = , where h = artesian pressure head
3x2=1xh
h = 6 m = 0.6 kg/cm2 or 6 T/m2 artesian pressure
Example 3 A clay layer 3.66 m thick rests beneath a deposit of submerged sand 7.92 m thick.
The top of the sand is located 3.05 m below the surface of a lake. The saturated unit weight of
the sand is 19.62 kN/m3 and of the clay is 18.36 kN/m3 Compute (a) the total vertical pressure,
(b) the pore water pressure, and (c) the effective vertical pressure at mid height of the clay layer
(Refer to Fig. Ex.3.)

Sol:
(a) Total pressure
The total pressure cr, over the midpoint of the clay is due to the saturated weights of clay and

63
Lecture Note on GTE-I

sand layers plus the weight of water over the bed of sand, that is

σ=3.66/2 x 18.36 +7.92 x 19.62 + 3.05 x 9.81 = 33.6 + 155.4 + 29.9 = 218.9 kN/m2
(b) Pore water pressure is due to the total water column above the midpoint.
That is
u =3.66/2 x 9.81 + 7.92 x 9.81 + 3.05 x 9.81 = 125.6 kN/m2
(c) Effective vertical pressure
σ-u = 218.9-125.6 = 93.3 kN/m2
Example 4
The water table is located at a depth of 3.0 m below the ground surface in a deposit of sand 11.0
m thick (Fig. Ex. 4). The sand is saturated above the water table. The total unit weight of the
sand is 20 kN/m3. Calculate the (a) the total pressure, (b) the pore water pressure and (c) the
effective pressure at depths 0, 3.0, 7.0, and 11.0 m from the ground surface, and draw the
pressure distribution diagram.

The calculation of total and effective stresses at different depth are shown in the table

Problem No.1
A clay stratum 8.0 m thick is located at a depth of 6 m from the ground surface. The natural
moisture content of the clay is 56% and G = 2.75. The soil stratum between the ground surface
and the clay consists of fine sand. The water table is located at a depth of 2 m below the ground

64
Lecture Note on GTE-I

surface. The submerged unit weight of fine sand is 10.5 kN/m3, and its moist unit weight above
the water table is 18.68 kN/m3. Calculate the effective stress at the center of the clay layer.

7.0 PERMEABILITY OF SOIL


Definition of Permeability:
It is defined as the property of a porous material which permits the passage or seepage of water
(or other fluids) through its interconnecting voids.
A material having continuous voids is called permeable. Gravels are highly permeable while stiff
clay is the least permeable, and hence such a clay may be termed impermeable for all practical
purpose.

The study of seepage of water through soil is important for the following engineering
problems:

1. Determination of rate of settlement of a saturated compressible soil layer.


2. Calculation of seepage through the body of earth dams and stability of slopes for highways.
3. Calculation of uplift pressure under hydraulic structure and their safety against pipin
4. Groundwater flow towards well and drainage of soil.

7.1 Darcy’s Law (1856) of Permeability:


For laminar flow conditions in a saturated soil, the rate of the discharge per unit time is
proportional to the hydraulic gradient.
q = kiA------------------ Eq.7.1

65
Lecture Note on GTE-I

v = q/A = ki … Eq.7.2
Where q = discharge per unit time
A = total cross-sectional area of soil mass, perpendicular to the direction of flow
i = hydraulic gradient
k = Darcy’s coefficient of permeability
v = velocity of flow or average discharge velocity

7.1.1 Coefficient of permeability (or) permeability

It is defined as the average velocity of flow that will occur through the total cross-sectional are of
soil under unit hydraulic gradient. The coefficient of permeability is denoted as K. It is usually
expressed as cm/sec (or) m/day (or) feet/day.

When hydraulic gradient is unity, k is equal to V. Thus, the coefficient of permeability, or simply
permeability is defined as the average velocity of flow that will occur through the total cross-
sectional area of soil under unit hydraulic gradient. Dimensions are same as of velocity, cm/sec.

7.1.2 Discharge and seepage velocities

Figure 7.1 shows a soil sample of length L and cross-sectional area A. The sample is placed in a
cylindrical horizontal tube between screens. The tube is connected to two reservoirs R1 and R2 in
which the water levels are maintained constant. The difference in head between R{ and R2 is h.
This difference in head is responsible for the flow of water. Since Darcy's law assumes no
change in the volume of voids and the soil is saturated, the quantity of flow past sections AA, BB
and CC should remain the same for steady flow conditions. We may express the equation of
continuity as follows

Fig.7.1: Flow of water through a sample soil

𝑞𝑎𝑎 = 𝑞𝑏𝑏 = 𝑞𝑐𝑐-----------------------Eq.7.3

66
Lecture Note on GTE-I

If soil mass is divided into two parts soil solid and void space, tjen the area are available for
passage of water is only through the area of voids A v. If vs is the velocity of flow in voids and v
is the average velocity across the section then we have

𝐴𝑣𝑣𝑠 = 𝐴𝑣----------------- Eq.7.4

Or 𝑣� =
𝐴

𝑣---------------------
� 𝐴
Eq.7.5
= =
𝑣 𝐴 1
1+𝑒 𝐴𝑣 𝑛
But since, Eq.7.6
So 𝑣� = =
𝑣 1+𝑒
𝑒
-----------------
𝑛
𝑣--------------------
Since�1+𝑒𝑒is always greater than unity, 𝑣 is always greater than v.
Eq.7.7

𝑒 𝑠
7.1.3 Validity of Darcy’s Law:
In accordance with the Darcy’s Law, the velocity of flow through soil mass is directly proportion
to the hydraulic gradient for laminar flow condition only. It is expected that the flow to be
always laminar in case of fine-grained soil deposits because of low permeability and hence low
velocity of flow.
So v=ki
Where i is the hydraulic gradient and k is a constant known as coefficient of permeability.
However, in case of sands and gravels flow will be laminar up to a certain value of velocity for
each deposit and investigations have been carried out to find a limit for application of Darcy’s
law.
According to researchers, flow through sands will be laminar and Darcy’s law is valid so long as

≤ 1-------------
𝑣𝐷𝑎𝛾𝑤
Reynolds number expressed in the form is less than or equal to unity as shown below

𝜂𝑔
Eq.7.8
Where v = velocity of flow in cm/sec
Da = size of particles (average) in cm.
It is found that the limiting value of Reynolds number taken as 1 is very approximate as its actual
value can have wide variation depending partly on the characteristic size of particles used in the
equation.
7.2 FACTORS AFFECTING PERMEABILITY ARE

1. Size of soil particle


2. Specific Surface Area of Soil Particle
3. Shape of soil particle
4. Void ratio
5. Soil structure
6. Degree of saturation
7. Water properties
8. Temperature
9. Adsorbed water
67
Lecture Note on GTE-I

10. Organic Matter


1. Size of Soil Particle
Permeability varies according to size of soil particle. If the soil is coarse grained, permeability is more
and if it is fine grained, permeability is low. The relation between coefficient of permeability (k)

𝑘𝛼𝐷2
and particle size (D) can be shown from equation (1) as follows.

2. Specific Surface Area of Particles


Specific surface area of soil particles also affects the permeability. Higher the specific surface area

1
lower will be the permeability.
𝐾𝛼
𝑆𝑝𝑒𝑐𝑖𝑓𝑖𝑐 𝑆𝑢𝑟𝑓𝑎𝑐𝑒 𝐴𝑟𝑒𝑎
3. Shape of Soil Particle
Rounded Particles will have more permeability than angular shaped. It is due to specific surface
area of angular particles is more compared to rounded particles.
4. Void Ratio
In general, Permeability increases with void ratio. But it is not applicable to all types of soils. For
example, Clay has high void ratio than any other types of soil but permeability for clays is very
low. This is due to, the flow path through voids in case of clays is extremely small such that
water cannot permit through this path easily.

Clay 𝑘𝛼
𝐶𝑒3
The relation between coefficient of permeability and void ratio can be expressed as
For
1+𝑒

Where, C = Shape of the flow path,


For
coarse grained soil, “C” can be neglected. Hence 𝑘 ∝
𝑒3
e = Void ratio.

1+𝑒
5. Soil Structure
Structure of any two similar soil masses at same void ratio need not be same. It
varies according to the level of compaction applied. If a soil contains flocculated structure, the
particles are in random orientation and permeability is more in this case.
If the soil contains dispersed structure, the particles are in face to face orientation hence,
permeability is very low. The permeability of stratified soil deposits also varies according to the
flow direction. If the flow is parallel, permeability is more. If it is perpendicular, permeability is
less.
6. Degree of Saturation
Partially saturated soil contains air voids which are formed due to entrapped air or gas released
from the percolating fluid or water. This air will block the flow path thereby reduces the
permeability. Fully saturated soil is more permeable than partially saturated soil.
7. Water Properties

68
Lecture Note on GTE-I

Various properties of water or fluid such as unit weight and viscosity also effects the
permeability. However, unit weight of water will not affect much since it does not change much
with temperature.
But when temperature is increased viscosity decreases rapidly. From equation (1), permeability

𝛾
increase when viscosity decreases.

𝑘𝛼
𝑤

8. Temperature 𝜇
Temperature also affects the permeability in soils. The permeability is inversely proportional to
the viscosity of the fluid. It is known that viscosity varies inversely to the temperature. Hence,
Permeability is directly related to temperature.
Greater the temperature, higher will be the permeability. That is the reason; seepage is more in

1
summer seasons than in winter.

𝑘𝛼 𝛼 𝑡𝑒𝑚𝑝𝑒𝑟𝑎𝑡𝑢𝑟𝑒
𝜇
9. Adsorbed Water
Adsorbed water is the water layer formed around the soil particle especially in the case of fine-
grained soils. This reduces the size of the void space by about 10%. Hence, permeability reduces.
10. Organic matter
Presence of organic matter decreases the permeability. This is due to blockage of voids by the
organic matter.
7.3 PERMEABILITY OF STRATIFIED SOIL DEPOSITS:
Natural soil deposits may exhibit stratification. Each layer may have its own coefficient of
permeability, assuming it to be homogeneous. The ‘average permeability’ of the entire deposit
will depend upon the direction of flow in relation to the orientation the bedding planes.
Two cases will be considered—the first one with flow perpendicular to the bedding planes and
the next with flow parallel to the bedding planes.
7.3.1 Flow Perpendicular to the Bedding Planes

Fig. 7.2: Flow perpendicular to bedding planes

69
Lecture Note on GTE-I

Let h1, h3, h3 ...hn be the thicknesses of each of the n layers which constitute the deposit, of total
thickness h. Let k1, k2, k3 ... kn be the Darcy coefficients of permeability of these layers
respectively.
In this case, the velocity of flow v, and hence the discharge q, is the same through all the layers,
for the continuity of flow.

∆ℎ = ∆ℎ1 + ∆ℎ2 + ∆ℎ3 ± − − − − ∆ℎ𝑛------------


Let the total head lost be Δh and the head lost in each of the layers be Δh2, Δh2, Δh3, ..
Eq.7.9

∆ℎ1
The hydraulic gradients are
𝑖1 = ⁄ where i is the hydraulic gradient for deposit
ℎ1
∆ℎ2
𝑖2 = ⁄ -----------
ℎ2

∆ℎ𝑛
𝑖𝑛 = ⁄ --------------------
ℎ𝑛
Eq.7.10
Since q is the same in all the layers, and area of cross-section of flow is the same, the velocity is
the same in all layers.

Now 𝑘𝑧𝑖 = 𝑘1𝑖1 = 𝑘2𝑖2=------ = 𝑘𝑛𝑖𝑛 = 𝑣---------------------


Let kz be the average permeability perpendicular to the bedding planes

∆ℎ ∆ℎ1 ∆ℎ2 ∆ℎ𝑛


Eq.7.11
∴ 𝑘𝑧 ⁄ = 𝑘1 ⁄ = 𝑘2 ⁄ = − − −− = 𝑘𝑛 ⁄ =𝑣
ℎ ℎ ℎ
Eq.7.12 ℎ
1 2 𝑛
Now substituting the expression for ∆ℎ1, ∆ℎ2----in terms of v in the equation for ∆ℎ , we get
𝑣ℎ⁄ 𝑣ℎ2 + − − − − +
𝑧 = + ⁄
� 𝑣ℎ1 𝑘 𝑣ℎ𝑛

-------- Eq.7.13
⁄ �𝑛
𝑘 1 2
Hence 𝑘 =



𝑧
------------------------ Eq.7.14
ℎ1 ℎ𝑛
ℎ2
+ ±−−−−−− +
𝑘1 𝑘2 𝑘𝑛

This is the equation for average permeability for flow perpendicular to the bedding planes.
7.3.2 Flow Parallel to the Bedding Planes

Fig. 7.3: Flow parallel to the bedding planes


70
Lecture Note on GTE-I

Let the flow be parallel to the bedding planes as shown in Figure 7.3

With the same notation as in the first case, the hydraulic gradient i will be the same for all the
layers as for the entire deposit. Since v = ki, and k is different for different layers, v will be
different for the layers, say, v1, v2, ... vn
Also v1=k1i, v2=k2i and so on
Considering unit dimension perpendicular to the plane of the paper, the areas of flow for each
layer will be plane of the paper, the areas of flow for each layer will be h1, h2, ... hn respectively,
and it is h for the entire deposit.
The discharge through the entire deposit is equal to the sum of the discharge through the
individual layers. Assuming kx to be the average permeability of the entire deposit parallel to the

𝑞 = 𝑞1 + 𝑞2 + − − +𝑞𝑛------------------------
bedding planes, and applying the equation:

We have , 𝑘𝑥𝑖ℎ = 𝑘1𝑖ℎ1 + 𝑘2𝑖ℎ2 + − − − + 𝑘𝑛𝑖ℎ𝑛---------------------------- Eq.7.16


Eq.7.15

∴ = [
𝑘1ℎ1+𝑘2ℎ2+ −−+𝑘𝑛ℎ𝑛
]-----------------------------------
𝑘𝑥 ℎ
Eq.7.17

Where h=h1+h2+------+hn----------------------------------------- Eq.7.18


In other words, kx is the weighted mean value, the weights being the thickness for each layer.
It can be shown that kx is always greater than kz for a given situation.

7.4 CONSTANT HEAD PERMEABILITY TEST


The constant head permeability test is a laboratory experiment conducted to determine the
permeability of soil. The soils that are suitable for this test are sand and gravels. Soils with silt
content cannot be tested with this method. The test can be employed to test granular soils either
reconstituted or disturbed.
What is Coefficient of Permeability?
The coefficient of permeability, k is defined as the rate of flow of water under laminar flow
conditions through a porous medium area of unit cross section under unit hydraulic gradient.
The coefficient of permeability (k) is obtained from the relation
𝑘= =
𝑞𝐿 𝑄𝐿
𝐴ℎ
Eq.7.19
------------------------
Where q= Discharge, Q = Total volume of water, t=time period, h= Head causing flow, L=
Length of specimen, A= Cross-sectional area.
Apparatus for Constant Head Permeability Test

1 Permeameter mould, internal diameter = 100mm, effective height =127.3 mm, capacity =
1000ml.
2 Detachable collar, 100mm diameter, 60mm height
3 Dummy plate, 108 mm diameter, 12mm thick,
4 Drainage base, having porous disc
71
Lecture Note on GTE-I

5 Drainage cap having porous disc with a spring attached to the top.
6 Compaction equipment such as Proctor’s rammer or a static compaction equipment, as
specified in IS: 2720 (Part VII)-1965.
7 Constant head water supply reservoir
8 Vacuum pump
9 Constant head collecting chamber
10 Stop watch
11 Large funnel
12 Thermometer
13 Weighing balance accuracy 0.1g
14 Filter paper.
Procedure
Specimen Preparation

1. Remove the collar of the mould. Measure the internal dimensions of the mould. Weigh
the mould with dummy plate to the nearest gram.
2. Apply a little grease on the inside to the mould. Clamp the mould between the base plate
and the extension collar and place the assembly on a solid base.
3. Take about 2.5kg of the soil sample, from a thoroughly mixed wet soil, in the mould.
Compact the soil at the required dry density using a suitable compacting device.
4. Take a small specimen of the soil in a container for the water content determination.
5. Remove the collar and base plate. Trim the excess soil level with the top of the mould.
6. Clean the outside of the mould and the dummy plate. Find the mass of the soil in the
mould.
7. The mould with the sample is now placed over the permeameter. This will have drainage
and cap discs properly saturated

Test Procedure

1. Through the top inlet of the constant head reservoir, the specimen is connected.
2. The bottom outlet is opened and a steady flow is established
3. For a particular time interval, the quantity of flow can be collected.
4. Measure the difference of head (h) in levels between the constant head reservoir and the
outlet in the base.
5. For the same interval, this is repeated three times.

72
Lecture Note on GTE-I

Fig.7.4: Constant Head Permeameter setup

Observation and Calculations


Initially, the flow is very slow. It later increases and will become constant. The constant head
permeability test is best for cohesionless soils.
Observations and Computation in Constant Head Pemeameter
The data that is obtained directly from the tests are:
1. Length (L) in cm
2. Area (A) in cm2
3. Constant Head (H) in cm
4. Discharge (Q) in liter

𝑘 = 𝑄 × 𝐿/(𝐴 × 𝑡 × ℎ)-----------
Calculate the coefficient of permeability (k) using the following equation:
Eq.7.20
where: k = coefficient of permeability,
Q= quantity of water discharged,
L = distance between manometers,
A = cross-sectional area of specimen, t =
total time of discharge
h = head required for flow
7.5 FALLING HEAD PERMEABILITY TEST

Falling head permeability test is one of several techniques by which the permeability of soil is
determined. It is used to evaluate the permeability of fairly less previous soil particularly for fine

73
Lecture Note on GTE-I

grained soil sample. Permeability is the measure of the ability of soil to allow water to flow its
pores or voids.

A falling head permeameter is shown in Fig.7.5. The soil sample is kept in a vertical cylinder of
cross-sectional area A. A transparent stand pipe of cross sectional area, a, is attached to the test
cylinder. The test cylinder is kept in a container filled with water, the level of which is kept
constant by overflows. Before the commencement of the test the soil sample is saturated by
allowing the water to flow continuously through the sample from the stand pipe. After saturation
is complete, the stand pipe is filled with water up to a height of h0 and a stop watch is started. Let
the initial time be t0. The time t1 when the water level drops from h0 to h1 is noted. The hydraulic
conductivity k can be determined on the basis of the drop in head (h 0 – h1) and the elapsed time
(t1 – t0) required for the drop as explained below.

Let h be the head of water at any time t. Let the head drop by an amount dh in time dt.

𝑑𝑄 = 𝑘𝑖𝐴𝑑𝑡 = 𝑘 𝐴𝑑𝑡-------------

The quantity of water flowing through the sample in time dt from Darcy's law is

𝐿
Eq.7.21
where, i = h/L the hydraulic gradient. The quantity of discharge dQ can be expressed as
dQ = -adh ---------------- Eq.7.22
Since the head decreases as time increases, dh is a negative quantity in Eq.7.22. & Eq. 7.21 can

= 𝑘𝐴𝑑𝑡/𝐿--------------
be equated to Eq. 7.22
−𝑎𝑑ℎ

Eq.7.23
The discharge Q in time (t1 – t0) can be obtained by integrating Eq. 2 or 3. Therefore, Eq.4 can

−𝑎 ∫ℎ1 𝑑ℎ= 𝐾𝐴∫ 𝑡𝑑𝑡-------------------------


be rearranged and integrated as follows
1 Eq.7.24
ℎ 𝑡
ℎℎ0 𝐿
Or 𝑎𝑙𝑛 = (𝑡 − 𝑡 )-----------------------
0
𝐾𝐴 0
Eq.7.25
ℎ1 𝐿 1 0
The general expression for k is
𝑙𝑛
ℎ0
𝐾 =𝐴(𝑡1−𝑡0 ℎ1 -----------------------
𝑎𝐿 Eq.7.26
)

𝒉𝟎
OR

𝑲 =
𝟐.𝟑𝒂𝑳
𝒍𝒐 𝟏 𝒉
𝒈
------------------- Eq.7.27
𝑨(𝒕𝟏−𝒕 𝟎 𝟏
𝟎)
Apparatus
1. Mould Assembly
The mould assembly including drainage base and drainage cap which need to be conform to IS:
11209-1985
2. Compaction Hammer
3. Set of Stand Pipes
Glass stand pipes varying in diameter from 5 to 20 mm, suitably mounted on stand or otherwise
fixed on wall
74
Lecture Note on GTE-I

4. Constant Head Tank

75
Lecture Note on GTE-I

Suitable water reservoir capable or supplying water to the permeameter under constant head for constant
head test arrangement is required.
5. Miscellaneous Apparatus
For instance, IS sieves, mixing pan, graduated cylinder, meter scale, stop watch, 75micron wire gauge,
thermometer, and a source of de-aired water.

Soil Specimen Preparation (Disturbed Soil Sample)

 Take 2.5-kg soil from a thoroughly mixed air-dried or oven-dried material and evaluate
its moisture content.
 Remove the collar of the mould. Measure the internal dimensions of the mould. Weigh
the mould with dummy plate to the nearest gram.
 Apply a little grease on the inside to the mould.
 Clamp the mould between the base plate and the extension collar and place the assembly
on a solid base.
 Place soil specimen in the mould, and compact it at the required dry density using a
suitable compacting device.
 Take a small specimen of the soil in a container for the water content determination.
 Remove the collar and base plate. Trim the excess soil level with the top of the mould.
 Clean the outside of the mould and the dummy plate.
 Find the mass of the soil in the mould.
 The mould with the sample is now placed over the permeameter. This will have drainage
and cap discs properly saturated

Undisturbed Soil Sample

 Trim the specimen in a form of cylinder not larger than 85cm in diameter, and having a
height equal to that of the mould.
 Place the specimen over porous disc of the drainage base fixed to the mould.
 Use impervious material like cement slurry to fill the space between mould and the
specimen.
 Fix the drainage cap over the top of the mould.
Procedure
1. Connect the specimen to the selected stand-pipe through the top inlet.
2. Open the bottom outlet and record the time interval required for the water level to fall
from a known initial head to a known final head as measured above the center of the
outlet.

76
Lecture Note on GTE-I

3. Refill the stand-pipe with water and repeat the test till three successive observations give
nearly same time interval; the time intervals being recorded for the drop in head from the
same initial to final values, as in the first determination.
4. Alternatively, after selecting the suitable initial and final heads, h 1, and h2, respectively,
observe the time intervals for the head to fall from h1 to √ℎ1ℎ2 and similarly
from√ℎ1ℎ2 to h2.
5. The time intervals should be the same; otherwise the observation shall be repeated after
refilling the stand-pipe.

Fig.7.5: Falling Head Permeameter Test Setup

Data Sheet
The following values shall be recorded in the data sheet of variable head permeability test: Length of
specimen (L)
Diameter of specimen (D)
Volume of specimen (V)
Water content (ω)
Diameter of stand pipe (d)
Area of stand pipe (a)
Specific gravity of solids (Gs)

At temperature T of water, the permeability 𝑘𝑇 is calculated using the following expression:


Calculations

𝑘𝑇 = [(2.30aL)/(At)]log10(h1/h2) ------------------------ Eq.7.28


Where
h1: initial head
77
Lecture Note on GTE-I

h2: final head


t: time
interval
a: cross-sectional area of the liquid stand pipe
A: cross-sectional area of the specimen
L: length of specimen.
Result
The permeability values at temperatures T and 27°C are reported as numbers with units as cm/s.
The state of the sample is also reported in terms of water content, void ratio and degree of
saturation.
Example 1: A sand sample of 35 cm2 cross sectional area and 20 cm long was tested in a
constant head permeameter. Under a head of 60 cm, the discharge was 120 ml in 6 min. The dry
weight of sand used for the test was 1120 g, and Gs = 2.68. Determine (a) the hydraulic
conductivity in cm/sec, (b) the discharge velocity, and (c) the seepage velocity.

Solution

Use Equation, k = QL/ hAt

where , Q = 120 ml, t = 6 min, A = 35 cm2 , L = 20 cm, and h = 60 cm. Substituting, we have

k = 3.174 x 10-3 cm/sec

Discharge velocity, v = ki = 3.174 x 10-3 x 60/20 = 9.52 x 10-3 cm/sec

7.6 DIRECT DETERMINATION OF “k” OF SOILS IN PLACE BY PUMPING TEST


The most reliable information concerning the permeability of a deposit of coarse grained
material below the water table can usually be obtained by conducting pumping tests in the field.
Although such tests have their most extensive application in connection with dam foundations,
they may also prove advisable on large bridge or building foundation jobs where the water table
must be lowered. The arrangement consists of a test well and a series of observation wells. The
test well is sunk through the permeable stratum up to the impermeable layer. A well sunk into a
water bearing stratum, termed an aquifer, and tapping free flowing ground water having a free
ground water table under atmospheric pressure, is termed a gravity or unconfined well. A well
sunk into an aquifer where the ground water flow is confined between two impermeable soil
layers, and is under pressure greater than atmospheric, is termed as artesian or confined well.
Observation wells are drilled at various distances from the test or pumping well along two
straight lines, one oriented approximately in the direction of ground water flow and the other at
right angles to it. A minimum of two observation wells and their distances from the test well are
needed. These wells are to be provided on one side of the test well in the direction of the ground
water flow.
The test consists of pumping out water continuously at a uniform rate from the test well until the

78
Lecture Note on GTE-I

water levels in the test and observation wells remain stationary. When this condition is achieved
the water pumped out of the well is equal to the inflow into the well from the surrounding strata.
The water levels in the observation wells and the rate of water pumped out of the well would
provide the necessary additional data for the determination of k. As the water from the test well
is pumped out, a steady state will be attained when the water pumped out will be equal to the
inflow into the well. At this stage the depth of water in the well will remain constant. The
drawdown resulting due to pumping is called the cone of depression. The maximum drawdown
DQ is in the test well. It decreases with the increase in the distance from the test well. The
depression dies out gradually and forms theoretically, a circle around the test well called the
circle of influence. The radius of this circle is called the radius of influence of the depression
cone.
7.6.1 Equation for k for an Unconfined Aquifer
Figure below gives the arrangement of test and observation wells for an unconfined aquifer.
Only two observation wells at radial distances of r1 and r2 from the test well are shown. When
the inflow of water into the test well is steady, the depths of water in these observation wells are
h1 and h2 respectively.
Let h be the depth of water at radial distance r. The area of the vertical cylindrical surface of

𝐴
The= 2𝜋𝑟ℎ-----------------Eq.7.29
radius r and depth h through which water flows is given as

hydraulic gradient is 𝑖 =
𝑑ℎ

𝑑𝑟
As per Darcy's law the rate of inflow into the well when the water levels in the wells remain
stationary is
q = kiA-------- Eq.7.30

Fig.7.6: Pumping test conducted on Unconfined Aquifer

𝑞 =𝑘 2𝜋𝑟ℎ----------------
𝑑ℎ
Substituting for A and i the rate of inflow across the cylindrical surface is

𝑑𝑟
Eq.7.31

=
2𝜋𝑘ℎ𝑑ℎ
Rearranging the terms, we have
𝑑𝑟

𝑟 𝑞
------------------ Eq.7.32

79
Lecture Note on GTE-I

The integral of the equation within the boundary limits is


𝑟2 𝑑𝑟 ℎ2
= ℎ𝑑ℎ---------------
2𝜋𝑘
∫𝑟 ∫ℎ
Eq.7.33
1 𝑟 𝑞 1

𝑘 = 2 2 𝑙𝑜𝑔 ------------
The equation for k2.3𝑞 𝑟2
after integration and rearranging is
𝜋(ℎ −ℎ )
Eq.7.34
2 1 10 𝑟1
Proceeding in the same way as before another equation for k in terms of r0, h0 and Ri can be

𝑘 = 𝑙𝑜𝑔 𝑖----------
2.3𝑞 to Fig.7.6)
established as (referring 𝑅
Eq.7.35
𝜋(𝐻2−ℎ2
0) 10 𝑟
0
If we write h0 = (H- D0) in the above Eq.7.35 , where D0 is the depth of maximum drawdown in

𝑙𝑜𝑔 𝑖--------------
𝑅
the test well, we have
𝑘 =
Eq.7.36
2.3𝑞
𝜋𝐷0(2𝐻− 10 𝑟
𝐷0) 0

Now the maximum yield from the well may be written from the above expression as
1
𝑞=
𝜋𝐷0(2𝐻−𝐷0)𝑘 𝑙𝑜
𝑔
𝑅 -------------
( 𝑖)
Eq.7.37
2.3
10 𝑟0
7.6.2 Equation for k in a Confined Aquifer
Figure 7.7 shows a confined aquifer with the test and observation wells. The water in the
observation wells rises above the top of the aquifer due to artesian pressure. When pumping from
such an artesian well two cases might arise. They are:
Case 1. The water level in the test well might remain above the roof level of the aquifer at steady
flow condition.

Fig.7.7: Pumping test conducted on Confined Aquifer


Case 2. The water level in the test well might fall below the roof level of the aquifer at steady
flow condition.
If H0 is the thickness of the confined aquifer and h0 is the depth of water in the test well at the

80
Lecture Note on GTE-I

steady flow condition Case 1 and Case 2 may be stated as


Case l. When h0>H0 AND Case 2. When h0<H0

81
Lecture Note on GTE-I

Case 1. When h0 > H0


In this case, the area of a vertical cylindrical surface of any radius r does not change, since the
depth of the water bearing strata is limited to the thickness H0. Therefore, the discharge surface

𝐴 = 2𝜋𝑟𝐻0-------------
area is

A gain writing 𝑖 = the flow equation as per Darcy’s


Eq.7.38
𝑑ℎ

𝑑ℎ 𝑑𝑟
𝑞 = 𝑘𝑖𝐴 = 𝑘 2𝜋𝑟𝐻 --------------
𝑑𝑟 0
Eq.7.39

ℎ2
integration of𝑟 the equation after rearranging the terms yields
𝑑ℎ =
𝑞 2 𝑑𝑟

The

ℎ1
-------------------- Eq.7.40
2𝜋𝑘𝐻0 𝑟1 𝑟

− ) = 2𝜋𝑘 𝑙𝑛 -------------------
𝑞 𝑟2

(ℎ2 ℎ1
Or Eq.7.41
𝐻0 𝑟1

𝑙𝑜𝑔 -------------
𝑟2
The equation for k is
𝑘 =
Eq.7.42
2.3𝑞
2𝜋𝐻0(ℎ2−ℎ 10 𝑟
1)
1

Alternate Equations

𝑙𝑜𝑔 ----------------
𝑟1
As before we can write the following equation for determining k
𝑘 =
Eq.7.43
2.3𝑞
2𝜋𝐻0(ℎ1−ℎ 10 𝑟
0)
0

𝑙𝑜𝑔
𝑅𝑖
𝑘 =
-------------------- Eq.7.44
2.3𝑞
2𝜋𝐻0(𝐻−ℎ 10 𝑟
0)
0

𝑘 = 𝑙𝑜𝑔
2.3𝑞 𝑅𝑖
----------------------- Eq.7.45
2𝜋𝐻0 10 𝑟
0
𝐷0
Case 2. When h0 < H0
Under the condition when h0 is less than H0, the flow pattern close to the well is similar to that of
an unconfined aquifer whereas at distances farther from the well the flow is artesian. Muskat

𝑙𝑜𝑔 𝑖----------------
𝑅
(1946) developed an equation to determine the hydraulic conductivity. The equation is
𝑘 =2𝜋(𝐻𝐻
2.3𝑞
0−𝐻 −
2
Eq.7.46
10 𝑟
2
ℎ ) 0
0 0
Example No.1
A pumping test was carried out for determining the hydraulic conductivity of soil in place. A
well of diameter 40 cm was drilled down to an impermeable stratum. The depth of water above
the bearing stratum was 8 m. The yield from the well was 4 m3/ min at a steady drawdown of 4.5

82
Lecture Note on GTE-I

m. Determine the hydraulic conductivity of the soil in m/day if the observed radius of influence
was 150m.

that 𝑘 = 𝑙𝑜 𝑅𝑖
Solution
𝑔
We know
2.3𝑞

𝜋𝐷0(2𝐻−𝐷0) 10 𝑟
0
Here q = 4 m /min = 4 x 60 x 24 m3/day
3

D0 = 4.5 m, H = 8 m, Ri = 150 m, r0 = 0.2 m


By substituting above values we get
k= 234.4 m/day

83
Lecture Note on GTE-I

Example No.2
A field pumping test was conducted from an aquifer of sandy soil of 4 m thickness confined
between two impervious strata. When equilibrium was established, 90 liters of water was
pumped out per hour.
The water elevation in an observation well 3.0 m away from the test well was 2.1 m and another
6.0 m away was 2.7 m from the roof level of the impervious stratum of the aquifer. Find the
value of k of the soil in m/sec.
Solution:
We know that k for confined aquifer is given as

2.3𝑞 𝑟2
𝑘 =
2𝜋𝐻 (ℎ 𝑙𝑜𝑔10 1
−ℎ
0 2
) 𝑟
1
q = 90 x 103 cm3/hr = 25 x 10-6 m3/sec
H0=4 m, r1=3 m and r2=6m
Substituting the above values we get k= 1.148 x 10-6 m/s

Problem No.1
Calculate the yield per hour from a well driven into a confined aquifer. The following data are
Available, height of original piezometric level from the bed of the aquifer, H = 30m thickness of
aquifer, H0 4m, the depth of water in the well at steady state, h0 = 6 m hydraulic conductivity of
soil = 0.079 m/min, radius of well, r0 = 0.25 m, radius of influence, Ri = 150 m

7.7 PUMPING IN TESTS


1 Open end Test
2 Packer Test
Open-end Tests
An open end pipe is sunk into the strata and the soil is taken out of the pipe. Clear water having
temperature slightly higher than the ground water is added through a metering system. During
84
Lecture Note on GTE-I

this stage flow should maintain gravity flow under constant head. Water may also be allowed to
entre the pipe under some pressure head.

𝑘=
𝑞
The permeability is calculated from the expression as
5.5𝑟 --------------

Eq.7.47
Where q= constant rate of flow
h= Difference head of flow (gravity head and pressure head)
r = radius if the casing
Packer Test
Packers are primarily used in bore holes for testing the permeability of rocks under applied
pressures. The apparatus used for the pressure test is comprised of a water pump, a manually
adjusted automatic pressure relief valve, pressure gage, a water meter and a packer assembly.
The packer assembly consists of a system of piping to which two expandable cylindrical rubber
sleeves, called packers, are attached. The packers which are provided by means of sealing a
limited section of bore hole for testing should have a length five times the diameter of the hole.
They may be of the pneumatically or mechanically expandable type. The former are preferred
since they adapt to an oversized hole whereas the latter may not. However, when pneumatic
packers are used, the test apparatus must also include an air or water supply connected, through a
pressure gage, to the packers by means of a higher pressure hose. The piping of a packer
assembly is designed to permit testing of either the portion of the hole between the packers or the
portion below the lower packer. The packers are usually set 50, 150 or 300 cm apart. The wider
spacing is used for rock which is more uniform. The short spacing is used to test individual joints
which may be the cause of high water loss in otherwise tight strata.
Two types of packer methods are used for testing of permeability. They are:
1. Single packer method.
2. Double packer method.
The single packer method is useful where the full length of the borehole cannot stand
uncased/un-grouted in soft rocks, such as soft sand stone, clay shale or due to the highly
fractured and sheared nature of the rocks, or where it is considered necessary to have
permeability values side by side with drilling. Where the rocks are sound and the full length of
the hole can stand without casing/grouting, the double packer method may be adopted. The
disadvantage of the double packer method is that leakage through the lower packer can go
unnoticed and lead to overestimation of water loss.
Single Packer Method
The method used for performing water percolation tests in a section of a drilled hole using a
single packer is shown in Fig. 3 a. In this method the hole should be drilled to a particular depth
desirable for the test. The core barrel should then be removed and the hole cleaned with water.
The packer should be fixed at the desired level above the bottom of the hole and the test
performed. Water should be pumped into the section under pressure. Each pressure should be
maintained until the readings of water intake at intervals of 5 min show a nearly constant reading
of water intake for one particular pressure. The constant rate of water intake should be noted.

85
Lecture Note on GTE-I

After performing the test the entire assembly should be removed. The drilling should then
proceed for the next test section.

Fig.7.8: Test sections for single and double Packer Method


Double Packer Method
In this method the hole is first drilled to the final depth and cleaned. The packer assembly may
be fixed at any desired test section as shown in Fig. 3 b.
Both packers are then expanded and water under pressure is introduced into the hole between the
packers. The tests are conducted as before.
Regardless of which procedure is used, a minimum of three pressures should be used for each
section tested. The magnitudes of these pressures are commonly 100, 200 and 300 kPa. (1,2 and
3 kg/cm2) above the natural piezometric level. However in no case should the excess pressure be
greater than about 20 kPa per meter of soil and rock overburden above the upper packer. The
limitation is imposed to insure against possible heavy damage to the foundation.
The formulae used to compute the permeability from pressure test data are (from US Bureau of

𝑘=
𝑞
Reclamation, 1968)
2𝜋𝐿 𝑙𝑛
𝐿 for 𝐿 ≥
𝐻
10𝑟0
and------------------ Eq.7.48
𝑟
0
𝑘=
𝑞
2𝜋𝐿 𝑠𝑖𝑛ℎ > 𝐿 ≥ 𝑟0-------------------
−1
𝐻 𝐿
10𝑟0
for Eq.7.49
where 2𝑟0
k = hydraulic conductivity
q = constant rate of flow into the hole
L = length of the test section
H = differential head on the test section
Example: A sample in a variable head permeameter is 8 cm in diameter and 10 cm high. The
permeability of the sample is estimated to be 10x10 -4 cm/sec. If it is desired that the head in the
stand pipe should fall from 24 cm to 12 cm in 3min, determine the size the stand pipe which
should be used.
Sol:
86
Lecture Note on GTE-I

Variable head Permeameter:

87
Lecture Note on GTE-I

Soil sample diameter=8 cm, Height of sample =10 cm


Estimated coefficient of permeability: K=10x10-4 cm/sec

We know 𝐾 = 2.303 𝑙𝑜𝑔 ( )


h1=24 cm and h2=12 cm,𝑎𝐿 time t=180ℎsec
1

𝐴𝑡 10 ℎ
2

𝜋 × 16 × 180
Substituting for various values, we get

𝑎 =
2
2.303 × 104(𝑙𝑜𝑔
1
0
2)
Hence a=1.305 cm
Diameter of stand pipe d= 1.29 cm

Solution:

2.3𝑞
𝑘 = 𝑙𝑜𝑔10𝑟2
𝜋(ℎ2 − ℎ2) 𝑟
2 70 1 1
= 𝑙𝑜𝑔 = 1.18 × 10−1mm/sec
2.3×2×100

𝜋(31.31−292)×60 10 15

8.0 SEEPAGE ANALYSIS


In this class we will discuss about the Laplace equation for determining two-dimensional flow of
soil elements.
The assumptions made in deriving the Laplace equation, the following may be stated as the
assumptions of Laplace equation:
1. The flow is two-dimensional.
2. The flow is steady and laminar.
3. Water and the soil are incompressible.
4. The soil mass is homogeneous and isotropic.
5. The soil is fully saturated and Darcy’s law is valid.

88
Lecture Note on GTE-I

Consider a soil element of infinitesimally small size of dx and dz in X- and Z-directions, respectively,
through which the flow is taking place, shown in Fig. 8.1. Consider unit length of the soil
element in the Y-direction.

Fig.8.1: Soil element with two dimensional flow


Let vx be the velocity of flow at the entry in the X-direction and vz the velocity of flow at entry in the Z-
direction. Then the velocity at the exit in the X-direction will be
vx + (∂/∂x)(vx).dx
and the velocity at the exit in the Z-direction will be –
vz + (∂/∂z)(vz).dz
Quantity of water entering the element in the X-direction is given by –
qix = vx.(dz.1) = vx.dz
Quantity of water entering the element in the Z-direction is given by –
qiz = vz.(dx.1) = vz. dx

𝑞 = [𝑣 + (𝑣 )𝑑𝑥] × (𝑑𝑧. 1) = [𝑣 + (𝑣 )𝑑𝑥] × (𝑑𝑧)---------------


𝛛
Similarly, quantity of water leaving the soil element𝛛 in the X-direction is
0𝑥 𝑥 𝑥 𝑥 𝑥
𝛛𝑥 𝛛𝑥
Eq.8.1
Quantity of water leaving the soil element in the Z-direction is –
𝑞0 = [𝑣� + (𝑣 )𝑑𝑧] × (𝑑𝑥. 1)� = [𝑣 + (𝑣 )𝑑𝑧] ×
𝛛 𝛛
𝛛 � 𝛛 �
(Assuming
𝑑𝑥. )---- that
Eq.8.2
𝑧 � 𝑧 � � 𝑧 �
the flow is steady and incompressible, the quantity of water entering the soil

∑ 𝑞𝑖 = ∑ 𝑞𝑜-----------------
element is equal to the quantity of water leaving the soil element –

≡ 𝑞𝑖𝑥 + 𝑞𝑖𝑧 = 𝑞0𝑥 + 𝑞0𝑧-----------------


Eq.8.3

≡ 𝑣 𝑑𝑧 + 𝑣 𝑑𝑥 = [𝑣 + (𝑣 )𝑑𝑥] × (𝑑𝑧) + [𝑣 + (𝑣 )𝑑𝑧] × (𝑑𝑥. )--------Eq.8.5


𝛛 𝛛
Eq.8.4
𝑥 𝑥 𝑥 𝑧 𝑧
𝑧 𝛛𝑥 𝛛𝑧

[𝑣 + (𝑣 )𝑑𝑥] × (𝑑𝑧) − 𝑣 𝑑𝑧 + [𝑣 + (𝑣 )𝑑𝑧] × (𝑑𝑥. ) − 𝑣 𝑑𝑥 = 0-----


𝛛 𝛛
Rearranging the terms, we have –

𝑥 𝑥 𝑥 𝑧 𝑧
𝛛𝑥 𝑧 𝛛𝑧
Eq.8.6

89
Lecture Note on GTE-I

≡ 𝑣 𝑑𝑧 + (𝑣 )𝑑𝑥𝑑𝑧 − 𝑣 𝑑𝑧 + 𝑣 𝑑𝑥 + (𝑣 )𝑑𝑧. 𝑑𝑥 − 𝑣 𝑑𝑥 = 0-----------------


𝛛 𝛛

𝑥 𝑥 𝑥 𝑧 𝑧
𝛛𝑥 𝑧 𝛛𝑧

(𝑣 )𝑑𝑥. 𝑑𝑧 + (𝑣 )𝑑𝑥. 𝑑𝑧 = 0------------------


𝛛
Eq.8.7
𝛛
𝛛𝑥 𝑥 𝑧
Eq.8.8
𝛛𝑧

(𝑣 ) + (𝑣 ) = 0------------------
𝛛
On simplifying, and dividing throughout with dz. dx, we have
𝛛
𝑥 𝑧
Eq.8.9
𝛛𝑥 𝛛𝑧
Let h be the total head at any point. Then the component of hydraulic gradient in the X-direction
will be
ix = – ∂h/∂x-------------- Eq.8.10
The negative sign is to indicate that the head decreases in the direction of flow. Similarly, the
component of hydraulic gradient in the Z-direction will be –
iz = -∂h/∂z------------------ Eq.8.11
By Darcy’s law, we know that –
vx = kx .ix = – kx . ∂h/∂x------------- Eq.8.12
Similarly
vz = kz .iz = – kz . ∂h/∂x-------------- Eq.8.13

(−𝑘 )+ (−𝑘 ) = 0-------------


𝛛ℎ
Substituting these𝛛values in𝛛ℎthe above equation, we have
𝛛

𝑥 𝑧 𝛛𝑧
Eq.8.14
𝛛𝑥 𝛛
𝛛𝑥 𝑧
Assuming that the soil mass is homogeneous; permeability will be same throughout in a given

−𝑘 ( )−𝑘 ( ) = 0------------
𝛛 𝛛ℎ 𝛛 𝛛ℎ
direction. Hence
𝑥 𝛛𝑥 𝑧 𝛛𝑧
Eq.8.15
𝛛𝑧
𝛛𝑥

≡ = 0-----------------
𝛛2ℎ
+
𝛛2ℎ
𝑥
𝑘
𝑧
2 𝑘
𝛛𝑧
Eq.8.16
𝛛𝑥

+ = 0----------------
If2 the soil
𝛛 ℎ 𝛛2ℎ
is isotropic, then kx = kz = k. Then, we have

𝛛𝑥2
Eq.8.17
𝛛𝑧2
This is the Laplace equation for two-dimensional flow. It says that the change of gradient in the
x--direction plus the change of gradient in the z-direction is zero. The solution of this equation
gives a family of curves meeting at right angles to each other. One family of these curves
represents flow lines and the other equipotential lines. The graphical representation of solution to
the Laplace equation which gives two sets of mutually perpendicular curves is known as Flow
net.
8.1 Velocity Potential
Velocity potential is a scalar function of space and time such that its derivative in any direction
gives the component of velocity in that direction. Thus –
ɸ = kh ------------- Eq.8.18
∂ɸ = ∂x = vx ----------------- Eq.8.19
∂ɸ = ∂z = vz-------------------- Eq.8.20
90
Lecture Note on GTE-I

( )+ ( ) = 0----------------------------
𝛛∅ 𝛛 𝛛∅
Substituting these in Laplace Equation, we have
𝛛

𝛛 𝛛 𝛛 𝛛𝑧
Eq.8.21
𝑥 𝑥 𝑧

91
Lecture Note on GTE-I

+ = 0------------------
or2
𝛛 ∅ 𝛛2∅
𝛛𝑥2
Eq.8.22
𝛛𝑧2
This is the Laplace equation in terms of velocity potential.
8.2 Stream Function
Stream function is a scalar function of space and time such that its derivative in any direction
gives the component of velocity in the perpendicular direction clockwise. Thus a positive stream
function when derived would give negative velocity and vice versa, as shown in Fig. 8.2.
Referring to Fig. 2, we have

Fig.8.2: Definition of Stream Function


∂Ψ/∂x = – vz -------------------- Eq.8.23
∂Ψ/∂z = vx --------------------- Eq.8.24
8.3 Anisotropic Soil
Soils in nature do possess permeabilities which are different in the horizontal and vertical
directions. The permeability in the horizontal direction is greater than in the vertical direction in
sedimentary deposits and in most earth embankments. In loess deposits the vertical permeability
is greater than the horizontal permeability. The study of flow nets would be of little value if this
variation in the permeability is not taken into account. Laplace equation applies for a soil mass

𝛛 ℎ + = 0----------------
𝛛2ℎ
where anisotropy exists. This equation may be written in the form
2
𝑘𝑧
𝛛𝑥
Eq.8.25
𝛛𝑧2
2
𝑘𝑥
If we consider a new coordinate variable xc measured in the same direction as x multiplied by a
constant, expressed by

𝑥
= √ -------------------
𝑘𝑧

𝑥𝑥
Eq.8.26
𝑐 𝑘𝑥

So the Laplace Equation can be now rewritten as


92
Lecture Note on GTE-I

+ = 0----------------
𝛛2ℎ 𝛛2ℎ
Eq.8.27
𝛛𝑥�
2
𝛛𝑧2

Now this equation is a Laplace equation in the coordinates xc and z. This equation indicates that
a cross-section through an anisotropic soil can be transformed to an imaginary section which
possesses the same permeability in all directions. The transformation of the section can be
𝑘𝑧
effected by multiplying the x-coordinates by √ and keeping the z-coordinates at the natural
𝑘𝑥
scale. The flow net can be sketched on this transformed section. The permeability to be used
with the transformed section is
𝑘𝑒 = √𝑘𝑧𝑘𝑥----------------- Eq.8.28
From the transformed section, the rate of seepage can be determined using Eq. 7.1 with
exception that ke is to be substituted for k (see Fig. 8.3):

Fig.8.3: Flow in anisotropic soil

8.4 FLOW NET


A flow net for an isometric medium is a network of flow lines and equipotential lines
intersecting at right angles to each other. The path which a particle of water follows in its course
of seepage through a saturated soil mass is called a flow line. Equipotential lines are lines that
intersect the flow lines at right angles. At all points along an equipotential line, the water would
rise in piezometric tubes to the same elevation known as the piezometric head.
8.4.1 FLOW NET CONSTRUCTION
Properties of a Flow Net
The properties of a flow net can be expressed as given below:
1. Flow and equipotential lines are smooth curves.
2. Flow lines and equipotential lines meet at right angles to each other.
3. No two flow lines cross each other.
4. Two flow or equipotential lines can not start from the same point.

93
Lecture Note on GTE-I

Fig. 8.4: Flow through a homogeneous stratum of soil


8.4.2 Boundary Conditions
Flow of water through earth masses is in general three dimensional. Since the analysis of three-
dimensional flow is too complicated, the flow problems are solved on the assumption that the
flow is two-dimensional. All flow lines in such a case are parallel to the plane of the figure, and
the condition is therefore known as two-dimensional flow. All flow studies dealt with herein are
for the steady state case. The expression for boundary conditions consists of statements of head
or flow conditions at all boundary points. The boundary conditions are generally four in number
though there are only three in some cases. The boundary conditions for the case shown in Fig.
8.4 are as follows:
1 . Line ab is a boundary equipotential line along which the head is hf
2. The line along the sheet pile wall is a flow boundary
3. The line xy is a boundary equipotential line along which the head is equal to hd
4. The line m n is a flow boundary (at depth H below bed level).
If we consider any flow line, say, p1, p2, p3 in Fig.8.4, the potential head at p1 is h1 and at p3 is hd. The
total head lost as the water flows along the line is h which is the difference between the upstream
and downstream heads of water. The head lost as the water flows from p1 to

94
Lecture Note on GTE-I

equipotential line k is ∆ℎ which the difference between the heads is shown by the piezometers.
This loss of head ∆ℎ is a fraction of the total head lost.
8.5 FLOW NET CONSTRUCTION

Flow nets are constructed in such a way as to keep the ratio of the sides of each block bounded
by two flow lines and two equipotential lines a constant. If all the sides of one such block are
equal, then the flow net must consist of squares. The square block referred to here does not
constitute a square according to the strict meaning of the word; it only means that the average
width of the square blocks is equal. For example, in Fig.8.4, the width a1 of block 1 is equal to its
length b1. The area bounded by any two neighboring flow lines is called a flow channel. If the
flow net is constructed in such a way that the ratio a/b remains the same for all blocks, then it can
be shown that there is the same quantity of seepage in each flow channel. In order to show this
consider two blocks 1 and 2 in one flow channel and another block 3 in another flow channel as
shown in Figure. Block 3 is chosen in such a way that it lies within the same equipotential lines
that bound the block 2. Darcy's law for the discharge through any block such as 1 per unit length

∆𝑞 = 𝑘𝑖𝑎 = 𝑘 𝑎 = 𝑘∆ℎ( )-------------------


∆ℎ 𝑎
of the section may be written as

𝑏 𝑏
Eq.8.29
where ∆ℎ represents the head loss in crossing the block. The expressions in this form for each of

∆𝑞 = 𝑘∆ℎ and ∆𝑞 = 𝑘∆ℎ


the three blocks 𝑎under
1
consideration are𝑎2
--------------------- Eq.8.30
1 𝑏 2 2 𝑏
1 1 2
In the above equation the value of hydraulic conductivity k remains the same for all the blocks. If

= =1
𝑎1 𝑎2
the blocks are all squares then

𝑏1 𝑏2
Since blocks 1 and 2 are in the same flow channel, we have∆𝑞1 = ∆𝑞2. Since blocks 2 and 3
are within the same equipotential lines we have ∆ℎ1 = ∆ℎ2 If these equations are inserted we

∆𝑞1 = ∆𝑞2 and ∆ℎ1 = ∆ℎ2


obtain the following relationship:

This proves that the same quantity flows through each block and there is the same head drop in
crossing each block if all the blocks are squares or possess the same ratio a/b. Flow nets are
constructed by keeping the ratio a/b the same in all figures. Square flow nets are generally used
in practice as this is easier to construct.
There are many methods that are in use for the construction of flow nets. Some of the important
methods are
1. Analytical method,
2. Electrical analog method,
3. Scaled model method,
4. Graphical method.
The analytical method, based on the Laplace equation although rigorously precise, is not
universally applicable in all cases because of the complexity of the problem involved. The

95
Lecture Note on GTE-I

mathematics involved even in some elementary cases is beyond the comprehension of many
design engineers. Although this approach is sometimes useful in the checking of other methods,
it is largely of academic interest.
The electrical analogy method has been extensively made use of in many important for design
problems. However, in most of the cases in the field of soil mechanics where the estimation of
seepage flows and pressures are generally required, a more simple method such as the graphical
method is preferred.
Scaled models are very useful to solve seepage flow problems. Soil models can be constructed to
depict flow of water below concrete dams or through earth dams. These models are very useful
to demonstrate the fundamentals of fluid flow, but their use in other respects is limited because
of the large amount of time and effort required to construct such models.
The graphical method developed by Forchheimer (1930) has been found to be very useful in
solving complicated flow problems. A. Casagrande (1937) improved this method by
incorporating many suggestions. The main drawback of this method is that a good deal of
practice and aptitude are essential to produce a satisfactory flow net. In spite of these drawbacks,
the graphical method is quite popular among engineers.
8.5.1 Graphical Method

The usual procedure for obtaining flow nets is a graphical, trial sketching method, sometimes
called the Forchheimer Solution. This method of obtaining flow nets is the quickest and the most
practical of all the available methods. A. Casagrande (1937) has offered many suggestions to the
beginner who is interested in flow net construction. Some of his suggestions are summarized
below:
1. Study carefully the flow net pattern of well-constructed flow nets.
2. Try to reproduce the same flow nets without seeing them.
3. As a first trial, use not more than four to five flow channels. Too many flow channels would
confuse the issue.
4. Follow the principle of 'whole to part', i.e., one has to watch the appearance of the entire flow
net and when once the whole net is found approximately correct, finishing touches can be given
to the details.
5. All flow and equipotential lines should be smooth and there should not be any sharp
transitions between straight and curved lines.
The above suggestions, though quite useful for drawing flow nets, are not sufficient for a
beginner. In order to overcome this problem, Taylor (1948) proposed a procedure known as the
procedure by explicit trials. Some of the salient features of this procedure are given below:
1. As a first step in the explicit trial method, one trial flow line or one trial equipotential
line is sketched adjacent to a boundary flow line or boundary equipotential.
2. After choosing the first trial line (say it is a flow line), the flow path between the line and the
boundary flow line is divided into a number of squares by drawing equipotential lines. These

96
Lecture Note on GTE-I

equipotential lines are extended to meet the bottom flow line at right angles keeping in view that the
lines drawn should be smooth without any abrupt transitions.
3. The remaining flow lines are next drawn, adhering rigorously to square figures.
4. If the first trial is chosen property, the net drawn satisfies all the necessary conditions.
Otherwise, the last drawn flow line will cross the bottom boundary flow line, indicating that the
trial line chosen is incorrect and needs modification.
5. In such a case, a second trial line should be chosen and the procedure repeated.
A typical example of a flow net under a sheet pile wall is given in Fig. 1. It should be
understood that the number of flow channels will be an integer only by chance. That means, the
bottom flow line sketched might not produce full squares with the bottom boundary flow line. In
such a case the bottom flow channel will be a fraction of a full flow channel. It should also be
noted that the figure formed by the first sketched flow line with the last equipotential line in the
region is of irregular form. This figure is called a singular square. The basic requirement for such
squares, as for all the other squares, is that continuous sub-division of the figures give an
approach to true squares. Such singular squares are formed at the tips of sheet pile walls also.
Squares must be thought of as valid only where the Laplace equation applies. The Laplace
equation applies to soils which are homogeneous and isotropic. When the soil is anisotropic, the
flow net should be sketched as before on the transformed section. The transformed section can be
obtained from the natural section explained earlier.
8.6 DETERMINATION OF QUANTITY OF SEEPAGE
Flow nets are useful for determining the quantity of seepage through a section. The quantity of seepage

∆𝑞 = 𝑘∆ℎ--------------
q is calculated per unit length of the section. The flow through any square can be written as
Eq.8.31
Let the number of flow channel and equipotential drops in a section be Nf and Nd, respectively. Since all

∆ℎ = -------------------

drops are equal, we can write
Eq.8.32
𝑁𝑑

𝑞 = 𝑁𝑓∆𝑞----------------------
Since the discharge in each flow channel is the same we can write,

Substituting for ∆𝑞 and ∆ℎ we have


Eq.8.33

𝑁𝑓
𝑞 = 𝑘ℎ ----------------- Eq.8.34
𝑁𝑑
This Eq. can also be used to compute the seepage through anisotropic sections by writing ke in place of
k, where ke is equal to √𝑘𝑥𝑘𝑧 , where kx and kz are the hydraulic conductivities in the x and z
directions, respectively. The validity of this relationship can be proved as follows.
Consider a figure bounded by flow and equipotential lines in which the flow is parallel to the x
direction. In Figure 8.5 the figure in question is drawn to a transformed scale in (b) and the same to the
natural scale in (a). In Fig. 8.5 (b) the permeability has the effective value ke in both the x and z
directions and the flow through the square according to
∆𝑞 = ∆ℎ𝑎
= ∆ℎ---
𝑘𝑒 𝑘
Eq.8.35
� 𝑎

97
Lecture Note on GTE-I

Fig.8.5: Flow through anisotropic soil


In Fig. 8.5 (a) the hydraulic conductivity kx in the horizontal section must apply because the flow
is horizontal and the sketch is to the natural scale. The flow equation is, therefore,
∆𝑞 = 𝑖𝐴 = 𝑥
∆ℎ𝑎

𝑘𝑥 𝑘
----------------- Eq.8.36
𝑘𝑥
𝑎×√
𝑘𝑧
8.7 DETERMINATION OF SEEPAGE PRESSURE
Flow nets are useful in the determination of the seepage pressure at any point along the flow
path.
Consider the cubical element 1 in Fig.8.4 (a) with all the sides equal to a. Let h1 be the
piezometric head acting on the face kt and h2 on face jo.

The total force on face yo = P2 = 𝑎2𝛾𝑤ℎ2-------


The total force on face kt = P1=𝑎2𝛾𝑤ℎ1------------------- Eq.8.37
Eq.8.38

𝑃1 − 𝑃2 = 𝑎2𝛾𝑤(ℎ1 − ℎ2)------------
The differential force acting on the element is

∆ℎ , we can write
Eq.8.39

𝑃 = 𝑎2𝛾 (∆ℎ) = 𝑎3 ∆ℎ=𝑎3𝛾 𝑖-------------


𝛾𝑤
Since (h1 – h2) is the head drop
3 𝑤
Eq.8.40
𝑎 𝑤

PS = iγw-----------------
3
Where, a is the volume of the element. The force per unit volume of the element is, therefore
Eq.8.41
This force exerts a drag on the element known as the seepage pressure. It has the dimension of
unit weight, and at any point its line of action is tangent to the flow line. The seepage pressure is
a very important factor in the stability analysis of earth slopes. If the line of action of the seepage
force acts in the vertical direction upward as on an element adjacent to point x, in Fig.8.6 (a), the
force that is acting downward to keep the element stable is the buoyant unit weight of the
element. When these two forces balance, the soil will just be at the point of being lifted up, and
there will be effectively no grain-to-grain pressures.

98
Lecture Note on GTE-I

Fig.8.6: Flow-net in anisotropic soil

8.8 DETERMINATION OF UPLIFT PRESSURES


Water that seeps below masonry dams or weirs founded on permeable soils exerts pressures on
the bases of structures. These pressures are called uplift pressures. Uplift pressures reduce the
effective weight of the structure and thereby cause instability. It is therefore very essential to
determine the uplift pressures on the base of dams or weirs accurately. Accurate flow nets should
be constructed in cases where uplift pressures are required to be determined. The method of
determining the uplift pressures can be explained as follows.
Consider a concrete dam as shown in Figure 8.7 a founded on a permeable foundation at a depth
D below the ground surface. The thickness of the permeable strata is H. The depth of water on
the upstream side is hl and on the downstream side is zero. Water flows from the upstream to the
downstream side. It is necessary to determine the uplift pressure on the base of the dam by means
of flow nets as shown in the figure.

of equipotential drops be Nd. The head lost per drop be ∆h =


h𝐹
The difference in head between the upstream and downstream water levels is hf Let the number
. As the water flows along the
Nd
side and base of the dam, there will be equal drops of head between the equipotential lines that
meet the dam as shown in the figure. A piezometer tube at point a (coinciding with the corner of
the dam in the figure) gives a pressure head ha. Now the uplift pressure at point a may be

𝑢𝑎 = ℎ𝑎𝛾𝑤 = (ℎ𝑓 + 𝐷 − ∆ℎ)𝛾𝑤-----------


expressed as
Eq.8.42
Similarly, the uplift pressure at any other point, say e (see the figure), may be estimated from the
expression

99
Lecture Note on GTE-I

𝑢𝑒 = (ℎ𝑙 + 𝐷 − 𝑛𝑑∆ℎ)𝛾𝑤-------------- Eq.8.43


where, nd = the number of equipotential drops to the point e.
Fig. 8.7.b shows the distribution of uplift pressure on the base of the dam.

Fig.8.7: Uplift pressure on the base of a concrete dam


Example 1
Two lines of sheet piles were driven in a river bed as shown in Fig. The depth of water over the
river bed is 8.20 ft. The trench level within the sheet piles is 6.6 ft below the river bed. The water
level within the sheet piles is kept at trench level by resorting to pumping. If a quantity of water
flowing into the trench from outside is 3.23 ft 3/hour per foot length of sheet pile, what is the
hydraulic conductivity of the sand? What is the hydraulic gradient immediately below the trench
bed?

100
Lecture Note on GTE-I

Solution:
Fig. gives the flow net and other details. The differential head between the bottom of trench and
the water level in the river is 14.8 ft.
Number of channels = 6
Number of equipotential drops =10

𝑁𝑓
Discharge

𝑞 = 𝑘ℎ
�𝑑
-4
From which, k= 1x10 ft/sec �
The distance between the last two equipotentials given is 2.95 ft. The calculated hydraulic

i = =0.50
∆h
gradient is

∆s
8.9 SEEPAGE FLOW THROUGH HOMOGENEOUS EARTH DAMS
In almost all problems concerning seepage beneath a sheet pile wall or through the foundation of
a concrete dam all boundary conditions are known. However, in the case of seepage through an
earth dam the upper boundary or the uppermost flow line is not known. This upper boundary is a
free water surface and will be referred to as the line of seepage or phreatic line. The seepage line
may therefore be defined as the line above which there is no hydrostatic pressure and below
which there is hydrostatic pressure. In the design of all earth dams, the following factors are very
important.
1. The seepage line should not cut the downstream slope.
2. The seepage loss through the dam should be the minimum possible.
The two important problems that are required to be studied in the design of earth dams are:
1. The prediction of the position of the line of seepage in the cross-section.
2. The computation of the seepage loss.
If the line of seepage is allowed to intersect the downstream face much above the toe, more or
less serious sloughing may take place and ultimate failure may result. This mishap can be
prevented by providing suitable drainage arrangements on the downstream side of the dam. The
section of an earth dam may be homogeneous or non-homogeneous. A homogeneous dam
contains the same material over the whole section and only one coefficient of permeability may
be assumed to hold for the entire section. In the non homogeneous or the composite section, two
or more permeability coefficients may have to be used according to the materials used in the
section. When a number of soils of different permeabilities occur in a cross-section, the
prediction of the position of the line of seepage and the computation of the seepage loss become
quite complicated.

101
Lecture Note on GTE-I

Fig.8.8: Basic parabola and the phreatic line for a homogeneous earth dam
It has been noticed from experiments on homogeneous earth dam models that the line of seepage
assumes more or less the shape of a parabola as illustrated in Fig. 8.8. In some sections a little
divergence from a regular parabola is required at the surfaces of entry and discharge of the line
of seepage. In some ideal sections where conditions are favorable the entire seepage line may be
considered as a parabola. When the entire seepage line is a parabola, all the other flow lines will
be confocal parabolas. The equipotential lines for this ideal case will be conjugate confocal
parabolas as shown in Fig.1. As a first step it is necessary to study the ideal case where the entire
flow net consists of conjugate confocal parabolas.
8.9.1 Flow net consisting of conjugate confocal parabolas
As a prelude to the study of an ideal flow net comprising of parabolas as flow and equipotential
line, it is necessary to understand the properties of a single parabola. The parabola ACV
illustrated in Fig.8.9, is defined as the curve whose every point is equidistant from a point F
called the focus and a line DG called the directrix. If we consider any point, say, A, on the curve,
we can write FA = AG, where the line AG is normal to the directrix. If F is the origin of
coordinates, and the coordinates of point A are (x, y), we can write
𝐴𝐹 = 𝑦√𝑥2 2+ 𝑦2 = 𝐴𝐺 = 𝑥 + 𝑦0------ Eq.8.44
2
−𝑦
Or 𝑥 = 0
----------------------------- Eq.8.45
2𝑦0
where, y0 = FD
Eq.8.45 is the equation of the basic parabola. If the parabola intersects the y-axis at C, we can
write
FC=CE = y0
Similarly for the vertex point V, the focal distance a0 is
𝐹𝑉 = 𝑉𝐷 𝑦0
= 𝑎0 =2
Figure 8.9 illustrates the ideal flow net consisting of conjugate confocal parabolas. All the
parabolas have a common focus F.
The boundary lines of such an ideal flow net are:
1 . The upstream face AB, an equipotential line, is a parabola.
102
Lecture Note on GTE-I

2. The downstream discharge face FV, an equipotential line, is horizontal.


3. ACV, the phreatic line, is a parabola.
4. BF, the bottom flow line, is horizontal.
The known boundary conditions are only three in number. They are, the two equipotential
lines AB and FV, and the bottom flow line BF. The top flow line ACV is the one that is unknown.
The theoretical investigation of Kozeny (1931) revealed that the flow net for such an ideal
condition mentioned above with a horizontal discharge face FV consists of two families of
confocal parabolas with a common focus F. Since the conjugate confocal parabolas should
intersect at right angles to each other, all the parabolas crossing the vertical line FC should have
their intersection points lie on this line.
Since the seepage line is a line of atmospheric pressure only, the only type of head that can exist
along it is the elevation head. Therefore, there must be constant drops in elevation between the
points at which successive equipotentials meet the top flow line, as shown in Fig.1.
In all seepage problems connected with flow through earth dams, the focus F of the basic
parabola is assumed to lie at the intersection of the downstream discharge face FV and the
bottom flow line BF as shown in Fig. 8.9. The point F is therefore known. The point A, which is
the intersection point of the top flow line of the basic parabola and the upstream water level, is
also supposed to be known. When the point A is known, its coordinates (d, K) with respect to the
origin F can be determined. With these two known points, the basic parabola can be constructed
as explained below. We may write
𝐴𝐹 = 𝐴𝐺 = √𝑑2 + ℎ2----------------------- Eq. 8.46
𝐹𝐷 = 2𝑎0 = √𝑑2 + ℎ2 − 𝑑----------------- Eq. 8.47
= √𝑑2 + ℎ2 − 𝑑---------------------
1

𝑎0
Or
2
Eq. 8.48

Fig.8.9: Ideal flow net consisting of conjugate confocal parabolas


103
Lecture Note on GTE-I

8.9.2 Seepage Loss Through the Dam


The seepage flow q across any section can be expressed according to Darcy's law as
q = kiA
Considering the section FC in Fig. 8.9, where the sectional area A is equal to y0, the hydraulic
gradient can be determined analytically as follows:
From Eq. (2), the equation of the parabola can be expressed as
𝑦 = √2𝑥𝑦0 +0 𝑦2-------------------------- Eq. 8.49
The hydraulic gradient i at any point on the seepage line in Fig. 8.9 can be expressed as

=
𝑑𝑦 ------------------------------ Eq.8.50
𝑦0
𝑑
𝑥 √2𝑥𝑦0+𝑦2
0

= = 1-----------------------------
𝑦0
For the point C which has coordinates (0, y0), the hydraulic gradient from Eq. (8.50) is
𝑑𝑦

𝑑
Eq.8.51
𝑥 √𝑦02

𝑞=𝑘 𝑦 = 𝑘𝑦 ------------------------
𝑑𝑦
Therefore, the seepage quantity across section FC is
Eq.8.52
𝑑𝑥 0 0
8.9.3 Seepage Through Homogeneous and Isotropic Earth Dams
Types of Entry and Exit of Seepage lines
The flow net consisting of conjugate confocal parabolas is an ideal case which is not generally
met in practice. Though the top flow line resembles a parabola for most of its length, the
departure from the basic parabola takes place at the faces of entry and discharge of the flow line.
The departure from the basic parabola depends upon the conditions prevailing at the points of
entrance and discharge of the flow line as illustrated in Fig. 8.10 from (a) to (e).

Fig.8.10 Types of entry and exit of seepage lines


The seepage line should be normal to the equipotential line at the point of entry as shown in Fig.
8.10 (a). However, this condition is violated in Fig. 8.10 (b), where the angle made by the
104
Lecture Note on GTE-I

upstream face AB with the horizontal is less than 90°. It can be assumed in this case the coarse
material used to support the face AB is highly permeable and does not offer any resistance for

105
Lecture Note on GTE-I

flow. In such cases AB taken as the upstream equipotential line. The top flow line cannot therefore be
normal to the equipotential line. However, this line possesses zero gradient and velocity at the
point of entry. This zero condition relieves the apparent inconsistency of deviation from a normal
intersection. The conditions prevailing at the downstream toe of the dam affect the type of exit of
the flow line at the discharge face. In Fig. 8.10 (c) the material at the toe is the same as in the
other parts of the dam whereas in (d) and (e) rock toe drains are provided. This variation in the
soil condition at the toe affects the exit pattern of the flow line. The flow line will meet the
discharge face FE tangentially in 8.10 (c). This has to be so because the particles of water as they
emerge from the pores at the discharge face have to conform as nearly as possible to the direction
of gravity. But in cases where rock toe drains are provided, the top flow line becomes tangential
to the vertical line drawn at the point of exit on the discharge face as shown in (d) and (e) of Fig.
8.10.
8.10 METHOD OF LOCATING SEEPAGE LINE
The general method of locating the seepage line in any homogeneous dam resting on an impervious
foundation may be explained with reference to Fig. 8.11 (a).
As explained earlier, the focus F of the basic parabola is taken as the intersection point of the bottom
flow line BF and the discharge face EF. In this case the focus coincides with the toe of the dam.
One more point is required to construct the basic parabola. Analysis of the location of seepage
lines by A. Casagrande has revealed that the basic parabola with focus F intersects the upstream
water surface at A such that AA'= 0.3m, where m is the projected length of the upstream
equipotential line A'B on the water surface. Point A is called the corrected entrance point.

Fig.8.11: Construction of seepage line

106
Lecture Note on GTE-I

The parabola APSV may now be constructed as per Eq. (2). The divergence of the seepage line
from the basic parabola is shown as AT1 and SD in Fig. 8.11 (a). For dams with flat slopes, the
divergences may be sketched by eye keeping in view the boundary requirements. The error
involved in sketching by eye, the divergence on the downstream side, might be considerable if
the slopes are steeper. Procedures have therefore been developed to sketch the downstream
divergence as explained below. As shown in Fig. 8.11 (a), E is the point at which the basic

DF as a. The values of ∆𝑎 and a + ∆a, vary with the angle, β, made by the discharge face with
parabola intersects the discharge face. Let the distance ED be designated as Aa and the distance

the horizontal measured clockwise. The angle may vary from 30° to 180°. The discharge face is
horizontal as shown in Fig. 8.10(e). Casagrande (1937) determined the ratios of ∆𝑎 for a
(𝑎+∆𝑎)
number of discharge slopes varying from 30° to 180° and the relationship is shown in a graphical

with F as the focus. With the known (a +∆𝑎) and the discharge face angle β, ∆𝑎 can be
form Fig. 8.11 (b). The distance (a +∆a) can be determined by constructing the basic parabola

determined from Fig. 8.11 (b). The point D may therefore be marked out at a distance of ∆𝑎
from
E. With the point D known, the divergence DS may be sketched by eye.
It should be noted that the discharge length a, is neither an equipotential nor a flow line, since it
is at atmospheric pressure. It is a boundary along which the head at any point is equal to the
elevation.
8.10.1 Analytical Solutions for Determining a and q

𝑎
Casagrande (1937) proposed the following equation for determining ‘a’ for β < 30°
=
𝑑
− √
𝑑2

𝑐𝑜𝑠2 − 𝑠𝑖𝑛2
𝑐𝑜𝑠𝛽 ℎ2 ---------------------- Eq. 8.53
𝛽 𝛽

L. Casagrande (1932) gave the following equation for ‘a’ when β lies between 30° and 90°.
𝑎 = √ℎ2 + 𝑑2 − √𝑑2 − ℎ2𝑐𝑜𝑡2𝛽-------------- Eq. 8.54
The discharge q per unit length through any cross-section of the dam may be expressed as

For β<300, 𝑞 = 𝑘𝑎𝑠𝑖𝑛𝛽𝑐𝑜𝑠𝛽-------------------


follows:

For 30 <β<90 , 𝑞 = 𝑘𝑎𝑠𝑖𝑛 𝛽------------------


Eq. 8.55
0 0 2
Eq. 8.56
8.11 PIPING FAILURE
Piping failures caused by heave can be expected to occur on the downstream side of a hydraulic
structure when the uplift forces of seepage exceed the downward forces due to the submerged
weight of the soil.
The mechanics of failure due to seepage was first presented by Terzaghi. The principle of this
method may be explained with respect to seepage flow below a sheet pile wall. Fig. 8.12 (a) is a
sheet pile wall with the flow net drawn. The uplift pressures acting on a horizontal plane ox can
be determined. The ordinates of curve C in Fig. 8.12 (b) represent the uplift pressure at any point
on the line ox. It is seen that the uplift pressure is greatest close to the wall and gradually
becomes less with an increase in the distance from the wall. When the upward forces of seepage
107
Lecture Note on GTE-I

on a portion of ox near the wall become equal to the downward forces exerted by the submerged
soil, the surface of the soil rises as shown in Fig. 8.12 (a). This heave occurs simultaneously with

108
Lecture Note on GTE-I

an expansion of the volume of the soil, which causes its permeability to increase. Additional
seepage causes the sand to boil, which accelerates the flow of water and leads to complete
failure. Terzaghi determined from model tests that heave occurs within a distance of about D/2
(where D is the depth of penetration of the pile) from the sheet pile and the critical section ox
passes through the lower edge of the sheet pile.

Fig. 8.12: Piping failure

8.11.1 Filter Requirements to Control Piping


Filter drains are required on the downstream sides of hydraulic structures and around drainage
pipes. A properly graded filter prevents the erosion of soil in contact with it due to seepage
forces.
To prevent the movement of erodible soils into or through filters, the pore spaces between the
filter particles should be small enough to hold some of the protected materials in place. Taylor
(1948) shows that if three perfect spheres have diameters greater than 6.5 times the diameter of a
small sphere, the small spheres can move through the larger as shown in Fig. 8.13 (a). Soils and
aggregates are always composed of ranges of particle sizes, and if pore spaces in filters are small
enough to hold the 85 per cent size (D85) of the protected soil in place, the finer particles will
also be held in place as exhibited schematically in Fig. 8.13(b).
The requirements of a filter to keep the protected soil particles from invading the filter
significantly are based on particle size. These requirements were developed from tests by
Terzaghi which were later extended by the U.S. Army Corps of Engineers (1953). The resulting
filter specifications relate the grading of the protective filter to that of the soil being protected by

≤ 4, 4 < ≤ 20 and 𝐷50𝑓𝑖𝑙𝑡𝑒𝑟 ≤ 25-------------------


𝐷15𝑓𝑖𝑙𝑡𝑒𝑟 𝐷15𝑓𝑖𝑙𝑡𝑒𝑟
the following;
Eq.8.57
𝐷85𝑠𝑜 𝐷15𝑠𝑜 𝐷50𝑠𝑜𝑖𝑙
𝑖𝑙 𝑖𝑙

109
Lecture Note on GTE-I

Fig.8.13: Requirements of a filter


The criteria may be explained as follows:
1.
The 15 per cent size (D15) of filter material must be less than 4 times the 85 per cent size (D85)
of a protected soil. The ratio of D15 of a filter to D85 of a soil is called the piping ratio.
2.
The 15 per cent size (D15) of a filter material should be at least 4 times the 15 per cent size
(D]5) of a protected soil but not more than 20 times of the latter.
3.
The 50 per cent size (D50) of filter material should be less than 25 times the 50 per cent size
(D50) of protected soil.
Experience indicates that if the basic filter criteria mentioned above are satisfied in every part of
a filter, piping cannot occur under even extremely severe conditions.
A typical grain size distribution curve of a protected soil and the limiting sizes of filter materials
for constructing a graded filter is given in Figure 8.14. The size of filter materials must fall
within the two curves C2 and C3 to satisfy the requirements.

Fig.8.14: Grain size distribution curves for graded filter and protected materials

110
Lecture Note on GTE-I

9.0 COMPACTION
In construction of highway embankments, earth dams and many other engineering structures,
loose soils must be compacted to improve their strength by increasing their unit weight;
Compaction - Densification of soil by removing air voids using mechanical equipment; the
degree of compaction is measured in terms of its dry unit weight.
Compaction, in general, is the densification of soil by removal of air, which requires mechanical
energy. The degree of compaction of a soil is measured in terms of its dry unit weight. When
water is added to the soil during compaction, it acts as a softening agent on the soil particles. The
soil particles slip over each other and move into a densely packed position. The dry unit weight
after compaction first increases as the moisture content increases.
When the moisture content is gradually increased and the same compactive effort is used for
compaction, the weight of the soil solids in a unit volume gradually increases.
Beyond a certain moisture content, any increase in the moisture content tends to reduce the dry
unit weight. This phenomenon occurs because the water takes up the spaces that would have
been occupied by the solid particles. The moisture content at which the maximum dry unit
weight is attained is generally referred to as the optimum moisture content.

9.1 Objectives for Compaction


Increasing the bearing capacity of foundations;
Decreasing the undesirable settlement of structures;
Control undesirable volume changes;
Reduction in hydraulic conductivity;
Increasing the stability of slopes.
In general, soil densification includes compaction and consolidation.
Compaction is one kind of densification that is realized by rearrangement of soil particles
without outflow of water. It is realized by application of mechanic energy. It does not involve
fluid flow, but with moisture changing altering.
Consolidation is another kind of densification with fluid flow away. Consolidation is primarily
for clayey soils. Water is squeezed out from its pores under load.
9.2 FACTORS AFFECTING COMPACTION
The compaction of soil depends on the following factors:
1 Water content
2 Compactive effort
3 Type of soil
4 Method of compaction
Water content: Water content has significant effect on compaction characteristics of soil. At low
water content, soil is stiff and soil grains offer more resistance to compaction. As the water
content increases, water films are formed around the soil grains and water around the soil grains
act as lubricant. Due to this, soil grains come close to each other and make a dense configuration.
At optimum moisture content, soil reaches the maximum unit weight as lubrication effect is the

111
Lecture Note on GTE-I

maximum at this stage. Further addition of more water replaces the soil grains. Thus, addition of
more water after optimum moisture content reduces the unit weight as unit weight of water is
less than the unit weight of soil grains. Because of the due to the addition of water dry unit
weight of soil increases up to a certain water content value (OMC) and beyond that addition of
more water decreases the dry unit weight of soil (as shown in Fig.9.1). This can be also
explained by concept of soil structure and electrical double layer theory.

Fig.9.1: Compaction curves for clay


Zero-air voids line: It means the soil is fully saturated, with the air is not present in the pores of
the soil mass. It is obtained from the compaction test. The zero-air void line drawn with the
results, obtained from the compaction test as shown below.

Compactive effort: As the compactive effort increases maximum dry unit weight increases in a
decreasing rate. However, the OMC value decreases as the compactive effort increases. Thus, if
the compactive effort increases the compaction curve is shifted to the top and to the left side as
shown in Figure.2. However, as the water content increases the effect of compactive effort on
dry unit weight decreases. If the peaks of the compaction curves for different compactive efforts
are joined the obtained line is called line of optimum (as shown in Figure 2). The line of
optimum is nearly parallel to line of zero air voids. Thus, the efficiency of the compaction does
not increase as the compactive effort increases.

112
Lecture Note on GTE-I

Fig.9.3: Effect of Compactive effort

From the preceding observation and Figure 9.3, we can see that
1. As the compaction effort is increased, the maximum dry unit weight of compaction is
also increased.
2. As the compaction effort is increased, the optimum moisture content is decreased to
some extent.
The preceding statements are true for all soils. Note, however, that the degree of compaction is
not directly proportional to the compaction effort.
Type of soil: Soil type has significant effect on compaction. It is observed that cohesive soil has
higher OMC value and poorly graded sands have lower dry unit weight as compared to the
coarse-grained, well graded soil with some percentage of fines. However, excessive amount of
fines reduces the maximum dry unit weight value of the soil. Maximum dry unit weight of sand
is obtained either completely dry or saturated condition. Compaction curve has very little
importance in case of sandy soils.

Method of compaction: It is observed that mode of compaction influences the shape of


compaction curves. The shape of compaction curves for standard proctor test and modified
proctor test are somewhat different. The shape of the compaction curve obtained from laboratory
tests and obtained from field by various compactive methods are also different.

9.3 EFFECT OF COMPACTION ON SOIL PROPERTIES


Lambe (1958a) studied the effect of compaction on the structure of clay soils. If clay is
compacted with a moisture content on the dry side of the optimum, as represented by point, it
will possess a flocculent structure. This type of structure results because, at low moisture
content, the diffuse double layers of ions surrounding the clay particles cannot be fully
developed; hence, the inter particle repulsion is reduced. This reduced repulsion results in a more
random particle orientation and a lower dry unit weight. When the moisture content of

113
Lecture Note on GTE-I

compaction is increased, as shown by point B, the diffuse double layers around the particles
expand, which increases the repulsion between the clay particles and gives a lower degree of
flocculation and a higher dry unit weight. A continued increase in moisture content expands the
double layers more. This expansion results in a continued increase of repulsion between the
particles and thus a still greater degree of particle orientation and a more or less dispersed
structure. However, the dry unit weight decreases because the added water dilutes the
concentration of soil solids per unit volume. At given moisture content, higher compactive effort
yields a more parallel orientation to the clay particles, which gives a more dispersed structure.
The particles are closer and the soil has a higher unit weight of compaction.
Compaction induces variations in the structure of cohesive soils. Results of these structural
variations include changes in hydraulic conductivity, compressibility, and strength. The
hydraulic conductivity, which is a measure of how easily water flows through soil, decreases
with the increase of moisture content. It reaches a minimum value at approximately the optimum
moisture content. Beyond the optimum moisture content, the hydraulic conductivity increases
slightly. The high value of the hydraulic conductivity on the dry side of the optimum moisture
content is due to the random orientation of clay particles that results in larger pore spaces.
Some expansive clays in the field do not stay compacted, but expand upon entry of water and
shrink with loss of moisture. This shrinkage and swelling of soil can cause serious distress to the
foundations of structures. Laboratory observations such as this will help soils engineers to adopt
a moisture content for compaction to minimize swelling and shrinkage.
When a soil is compacted, it changes its engineering properties and thereby behaves differently.
Some of the engineering properties which changes on application of compactive effort is briefly
described below.
1. PERMEABILITY
The effect of compaction is to decrease the permeability. In the case of fine grained soils it has
been found that for the same dry density soil compacted wet of optimum will be less permeable
than that of compacted dry of optimum.
2. COMPRESSIBILITY
In case of soil samples initially saturated and having same void ratio, it has been found that in
low pressure range a wet side compacted soil is more compressible than a dry side compacted
soil, and vice versa in high pressure range.
3. PORE PRESSURE
In undrained shear test conducted on saturated samples of clay it has been found that lower pore
pressures develop at low strains when the sample is compacted dry of optimum, compared to the
case when the sample is compacted wet of optimum. But at high strains in both types of samples
the development of pore pressure is same for same density and water content.
4. STRESS-STRAIN RELATION
Samples compacted dry of optimum produce much steeper stress-starin curves with peaks at low
strains, whereas samples compacted wet of optimum, having the same density, produce much
flatter stress-strain curves with increase in stress even at high strains.

114
Lecture Note on GTE-I

5. SHRINKAGE AND SWELLING


At same density a soil compacted dry of optimum shrinks appreciably less than that of
compacted wet of optimum. Also the sol compacted dry of optimum exhibits greater swelling
characteristics than samples o f the same density compacted wet of optimum.

9.4 COMPACTION TESTS


Compaction test is used to establish the relationship between the moisture content and density of
soils compacted in a mould of a given size with a 2.5 kg rammer dropped from a height of 30
cm. the results obtained from this test will be helpful in increasing the bearing capacity of
foundations, Decreasing the undesirable settlement of structures, Control undesirable volume
changes, Reduction in hydraulic conductivity, Increasing the stability of slope sand so on.

APPARATUSREQUIRED:

1. Proctor mould having a capacity of 944 cc with an internal diameter of 10.2 cm and a height
of 11.6 cm. The mould shall have a detachable collar assembly and a detachable base plate.

2. Rammer: A mechanical operated metal rammer having a 5.08 cm diameter face and a weight
of 2.5 kg. The rammer shall be equipped with a suitable arrangement to control the height of
drop to a free fall of 30 cm.

3. Sample extruder, mixing tools such as mixing pan, spoon, towel, and spatula.

4. A balance of 15 kg capacity, Sensitive balance, Straight edge, Graduated cylinder, Moisture


tins.

PROCEDURE:

1. Take a representative oven-dried sample, approximately 5 kg in the given pan. Thoroughly


mix the sample with sufficient water to dampen it with approximate water content of 4-6 %.

2. Weigh the proctor mould without base plate and collar. Fix the collar and base plate. Place the
soil in the Proctor mould and compact it in 3 layers giving 25 blows per layer with the 2.5 kg
rammer falling through. The blows shall be distributed uniformly over the surface of each layer.
3. Remove the collar; trim the compacted soil even with the top of mould using a straight edge
and weigh to find the bulk density,

4. Divide the weight of the compacted specimen by 944 cc and record the result as the bulk
density.

5. Remove the sample from mould and slice vertically through and obtain a small sample for
water content.

115
Lecture Note on GTE-I

6. Thoroughly break up the remainder of the material until it will pass through 4.25mm sieve as
judged by the eye. Add water in sufficient amounts to increase the moisture content of the soil
sample by one or two percentage points and repeat the above procedure for each increment of
water added. Continue this series of determination until there is either a decrease or no change in
the wet unit weight of the compacted soil.

Fig.9.4: Typical mould section with dimension


3
Volume cm A dia B C dia D E dia F
1000 100 127.3 110 150 120 180φ or 150 sq.
2250 150 127.3 160 200 170 230φ or 200 sq.

Calculation
First, the compaction water content (w) of the soil sample is calculated using the average of the three
measurements obtained (top, middle and bottom part of the soil mass). Subsequently, the dry
unit weight (γd) is calculated as follows:
𝛾
= 1+
𝛾 --------------- Eq.9.1
𝑑 𝑤

116
Lecture Note on GTE-I

Fig.9.5: Typical section of hammer


9.4.1 The Compaction Curve
The weight of each specimen is used to calculate wet unit weights and the oven-dried moistures
are used to determine a dry unit weight for each point. The results are plotted on a graph as dry
unit weight vs. moisture content and will show the curvilinear relationship that allows the
maximum dry weight and optimum moisture for each type of soil to be established. A curve
showing values at 100% saturation, or zero air voids is plotted on the same graph as detailed in
the test method. This curve helps define the shape of the wet side of the maximum density curve.
When performing a field density test, the dry unit weight is compared to the maximum dry
weight of the Proctor tests to express a percent of compaction. When the required densities are
hard to reach in the field, the moisture contents should be compared to the optimum moisture
content from the lab tests. If the differences are more than 2 to 3%, it will help to adjust the field
moistures accordingly by aerating/discing the soil, or by adding water.

Fig.9.1: Compaction curve for Standard Proctor Test

117
Lecture Note on GTE-I

9.5 MODIFIED PROCTOR TEST


Objective
To obtain the graphical relationship of the “dry density’ to “moisture content” in the form of
“compaction curve“, for determining the values of Optimum Moisture Content (OMC) and
Maximum Dry Density (MDD).
Apparatus Required
Proctor Mould & Metal Rammer
Metal mould (volume = 1000 cm3 for 100 mm diameter mould and volume= 2250 cm 3 for 150
mm diameter mould (as per IS: 10074-1982) & Metal rammer conforming to IS: 9189-1979.
(weight = 4.9 kg)
Balance
Balance: one of capacity 10 kg and least count 1g and other of 200 g capacity and sensitivity
0.01 g
Sieves: 4.75mm, 19 mm and 37.5 mm I.S. Sieves conforming to IS: 460 (Part 1) –
1985 Oven: Thermostatically controlled to maintain temperature between 1050 to
1100C Steel Straight Edge: For trimming the protruded excessive soil of the mould
Airtight Container: Taking sample for determination of Moisture Content
Procedure

1. Take a representative portion of air-dried soil large enough to provide about 5 kg of


material passing through 4.25 mm sieve.

2. Add suitable amount of water with the soil and mix it thoroughly. For sandy and gravelly
soil add 3% to 5% of water. For cohesive soil the amount of water to be added should be
12% to 16% below the plastic limit.

3. Weigh the mould with base plate attached, to the nearest 1g and record the weight as W1.
Attach the extension collar with the mould. Compact the moist soil into the mould in five
layers of approximately equal mass, each layer being given 25 blows, with the help of 4.9
kg rammer, dropped from a height of 450mm above the soil. The blows must be
distributed uniformly over the surface of each layer. The operator shall ensure that the
tube of the rammer is kept clear of soil so that the rammer always falls freely.

4. After completion of the compaction operation, remove the extension collar and level
carefully the top of the mould by means of straightedge. Weigh the mould with the
compacted soil to the nearest 1 g and record this weight as W2.

5. Remove the compacted soil from the mould and place it on the mixing tray. Determine
the water content of a representative sample of the specimen. Record the moisture content
as ‘M’.

6. The remainder of the soil shall be broken up and repeat Steps (iii) to (v) above, by adding
suitable increment of water to the soil. For sandy and gravelly soils the increment is

118
Lecture Note on GTE-I

generally 1% to 2% and for cohesive soils the increment is generally 2% to 4%. The total
number of determinations made shall be at least five, and the moisture contents should be
such that the optimum moisture content, at which the maximum dry density occurs, is
within that range.

7. For compacting soil containing coarse material up to 37.5 mm size, the 2250 cm3 mould
should be used. A sample weighing about 30 kg and passing the 37.5 mm IS sieve is used
for the test. Soil is compacted in five layers, each layer being given 55 blows of the 4.9
kg rammer.

NOTE 1: It is important that the water is mixed thoroughly and adequately with the soil, since
inadequate mixing gives rise to variable test results. This is particularly important with
cohesive souls when adding a substantial quantity of water to the air-dried soil. With clays of
high plasticity, or where hand mixing is employed, it may be difficult to distribute the water
uniformly through the air-dried soil by mixing alone, and it may be necessary to store the
mixed sample in a sealed container for a minimum period of about 16 hours before continuing
with the test.
NOTE 2: It is necessary to control the total volume of soil compacted, since it has been found that if the
amount of soil struck off after removing the extension is too great, the test results will be
inaccurate.
NOTE 3: The water added for each stage of the test should be such that a range of moisture contents is
obtained which includes the optimum moisture. In general, increments of 1 to 2 percent are
suitable for sandy and gravelly soils and of 2 to 4 percent for cohesive soils. To increase the
accuracy of the test it is often advisable to reduce the increments of water in the region of the
optimum moisture content.

Calculation

1. Bulk density, γm in g/cm3 of each compacted specimen is calculated from the


following equation.
γm = (W2-W1)/Vm------------------------ Eq.9.2
Where,
W1 = Weight in g of mould + base plate
W2 = Weight in g of mould + base plate + soil Vm =
Volume of mould i.e. 1000 cm3.
2. Dry density, γd in g/cm3 of each compacted specimen is calculated from the following
equation.
γd = 100 γm/ (100+w)----------------- Eq.9.3

119
Lecture Note on GTE-I

3. Where,
γm = Bulk density of soil in g/cm3.
w = Moisture content of soil

Graph

The dry densities γd obtained in a series of determinations is plotted against the corresponding
moisture content ‘w’. A smooth curve is then drawn through the resulting points and the position
of the maximum on this curve is determined, which is called maximum dry density (M.D.D).
And the corresponding moisture content is called optimum moisture content (O.M.C.).

Fig. 9.2: Comparison of Compaction curve for Standard and Modified Proctor Test

9.6 SATURATION (ZERO-AIR-VOIDS) LINE

A line showing the relation between water content and dry density at a constant degree of
saturation S may be established from the equation:
𝛾 =
𝐺𝛾𝑤
----------------- Eq.9.4
𝜔𝐺
𝑑 1+ 𝑆

Substituting S = 95%, 90%, and so on, one can arrive at γd-values for different values of water
content in %. The lines thus obtained on a plot of γd versus w are called 95% saturation line, 90%
saturation line and so on as shown in Fig.9.2.
If one substitutes S = 100% and plots the corresponding line, one obtains the theoretical
saturation line, relating dry density with water content for a soil containing no air voids. It is said
to be ‘theoretical’ because it can never be reached in practice as it is impossible to expel the pore
air completely by compaction.
We then use

𝛾 = 𝐺𝛾𝑤
---------------- Eq.9.5
𝜔𝐺
𝑑 1+100

120
Lecture Note on GTE-I

9.7 IN-SITU OR FIELD COMPACTION

For any type of construction job which requires soil to be used as a foundation material or as a
construction material, compaction in-situ or in the field is necessary.

The construction of a structural fill usually consists of two distinct operations—placing and
spreading in layers and then compaction. The first part assumes greater significance in major
jobs such as embankments and earth dams where the soil to be used as a construction material
has to be excavated from a suitable borrow area and transported to the work site. In this phase
large earth moving equipment such as self-propelled scrapers, bulldozers, graders and trucks are
widely employed.

The thickness of layers that can be properly compacted is known to be related to the type of soil
and method or equipment of compaction. Generally speaking, granular soils can be adequately
compacted in thicker layers than fine-grained soils and clays; also, for a given soil type, heavy
compaction equipment is capable of compacting thicker layers than light equipment. Although
the principle of compaction in the field is relatively simple, it may turn out to be a complex
process if the soil in the borrow area is not at the desired optimum moisture content for
compaction. The existing moisture content is to be determined and water added, if necessary.
Addition of water to the soil is normally done either during excavation or transport and rarely on
the construction spot; however, water must be added before excavation in the case of clayey
soils. In case the soil has more moisture content than is required for proper compaction, it has to
be air-dried after excavation and compacted as soon as the desired moisture content is attained.

Soil compaction or densification can be achieved by different means such as tamping action,
kneading action, vibration, and impact. Compactors operating on the tamping, kneading and
impact principle are effective in the case of cohesive soils, while those operating on the
kneading, tamping and vibratory principle are effective in the case of cohesionless soils.

The primary types of compaction equipment are:


1 Rollers,
2 Rammers and
3 Vibrators.
Of these, by far the most common are rollers.
Rollers are further classified as follows:
(a) Smooth-wheeled rollers,
(b) Pneumatic-tyred rollers,
(c) Sheeps -foot rollers, and
(d) Grid rollers.
Vibrators are classified as:

(a) Vibrating drum,

121
Lecture Note on GTE-I

(b) Vibrating pneumatic tyre


(c) Vibrating plate, and
(d) Vibro-flot
The maximum dry density sought to be achieved in-situ is specified usually as a certain
percentage of the value obtainable in the laboratory compaction test. Thus control of compaction
in the field requires the determination of in-situ unit weight of the compacted fill and also the
moisture content.
The methods available for the determination of in-situ unit weight are:
(a) Sand-replacement method,
(b) Core-cutter method
(c) Volumenometer method,
(d) Rubber balloon method
(e) Nuclear method,
(f) Proctor plastic needle method.
Rapid methods of determination of moisture content such as the speedy moisture tester are dopted
in this connection. Some of the above aspects are dealt with in the following subsections.

9.7.1 Types of In-situ Compaction Equipment


Certain types of in-situ compaction equipment are described below:
Rollers
(a) Smooth-wheeled rollers: This type imparts static compression to the soil. There may be two
or three large drums; if three drums are used, two large ones in the rear and one in the front is the
common pattern. The compaction pressure is relatively low because of a large contact area. This
type appears to be more suitable for compacting granular base courses and paving mixtures for
highway and airfield work rather than for compacting earth fill. The relatively smooth surface
obtained acts as a sort of a ‘seal’ at the end of a day’s work and drains off rain water very well.
The roller is self-propelled by a diesel engine and has a weight distribution that can be altered by
the addition of ballast to the rolls. The common weight is 80 kN to 100 kN (8 to 10 t), although
the range may be as much as 10 kN to 200 kN (1 to 20 t). The pressure may be of the order of
300 N (30 kg) per lineal cm of the width of rear rolls. The number of passes varies with the
desired compaction; usually eight passes may be adequate to achieve the equivalent of standard
Proctor compaction.
(b) Pneumatic-tyred rollers
This type compacts primarily by kneading action. The usual form is a box or container— mounted on
two axles to which pneumatic-tyred wheels are fitted; the front axle will have one wheel less
than the rear and the wheels are mounted in a staggered fashion so that the entire width between
the extreme wheels is covered. The weight supplied by earth ballast or other material placed in
the container may range from 120 kN (12 t) to 450 kN (45 t), although an exceptionally heavy
capacity of 2000 kN (200 t) may be occasionally used. Some equipment is provided with a
“Wobble-wheel” effect, a design in which a slightly weaving path is tracked by

122
Lecture Note on GTE-I

the travelling wheels; this facilitates the exertion of a steady pressure on uneven ground, which is
very useful in the initial stages of a fill. The weight of the roller as well as the contact pressure is
an important parameter for the performance; the latter may be varied from 0.20 to 1 N/mm2 (2 to
10 kg/cm2) through the adjustment of air pressure in the tyres. Although this type has originated
as a towed unit, self propelled units are also available. The number of passes required is similar
to that with smooth wheeled-rollers. This type is suitable for compacting most types of soil and
has particular advantages with wet cohesive materials.
(c) Sheep foot rollers: This type of roller consists of a hollow steel drum provided with
projecting studs or feet; the compaction is achieved by a combination of tamping and kneading.
The drum can be filled with water or sand to provide and control the dead weight. As rolling is
done, most of the roller weight is imposed through the projecting feet. It is generally used as a
towed assembly with the drums mounted either singly or in pairs; self-propelled units are also
available. The feet are usually club-shaped (100 × 75 mm) or tapered (57 × 57 mm), the number
on a 50 kN (5 t) roller ranging from 64 to 88. The contact pressures of the feet may range from
700 kN/m2 (7 kg/cm2) to 4200 kN/m2 (42 kg/cm2) and weight per drum from 25 kN (2.5 t) to 130
kN (13 t). Initially, the projections sink into the loose soil and compact the soil near the lowest
portion of the layer. In subsequent passes with the roller, the zone of compaction continues to
rise until the surface is reached, when the roller is said to “Walk-out”. The length of the studs,
the contact area and the weight of roller are related to the roller performance. This type of roller
is found suitable for cohesive soils. It is unsuitable for granular soils as the studs tend to loosen
these continuously. The tendency of void formation is more in soils compacted with sheeps foot
rollers.

Fig. 9.3: Schematic diagram of Sheep Foot Roller

(d) Grid rollers: This type consists of rolls made from 38 mm steel bars at 130 mm centres, with
spaces of 90 mm square. The weight of the roller ranges from 55 kN (5.5 t) to 110 kN (11 t).
This is usually a towed unit which is suitable for many types of soil including wet clays and silts.
Rammers
This type includes the dropping type and pneumatic and internal commission type, which are
also called ‘frog rammers’. They weigh up to about 1.5 kN (150 kg) and even as much as 10 kN
(1 t) occasionally. This type may be used for cohesionless soils, especially in small restricted and
confined areas such as beds of drainage trenches and back fills of bridge abutments.

123
Lecture Note on GTE-I

Vibrators
These are vibrating units of the out-of-balance weight type or the pulsating hydraulic type. Such
a type is highly effective for cohesionless soils. Behind retaining walls where the soil is
confined, the backfill, much deeper in thickness, may be effectively compacted by vibration type
of compactors.
A few of this type are dealt with below:
(a) Vibrating drum: A separate motor drives an arrangement of eccentric weights so as to cause a
high-frequency, low-amplitude, vertical oscillation to the drum. Smooth drums as well as sheep
foot type of drums may be used. Layers of the order of 1 meter deep could be compacted to high
densities.
(b) Vibrating pneumatic tyre: A separate vibrating unit is attached to the wheel axle. The ballast
box is suspended separately from the axle so that it does not vibrate. A 300 mm thick layer of
granular soil will be satisfactorily compacted after a few passes.
(c) Vibrating plate: This typically consists of a number of small plates, each of which is operated
by a separate vibrating unit. These have a limited depth of effectiveness and hence are used in
compacting granular base courses for highway and airfield pavements.
(d) Vibroflot: A method suited for compacting thick deposits of loose sandy soil is called the
‘vibro flotation’ process. The improvement of density is restricted to the surface zone in the case
of conventional compaction equipment. The vibroflotation method first compacts deep zone in
the soil and then works its way towards the surface. A cylindrical vibrator weighing about 20 kN
(2 t) and approximately 400 mm in diameter and 2 m long, called the ‘Vibroflot’, is suspended
from a crane and is jetted to the depth where compaction is to start. The jetting consists of a
water jet under pressure directed into the earth from the tip of the vibroflot; as the sand gets
displaced, the vibroflot sinks into the soil. Depths up to 12 m can be reached. After the vibroflot
is sunk to the desired depth, the vibrator is activated. The compaction of the soil occurs in the
horizontal direction up to as much as 1.5 m outward from the vibroflot. Vibration continues as
the vibroflot is slowly raised toward the surface. As this process goes on, additional sand is
continually dropped into the space around the vibroflot to fill the void created. To densify the
soil in a given site, locations at approximately 3-m spacings are chosen and treated with vibro
flotation.
9.8 CONTROL OF COMPACTION IN THE FIELD
Control of compaction in the field consists of checking the water content in relation to the
laboratory optimum moisture content and the dry unit weight achieved in-situ in relation to the
laboratory maximum dry unit weight from a standard compaction test. Typically, each layer is
tested at several random locations after it has been compacted.
Several methods are available for the determination of in-situ unit weight and moisture content
The common approaches for the determination of unit weight are the core-cutter method and
sand-replacement method. A faster method is what is known as the Proctor needle method,
which may be used for the determination of in-situ unit weight as well as in-situ moisture
content. The required density can be specified either by ‘relative compaction’ (also called

124
Lecture Note on GTE-I

‘degree of compaction’) or by the final air-void content. Relative compaction means the ratio of
the insitu dry unit weight achieved by compaction to the maximum dry unit weight obtained
from an appropriate standard compaction test in the laboratory. Usually, the relative compaction
of 90 to 100% (depending upon the maximum laboratory value), corresponding to about 5 to
10% air content, is specified and sought to be achieved. Typical values of dry unit weights
achieved may be as high as 22.5 kN/m 3 (2250 kg/m3) for well-graded gravel and may be as low
as 14.4 kN/m3 (1440 kg/m3) for clays. Approximate ranges of optimum moisture content may be
6 to 10% for sands, 8 to 12% for sand-silt mixtures, 11 to 15% for silts and 13 to 21% for clays
(as got from modified AASHO tests).
A variation of 5 to 10% is allowed in the field specification of dry unit weight at random
locations, provided the average is about the specified value.
9.8.1 Proctor Needle
The Proctor needle approach given here is an efficient and fast one for the simultaneous
determination of in-situ unit weight and in-situ moisture content, it is also called ‘penetration
needle’. The apparatus basically consists of a needle attached to a spring-loaded plunger through
a shank. An array of interchangeable needle tips is available, ranging from 6.45 to 645 mm 2, to
facilitate the measurement of a wide range of penetration resistance values. A calibration of
penetration against dry unit weight and water content is obtained by pushing the needle into
specially prepared samples for which these values are known and noting the penetration. The
penetration of the needle and the penetration resistance (load applied) may be shown on a
graduated scale on the shank and the stem of handle respectively.
The procedure for the use of the Proctor ‘plasticity’ needle, as it is called, is obvious. The spring-
loaded plunger is pressed into the compacted layer in the field with an appropriate plasticity
needle. The penetration resistance is recorded for a standard depth of penetration at a standard
time-rate of penetration. Against this penetration resistance, the corresponding values of water
content and dry unit weight are obtained from the calibration curve.
The size of the needle to be chosen depends upon the type of soil such that the resistance to be
read is neither too large nor too small. The Tennessee Valley Authority (TVA) engineers had
devised a similar device, which is called the TVA ‘Penetrometer’.
9.8.2 Special Compaction Techniques
Several special types of compaction techniques have been developed for deep compaction of in-
place soils, and these techniques are used in the field for large-scale compaction works. Among
these, the popular methods are vibroflotation, dynamic compaction, and blasting. Details of these
methods are provided in the following sections.
9.8.3 Vibro-flotation
Vibro-flotation is a technique for in situ densification of thick layers of loose granular soil
deposits. It was developed in Germany in the 1930s. The first vibroflotation device was used in
the United States about 10 years later. The process involves the use of a Vibroflot Unit.

125
Lecture Note on GTE-I

This vibrating unit has an eccentric weight inside it and can develop a centrifugal force, which
enables the vibrating unit to vibrate horizontally. There are openings at the bottom and top of the
vibrating unit for water jets.
The vibrating unit is attached to a follow-up pipe. Figure 9.4 shows the entire assembly of
equipment necessary for conducting the field compaction.

Fig.9.4: Vibro-flotation Unit


The entire vibro-flotation compaction process in the field can be divided into four stages as
described in Figure 9.5.

Fig.9.5: Various stages of Vibroflotation

126
Lecture Note on GTE-I

Stage 1: The jet at the bottom of the Vibroflot is turned on and lowered into the ground.
Stage 2: The water jet creates a quick condition in the soil and it allows the vibrating unit to sink
into the ground.
Stage 3: Granular material is poured from the top of the hole. The water from the lower jet is
transferred to the jet at the top of the vibrating unit.
This water carries the granular material down the hole.
Stage 4: The vibrating unit is gradually raised in about 0.3 m lifts and held vibrating for about 30
seconds at each lift. This process compacts the soil to the desired unit weight.
9.9 DYNAMIC COMPACTION
Dynamic compaction is a technique that has gained popularity in the United States for the
densification of granular soil deposits. This process consists primarily of dropping a heavy
weight repeatedly on the ground at regular intervals. The weight of the hammer used varies over
a range of 80 to 360 kN, and the height of the hammer drop varies between 7.5 and 30.5 m. The
stress wave generated by the hammer helps in the densification.

The degree of compaction achieved at a given site depends on the following three factors:
1. Weight of hammer
2. Height of hammer drop
3. Spacing of locations at which the hammer is dropped
Figure 9.6, shows a dynamic compaction in progress. A site immediately after the completion is
shown in Figure 9.6 b.

Fig.9.6 a Dynamic compaction in progress, b Site after Compaction

127
Lecture Note on GTE-I

Fig.9.7: Different shapes of soil after dynamic compaction

Leonards, Cutter, and Holtz (1980) suggested that the significant depth of influence for
compaction can be approximated by using the equation
𝐷=
1
ℎ------------------
√𝑊
Eq.9.6
2 𝐻
where D =significant depth of densification (m)
WH = dropping weight (metric ton)
h = height of drop (m)
In English units, the preceding equation takes the form
𝐷 = 0.61√𝑊𝐻ℎ------------------------- Eq.9.7
where the units of D and h are ft, and the unit of WH is kip.

Fig.9.8: Design chart

In 1992, Poran and Rodriguez suggested a rational method for conducting dynamic compaction
for granular soils in the field. According to their method, for a hammer of width D having a
weight WH and a drop h, the approximate shape of the densified area will be of the type shown in
Figure 9.7, a (i.e., a semiprolate spheroid).

128
Lecture Note on GTE-I

Note that in this figure b = DI (where DI is the significant depth of densification). Figure 9.8
gives the design chart for a/D and b/D versus NWHh/Ab (D is the width of the hammer if not
circular in cross section; A = area of cross section of the hammer; and N is the number of
required hammer drops). This method uses the following steps.

Step 1: Determine the required significant depth of densification, DI ( b).

Step 2: Determine the hammer weight (W H), height of drop (h), dimensions of the cross section,
and thus, the area A and the width D.

Step 3: Determine DI/D = (b/D.

Step 4: Use Figure 6.39 and determine the magnitude of NWHh/Ab for the value of b/D obtained
in step 3.

Step 5: Since the magnitudes of W H, h, A, and b are known (or assumed) from step 2, the
number of hammer drops can be estimated from the value of NWH h/Ab obtained from step 4.

Step 6: With known values of NWH h/Ab, determine a/D and thus a from Figure 9.8.

Step 7: The grid spacing, Sg, for dynamic compaction may now be assumed to be equal to or
somewhat less than a. (See Fig. 9.9)

Fig. 9.9: Approximate grid spacing for dynamic compaction

9.10 BLASTING

Blasting is a technique that has been used successfully in many projects (Mitchell, 1970) for the
densification of granular soils. The general soil grain sizes suitable for compaction by blasting
are the same as those for compaction by vibroflotation.

The process involves the detonation of explosive charges, such as 60% dynamite at a certain
depth below the ground surface in saturated soil. The lateral spacing of the charges varies from
about 3 to 9 m. Three to five successful detonations are usually necessary to achieve the desired
compaction. Compaction (up to a relative density of about 80%) up to a depth of about 18 m

129
Lecture Note on GTE-I

over a large area can easily be achieved by using this process. Usually, the explosive charges are
placed at a depth of about two-thirds of the thickness of the soil layer desired to be compacted.
The sphere of influence of compaction by a 60% dynamite charge can be given as follows
(Mitchell, 1970):

r = √
Wex

C
--------------------- Eq.9.8

where , r = sphere of influence


Wex =weight of explosive 60% dynamite
C = 0.0122 when Wex is in kg and r is in m

Example 9.1: An earth embankment is compacted at a water content of 18% to a bulk density of
19.2 kN/m3. If the specific gravity of the sand is 2.7, find the void ratio and the degree of
saturation of the compacted embankment.
Sol:

Problem 1: The soil in a borrow pit has a void ratio of 0.90. A fill-in-place volume of 20,000 m3
is to be constructed with an in-place dry density of 18.84 kN/m 3. If the owner of borrow area is
to be compensated at Rs.1.50 per cubic metre of excavation, determine the cost of compensation.

130
Lecture Note on GTE-I

10.0 CONSOLIDATION AND COMPRESSION OF SOILS


When a soil layer is subjected to vertical stress, volume change can take place through
rearrangement of soil grains, and some amount of grain fracture may also take place. The volume
of soil grains remains constant, so change in total volume is due to change in volume of water. In
saturated soils, this can happen only if water is pushed out of the voids. The movement of water
takes time and is controlled by the permeability of the soil and the locations of free draining
boundary surfaces.
It is necessary to determine both the magnitude of volume change (or the settlement) and the
time required for the volume change to occur. The magnitude of settlement is dependent on the
magnitude of applied stress, thickness of the soil layer, and the compressibility of the soil.
When soil is loaded undrained, the pore pressure increases. As the excess pore pressure
dissipates and water leaves the soil, settlement takes place. This process takes time, and the rate
of settlement decreases over time. In coarse soils (sands and gravels), volume change occurs
immediately as pore pressures are dissipated rapidly due to high permeability. In fine soils (silts
and clays), slow seepage occurs due to low permeability.
Elastic settlement is on account of change in shape at constant volume, i.e. due to vertical
compression and lateral expansion. Primary consolidation (or simply consolidation) is on
account of flow of water from the voids, and is a function of the permeability and
compressibility of soil. Secondary compression is on account of creep-like behaviour.
Primary consolidation is the major component and it can be reasonably estimated. A general
theory for consolidation, incorporating three-dimensional flow is complicated and only
applicable to a very limited range of problems in geotechnical engineering. For the vast majority
of practical settlement problems, it is sufficient to consider that both seepage and strain take
place in one direction only, as one-dimensional consolidation in the vertical direction.
10.1 Compressibility Characteristics
Soils are often subjected to uniform loading over large areas, such as from wide foundations, fills
or embankments. Under such conditions, the soil which is remote from the edges of the loaded
area undergoes vertical strain, but no horizontal strain. Thus, the settlement occurs only in one-
dimension.
The compressibility of soils under one-dimensional compression can be described from the
decrease in the volume of voids with the increase of effective stress. This relation of void ratio
and effective stress can be depicted either as an arithmetic plot or a semi-log plot.

In the arithmetic plot as shown, as the soil compresses, for the same increase of effective
stress σ', the void ratio reduces by a smaller magnitude, from ∆e1 to ∆e2. This is on account of an
increasingly denser packing of the soil particles as the pore water is forced out. In fine soils, a
much longer time is required for the pore water to escape, as compared to coarse soils.

It can be said that the compressibility of a soil decreases as the effective stress increases. This
can be represented by the slope of the void ratio – effective stress relation, which is called
the coefficient of compressibility, av.

131
Lecture Note on GTE-I

Fig.10.1: Variation of effective stress with void ratio


=−
∆𝑒----------------------

𝑎𝑣 ∆𝜎′
For a small range of effective stress, Eq.10.1

=−
𝑑𝑒________________________________

𝑎𝑣
𝑑𝜎′
or, Eq.10.2

The -ve sign is introduced to make av a positive parameter.


If e0 is the initial void ratio of the consolidating layer, another useful parameter is the coefficient
of volume compressibility, mv, which is expressed as
𝑚
= 1+𝑒
𝑎𝑣 Eq.10.3
𝑣 0
It represents the compression of the soil, per unit original thickness, due to a unit increase of
pressure.
10.2 NORMALLY CONSOLIDATED AND OVER CONSOLIDATED

If the current effective stress, 𝝈′ , is equal (note that it cannot be greater than) to the pre-
consolidation stress, then the deposit is said to be normally consolidated (NC). If the current
effective stress is less than the pre-consolidation stress, then the soil is said to be over-
consolidated (OC).
The figure shows the relation of void ratio and effective stress of a clay soil as a semi-log plot.

132
Lecture Note on GTE-I

Fig.10.2: Variation of effective in logarithmic scale

133
Lecture Note on GTE-I

OP corresponds to initial loading of the soil. PQ corresponds to unloading of the


soil. QFR corresponds to a reloading of the soil. Upon reloading beyond P, the soil continues

The pre consolidation stress, 𝝈𝒑′ , is defined to be the maximum effective stress experienced by
along the path that it would have followed if loaded from O to R continuously.

𝒄
the soil. This stress is identified in comparison with the effective stress in its present state. For
soil at state Q or F, this would correspond to the effective stress at point P.
It may be seen that for the same increase in effective stress, the change in void ratio is much less
for an over-consolidated soil (from e0 to ef), than it would have been for a normally consolidated
soil as in path OP. In unloading, the soil swells but the increase in volume is much less than the
initial decrease in volume for the same stress difference.
The distance from the normal consolidation line has an important influence on soil behaviour.
This is described numerically by the over-consolidation ratio (OCR), which is defined as the
ratio of the pre-consolidation stress to the current effective stress.
𝜎′
𝑂𝐶𝑅
𝑝𝑐

=
𝜎′
--------------------- Eq.10.4

Note that when the soil is normally consolidated, OCR = 1


Settlements will generally be much smaller for structures built on overconsolidated soils. Most
soils are over-consolidated to some degree. This can be due to shrinking and swelling of the soil
on drying and rewetting, changes in ground water levels, and unloading due to erosion of
overlying strata.
For NC clays, the plot of void ratio versus log of effective stress can be approximated to a
straight line, and the slope of this line is indicated by a parameter termed as compression
index, Cc.

=

𝜎
---------------------- Eq.10.5
𝑐 ′2
∆𝑒

𝜎1
𝑙𝑜𝑔1 ′
0

Estimation of Pre-Consolidation Stress

134
Lecture Note on GTE-I

Fig.10.3: Graphical method to determine pre consolidation stress

135
Lecture Note on GTE-I

It is possible to determine the pre-consolidation stress that the soil had experienced. The soil
sample is to be loaded in the laboratory so as to obtain the void ratio - effective stress
relationship. Empirical procedures are used to estimate the pre-consolidation stress, the most
widely used being Casagrande's construction which is illustrated as
The steps in the construction are:

• Draw the graph using an appropriate scale.


• Determine the point of maximum curvature A.
• At A, draw a tangent AB to the curve.
• At A, draw a horizontal line AC.
• Draw the extension ED of the straight line portion of the curve.
• Where the line ED cuts the bisector AF of angle CAB, that point corresponds to the pre-
consolidation stress.

The total stress increases when additional vertical load is first applied. Instantaneously, the pore
water pressure increases by exactly the same amount. Subsequently there will be flow from
regions of higher excess pore pressure to regions of lower excess pore pressure causing
dissipation. The effective stress will change and the soil will consolidate with time. This is
shown schematically.

Fig.10.4: Changes of effective stress, pore water pressure and consolidation

136
Lecture Note on GTE-I

10.3 SPRING ANALOGY TO EXPLAIN CONSOLIDATION THEORY


A mechanistic model for the phenomenon of consolidation was given by Taylor (1948), by
which the process can be better understood. This model, with slight modifications, is presented
in Fig. 10.5 and is explained below:
A spring of initial height Hi is surrounded by water in a cylinder. The spring is analogous to the
soil skeleton and the water to the pore water. The cylinder is fitted with a piston of area A
through which a certain load may be transmitted to the system representing a saturated soil. The
piston, in turn, is fitted with a vent, and a valve by which the vent may be opened or closed.
Referring to Fig.10.5 (a), let a load P be applied on the piston. Let us assume that the valve of
the vent is open and no flow is occurring. This indicates that the system is in equilibrium under
the total stress P/A which is fully borne by the spring, the pressure in the water being zero.

Fig.10.5: Spring Model for Consolidation of Soil mass


Referring to Fig.10.5 (b), let us apply an increment of load δP to the piston, the valve being kept
closed. Since no water is allowed to flow out, the piston cannot move downwards and compress
the spring; therefore, the spring carries the earlier stress of P/A, while the water is forced to carry
the additional stress of δP/A imposed on the system, the sum counteracting the total stress
imposed. This additional stress δP/A in the water in known as the hydrostatic excess pressure.
Referring to Fig.10.5 (c), let us open the valve and start reckoning time from that instant. Water
just starts to flow under the pressure gradient between it and the atmosphere seeking to return to
its equilibrium or atmospheric pressure. The excess pore pressure begins to diminish, the spring
starts getting compressed as the piston descends consequent to expulsion of pore water. It is just

137
Lecture Note on GTE-I

the beginning of transient flow, simulating the phenomenon of consolidation. The openness of
the valve is analogous to the permeability of soil.
Referring to Fig.10.5 (d), flow has occurred to the extent of dissipating 50% of the excess pore
pressure. The pore water pressure at this instant is half the initial value, i.e., 1/2(δP/A). This
causes a corresponding increase in the stress in the spring of ½ (δP/A), the total stress remaining
constant at [(P/A) + (δP/A)]. This stage refers to that of “50% consolidation”.
Referring to Fig.10.5 (e), the final equilibrium condition is reached when the transient flow
situation ceases to exist, consequent to the complete dissipation of the pore water pressure. The
spring compresses to a final height Hf < Hi, carrying the total stress of (P + δP)/A, all by itself,
since the excess pore water pressure has been reduced to zero, the pressure in it having equalled
the atmospheric. The system has reached the equilibrium condition under the load (P + δP). This
represents “100% consolidation” under the applied load or stress increment. We may say that the
“soil” has been consolidated to an effective stress of (P + δP)/A.
In this mechanistic model, the compressible soil skeleton is characterised by the spring and the
pore water by the water in the cylinder. The more compressible the soil, the longer the time
required for consolidation; the more permeable the soil, the shorter the time required. There is
one important aspect in which this analogy fails to simulate consolidation of a soil. It is that the
pressure conditions are the same throughout the height of the cylinder, whereas the consolidation
of a soil begins near the drainage surfaces and gradually progresses inward. In may be noted that
soil consolidates only when effective stress increases; that is to say, the volume change
behaviour of a soil is a function of the effective stress and not the total stress.
Similar arguments may be applied to the expansion characteristics under the decrease of load.
An alternative mechanical analogy to the consolidation process is shown in Fig. 10.6. A cylinder
is fitted with a number of pistons connected by springs to one another. Each of the compartments
thus formed is connected to the atmosphere with the aid of standpipes. The cylinder is full of
water and is considered to be airtight. The pistons are provided with perforations through which
water can move from one compartment to another. The topmost piston is fitted with valves
which may open or close to the atmosphere. It is assumed that any pressure applied to the top
piston gets transmitted undiminished to the water and springs.
Initially, the cylinder is full of water and weights of the pistons are balanced by the springs; the
water is at atmospheric pressure and the valves may be open. The water level stands at the
elevation PP in the standpipes as shown. The valves are now closed, the water level continuing
to remain at PP. An increment of pressure∆𝜎 is applied on the top piston. It will be observed
that the water level rises instantaneously in all the stand pipes to an elevation
QQ, above PP by a height h = Δσ/γw. Let all the valves be opened simultaneously with the
application of the pressure increment, the time being reckoned from that instant. The height of
the springs remains unchanged at that instant and the applied increment of pressure is fully taken
up by water as the hydrostatic excess pressure over and above the atmospheric. An equal rise of
water in all the standpipes indicates that the hydrostatic excess pressure is the same in all
compartments immediately after application of pressure. As time elapses, the water level in the

138
Lecture Note on GTE-I

pipes starts falling, the pistons move downwards gradually and water comes out through the
open valves.

Fig.10.6: Mechanical Analogy to Consolidation process


At any time t = t1, the water pressure in the first compartment is least and that in the last or the
bottommost is highest, as indicated by the water levels in the standpipes. The variation of
hydrostatic excess pressure at various points in the depth of the cylinder, as shown by the dotted
lines, varies with time. Ultimately, the hydrostatic excess pressure reduces to zero in all
compartments, the water levels in the standpipes reaching elevation PP; this theoretically
speaking, is supposed to happen after the lapse of infinite time. As the hydrostatic excess
pressure decreases in each compartment, the springs in each compartment experience a
corresponding pressure and get compressed. For example, at time t = t1, the hydrostatic excess
pressure in the first compartment is given by the head PJ; the pressure taken by the springs is
indicated by the head JQ, the sum of the two at all times being equivalent to the applied pressure
increment; that is to say, it is analogous to the effective stress principle: σ = σ + u, the pressure
transferred to the springs being analogous to intergranular or effective stress in a saturated soil,
and the hydrostatic excess pressure to the neutral pressure or excess pore water pressure.
Since water is permitted to escape only at one end, it is similar to the case of a single drainage
face for a consolidating clay sample. The distribution of hydrostatic excess pressure will be
symmetrical about mid-depth for the situation of a double drainage face, the maximum occurring
at mid-depth and the minimum or zero values occurring at the drainage faces.

10.4 TERZAGHI’S THEORY OF ONE-DIMENSIONAL CONSOLIDATION


Terzaghi (1925) advanced his theory of one-dimensional consolidation based upon the following
assumptions, the mathematical implications being given in parentheses:
1. The soil is homogeneous (kz is independent of z).
2. The soil is completely saturated (S = 100%).

139
Lecture Note on GTE-I

3. The soil grains and water are virtually incompressible (γw is constant and volume change of
soil is only due to change in void ratio).
4. The behaviour of infinitesimal masses in regard to expulsion of pore water and consequent
consolidation is no different from that of larger representative masses (Principles of calculus may
be applied).
5. The compression is one-dimensional (u varies with z only).
6. The flow of water in the soil voids is one-dimensional, Darcy’s law being valid.
𝛛𝑣𝑦
= = 0 and 𝑣 = 𝑘
𝛛𝑣𝑥 𝛛ℎ
--------------- Eq.10.6
𝛛 𝛛 𝑧 𝑧 𝛛𝑧
𝑥 𝑦

Also, flow occurs on account of hydrostatic excess pressure (h = u/γw).

7. Certain soil properties such as permeability and modulus of volume change are constant; these
actually vary somewhat with pressure. (k and mv are independent of pressure).
8. The pressure versus void ratio relationship is taken to be the idealised one, av is constant).
9. Hydrodynamic lag alone is considered and plastic lag is ignored, although it is known to exist.
(The effect of k alone is considered on the rate of expulsion of pore water).
The first three assumptions represent conditions that do not vary significantly from actual
conditions.
The fourth assumption is purely of academic interest and is stated because the differential
equations used in the derivation treat only infinitesimal distances. It has no significance for the
laboratory soil sample or for the field soil deposit.
The fifth assumption is certainly valid for deeper strata in the field owing to lateral confinement
and is also reasonably valid for an oedometer sample.
The sixth assumption regarding flow of pore water being one-dimensional may be taken to be
valid for the laboratory sample, while its applicability to a field situation should be checked.
However, the validity of Darcy’s law for flow of pore water is unquestionable.
The seventh assumption may introduce certain errors in view of the fact that certain soil
properties which enter into the theory vary somewhat with pressure but the errors are considered
to be of minor importance.
The eighth and ninth assumptions lead to the limited validity of the theory. The only justification
for the use of the eighth assumption is that, otherwise, the analysis becomes unduly complex.
The ninth assumption is necessitated because it is not possible to take the plastic lag into account
in this theory. These two assumptions also may be considered to introduce some errors.
Now let us see the derivation of Terzaghi’s theory with respect to the laboratory oedometer
sample with double drainage as shown in Fig. 10.7.

140
Lecture Note on GTE-I

Fig.10.7: Consolidation of clay soil sample with double drainage


Let us consider a layer of unit area of cross-section and of elementary thickness dz at depth z
from the pervious boundary. Let the increment of pressure applied be Δσ. immediately on
application of the pressure increment, pore water starts to flow towards the drainage faces. Let
∂h be the head lost between the two faces of this elementary layer, corresponding to a decrease
of hydrostatic excess pressure ∂u.

𝑘 +𝑘 [𝑒 + 𝑆 ]------------------
2
Equation 2,𝛛for flow of𝛛2water
ℎ ℎ 1 𝛛𝑠 𝛛𝑒
through soil, holds here also,
𝑥 𝛛𝑥2 𝑧
= Eq.10.7
1+ 𝛛 𝛛𝑡
𝛛𝑧2 𝑒 𝑡

𝑘 [𝑒 + 𝑆 ]------------------
2
For 𝛛one-dimensional
ℎ 1 𝛛𝑠
flow𝛛𝑒situation, this reduces to:
𝑧
= Eq.10.8
1+ 𝛛 𝛛𝑡
𝛛𝑧2 𝑒 𝑡
During the process of consolidation, the degree of saturation is taken to remain constant at 100%,
while void ratio changes causing reduction in volume and dissipation of excess hydrostatic

S=0 or unity and 𝛛𝑠 = 0


pressure through expulsion of pore water; that is,

𝛛𝑡1
Hence 𝑘 [ ] = [ ]------------------
𝛛2 ℎ 𝛛𝑒 𝛛 𝑒
𝑧 𝛛𝑧2
=− Eq.10.9
1+ 𝛛 𝛛𝑡 1+𝑒
𝑒 𝑡
Negative sign denoting decrease of e for increase of h.

grains are virtually incompressible, 𝛛 [ ] represents time-rate of volume change per unit
𝑒
Since volume decrease can be due to a decrease in the void ratio only as the pore water and soil

𝛛 1+𝑒
volume 𝑡
The flow is only due to the hydrostatic excess pressure, ℎ =
𝑢

𝛾𝑤
=−
2
𝑘 𝛛 𝑢 𝛛𝑉
𝛾𝑤
So -------------- Eq.10.10
𝛛𝑡
𝛛𝑧2
Here k is the permeability of soil in the direction of flow, and ∂V represents the change in
volume per unit volume. The change in hydrostatic excess pressure, ∂u, changes the intergranular
or effective stress by the same magnitude, the total stress remaining constant.
The change in volume per unit volume, ∂V, may be written, as per the definition of the modulus
141
Lecture Note on GTE-I

𝜕𝑉 = 𝑚𝑣𝜕𝜎 = −𝑚𝑣𝜕𝑢----------------
of volume change, mv
Eq.10.11

142
Lecture Note on GTE-I

Negative sign is used, since increase in stress reduces pore water pressure.

=
Differentiating both sides with respect to time,
𝛛𝑉
𝛛𝑢
−𝑚 � 𝛛𝑡
-------------------- Eq.10.12
𝛛𝑡 �

=
𝑘 𝛛2 𝑢
From Eq. 10.11 and 10.12, we get
𝛛𝑢
---------------- Eq.10.13
𝛛𝑡 𝛾𝑤𝑚𝑣 𝛛𝑧2

= 𝑐
2𝛛 𝑢
This is written as:
𝛛𝑢
𝑣 𝛛𝑧2
---------------- Eq.10.14
𝛛𝑡

= 𝛾𝑤𝑚𝑣
𝑘
𝑐𝑣
Where

cv is known as the “Coefficient of consolidation”. u represents the hydrostatic excess pressure at


a depth z from the drainage face at time t from the start of the process of consolidation.
The coefficient of consolidation may also be written in terms of the coefficient of compressibility
𝑐 =
𝑘(1+𝑒0)

= 𝛾𝑤𝑚
𝑘 ------------- Eq.10.15
𝑣 𝑎𝑣𝛾𝑤
𝑣
Equation 10.14 is the basic differential equation of consolidation according to Terzaghi’s theory
of one-dimensional consolidation. The coefficient of consolidation combines the effect of
permeability and compressibility characteristics on volume change during consolidation. Its units
can be shown to be mm2/s or L2 T-1.
The initial hydrostatic excess pressure, ui, is equal to the increment of pressure Δσ, and is the
same throughout the depth of the sample, immediately on application of the pressure, and is
shown by the heavy line in Fig. 10.7 (b). The horizontal portion of the heavy line indicates the
fact that, at the drainage face, the hydrostatic excess pressure instantly reduces to zero,
theoretically speaking. Further, the hydrostatic excess pressure would get fully dissipated
throughout the depth of the sample only after the lapse of infinite time*, as indicated by the
heavy vertical line on the left of the figure. At any other instant of time, the hydrostatic excess
pressure will be maximum at the farthest point in the depth from the drainage faces, that is, at the
middle and it is zero at the top and bottom. The distribution of the hydrostatic excess pressure
with depth is sinusoidal at other instants of time, as shown by dotted lines. These curves are
called “Isochrones”.
10.4.1 Alternative Method:

=
1 𝛛𝑢
With reference to Fig.10.7, the hydraulic gradient i1 at depth z
𝛛ℎ
--------------- Eq.10.16
𝛛𝑧 𝛾𝑤 𝛛𝑧
The hydraulic gradient i at depth z+𝜕𝑧 = ( +2𝛛 𝑑𝑧)-----------
1 𝛛𝑢 𝑢
Eq.10.17
𝛾𝑤 𝛛𝑧 𝛛𝑧2
2

Rate of inflow per unit area = Velocity at depth z = k.i1, by Darcy’s law.

Water lost per unit time = k(i – i ) = 𝑘 (𝛛 𝑑𝑧)-----------------


Rate of outflow per unit area = velocity at2 (z + dz) = k.i2
𝑢
Eq.10.18
𝛾𝑤 𝛛𝑧2
2 1

143
Lecture Note on GTE-I

This should be the same as the time-rate of volume decrease.


Volumetric strain = mv.Δσ = – mv∂(σ – u), by definition of the modulus of volume change, mv

144
Lecture Note on GTE-I

∴ Change of volume = – mv∂(σ – u).dz, since the elementary layer of thickness dz and unit cross-
(The negative sign denotes decrease in volume with increase in pressure).

sectional area in considered.


𝛛
(𝜎 − 𝑢)𝑑𝑧-----------------
𝑚
� 𝛛𝑡
Time-rate of change of volume = − Eq.10.19

But 𝛛𝜎= 0, since σ is constant.


𝛛𝑡

∴ Time-rate of change of 𝛛𝑢
� 𝛛𝑡
𝑑𝑧----------------- Eq.10.20
volume=+𝑚 �

𝑑𝑧---------------
𝑘 𝛛2𝑢 𝛛𝑢
Equating this to water lost per unit time
𝑣 𝛛𝑡
dz=-𝑚 Eq.10.21
𝛾𝑤𝑚𝑣 𝛛𝑧2
= 𝑐
𝛛𝑢 𝛛 𝑢
2

𝑣 𝛛𝑧2
---------------------- Eq.10.22
𝛛𝑡

= 𝛾𝑤𝑚
𝑘
𝑐𝑣
Where ---------------------------------------- Eq.10.23
𝑣
10.5 SOLUTION OF TERZAGHI’S EQUATION FOR ONE-DIMENSIONAL
CONSOLIDATION
Terzaghi solved the differential equation for a set of boundary conditions which have utility in
solving numerous engineering problems and presented the results in graphical form using
dimensional parameters.
The following are the boundary conditions:
1. There is drainage at the top of the sample: At z = 0, u = 0, for all t.
2. There is drainage at the bottom of the sample: At z = 2H, u = 0, for all t.
3. The initial hydrostatic excess pressure ui is equal to the pressure increment, Δσ u = ui = Δσ, at
t = 0.
Terzaghi chose to consider this situation where u = ui initially throughout the depth, although
solutions are possible when ui varies with depth in any specified manner. The thickness of the
sample is designated by 2H, the distance H thus being the length of the longest drainage path,
i.e., maximum distance water has to travel to reach a drainage face because of the existence of
two drainage faces. (In the case of only one drainage face, this will be equal to the total thickness
of the clay layer).
The general solution for the above set of boundary conditions has been obtained on the basis of

𝑢 = 𝑓(𝑧, 𝑛𝜋𝑧 𝑡) = ∑∞
1 2𝐻
[ 𝑢 𝑠𝑖𝑛
𝑛𝜋𝑧
𝑑𝑧] (𝑠𝑖𝑛 )𝑒
separation of variables and Fourier series expansion and is as follows:
−𝑛 𝜋 𝐶𝑣𝑡/4𝐻
2 2 2


-- Eq.10.24
𝑛=1
𝐻 0 2 2𝐻
𝐻
𝑖
This solution enables the hydrostatic excess u to be computed for a soil mass under any initial
system of stress ui, at any depth z, and at any time t.

𝑢 = ∑∞ [ 𝑖 (1 − 𝑐𝑜𝑠𝑛𝜋) (𝑠𝑖𝑛 ) 𝑒−𝑛 𝜋 𝑐𝑣𝑡/4𝐻 ]-------------------


2𝑢 𝑛𝜋𝑧
In particular, if ui is considered constant with respect to depth, this equation reduces to
2 2 2

𝑛=1
Eq.10.25
145
Lecture Note on GTE-I

𝑛𝜋 2𝐻
When n is even, (1 – cosnπ) vanishes; when n is odd, this factor becomes 2. Therefore it is

𝑢 = ∑∞ [ 4𝑢𝑖 (𝑠𝑖𝑛2𝑚+1) 𝑒 ]--------------


convenient to replace n by (2m + 1), m being an integer. Thus, we have
𝑚=1 (2𝑚+1)
Eq.10.26
2𝐻
−(2𝑚+1) 𝜋 𝑐𝑣𝑡/4𝐻2
2 2
𝜋

146
Lecture Note on GTE-I

Three-dimensionless parameters are introduced for convenience in presenting the results in a


form usable in practice. The first is z/H, relating to the location of the point at which
consolidation is considered, H being the maximum length of the drainage path. The second is the
consolidation ratio, Uz, to indicate the extent of dissipation of the hydrostatic excess pressure in

𝑈 = 𝑖 = (1 − )-------------------
relation𝑢to−𝑢the initial value:
𝑢
Eq.10.27
𝑢𝑖 𝑢 𝑖
𝑧
The subscript z is significant, since the extent of dissipation of excess pore water pressure is
different for different locations, except at the beginning and the end of the consolidation process.
The third dimensionless parameter, relating to time, and called ‘Time-factor’, T, is defined as

𝑇 =
𝑐𝑣𝑡
follows:
-------------------- Eq.10.28
𝐻2
Where cv is the coefficient of consolidation,
H is the length drainage path, and t is the elapsed time from the start of consolidation process.
In the context of consolidation process at a particular site, cv and H are constants, and the time

𝑢 = ∑ (𝑠𝑖𝑛 ) 𝑒−𝑀 𝑇-------------------------------------


2𝑢𝑖 𝑀𝑍
factor is directly proportional to time. Introducing the time factor we have
2

𝑀 𝐻
Eq.10.29

𝑈𝑍 =−𝑀 12− = 1 − ∑∞ (𝑠𝑖𝑛


𝑢 2
Introducing the consolidation ratio, Uz, we have:
)𝑒
𝑀 𝑇
---------------- Eq.10.30
𝑚=0 𝑀 𝐻
𝑍 𝑢𝑖
The following approximate expressions have been found to yield values for T with good degree
of precision:
When U < 60%, T = (π/4) U2----------------------- Eq.10.31
When U > 60%, T = – 0.9332 log10 (1 – U) – 0.0851---------- Eq.10.32

10.6 LABORATORY CONSOLIDATION TESTS

The one-dimensional consolidation testing procedure was first suggested by Terzaghi. This test
is performed in a consolidometer (sometimes referred to as an oedometer). The schematic
diagram of a consolidometer is shown in Figure 10.8(a). Figure 10.8(b) shows a photograph of a
consolidometer. The soil specimen is placed inside a metal ring with two porous stones, one at
the top of the specimen and another at the bottom. The specimens are usually 64 mm in diameter
and 25 mm thick. The load on the specimen is applied through a lever arm, and compression is
measured by a micrometer dial gauge. The specimen is kept under water during the test. Each
load usually is kept for 24 hours. For consolidation testing, it is generally desirable that the
applied pressure at any loading stage be double than that at the preceding stage. The test may,
therefore, be continued using a loading sequence which would successively apply stress of 0.1,
0.2, 0.4, 0.8, 1.6, 3.2, etc, kgf/cm 2 on the soil specimen. For each loading increment, after
application of load, readings of the dial gauge shall be taken using a time sequence such as 0,
0.25, 1, 2.25, 4,6·25,9, 12·25, 16, 20.25, 25, 36, 49, 64, 81, 100, 121, 144, 169, 196, 225 min etc,
up to 24. These time sequences facilitate plotting of thickness or change of thickness of specimen
147
Lecture Note on GTE-I

against square root of time or against logarithm of time. The loading increment shall be left at
least until the slope of the characteristic linear secondary compression portion of thickness
versus log time plot is apparent or until the end of pre-consolidation is Indicated on a square roof
of time plot. A period of 24 h will usually be sufficient, but longer times may be required. In
every case, the same load increment duration shall be used for all load increments during a
consolidation test.

On completion of the final loading stage, the specimen shall be unloaded by pressure decrements
which decrease the load to one-fourth of the last load. Dial gauge readings may be taken as
necessary during each stage of unloading. If desired, the time intervals used during the
consolidation increments may be adopted; usually it is possible to proceed much more rapidly.
After the completion of test, dry weight of the soil sample is taken.

Figure 1(c) shows a consolidation test in progress. The general shape of the plot of deformation
of the specimen against time for a given load increment is shown in Figure 10.9. From the plot,
we can observe three distinct stages, which may be described as follows:
Stage I: Initial compression, which is caused mostly by preloading
Stage II: Primary consolidation, during which excess pore water pressure gradually is transferred
into effective stress because of the expulsion of pore water

Stage III: Secondary consolidation, which occurs after complete dissipation of the excess pore
water pressure, when some deformation of the specimen takes place because of the plastic
readjustment of soil fabric

Fig. 10.8 (a): Schematic diagram of a consolidometer

148
Lecture Note on GTE-I

Fig. 10.8 (b): Photograph of a consolidometer Fig. 10.8 (c): Consolidation test in progress

Fig. 10.9: Shape of the plot of deformation of the specimen

10.6.1 Void Ratio–Pressure Plots

149
Lecture Note on GTE-I

After the time–deformation plots for various loadings are obtained in the laboratory, it is
necessary to study the change in the void ratio of the specimen with pressure. Following is a
step-by-step procedure for doing so:

Fig. 10.10: Change of height of specimen in one-dimensional consolidation test


Step 1: Calculate the height of solids, Hs, in the soil specimen (Figure 10.10) using the equation

= 𝐴𝐺𝑆 = 𝐴𝐺𝑆
𝑊𝑆 𝑀𝑆 --------- Eq.10.33
𝑆
𝛾𝑤 𝜌𝑤

where Ws _ dry weight of the specimen


Ms = dry mass of the specimen
A = area of the specimen

𝛾𝑤 = unit weight of water


Gs = specific gravity of soil solids

𝜌𝑤 =density of water

𝐻𝑉 = 𝐻 − 𝐻𝑆----------------------
Step 2: Calculate the initial height of voids as
Eq.10.34
Where, HV = initial height of the specimen.

𝑒 = = =
Step 3:𝑉Calculate
𝐻𝑉 𝐴 𝐻𝑉
the initial void ratio, of the specimen, using the equation
𝑉
---------------------- Eq.10.35
𝐻𝑆 𝐻𝑆
0
𝑉𝑆 𝐴

deformation ∆H1, calculate the change in the void ratio as


Step 4: For the first incremental loading, s1 (total load/unit area of specimen), which causes a

∆𝑒 =
∆𝐻1
---------------- Eq.10.36
𝐻𝑆
1

∆H1 is obtained from the initial and the final dial readings for the loading).
It is important to note that, at the end of consolidation, total stress 𝜎1 is equal to effective stress

𝑒1 = 𝑒0 − ∆𝑒1------------------
Step 5: Calculate the new void ratio after consolidation caused by the pressure increment as

For the next loading, 𝜎2 (note: 𝜎2 equals the cumulative load per unit area of specimen), which
Eq.10.37

causes additional deformation ∆H2, the void ratio at the end of consolidation can be calculated
as
𝑒 = −
∆𝐻2

2 𝑒1
----------------- Eq.10.38
𝐻2

150
Lecture Note on GTE-I

At this time, 𝜎2′ effective stress, proceeding in a similar manner, one can obtain the void ratios at
the end of the consolidation for all load increments.

151
Lecture Note on GTE-I

The effective stress 𝜎′ and the corresponding void ratios (e) at the end of consolidation are
plotted on semi logarithmic graph paper. The typical shape of such a plot is shown in Figure
10.11

Fig. 10.11: Typical plot of e against log 𝜎′

10.6.2 Evaluation of coefficient of consolidation from Oedometer test data


The coefficient of consolidation, Cv, in any stress range of interest, may be evaluated from its
definition, however it can be evaluated by experimentally determining the parameters k, av and e0
for the stress range under consideration. k may be got from a permeability test conducted on the
oedometer sample itself, after complete consolidation under the particular stress increment av
and e0 may be obtained from the oedometer test data, by plotting the e – σ curve. However,
consolidation equation is rarely used for the determination of cv. Instead, cv is evaluated from the
consolidation test data by the use of characteristics of the theoretical relationship between the
time factor T, and the degree of consolidation, U. These methods are known as ‘fitting methods’,
as one tries to fit in the characteristics of the theoretical curve with the experimental or
laboratory curve.
The more generally used fitting methods are the following:
(a) The square root of time fitting method
(b) The logarithm of time fitting method
These two methods will be presented in the following sub-sections.

Fig. 10.12: Time versus reduction in sample thickness for a load-increment

152
Lecture Note on GTE-I

10.6.3 The Square Root of Time Fitting Method


This method has been devised by D.W. Taylor (1948). The coefficient of consolidation is the soil
property that controls the time-rate or speed of consolidation under a load-increment. The
relation between the sample thickness and elapsed time since the application of the loading
increment is obtainable from an oedometer test and is somewhat as shown in Figure 10.13 for a
typical load-increment. This figure depicts change in sample thickness with time essentially due
to consolidation only the elastic compression which occurs almost instantaneously on application
of load increment is shown. The effect of prolonged compression that occurs after 100%
dissipation of excess pore pressure is not shown or is ignored; this effect is known as ‘Secondary
consolidation which is briefly presented in the following section. The curves of Figs.10.13 and
the theoretical curves bear striking similarity; in fact, one should expect it if Terzaghi’s theory is
to be valid for the phenomenon of consolidation. This similarity becomes more apparent if the
curves are plotted with square root of time/time factor as the function, as shown in Fig. 10.13 (a)
and (b). The theoretical curve on the square root plot is straight line up to about 60%
consolidation with a gentle concave upwards curve thereafter. If another straight line, shown
dotted, is drawn such that the abscissae of this line are 1.15 times those of the straight line
portion of the theoretical curve, it can be shown to cut the theoretical curve at 90% consolidation.
This may be established from the values of T at various values of U that is, the value of T at 90%
consolidation is 1.15 times the abscissa of an extension of the straight line portion of the U
versus T relation. This property is used for ‘fitting’ the theoretical curve to the laboratory curve.
The laboratory curve shows a sudden initial compression, called ‘elastic compression’ which
may be partly due to compression of gas in the pores. The corrected zero point at zero time is
obtained by extending the straight line portion of the laboratory plot backward to meet the axis
showing the sample thickness/dial gauge reading. The so-called ‘primary compression’ or
‘primary consolidation’ is reckoned from this corrected zero. A dashed line is constructed from
the corrected zero such that its abscissae are 1.15 times those of the straight line portion of the
laboratory plot. The intersection of the dashed line with the laboratory plot identifies the point
representing 90% consolidation in the sample. The time corresponding to this can be read off
from the laboratory plot. The point corresponding to 100% primary consolidation may be easily
extrapolated on this plot.

The coefficient of consolidation, 𝑐𝑣, may be obtained from


𝑐 = 𝑇90𝐻2

𝑡90
---------------------------------- Eq.10.39
𝑣
where t90 is read off from Fig. 10.7 (a)
T90 is 0.848 from Terzaghi’s theory
H is the drainage path, which may be taken as half the thickness of the sample for double
drainage conditions,

153
Lecture Note on GTE-I

Fig. 10.7: Square root of time fitting method (After Taylor, 1948)

10.6.4 The Logarithm of Time Fitting Method


This method was devised by A. Casagrande and R.E. Fadum (1939). The point corresponding to
100 per cent consolidation curve is plotted on a semi-logarithmic scale, with time factor on a
logarithmic scale and degree of consolidation on arithmetic scale, the intersection of the tangent
and asymptote is at the ordinate of 100% consolidation. A comparison of the theoretical and
laboratory plots in this regard is shown in Figs. 10.8 (a) and (b).

Fig. 10.8(a): Sample thickness/Dial gauge reading versus logarithm of time (Laboratory curve)

154
Lecture Note on GTE-I

Fig. 10.8 (b) Logarithm of time fitting method (After A. Casagrande, 1939)
Since the early portion of the curve is known to approximate a parabola, the corrected zero point
may be located as follows: The difference in ordinates between two points with times in the ratio
of 4 to 1 is marked off; then a distance equal to this difference may be stepped off above the
upper points to obtain the corrected zero point. This point may be checked by more trials, with
different pairs of points on the curve. After the zero and 100% primary compression points are
located, the point corresponding to 50% consolidation and its time may easily be obtained and
the coefficient of consolidation computed from:
𝑐 =
𝑇50𝐻2

𝑡50
----------------- Eq.10.40
𝑣
Where t50 is read off from Fig. 10.8 (a)
T50 = 0.197 from Terzaghi’s theory, and H is the drainage path as stated in the previous
subsection.

10.7 SECONDARY CONSOLIDATION


The time-settlement curve for a cohesive soil has three distinct parts as illustrated in Figure 10.9.
When the hydrostatic excess pressure is fully dissipated, no more consolidation should be
expected. However, in practice, the decrease in void ratio continues, though very slowly, for a
long time after this stage, called ‘Primary Consolidation’. The effect or the phenomenon of
continued consolidation after the complete dissipation of excess pore water pressure is termed
‘Secondary Consolidation’ and the resulting compression is called ‘Secondary Compression’.
During this stage, plastic readjustment of clay platelets takes place and other effects as well as
colloidal-chemical processes and surface phenomena such as induced electrokinetic potentials
occur. These are, by their very nature, very slow.

155
Lecture Note on GTE-I

Fig.10.9: Time settlement curve for cohesive soil


Secondary consolidation is believed to come into play even in the range of primary
consolidation, although its magnitude is small, because of the existence of a plastic lag right
from the beginning of loading. However, it is almost impossible to separate this component from
the primary compression. Since dissipation of excess pore pressure is not the criterion here,
Terzaghi’s theory is inapplicable to secondary consolidation. The fact that experimental time
compression curves are in agreement with Terzaghi’s theoretical curve only up to about 60%
consolidation is, in itself, an indication of the manifestation of secondary consolidation even
during the stage of primary consolidation.
Secondary consolidation of mineral soils is usually negligible but it may be considerable in the
case of organic soils due to their colloidal nature. This may constitute a substantial part of total
compression in the case of organic soils, micaceous soils, loosely deposited clays, etc. A possible
disintegration of clay particles is also mentioned as one of the reasons for this phenomenon.
Secondary compression is usually assumed to be proportional to the logarithm of time. Hence,
the secondary compression can be identified on a plot of void ratio versus logarithm of time (Fig.
10.10).
Secondary compression appears as a straight line sloping downward or, in some cases, as a
straight line followed by a second straight line with a flatter slope. The void ratio, ef, at the end
of primary consolidation can be found from the intersection of the backward extension of the
secondary line with a tangent drawn to the curve of primary compression, as shown in the figure.
The rate of secondary compression depends upon the increment of stress and the characteristics
of the soil.
The equation for the rate of secondary compression may be approximated as follows:
Δe = – α .𝑙𝑜𝑔10(t2/t1) --------------------- Eq.10.41
Here, t1 is the time required for the primary compression to be virtually complete, t2 any later
time, and is Δe is the corresponding change in void ratio. This means that the secondary
compression which occurs during the hydrodynamic phase is ignored, but the error is not
probably serious. α is a coefficient expressing the rate of secondary compression

156
Lecture Note on GTE-I

Fig.10.10: Voids ratio versus logarithmic of time


.

secondary compression’, 𝐶𝛼, in terms of strain or percentage of settlement as follows:


Another way of expressing the time–rate of secondary compression is through the ‘coefficient of

𝑒 = 𝑙𝑜𝑔1 ( )----------------------
𝑡2

−𝐶𝛼
Eq.10.42
𝑡1
0

In other words, 𝐶𝛼 may be taken to be the slope of the straight line representing the secondary

The relation between α and 𝐶𝛼 is


compression on a plot of strain versus logarithm of time.

𝐶
= 1+
𝛼 --------------------- Eq.10.43
𝛼 𝑒
Generally α and 𝐶𝛼 increase with increasing stress
Some common values of 𝐶𝛼 are given below
Sl. No Nature of Soil 𝐶𝛼 – Value
1 Over consolidated days 0.0005 to 0.0015
2 Normally consolidated days 0.005 to 0.030
3 Organic soils, peats 0.04 to 0.10

10.8 CALCULATION OF SETTLEMENT FROM ONE-DIMENSIONAL PRIMARY


CONSOLIDATION
With the knowledge gained from the analysis of consolidation test results, we can now proceed
to calculate the probable settlement caused by primary consolidation in the field, assuming one-
dimensional consolidation.
157
Lecture Note on GTE-I

average effective overburden pressure, Because of an increase of effective pressure, 𝜎′, let the
Let us consider a saturated clay layer of thickness H and cross-sectional area A under an existing

∆𝑉 = 𝑉0 − 𝑉1 = 𝐻𝐴 − (𝐻 − 𝑆𝐶)𝐴--------
primary settlement be Sc. Thus, the change in volume (Figure 10.11) can be given by

Hence ∆𝑉 = 𝑆𝐶𝐴-----------------------------
Eq.10.44
Eq.10.43

volume is equal to the change in the volume of voids, ∆𝑉𝑉. Hence,


Where, V0 and V1 are the initial and final volumes, respectively. However, the change in the total

∆𝑉 = 𝑆𝐶𝐴 = 𝑉𝑉0 − 𝑉𝑉1 = ∆𝑉𝑉-------------


Where, 𝑉𝑉0and 𝑉𝑉1are the initial and final void volumes, respectively. From the definition
Eq. 10.44

∆𝑉 = ∆𝑒𝑉𝑆------------------------------------
of void ratio, it follows that

Where, ∆e
Eq. 10.45

𝑉 = =
𝑉0
is the𝐴𝐻change of void ratio. But
𝑆
----------------------------- Eq. 10.46
1+𝑒0 1+𝑒0
Where, 𝑒0 is initial void ratio at volume V0. Thus, from Eqs. (10.43) through (10.46),
∆𝑉 = 𝑆 𝐴 = ∆𝑒𝑉 = 𝐴𝐻 ∆𝑒----------------
𝐶 𝑆
Eq.10.47
1+𝑒0

𝑆 =𝐻
∆𝑒
----------------------------------- Eq. 10.48
𝐶 1+𝑒0

Fig.10.11: Settlement caused by one dimensional consolidation


that exhibit′ a linear e-𝑙𝑜𝑔𝜎′relationship (see Figure 10.12),
∆𝑒 = 𝑐𝑐[log10(𝜎 + ∆𝜎 ) − 𝑙𝑜𝑔10𝜎 ]--------------
′ ′
For normally consolidated clays
0 0
Eq.10.49
Where, cc is the slope of the e-log 𝜎′ plot and is defined as the compression index. Substitution
of Eq. (10.49) into Eq. (10.48) gives

158
Lecture Note on GTE-I

𝑆𝐶
𝐻𝑙𝑜𝑔10+∆𝜎
( 0 )-------------------------
′ ′
𝑐𝑐 𝜎
= 1+𝑒 𝜎
′0
Eq.10.50
0

Fig.10.12: Consolidation characteristics of normally consolidated clays

In over-consolidated clays (see Figure 10.13), for field e-log 𝜎′ variation will be along the line
Fig.10.13: Consolidation characteristics of over consolidated clays

hj, the slope of which will be approximately equal to that for the laboratory rebound curve. The

∆𝑒 = 𝑐𝑠[log10(𝜎 + ∆𝜎 ) − 𝑙𝑜𝑔10𝜎 ]--------


slope of the rebound ′curve Cs′ is referred to ′as the swell index; so
0 0
Eq.10.51
Hence we obtain
159
Lecture Note on GTE-I

𝑆𝐶
𝐻𝑙𝑜𝑔10+∆𝜎
( 0 )-------------------
′ ′
𝑐𝑠 𝜎
= 1+𝑒 𝜎
′0
Eq.10.52

𝜎′ + ∆𝜎′ > 𝜎′ then


0

0 𝑐 ′
If
𝑐𝑠 𝜎 𝜎′ +∆𝜎′

𝑆𝐶 𝐻𝑙𝑜𝑔10 ( ) + 𝐻𝑙𝑜𝑔10( )-
𝑐 𝑐 𝑐 0
=
Eq.10.53
1+𝑒0 𝜎0

1+𝑒0 � �′
� �
10.9 CORRELATIONS FOR COMPRESSION INDEX (CC)
The compression index for the calculation of field settlement caused by consolidation can be
determined by graphic construction (as shown in Figure 10.13) after one obtains the laboratory
test results for void ratio and pressure.
Skempton (1944) suggested the following empirical expression for the compression index for

𝑐𝑐 = 0.009(𝐿𝐿 − 10%)---------------
undisturbed clays:
Eq.10.54
Where LL is the liquid limit
Several other correlations for the compression index are also available. They have been
developed by tests on various clays. Some of these correlations are given in Table 1.

On the basis of observations on several natural clays, Rendon-Herrero (1983) gave the

𝑐 = 0.14𝐺 ( ) -------------
1+𝑒𝑜 2.38
relationship for the compression index in the form
Eq.10.55
𝑠 𝐺𝑠
𝑐

𝑐 = 0.2343 ( ) 𝐺 --------------
𝐿𝐿%
Nagaraj and Murty (1985) expressed the compression index as
𝑐 𝑠
Eq.10.56
100
More recently, Park and Koumoto (2004) expressed the compression index by the following
relationship:
𝑐
= 371.747−4.275 -----------------
𝜂0 Eq.10.57
𝑐
𝜂0
where 𝜂0 = in situ porosity of the soil.

10.10 CORRELATIONS FOR SWELL INDEX (CS)


The swell index is appreciably smaller in magnitude than the compression index and generally
160
Lecture Note on GTE-I

𝑐 = 𝑡𝑜 𝑐 ------------------
can be 1determined
1
from laboratory tests. In most cases,
𝑠 𝑐
Eq.10.58
5 10

161
Lecture Note on GTE-I

𝑐 = 0.0463 ( ) 𝐺 ---------------
𝐿𝐿%
The swell index was expressed by Nagaraj and Murty (1985) as
𝑠 𝑠
Eq.10.59
100
Based on the modified Cam clay model, Kulhawy and Mayne (1990) have shown that
𝑐 =
𝑃𝐼

370
-------------------- Eq.10.60
𝑠
Where PI= Plasticity Index

Sol:

Sol:

148
Lecture Note on GTE-I

Sol:

Sol:

149
Lecture Note on GTE-I

Sol:

150
Lecture Note on GTE-I

Fig.1: Soil Profile at Building Site

Sol:

151
Lecture Note on GTE-I

Example 7:

Sol:

152
Lecture Note on GTE-I

Example 8:

Sol:

153
Lecture Note on GTE-I

Example 9:

Sol:

154
Lecture Note on GTE-I

Example 10

Sol:

155
Lecture Note on GTE-I

Example 11:

Sol:

156
Lecture Note on GTE-I

∆𝒆 𝟏 + 𝒆𝟏 + ∆𝒆
Now
𝟏+
=
∆𝑯 𝒆𝟎 𝑯𝟎
=
𝑯𝟎

∆𝑒
= 19.0
1.541+∆𝑒
3.52
Hence
0
OR ∆𝑒 = 0.350 and 𝑒0 = 0.541 + 0.350 = 0.891
In general, the relationship between ∆𝑒 and ∆𝐻 is given by
∆e 1.891
∆H 19.0
=

157
Lecture Note on GTE-I

158
Lecture Note on GTE-I

11.0 SHEAR STRENGTH OF SOILS

INTRODUCTION
Shearing Strength’ of a soil is perhaps the most important of its engineering properties. This is
because all stability analyses in the field of geotechnical engineering, whether they relate to
foundation, slopes of cuts or earth dams, involve a basic knowledge of this engineering property
of the soil. ‘Shearing strength’ or merely ‘Shear strength’ may be defined as the resistance to
shearing stresses and a consequent tendency for shear deformation.
Shearing strength of a soil is the most difficult to comprehend in view of the multitude of factors
known to affect it. A lot of maturity and skill may be required on the part of the engineer in
interpreting the results of the laboratory tests for application to the conditions in the field.
Basically speaking, a soil derives its shearing strength from the following:
(1) Resistance due to the interlocking of particles.
(2) Frictional resistance between the individual soil grains, which may be sliding friction, rolling
friction, or both.
(3) Adhesion between soil particles or ‘cohesion’.
Granular soils of sands may derive their shear strength from the first two sources, while cohesive
soils or clays may derive their shear strength from the second and third sources.
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 fall into the
second of the three categories just mentioned, from the point of view of shearing strength. The
shear strength of a soil cannot be tabulated in codes of practice since a soil can significantly
exhibit different shear strengths under different field and engineering conditions.

11.1 Internal Friction within Granular Soil Masses

In granular or cohesionless soil masses, the resistance to sliding on any plane through the point
within the mass is similar to that discussed in the previous sub-section; the friction angle in this
case is called the ‘angle of internal friction’. However, the frictional resistance in granular soil
masses is rather more complex than that between solid bodies, since the nature of the resistance
is partly sliding friction and partly rolling friction. Further, a phenomenon known as
‘interlocking’ is also supposed to contribute to the shearing resistance of such soil masses, as part
of the frictional resistance.
The angle of internal friction, which is a limiting angle of obliquity and hence the primary
criterion for slip or failure to occur on a certain plane, varies appreciably for a given sand with
the density index, 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, since it is almost so for a given sand at a given density.

159
Lecture Note on GTE-I

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.
11.2 PRINCIPAL PLANES AND PRINCIPAL STRESSES—MOHR’S CIRCLE
At a point in a stressed material, every plane will be subjected, in general, to a normal or direct
stress and a shearing stress. In the field of geotechnical engineering, compressive direct stresses
are usually considered positive, while tensile stresses are considered negative. A ‘Principal
plane’ is defined as a plane on which the stress is wholly normal, or one which does not carry
shearing stress. From mechanics, it is known that there exists three principal planes at any point
in a stressed material. The normal stresses acting on these principal planes are known as the
‘principal stresses’. The three principal planes are to be mutually perpendicular. In the order of
decreasing magnitude the principal stresses are designated the ‘major principal stress’, the
‘intermediate principal stress’ and the ‘minor principal stress’, the corresponding principal planes
being designated exactly in the same manner. It can be engineering by two-dimensional analysis,
the intermediate principal stress being commonly ignored.
Let us consider an element of soil whose sides are chosen as the principal planes, the major and
the minor, as shown in Fig. 11.1(a):

Fig. 11.1: Stresses on a plane inclined to the principal planes


Let O be any point in the stressed medium and OA and OB be the major and minor principal
planes, with the corresponding principal stresses σ1 and σ3. The plane of the figure is the
intermediate principal plane. Let it be required to determine the stress conditions on a plane
normal to the figure, and inclined at an angle θ to the major principal plane, considered positive
when measured counter-clockwise. If the stress conditions are uniform, the size of the element is
immaterial. If the stresses are varying, the element must be in finitestinal in size, so that the
variation of stress along a side needed to be considered.
Let us consider the element to be of unit thickness perpendicular to the plane of the figure, AB
being l. The forces on the sides of the element are shown dotted and their components parallel
and perpendicular to AB are shown by full lines. Considering the equilibrium of the element and
resolving all forces in the directions parallel and perpendicular to AB, the following equations

𝜎𝜃 = 𝜎1𝑐𝑜𝑠2𝜃 + 𝜎3𝑠𝑖𝑛2𝜃 = 𝜎3(𝜎1 − 𝜎3)𝑐𝑜𝑠2𝜃


may be obtained:

160
Lecture Note on GTE-I

=𝜎1+𝜎3 + ( ) 𝑐𝑜𝑠2𝜃-----------
𝜎1−𝜎3

2 2
Eq.11.1

𝜏 = 𝑠𝑖𝑛2𝜃--------------
𝜎1−𝜎3

2
Eq.11.2
𝜃

Fig. 11.2: Mohr’s circle for the stress conditions illustrated in Fig. 11.1

intermediate principal plane may be expressed in terms of 𝜎1 𝜎3 , and θ. Otto Mohr (1882)
Thus it may be noted that the normal and shearing stresses on any plane which is normal to the

represented these results graphical in a circle diagram, which is called Mohr’s circle. Normal
stresses are represented as abscissae and shear stresses as ordinates. If the coordinates, σ ɵ and τɵ

in Fig.11.2. This circle has its centre on the axis and cuts it at values 𝜎3 and 𝜎1. This circle is
represented by Eqs.1 and 2 are plotted for all possible values of θ, the locus is a circle as shown

known as the Mohr’s circle.


The Mohr’s circle diagram provides excellent means of visualisation of the orientation of
different planes. Let a line be drawn parallel to the major principal plane through D, the
coordinate of which is the major principal stress. The intersection of this line with the Mohr’s
circle, Qp is called the ‘Origin of planes’. If a line parallel to the minor principal plane is drawn
through E, the co-ordinate of which is the minor principal stress, it will also be observed to pass
through Op; the angle between these two lines is a right angle from the properties of the circle.
Likewise it can be shown that any line through Op, parallel to any arbitrarily chosen plane,
intersects the Mohr’s circle at a point the co-ordinates of which represent the normal and shear
stresses on that plane. Thus the stresses on the plane represented by AB in Fig.11.1 (a), may be
obtained by drawing Op parallel to AB, that is, at an angle θ with respect to OpD, the major
principal plane, and measuring off the co-ordinates of C, namely σθ and τθ.
Since angle COpD = θ, angle CFD = 2θ, from the properties of the circle. From the geometry of

𝜎𝜃 = 𝑀𝐺 = 𝑀𝐹 + 𝐹𝐺
the figure, the co-ordiantes of the point C, are established as follows:

𝜏𝜃 𝜎1 + 𝜎3
+(
2
= 𝐶𝐺 = 𝑠𝑖𝑛2𝜃
𝜎1−𝜎3

2
And
161
Lecture Note on GTE-I

𝜎1 − 𝜎3
) 𝑐𝑜𝑠2𝜃
2

162
Lecture Note on GTE-I

These are the same as in Eqs. 11.1 and 11.2, which prove our statements. In the special case
where the major and minor principal planes are vertical and horizontal respectively, or vice-
versa, the origin of planes will be D or E, as the case may be. In other words, it will lie on the σ-
axis. A few important basic facts and relationships may be directly obtained from the Mohr’s
circle:
1. The only planes free from shear are the given sides of the element which are the principal
planes. The stresses on these are the greatest and smallest normal stresses.
2. The maximum or principal shearing stress is equal to the radius of the Mohr’s circle, and

𝜏𝑚𝑎 =
it occurs on planes inclined at 45° to the principal planes.
𝜎1−𝜎3

2
------------------- Eq.11.3
𝑥
3. The normal stresses on planes of maximum shear are equal to each other and is equal to
half the sum of the principal stresses.
𝜎=
(𝜎1+𝜎3)

2
---------------- Eq.11.4
𝑐
4. Shearing stresses on planes at right angles to each other are numerically equal and are of
an opposite sign. These are called conjugate shearing stresses.

MG = 2MF = 𝜎1 + 𝜎3 . If we designate the normal stress on a plane perpendicular to the plane


5. The sum of the normal stresses on mutually perpendicular planes is a constant (MG′ +

on which it is 𝜎𝜃 �as 𝜎′
𝜎𝜃 + 𝜎 �

= 𝜎1 + 𝜎3--------------

Of the two stresses 𝜎𝜃 and �𝜎′ the one which makes the smaller angle with σ1 is the greater of the
Eq.11.5

6. The resultant stress, 𝜎𝑟, on any plane is


two.

𝜎𝑟 = √𝜎2 + 𝜏2 -----------------
𝜃 𝜃
Eq.11.6

𝛽 = 𝑡𝑎𝑛−1 ( 𝜃)----------------
𝜏
and has an obliquity, β, which is equal to
Eq.11.7
𝜎𝜃
7. Stresses on conjugate planes, that is, planes which are equally inclined in different
directions with respect to a principal plane are equal. (This is indicated by the co-ordinates of C
and C1 in Fig. 11.2).
8. When the principal stresses are equal to each other, the radius of the Mohr’s circle
becomes zero, which means that shear stresses vanish on all planes. Such a point is called an
isotropic point.
9. The maximum angle of obliquity, βm, occurs on a plane inclined at
𝜃𝑐 = 450+𝛽𝑚

2
--------------- Eq.11.8
𝑟
This may be obtained by drawing a line which passes through the origin and is tangential to the
Mohr’s circle. The co-ordinates of the point of tangency are the stresses on the plane of
maximum obliquity; the shear stress on this plane is obviously less than the principal or
maximum shear stress. On the plane of principal shear the obliquity is slightly smaller than βm.
163
Lecture Note on GTE-I

maximum shear, since the criterion of slip is limiting obliquity. When 𝛽𝑚 approaches and equals
It is the plane of maximum obliquity which is most liable to failure and not the plane of

the angle of internal friction, φ, of the soil, failure will become incipient. Mohr’s circle affords
an easy means of obtaining all important relationships. The following are a few such
relationships
𝑠𝑖𝑛𝛽 = ( )-----------------
𝜎1−𝜎3
Eq.11.9
𝑚 𝜎1+𝜎3

= ( )----------------
𝜎1 1+𝑠𝑖𝑛𝛽𝑚
Eq.11.10
𝜎3 1−𝑠𝑖𝑛𝛽𝑚
In case the normal and shearing stresses on two mutually perpendicular planes are known, the
principal planes and principal stresses may be determined with the aid of the Mohr’s circle
diagram, as shown in Fig. 11.3(a). The shearing stresses on two mutually perpendicular planes
are equal in magnitude by the principle of complementary shear

(a) General two-dimensional stress system

(b) Mohr’s circle for general two-dimensional stress system


Fig. 11.3: Determination of principal planes and principal stresses from Mohr’s circle

stresses 𝜎𝑥 and 𝜎𝑦 on mutually perpendicular planes and shear stresses 𝜏𝑥𝑦on these planes,
Figure 11.3 (a) shows an element subjected to a general two-dimensional stress system, normal

as
164
Lecture Note on GTE-I

indicated. Fig. 11.3 (b) shows the corresponding Mohr’s circle, the construction of which is
obvious.
From a consideration of the equilibrium of a portion of the element, the normal and shearing
stress components, 𝜎𝜃 and 𝜏𝜃, respectively, on a plane inclined at an angle θ, measured counter-
𝜎𝑥+𝜎𝑦 𝜎𝑥−𝜎𝑦
𝜎 = +( ) 𝑐𝑜𝑠2𝜃 + 𝜏 𝑠𝑖𝑛2𝜃---------
clockwise with respect to the plane on which σx acts, may be obtained as follows:

𝜃 𝑥𝑦
Eq.11.11
2 2
𝜎𝑥−𝜎𝑦
𝜏 = 𝑠𝑖𝑛2𝜃 𝑥 𝑐𝑜𝑠2𝜃---------------
𝜃 − 𝜏
Eq.11.12
𝑦
2
Squaring and adding these Eqs, we obtain
2
[𝜎𝜃 𝜎𝑥+𝜎𝑦 𝜎𝑥−𝜎

2 + 𝜏2 = + 𝜏2 ----------
)
𝑦 2
] (
Eq.11.13
2 � 𝑥
, 0 and radius √(
� 𝜎
𝑦𝑥+𝜎𝑦 𝜎𝑥−𝜎𝑦
2
2
) + 𝜏𝑥𝑦
This represents a circle with centre
2 2
Once the Mohr’s circle is constructed, the principal stresses σ1 and σ3, and the orientation of the
principal planes may be obtained from the diagram. The shearing stress is to be plotted upward
or downward according as it is positive or negative. It is common to take a shear stress which
tends to rotate the element counter-clockwise, positive.
It may be noted that the same Mohr’s circle and hence the same principal stresses are obtained,
irrespective of how the shear stresses are plotted. (The centre of the Mohr’s circle, C, is the mid-

being σy and 𝜏𝑥𝑦


point of DE, with the co-ordinates and 0; the radius of the circle is CG), the co-ordinates of G

+ √(𝜎 − 𝜎 ) + 4𝜏2 ---------------


The following relationships are also easily obtained
𝜎𝑥+𝜎𝑦 2
𝜎 =
1
1 𝑥 𝑦 𝑥𝑦
Eq.11.14
2
2

√(𝜎 − 𝜎 ) + 4𝜏2 -----------------


𝜎𝑥+𝜎𝑦 2
𝜎 = −
1
3 𝑥 𝑦 𝑥𝑦
Eq.11.15
2
2
The direction of principal stresses is expressed as given below
tan(2𝜃)1 =
2𝜏𝑥𝑦
--------------- Eq.11.16
.3 𝜎𝑥−𝜎𝑦

𝜏 = √(
1 𝜎𝑥−𝜎𝑦
2
2
) + 4𝜏𝑥𝑦
--------------- Eq.11.17
𝑚𝑎𝑥 2
2
Invariably, the vertical stress will be the major principal stress and the horizontal one the minor
principal stress in geotechnical engineering situations.
11.2.1 Mohr’s Strength Theory
We have seen that the shearing stress may be expressed as τ = σ tan β on any plane, where β is
the angle of obliquity. If the obliquity angle is the maximum or has limiting value φ, the shearing
165
Lecture Note on GTE-I

stress is also at its limiting value and it is called the shearing strength, s. For a cohesionless soil
the shearing strength may be expressed as:
s = σ tan φ ... Eq.11.18
If the angle of internal friction φ is assumed to be a constant, the shearing strength may be
represented by a pair of straight lines at inclinations of + φ and – φ with the σ-axis and passing

166
Lecture Note on GTE-I

through the origin of the Mohr’s circle diagram. A line of this type is called a Mohr envelope.
The Mohr envelopes for a cohesionless soil, as shown in Fig. 11.4, are the straight lines OA and
OA′.

Fig.11.4: Mohr’s strength theory—Mohr envelopes for cohesionless soil


If the stress conditions at a point are represented by Mohr’s circle I, the shear stress on any plane
through the point is less than the shearing strength, as indicated by the line BCD;
BC represents the shear stress on a plane on which the normal stress is given by OD.BD,
representing the shearing strength for this normal stress, is greater than BC.
The stress conditions represented by the Mohr’s Circle II, which is tangential to the Mohr’s
envelope at F, are such that the shearing stress, EF, on the plane of maximum obliquity is equal
to the shearing strength. Failure is incipient on this plane and will occur unless the normal stress
on the critical plane increases.
It may be noted that it would be impossible to apply the stress conditions represented by Mohr’s
circle III (dashed) to this soil sample, since failure would have occurred even by the time the
shear stress on the critical plane equals the shearing strength available on that plane, thus
eliminating the possibility of the shear stress exceeding the shearing strength.
The Mohr’s strength theory, or theory of failure or rupture, may thus be stated as follows:
The stress condition given by any Mohr’s circle falling within the Mohr’s envelope represents a
condition of stability, while the condition given by any Mohr’s circle tangent to the Mohr’s
envelope indicates incipient failure on the plane relating to the point of tangency. The Mohr’s
envelope may be treated to be a property of the material and independent of the imposed stresses.
Also, the Mohr’s circle of stress depends only upon the imposed stresses and has nothing to do
with the nature and properties of the material.
To emphasize that the stresses in Eq. 11.18 are those on the plane on which failure is incipient,

𝑠 = 𝜎𝑓tan∅---------------------------------------------
we add the subscript f to σ:
Eq.11.19
It is possible to express the strength in terms of normal stress on any plane, with the aid of the

𝜎𝑓 = 𝜎3(1 + 𝑠𝑖𝑛∅ ) = 𝜎1(1 − 𝑠𝑖𝑛∅)---------------


Mohr’s circle of stress. Some common relationships are :

𝑠 = 𝜎 𝑡𝑎𝑛∅ = 𝑐𝑜𝑠∅---------------------------
𝜎1−𝜎3
Eq.11.20
Eq.11.21
𝑓 2

167
Lecture Note on GTE-I

The primary assumptions in the Mohr’s strength theory are that the intermediate principal stress
has no influence on the strength and that the strength is dependent only upon the normal stress on
the plane of maximum obliquity. However, the shearing strength, in fact, does depend to a small
extent upon the intermediate principal stress, density speed of application of shear, and so on.
But the Mohr theory explains satisfactorily the strength concept in soils and hence is in vogue. It
may also be noted that the Mohr envelope will not be a straight line but is actually slightly
curved since the angle of internal friction is known to decrease slightly with increase in stress.
11.2.3 Mohr-Coulomb Theory
The Mohr-Coulomb theory of shearing strength of a soil, first propounded by Coulomb (1976)
and later generalised by Mohr, is the most commonly used concept. The functional relationship
between the normal stress on any plane and the shearing strength available on that plane was
assumed to be linear by Coulomb; thus the following is usually known as Coulomb’s law:
s = c + σ tan φ ------------- Eq. 11.22
where c and φ are empirical parameters, known as the ‘apparent cohesion’ and ‘angle of shearing
resistance’ (or angle of internal friction), respectively. These are better visualised as ‘parameters’
and not as absolute properties of a soil since they are known to vary with water content,
conditions of testing such as speed of shear and drainage conditions, and a number of other
factors besides the type of soil.
Coulomb’s law is merely a mathematical equation of the failure envelope shown in Figure 11.5
(a); Mohr’s generalisation of the failure envelope as a curve which becomes flatter with
increasing normal stress is shown in Fig. 11.5 (b).

Fig. 11.5: Mohr-Coulomb Theory—failure envelopes


The envelopes are called ‘strength envelopes’ or ‘failure envelopes’. The meaning of an
envelope has already been given in the previous section; if the normal and shear stress
components on a plane are plotted on to the failure envelope, failure is supposed to be incipient
and if the stresses plot below the envelope, the condition represents stability. And, it is
impossible that these plot above the envelope, since failure should have occurred previously
Coulomb’s law is also written as follows to indicate that the stress condition refers to that on the
plane of failure:
s = c + σf tan φ ----------- Eq. 11.23

168
Lecture Note on GTE-I

In a different way, 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 (circle I); otherwise, it would wholly lie below the envelopes as
shown in circle II, Fig.11.5 (b).
The Coulomb envelope in special cases may take the shapes given in Fig. 11.6 (a) and (b); for a
purely cohesionless or granular soil or a pure sand, it would be as shown in Fig. 11.6 (a) and for
a purely cohesive soil or a pure clay, it would be as shown in Fig. 11.6 (b).

Fig. 11.6: Coulomb envelopes for pure sand and for pure clay

11.3 SHEARING STRENGTH—A FUNCTION OF EFFECTIVE STRESS


Equation 11.23 apparently indicates that the shearing strength of a soil is governed by the total
normal stress on the failure plane. However, according to Terzaghi, it is the effective stress on
the failure plane that governs the shearing strength and not the total stress.
It may be expected intuitively that the denser a soil, the greater the shearing strength. It has been
learnt in chapter seven that a soil deposit becomes densest under any given pressure after the
occurrence of complete consolidation and consequent dissipation of pore water
Thus, complete consolidation, dependent upon the dissipation of pore water pressure and hence
upon the increase in the effective stress, leads to increase in the shearing strength of a soil. In
other words, it is the effective stress in the case of a saturated soil and not the total stress which
is relevant to the mobilisation of shearing stress.
Further, the density of a soil increase when subjected to shearing action, drainage being allowed
simultaneously. Therefore, even if two soils are equally dense on having been consolidated to the
same effective stress, they will exhibit different shearing strengths if drainage is permitted during
shear for one, while it is not for the other.
These ideas lead to a statement that ‘‘the strength of a soil is a unique function of the effective
stress acting on the failure plane’’.

s = c′ + 𝜎𝑓 tan φ′ --------------
Eq. 11.23 may now be modified to read:
Eq. 11.24
Where, c′ and φ′ are called the effective cohesion and effective angle of internal friction,
respectively, since they are based on the effective normal stress on the failure plane.

169
Lecture Note on GTE-I

Collectively, they are called ‘effective stress parameters’, while c and φ of Eq. 24 are called
‘‘total stress parameters’’.

11.4 SHEARING STRENGTH TESTS

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
1. Direct Shear Test
2. Triaxial Compression Test
3. Unconfined Compression Test
4. Laboratory Vane Shear Test
5. Torsion Test
6. Ring Shear Tests
Field Tests
1. Vane Shear Test
2. Penetration Test
The first three tests among the laboratory tests are very commonly used, while the fourth is
gaining popularity owing to its simplicity. The fifth and sixth are mostly used for research
purposes and hence are not dealt with here.
The principle of the field vane test is the same as that of the laboratory vane shear test, except
that the apparatus is bigger in size for convenience of field use. The penetration test involves the
measurement of resistance of a soil to penetration of a cone or a cylinder, as an indication of the
shearing strength. This procedure is indirect and rather empirical in nature although correlations
are possible. The field tests are also not considered here. The details of the test procedures are
available in the relevant I.S. codes or any book on laboratory testing, such as Lambe (1951).
11.4.1 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 shown schematically in Figure 11.7.
The soil is gripped in perforated metal grilles, behind which porous discs can be placed if
required to allow the specimen to drain. For undrained tests, metal plates and solid metal grilles
may be used. The usual plan size of the specimen is 60 mm square; but a larger size such as 300
mm square or even more, is employed for testing larger size granular material such as gravel.
The minimum thickness or height of the specimen is 20 mm. 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
throughout 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.

170
Lecture Note on GTE-I

Fig. 11.7: Direct Shear Test Device


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. The principles of these two types of devices are
illustrated schematically in Fig. 11.7 (b) and (c), respectively. In the stress-controlled type, the
shear stress, which is the 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 increment 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.
This shearing force is measured with the aid of a proving ring—a steel ring that has been

171
Lecture Note on GTE-I

carefully machined, balanced and calibrated. The deflection of the annular ring is measured with
the aid of a dial gauge set inside the ring, the causative force being got for any displacement by
means of the calibration chart supplied by the manufacturer.
The shear displacement is measured again 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. The strain-controlled type is very widely
used. The strain is taken as the ratio of the shear displacement to the thickness of the sample. The
proving ring readings may be taken at fixed displacements or even at fixed intervals of time as
the rate of strain is made constant by an electric motor. A sudden drop in the proving ring
reading or a leveling-off in successive readings indicates shear failure of the soil specimen.
The shear strain may be plotted against the shear stress; it may be plotted versus the ratio of the
shearing stress on normal stress; and it may also be plotted versus volume change.
Each plot may yield information useful in one way or the other. The stresses may be obtained
from the forces by dividing them by the area of cross-section of the sample. The stress-
conditions on the failure plane and the corresponding Mohr’s circle for direct shear test are
shown in Fig. 11.8 (a) and (b) respectively.

Fig.11.8: Mohr’s circle representation of stress condition for direct shear test results

The failure plane is predetermined as the horizontal plane here. Several specimens are tested
under different normal loads and the results plotted to obtain failure envelopes.
The direct shear test is a relatively simple test. Quick drainage, i.e., quick dissipation of pore
pressures is possible since the thickness of the specimen is small. However, the test suffers from
the following inherent disadvantages, which limit its application.
1. The stress conditions are complex primarily because of the non-uniform distribution of normal
and shear stresses on the plane.
2. There is virtually no control of the drainage of the soil specimen as the water content of a
saturated soil changes rapidly with stress.
3. The area of the sliding surface at failure will be less than the original area of the soil specimen
and strictly speaking, this should be accounted for.
4. The ridges of the metal gratings embedded on the top and bottom of the specimen, causes
distortion of the specimen to some degree.

172
Lecture Note on GTE-I

5. The effect of lateral restraint by the side walls of the shear box is likely to affect the results.
6. The failure plane is predetermined and this may not be the weakest plane. In fact, this is the
most important limitation of the direct shear test.
11.4.2 Triaxial Shear Test
The triaxial shear test is one of the most reliable methods available for determining shear strength
parameters. It is used widely for research and conventional testing. A diagram of the triaxial test
layout is shown in Figure 11.9.
In this test, a soil specimen about 36 mm in diameter and 76 mm (3 in.) long generally is used. The
specimen is encased by a thin rubber membrane and placed inside a plastic cylindrical chamber
that usually is filled with water or glycerine. The specimen is subjected to a confining pressure
by compression of the fluid in the chamber. (Note: Air is sometimes used as a compression
medium.) To cause shear failure in the specimen, one must apply axial stress (sometimes called
deviator stress) through a vertical loading ram.
This stress can be applied in one of two ways:
1. Application of dead weights or hydraulic pressure in equal increments until the specimen fails.
(Axial deformation of the specimen resulting from the load applied through the ram is measured
by a dial gauge.)
2. Application of axial deformation at a constant rate by means of a geared or hydraulic loading
press. This is a strain-controlled test.
The axial load applied by the loading ram corresponding to a given axial deformation is measured by a
proving ring or load cell attached to the ram.

Fig.11.9: Triaxial experimental set up

173
Lecture Note on GTE-I

Test Procedure
The essential steps in the conduct of the test are as follows:
(i) A saturated porous stone is placed on the pedestal and the cylindrical soil specimen is placed
on it.
(ii) The specimen is enveloped by a rubber membrane to isolate it from the water with which the
cell is to be filled later; it is sealed with the pedestal and top cap by rubber ‘‘O’’ rings.
(iii) The cell is filled with water and pressure is applied to the water, which in turn is transmitted
to the soil specimen all-round and at top. This pressure is called ‘cell pressure’, ‘chamber
pressure’ or ‘confining pressure’.
(iv) Additional axial stress is applied while keeping the cell pressure constant. This introduces
shearing stresses on all planes except the horizontal and vertical planes, on which the major,
minor and intermediate principal stresses act, the last two being equal to the cell pressure on
account of axial symmetry.
(v) The additional axial stress is continuously increased until failure of the specimen occurs.
(What constitutes failure is often a question of definition and may be different for different kinds
of soils. This aspect would be discussed later on).
A number of observations may be made during a Triaxial compression test regarding the
physical changes occurring in the soil specimen:
(a) As the cell pressure is applied, pore water pressure develops in the specimen, which can be
measured with the help of a pore pressure measuring apparatus, such as Bishop’s pore pressure
device (Bishop, 1960), connected to the pore pressure line, after closing the valve of the drainage
line.
(b) If the pore pressure is to be dissipated, the pore water line is closed, the drainage line opened
and connected to a burette. The volume decrease of the specimen due to consolidation is
indicated by the water drained into the burette.
(c) The axial strain associated with the application of additional axial stress can be measured by
means of a dial gauge, set to record the downward movement of the loading piston.
(d) Upon application of the additional axial stress, some pore pressure develops. It may be
measured with the pore pressure device, after the drainage line is closed. On the other hand, if it
is desired that any pore pressure developed be allowed to be dissipated, the pore water line is
closed and the drainage line opened as stated previously.
(e) The cell pressure is measured and kept constant during the course of the test.
(f) The additional axial stress applied is also measured with the aid of a proving ring and dial
gauge.
Thus the entire triaxial test may be visualised in two important stages:
(i) The specimen is placed in the triaxial cell and cell pressure is applied during the first stage.
(ii) The additional axial stress is applied and is continuously increased to cause a shear failure,
the potential failure plane being that with maximum obliquity during the second stage.

174
Lecture Note on GTE-I

11.4.3 Area Correction for the Determination of Additional Axial Stress or Deviatoric Stress
The additional axial load applied at any stage of the test can be determined from the proving ring
reading. During the application of the load, the specimen undergoes axial compression and
horizontal expansion to some extent. Little error is expected to creep in if the volume is supposed
to remain constant, although the area of cross-section varies as axial strain increases.
The assumption is perfectly valid if the test is conducted under undrained conditions, but, for drained
conditions, the exact relationship is somewhat different.
If A0, h0 and V0 are the initial area of cross-section, height and volume of the soil specimen respectively,
and if A, h, and V are the corresponding values at any stage of the test, the corresponding

𝐴(ℎ0 − ∆ℎ) = 𝑉 = 𝑉0 + ∆𝑉-----------------------


changes in the values being design

Hence 𝐴 =
𝑉0+∆𝑉
Eq.11.25
------------------------------- Eq.11.26
ℎ0−∆ℎ
But, for axial 𝑉 ∆𝑉
compression, Δh∆𝑉
is known to be negative.
0(1+ ) 𝐴0(1+

𝐴0+∆𝑉
=
𝑉0 𝑉0

=
𝑉 ---------------------------- Eq.11.27
1−𝜖𝑎
=∆ℎ
ℎ0(1− )
ℎ0−∆ℎ ℎ0

an undrained test, 𝐴 =
since the axial strain, εa = Δh/h0------------------------- Eq.11.28
For
𝐴0 ------------------------- Eq.11.29
1−𝜖𝑎
since ΔV = 0.
1
This is called the ‘Area correction’ and
1−𝜖𝑎
is the correction factor.

𝐴 =𝐴0
A more accurate expression for the corrected area is given by
=
𝑉0+∆𝑉

1−𝜖
------------------------ Eq.11.30
𝑎
ℎ0−∆ℎ

Once the corrected area is determined, the additional axial stress or the deviator stress, Δσ, is

∆𝜎 = 𝜎1 − 𝜎3=Axial load (from proving ring reading)/Corrected area


obtained as

The cell pressure or the confining pressure, σc, itself being the minor principal stress, σ3, this is

𝜎1 = 𝜎3 + ∆𝜎------------------------
constant for one test; however, the major principal stress, σ1, goes on increasing until failure.
Eq.11.31
11.4.4 Mohr’s Circle for Triaxial Test
The stress conditions 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 Figure 11.10:

175
Lecture Note on GTE-I

Fig. 11.10: Mohr’s circle during Triaxial test

176
Lecture Note on GTE-I

The cell pressure, 𝜎𝑐 which is also the minor principal stress is constant and 𝜎11, 𝜎12, 𝜎13, 𝜎1𝑓
are the major principal stresses at different stages of loading and at failure. The Mohr’s 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 for one particular value of cell
pressure will be as shown in Fig. 11.11.

Fig. 11.11: Mohr’s circle at failure for a particular cell pressure


The Mohr’s circles at failure for one particular cell pressure are shown for the three typical cases
of a general c–φ soil, a φ-soil and a c-soil in Figs. 11.11 (a), (b), and (c) respectively.
With reference to Fig. 11.11 (a), the relationship between the major and minor principal stresses
at failure may be established from the geometry of the Mohr’s circle, as follows:

∴ α = 45° + φ/2------------------
From ΔDCG, 2α = 90° + φ
Eq.11.32
Again from ΔDCG
𝜎1−𝜎3
𝑠𝑖𝑛∅ = = =
𝐷𝐶 𝐷𝐶
( )
2 𝜎 + -------------------------------
𝐺𝐶 𝑐𝑐𝑜𝑡∅+ 1
Eq.11.33
𝜎3
𝐺𝑀+𝑀𝐶 2
𝜎1 − 𝜎3 = 2𝑐𝑐𝑜𝑡∅ + (𝜎1 + 𝜎3)𝑠𝑖𝑛∅------------------------ Eq.11.34
= σ3 tan2 (45° + φ/2) + 2c tan(450 + )------------------------

𝜎1
Or, Eq.11.35
2

177
Lecture Note on GTE-I

Or , 𝜎1 = 𝜎3 tan2 α + 2c tan α ------------------------- Eq.11.36

178
Lecture Note on GTE-I

This is also written as


𝜎1 = 𝜎3𝑁∅ + 2𝑐√𝑁∅----------------------- Eq.11.37
= tan2 α = tan2(45° + ∅) -----------------------------
𝑁∅
2
where, Eq.11.38

The above equations define the relationship between the principal stresses at failure. This state of
stress is defined as ‘Plastic equilibrium condition’, when failure is imminent.

sets are necessary to evaluate the parameters c and ∅ conventionally, three or more such sets are
From one test, a set of σ1 and σ3 is known; however, it can be seen from that at least two such

used from a corresponding number of tests.


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 this 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 the cohesion and the
angle of slope of this line with σ-axis gives the angle of internal friction, as shown in Fig. 11.12.

Fig.11.12: Mohr’s circle at failure for different cell pressures


11.4.5 Merits of Triaxial Compression Test
The following are the significant points of merit of triaxial compression test:
(1) Failure occurs along the weakest plane unlike along the predetermined plane in the case of
direct shear test.
(2) The stress distribution on the failure plane is much more uniform than it is in the direct shear
test: the failure is not also progressive, but the shear strength is mobilised all at once. Of course,
the effect of end restraint for the sample is considered to be a disadvantage; however, this may
not have pronounced effect on the results since the conditions are more uniform to the desired
degree near the middle of the height of the sample where failure usually occurs.
(3) Complete control of the drainage conditions is possible with the triaxial compression test;
this would enable one to simulate the field conditions better.
(4) The possibility to vary the cell pressure or confining pressure also affords another means to
simulate the field conditions for the sample, so that the results are more meaningfully interpreted.
(5) Precise measurements of pore water pressure and volume changes during the test are
possible.
(6) The state of stress within the specimen is known on all planes and not only on a
predetermined failure plane as it is with direct shear tests.
(7) The state of stress on any plane is capable of being determined not only at failure but also at
any earlier stage.
179
Lecture Note on GTE-I

(8) Special tests such as extension tests are also possible to be conducted with the triaxial testing
apparatus.
(9) It provides an ingenious and a symmetrical three-dimensional stress system better suited to
simulate field conditions.

11.5 TYPES OF SHEAR TESTS BASED ON DRAINAGE CONDITIONS


Before considering various methods of conducting shearing strength tests on a soil, it is necessary to
consider the possible drainage conditions before and during the tests since the results is
significantly affected by these.
A cohesionless or a coarse-grained soil may be tested for shearing strength either in the dry condition or
in the saturated condition. A cohesive or fine-grained soil is usually tested in the saturated
condition. Depending upon whether drainage is permitted before and during the test, shear tests
on such saturated soils are classified as follows:
Unconsolidated Undrained Test
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,
because of the shortness of drainage path. However, undrained tests are often performed only on
soils of low permeability. This is the most unfavourable condition which might occur in
geotechnical engineering practice and hence is simulated in shear testing. Since a relatively small
time is allowed for the testing till failure, it is also called the ‘Quick test.’ It is designated UU, Q,
or Qu test.
Consolidated Undrained Test
Drainage is permitted fully in this type of test during the application of the normal stress and no drainage
is permitted during the application of the shear stress. Thus volume changes do not take place
during shear and excess pore pressure develops. Usually, after the soil is consolidated under the
applied normal stress to the desired degree, 5 to 10 minutes may be adequate for the test. This
test is also called ‘consolidated quick test’ and is designated CU or Qc test, These conditions are
also common in geotechnical engineering practice.
11.5.1 Consolidated-Drained Triaxial Test

In the CD test, the saturated specimen first is subjected to an all around confining pressure, 𝜎3, by
compression of the chamber fluid. As confining pressure is applied, the pore water pressure of
the specimen increases by uc (if drainage is prevented). This increase in the pore water pressure

𝐵 = ------------------
𝑢𝑐
can be expressed as a non dimensional parameter in the form
Eq.11.39
𝜎3
Where, B= Skempton’s pore pressure parameter (Skempton, 1954).

180
Lecture Note on GTE-I

For saturated soft soils, B is approximately equal to 1; however, for saturated stiff soils, the
magnitude of B can be less than 1. Black and Lee (1973) gave the theoretical values of B for
various soils at complete saturation.
Now, if the connection to drainage is opened, dissipation of the excess pore water pressure, and
thus consolidation, will occur. With time, uc will become equal to 0. In saturated soil, the change
in the volume of the specimen (∆Vc) that takes place during consolidation can be obtained from

Next, the deviator stress, ∆𝜎𝑑 on the specimen is increased very slowly (Figure 11.13b). The
the volume of pore water drained (Figure 11.13a).

drainage connection is kept open, and the slow rate of deviator stress application allows
complete dissipation of any pore water pressure that developed as a result (∆𝑢𝑑=0)
A typical plot of the variation of deviator stress against strain in loose sand and normally

over consolidated clay. The volume change, ∆ Vd, of specimens that occurs because of the
consolidated clay is shown in Figure 11.13b. Figure 1c shows a similar plot for dense sand and

application of deviator stress in various soils is also shown in Figures 11.13d and e.

Fig. 11.13: Consolidated-drained triaxial test: (a) volume change of specimen caused by
chamber-confining pressure; (b) plot of deviator stress against strain in the vertical direction for
loose sand and normally consolidated clay; (c) plot of deviator stress against strain in the vertical
direction for dense sand and over consolidated clay; (d) volume change in loose sand and
normally consolidated clay during deviator stress application; (e) volume change in dense sand
and over consolidated clay during deviator stress application

181
Lecture Note on GTE-I

Total and effective confining stress = 𝜎3 = 𝜎


Because the pore water pressure developed during the test is completely dissipated, we have

3

Total and effective axial stress at failure=𝜎3 + ∆𝜎𝑑𝑓 = 𝜎1 1= 𝜎
In a triaxial test,𝜎1′ is the major principal effective stress at failure and is the minor principal
effective stress at failure. Several tests on similar specimens can be conducted by varying the
confining pressure. With the major and minor principal stresses at failure for each test the
Mohr’s circles can be drawn and the failure envelopes can be obtained. Figure 11.14 shows the
type of effective stress failure envelope obtained for tests on sand and normally consolidated
clay.

Fig.11.14: Effective stress failure envelope from drained tests on sand and normally consolidated
clay
The coordinates of the point of tangency of the failure envelope with a Mohr’s circle (that is,
point A) give the stresses (normal and shear) on the failure plane of that test specimen.

𝑠𝑖𝑛∅ = ----------------------

For normally consolidated clay, referring to Figure 2
′ 𝑂𝐴
𝑂𝑂′
Eq.11.40

OR 𝑠𝑖𝑛∅′=𝜎
′ ′
𝜎1 3
−𝜎

+𝜎1
--------------------------- Eq.11.41
1 3 𝜎′−𝜎′
Hence ∅′ = 𝑠𝑖𝑛−1𝜎
( ′ 1+𝜎
3
)----------------------
1 Eq.11.42
1 3

Also, the failure plane will be inclined at an angle of 𝜃 = 45 + to the major principal plane, as

2
shown in Figure 11.14.

pressure of 𝜎𝑐 = 𝜎′ and is allowed to swell by reducing the chamber pressure to 𝜎 = 𝜎′ . The


Over consolidation results when a clay initially is consolidated under an all-around chamber
𝑐 3 3
failure envelope obtained from drained triaxial tests of such over consolidated clay specimens
shows two distinct branches (ab and bc in Figure 11.15). The portion ab has a flatter slope with a

𝜏𝑓 = 𝑐′ + 𝜎′𝑡𝑎𝑛∅′--------------------
cohesion intercept, and the shear strength equation for this branch can be written as
Eq.11.43

182
Lecture Note on GTE-I

Fig. 11.15: Effective stress failure envelope for over consolidated clay
A consolidated-drained triaxial test on a clayey soil may take several days to complete. This
amount of time is required because deviator stress must be applied very slowly to ensure full
drainage from the soil specimen. For this reason, the CD type of triaxial test is uncommon.

11.5.2 Consolidated-Undrained Triaxial Test

saturated soil specimen is first consolidated by an all-around chamber fluid pressure, 𝜎3, that
The consolidated-undrained test is the most common type of triaxial test. In this test, the

results in drainage (Figures 11.16 a and b).

Fig.11.16: Consolidated-undrained test: (a) specimen under chamber-confining pressure; (b)


volume change in specimen caused by confining pressure; (c) deviator stress application; (d)

183
Lecture Note on GTE-I

deviator stress against axial strain for loose sand and normally consolidated clay; (e) deviator
stress against axial strain for dense sand and over consolidated clay; (f) variation of pore water
pressure with axial strain for loose sand and normally consolidated clay; (g) variation of pore
water pressure with axial strain for dense sand and over consolidated clay.

deviator stress, ∆𝜎𝑑, on the specimen is increased to cause shear failure (Figure 11.16c). During
After the pore water pressure generated by the application of confining pressure is dissipated, the

permitted, the pore water pressure, ∆𝑢𝑑 , will increase. During the test, simultaneous
this phase of the test, the drainage line from the specimen is kept closed. Because drainage is not

measurements of ∆𝜎𝑑, and ∆𝑢𝑑 are made. The increase in the pore water pressure, ∆𝑢𝑑can be

𝐴 = 𝑑----------------------
∆𝑢
expressed in a nondimensional form as
Eq.11.44
∆𝜎𝑑
where 𝐴̅ = Skempton’s pore pressure parameter (Skempton, 1954).
The general patterns of variation of ∆𝜎𝑑 and ∆𝑢𝑑with axial strain for sand and clay soils
are shown in Figures 4 d through 4g. In loose sand and normally consolidated clay, the pore
water pressure increases with strain. In dense sand and over consolidated clay, the pore water
pressure increases with strain to a certain limit, beyond which it decreases and becomes
negative (with respect to the atmospheric pressure). This decrease is because of a tendency of
the soil to dilate. Unlike the consolidated-drained test, the total and effective principal stresses
are not the same in the consolidated-undrained test. Because the pore water pressure at failure
is measured in this test, the principal stresses may be analyzed as follows:
• Major principal stress at failure (total):
• Major principal stress at failure (effective):
• Minor principal stress at failure (total):
• Minor principal stress at failure (effective):

𝜎1 − 𝜎3 = 𝜎′ − 𝜎 ′ --------------------------
The preceding derivations show that
1 3
Eq.11.45
Tests on several similar specimens with varying confining pressures may be conducted to
determine the shear strength parameters. Figure 11.17 shows the total and effective stress

Fig. 11.17: Total and effective stress failure envelopes for consolidated undrained triaxial tests

184
Lecture Note on GTE-I

Mohr’s circles at failure obtained from consolidated-undrained triaxial tests in sand and normally
consolidated clay. Note that A and B are two total stress Mohr’s circles obtained from two tests.
C and D are the effective stress Mohr’s circles corresponding to total stress circles A and B,
respectively. The diameters of circles A and C are the same; similarly, the diameters of circles B
and D are the same. In Figure 11.17, the total stress failure envelope can be obtained by drawing
a line that touches all the total stress Mohr’s circles. For sand and normally consolidated clays,
this will be approximately a straight line passing through the origin and may be expressed by the

𝜏𝑓 = 𝜎𝑡𝑎𝑛∅--------------------
equation

where 𝜎 = total stress


Eq.11.46

∅ =the angle that the total stress failure envelope makes with the normal stress axis, also known
as the consolidated-undrained angle of shearing resistance

∅ = 𝑠𝑖𝑛−1 ( )--------------------
𝜎1−𝜎3
For sand and normally consolidated clay, we can write
Eq.11.47
𝜎1+𝜎3

∅′
= 𝑠𝑖𝑛−1 (
𝜎1−𝜎3
And

𝜎1+𝜎3−2(∆𝑢
)-------------------- Eq.11.48
𝑑)

Fig. 11.18: Total stress failure envelope obtained from consolidated-undrained tests in over
consolidated clay
Again referring to Fig.11.17, we see that the failure envelope that is tangent to all the effective

𝜏𝑓 = 𝜎′𝑡𝑎𝑛∅′-----------------------
stress Mohr’s circles can be represented by the equation
Eq.11.48
which is the same as that obtained from consolidated-drained tests.
In over consolidated clays, the total stress failure envelope obtained from consolidated undrained

𝜏𝑓 = 𝜎𝑡𝑎𝑛∅1------------------------
tests will take the shape shown in Fig.11.18. The straight line is represented by the equation
Eq.11.49
and the straight line follows the relationship given by the above equation. The effective stress
failure envelope drawn from the effective stress Mohr’s circles will be similar to that shown in
Figure 11.15.
Consolidated-drained tests on clay soils take considerable time. For this reason, consolidated-
undrained tests can be conducted on such soils with pore pressure measurements to obtain the
drained shear strength parameters. Because drainage is not allowed in these tests during the
185
Lecture Note on GTE-I

application of deviator stress, they can be performed quickly. Skempton’s pore water pressure
∆𝑢defined
̅
𝐴̅ = 𝐴̅̅ =
𝑑𝑓
parameter was as follows. At failure, the parameter can be written as
𝑓
----------------- Eq.11.50
∆𝜎
�𝑓

11.5.3 Unconsolidated-Undrained Triaxial Test

application of chamber pressure 𝜎3. The test specimen is sheared to failure by the application of
In unconsolidated-undrained tests, drainage from the soil specimen is not permitted during the

deviator stress, ∆𝜎𝑑, and drainage is prevented. Because drainage is not allowed at any stage,
the test can be performed quickly. Because of the application of chamber confining pressure, 𝜎3,
the pore water pressure in the soil specimen will increase by uc. A further increase in the pore
water pressure (∆𝑢𝑑) will occur because of the deviator stress application. Hence, the total pore

𝑢 = 𝑢𝑐 + ∆𝑢𝑑--------------------------
water pressure u in the specimen at any stage of deviator stress application can be given as

But we know that 𝑢𝑐 = 𝐵𝜎3 and ∆𝑢𝑐 = 𝐴̅∆𝜎𝑑


Eq.11.51

𝑢 = 𝐵𝜎3 +𝐴̅∆𝜎𝑑----------------------
Now substituting above relation, we obtain
Eq.11.52

concept for cohesive soils if the soil is fully saturated. The added axial stress at failure ∆𝜎𝑑𝑓 is
This test usually is conducted on clay specimens and depends on a very important strength

practically the same regardless of the chamber confining pressure. This property is shown in

hence is called a ∅ = 0 condition. So we get


Figure 7. The failure envelope for the total stress Mohr’s circles becomes a horizontal line and

𝜏𝑓 = 𝑐 = 𝑐𝑢---------------------------- Eq.11.53

Note that the ∅ = 0 concept is applicable to only saturated clays and silts.
Where, cu is the undrained shear strength and is equal to the radius of the Mohr’s circles.

Fig. 11.19: Mohr’c circle for total and effective stress and failure envelope ∅ = 0
The reason for obtaining the same added axial stress ∆𝜎𝑑𝑓 regardless of the confining
pressure can be explained as follows. If clay specimen (No. I) is consolidated at a chamber

186
Lecture Note on GTE-I

pressure 𝜎3and

187
Lecture Note on GTE-I

then sheared to failure without drainage, the total stress conditions at failure can be represented

equal to ∆𝑢𝑑𝑓. Thus, the major and minor principal effective stresses at failure are, respectively,
by the Mohr’s circle P in Fig.11.19. The pore pressure developed in the specimen at failure is

𝜎1′ = [𝜎3 + ∆𝜎𝑑𝑓] − ∆𝑢𝑑𝑓=𝜎1 − ∆𝑢𝑑𝑓--------------------


𝜎3 = 𝜎3 − ∆𝑢𝑑𝑓--------------------------
Eq.11.54

Eq.11.55
Q is the effective stress Mohr’s circle drawn with the preceding principal stresses.

Any value of 𝜎3 could have been chosen for testing the specimen. In any case, the deviator stress
Note that the diameters of circles P and Q are the same.

∆𝜎𝑑𝑓 to cause failure would have been the same as long as the soil was fully saturated and fully
undrained during both stages of the test.

11.6 UNCONFINED COMPRESSION TEST


This is a special case of a triaxial compression test; the confining pressure being zero. A
cylindrical soil specimen, usually of the same standard size as that for the triaxial compression, is
loaded axially by a compressive force until failure takes place. Since the specimen is laterally
unconfined, the test is known as ‘unconfined compression test’. No rubber membrane is
necessary to encase the specimen. The axial or vertical compressive stress is the major principal
stress and the other two principal stresses are zero. This test may be conducted on undisturbed or
remoulded cohesive soils. It cannot be conducted on coarse-grained soils such as sands and
gravels as these cannot stand without lateral support. Also the test is essentially a quick or
undrained one because it is assumed that there is no loss of moisture during the test, which is
performed fairly fast. Owing to its simplicity, it is often used as a field test, besides being used in
the laboratory. The failure plane is not predetermined and failure takes place along the weakest
plane. The test specimen is loaded through a calibrated spring by a simple manually operated
screw jack at the top of the machine. Different springs with stiffness values ranging from 2 to 20
N/mm may be used to test soils of varying strengths.
The graph of load versus deformation is traced directly on a sheet of paper by means of an
autographic recording arm. For any vertical or axial strain, the corrected area can be computed,
assuming no change in volume. The axial stress is got by dividing the load by the corrected area.
The apparatus is shown in Fig. 11.20.
The specimen is placed between two metal cones attached to two horizontal plates, the upper
plate being fixed and the lower one sliding on vertical rods. The spring is supported by a plate
and a screw on either side. The plate is capable of being raised by turning a handle so as to apply
a compressive load on the soil specimen.
The stress-strain diagram is plotted autographically. The vertical movement of the pen relative to
the chart is equal to the extension of the spring, and hence, is proportional to the load. As the
lower plate moves upwards, the upper one swings sideways, the weighted arms bearing on a
stop. The lateral movement of the pen is thus proportional to the axial strain of the soil specimen.
The area of cross-section increases as the specimen gets compressed. A transparent calibrated
mask is used to read the stress direct from the chart.

188
Lecture Note on GTE-I

Fig. 11.20: Unconfined Compression Apparatus


Alternatively, a loading frame with proving ring and a dial gauge for measuring the axial
compression of the specimen may also be used. The maximum compressive stress is that at the
peak of the stress-strain curve. If the peak is not well-defined, an arbitrary strain value such as
20% is taken to represent failure.

11.7.1 Mohr’s circle for Unconfined Compression Test

Fig. 11.21: Mohr’s circle for Unconfined Compression Test


The Mohr’s circles for the unconfined compression test are shown in Fig. 11.21.
𝜎 = 2𝑐𝑡𝑎𝑛(45 + )--------------

2
Eq.11.56
1
The two unknowns–c and ϕ–cannot be solved since any number of unconfined compression tests
would give only one value for σ 1. Therefore, the unconfined compression test is mostly found
useful in the determination of the shearing strength of saturated clays for which ϕ is negligible or

𝜎1 = 𝑄𝑢 = 2𝑐-----------------------
zero, under undrained conditions. In such a case, the above equation reduces to
Eq.11.57
189
Lecture Note on GTE-I

Where, Qu is the unconfined compression strength.


Thus, the shearing strength or cohesion value for saturated clay from unconfined compression
test is taken to be half the unconfined compression strength.

11.8 VANE SHEAR TEST


If suitable undisturbed for remoulded samples cannot be got for conducting triaxial or
unconfined compression tests, the shear strength is determined by a device called the Shear
Vane.
The vane shear test may also conduct in the laboratory. The laboratory shear vane will be usually
smaller in size as compared to the field vane. The shear vane usually consists of four steel plates
welded orthogonally to a steel rod, as shown in Fig. 11.22.

Fig. 11.22: Laboratory shear vane


The applied torque is measured by a calibrated torsion spring, the angle of twist being read on a
special gauge. A uniform rotation of about 1° per minute is used. The vane is forced into the soil
specimen or into the undisturbed soil at the bottom of a bore-hole in a gentle manner and torque
is applied. The torque is computed by multiplying the angle of twist by the spring constant.
The shear strength s of the clay is given by
𝑇
𝑠 2 𝐻 𝐷 ---------------------
=
Eq.11.58
𝜋𝐷 ( + )
2 6
If both the top and bottom of the vane are used in shearing the soil.
Here, T = torque
D = diameter of the vane
H = height of the vane
If only one end of the vane partakes in shearing the soil, then
𝑠
𝑇
2 𝐻 𝐷
=
---------------------------- Eq.11.59
𝜋𝐷 ( + )
2 12

190
Lecture Note on GTE-I

The above Equation may be derived as follows:


The shearing resistance is mobilised at failure along a cylindrical surface of diameter D, the
diameter of the vane, as also at the two circular faces at top and bottom.
The shearing force at the cylindrical surface = πD.H.s-------------------------- Eq.11.60
Where s is the shearing strength of the soil. The moment of this force about the axis of the vane
contributes to the torque and is given by
π D2/2.s.h--------------------------- Eq.11.61
For the circular faces at top or bottom, considering the shearing strength of a ring of thickness dr
at a radius r, the elementary torque is (2π r dr). s. r and the total for one face is
𝐷
2∫0 2𝜋𝑠𝑟2 𝑑𝑟--------------------
2
Eq.11.60
The total shear strength developed will be equal to the sum of Eq.11.60 and 61. The maximum
moment of the total shear resistance about the axis of torque equals the torque produced T at
failure. Hence
𝐷 𝐻𝐷2 𝐷 𝐻 𝐷
𝐷2
𝑇 = π . 𝑠. ℎ + 2 ∫ 2𝜋𝑠𝑟2 𝑑𝑟 = + ] = 𝜋𝐷 𝑠 [ ]-----------
3
2 2
2 𝜋𝑠 [ + 6
Eq.11.61
0 2 6 2

𝑇 = 𝜋𝐷2𝑠 [ + ]----------------
𝐻 𝐷
If only the bottom end partake in the shearing the above equation takes as

2 12
Eq.11.62
Regarding the shearing stress distribution on the soil cylinder, it is assumed uniform on the
cylindrical surface but it is triangular over the shear end faces, varying from zero at the axis of
the vane device, to maximum at the edge, as shown in Fig. 11.23.

Fig. 11.23: Shearing Distribution on the Sides and Faces of Soil Cylinder in the Vane Shear Test

The vane shear test is particularly suited for soft clays and sensitive clays for which suitable
cylindrical specimens cannot be easily prepared.

11.9 SHEARING CHARACTERISTICS OF SANDS


The shearing strength in sand may be said to consist of two parts, the internal frictional
resistance between grains, which is a combination of rolling and sliding friction and another part
known as ‘interlocking’. Interlocking, which means locking of one particle by the adjacent ones,
resisting movements, contributes a large portion of the shearing strength in dense sands, while it
does not occur in loose sands. The Mohr strength theory is not invalidated by the occurrence of

191
Lecture Note on GTE-I

interlocking. The Mohr envelopes merely show large ordinates and steeper slopes for dense soils
than for loose ones.
The angle of internal friction is a measure of the resistance of the soil to sliding along a plane.
This varies with the density of packing, characterised by density index, particle shape and
roughness and particle size distribution. Its value increases with density index, with the
angularity and roughness of particles and also with better gradation. This is influenced to some
extent by the normal pressure on the plane of shear and also the rate of application of shear.
The ‘angle of repose’ is the angle to the horizontal at which a heap of dry sand, poured freely
from a small height, will stand without support. It is approximately the same as the angle of
friction in the loose state.
Some clean sands exhibit slight cohesion under certain conditions of moisture content, owing to
capillary tension in the water contained in the voids. Since this is small and may disappear with
change in water content, it should not be relied upon for shear strength. On the other hand, even
small percentages of silt and clay in sand give it cohesive properties which may be sufficiently
large so as to merit consideration.
Unless drainage is deliberately prevented, a shear test on a sand will be a drained one as the high
value of permeability makes consolidation and drainage virtually instantaneous. A sand can be
tested either in the dry or in the saturated condition. If it is dry, there will be no pore water
pressures and if it is saturated, the pore water pressure will be zero due to quick drainage. In
either case, the inter granular pressure will be equal to the applied stress. However, there may be
certain situations in which significant pore pressures are developed, at least temporarily, in
sands. For example, during earth-quakes, heavy blasting and operation of vibratory equipment
instantaneous pore pressures are likely to develop due to large shocks or dynamic loads. These
may lead to the phenomenon of ‘liquefaction’ or sudden and total loss of shearing strength,
which is a grave situation of lack of stability.

11.9.1 Stress-strain Behaviour of Sands


The stress-strain behaviour of sands is dependent to a large extent on the initial density of
packing, as characterised by the density index. This is represented in Fig. 11.24.
It can be observed from Fig. 11.24 (a), the shear stress (in the case of direct shear tests) or
deviator stress (in the case of triaxial compression tests) builds up gradually for an initially loose
sand, while for an initially dense sand, it reaches a peak value and decreases at greater values of
shear/axial strain to an ultimate value comparable to that for an initially loose specimen. The
behaviour of a medium-dense sand is intermediate to that of a loose sand and a dense sand.
Intuitively, it should be expected that denser the sand is, the stronger it is. The hatched portion
represents the additional strength due to the phenomenon of interlocking in the case of dense
sands.

192
Lecture Note on GTE-I

Fig.11.24: Stress-strain Characteristics of Sand


The volume change characteristics of sands are another interesting feature, as depicted in Fig.
11.24(b). An initially dense specimen tends to increase in volume and become loose with
increasing values of strain, while an initially loose specimen tends to decrease in volume and
become dense. This is explained in terms of the rearrangement of particles during shear.
The changes in pore water pressure during undrained shear, which is rather not very common
owing to high permeability of sands, are depicted in Fig. 11.25 (c). Positive pore pressures
develop in the case of an initially loose specimen and negative pore pressures develop in the case
of an initially dense specimen.

11.9.2 Shearing Strength of Sands


The shearing strength of cohesionless soils has been established to depend primarily upon the
angle of internal friction which itself is dependent upon a number of factors including the normal
pressure on the failure plane. The nature of the results of the shear tests will be influenced by the
type of test—direct shear or triaxial compression, by the fact whether the sand is saturated or dry
and also by the nature of stresses considered—total or effective.
Each direct shear test is usually conducted under a certain normal stress. Each stress strain
diagram therefore reflects the beahaviour of a specimen under a particular normal stress. A
number of specimens are tested under different normal stresses. It is to be noted that only the
effective normal stress is capable of mobilising shear strength. The results when plotted appear
as shown in Fig. 11.25.
It may be observed from Fig. 11.25 (a) that the greater the effective normal pressure during
shear, the greater is the shearing stress at failure or shearing strength. The shear strength plotted
against effective normal pressure gives the Coulomb strength envelope as a straight line, passing

193
Lecture Note on GTE-I

through the origin and inclined at the angle of internal friction to the normal stress axis. It is
shown in Fig. 6 (b).

Fig. 11.25: Shear Characteristics of Sand in Direct Test


It may be observed from Fig. 11.25 (a) that the greater the effective normal pressure during
shear, the greater is the shearing stress at failure or shearing strength. The shear strength plotted
against effective normal pressure gives the Coulomb strength envelope as a straight line, passing
through the origin and inclined at the angle of internal friction to the normal stress axis. It is
shown in Fig. 11.25 (b). The failure envelope obtained from ultimate shear strength values is
assumed to pass through the origin for dry cohesionless soils. The same is true even for saturated
sands if the plot is made in terms of effective stresses. In the case of dense sands, the values of φ
obtained by plotting peak strength values will be somewhat greater than those from ultimate
strength values.
Ultimate values of φ may range from 29 to 35° and peak values from 32 to 45° 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.

Fig. 11.26: Failure Envelope of sand indicating Apparent Cohesion


If the sand is moist, the failure envelope does not pass through the origin as shown in Fig. 11.26.
The intercept on the shear stress axis is referred to as the ‘apparent cohesion’, attributed to
factors such as surface tension of the moisture films on the grains. The extra strength would be
lost if the soil were to dry out or to become saturated or submerged. For this reason the extra
shear strength attributed to apparent cohesion is neglected in practice.
In the case of triaxial compression tests, different tests with different cell pressure are to be
conducted to evaluate the shearing strength and the angle of internal friction. In each test, the

194
Lecture Note on GTE-I

axial normal stress is gradually increased keeping the cell pressure constant, until failure occurs.
The value of φ is obtained by plotting the Mohr Circles and the corresponding Mohr’s envelope.
The failure envelope obtained from a series of drained triaxial compression tests on saturated
sand specimens initially at the same density index is approximately a straight line passing
through the origin, as shown in Fig. 11.27.

Fig.11.27: Drained Triaxial Test on Saturated Sand


Similar results are obtained when undrained triaxial compression tests are conducted with pore
pressure measurements on saturated sand samples and Mohr’s circles are plotted in terms of
effective stresses. However, if Mohr’s circles are plotted in terms of total stresses, the shape of
envelopes will be similar to those for a purely cohesive soil. The failure envelope will be
approximately horizontal with an intercept on the shearing stress axis, indicating the so called
‘apparent cohesion’, as shown in Fig. 11.28.

Fig. 11.28: Un-drained Triaxial Test on Saturated Sand (Total Stress)


Example 1: The stresses at failure on the failure plane in a cohesionless soil mass were:
Shear stress = 4 kN/m2; normal stress = 10 kN/m2. Determine the resultant stress on the failure
plane, the angle of internal friction of the soil and the angle of inclination of the failure plane to
the major principal plane.
Sol: Resultant stress = √𝜎2 + 𝜏2
= √100 + 16= 10.77 kN/m2

∅= 21° 48’
tan∅= 4/10 = 0.4

𝜃 = 45° +21048’/2
= 55°54’

195
Lecture Note on GTE-I

Graphical solution (Fig. 1):


The procedure is first to draw the σ-and τ-axes from an origin O and then, to a suitable scale,set-
off point D with coordinates (10,4), Joining O to D, the strength envelope is got. The MohrCircle
should be tangential to OD to D. DC is drawn perpendicular to OD to cut OX in C, which is the
centre of the circle. With C as the centre and CD as radius, the circle is completed to cut OX in A
and B.

Fig.1

196
Lecture Note on GTE-I

Sol:

197
Lecture Note on GTE-I

Sol:

198
Lecture Note on GTE-I

Graphical Sol:

Sol:

199

You might also like